Vous êtes sur la page 1sur 1019

FRONTIERS IN OFFSHORE GEOTECHNICS

ISFOG 2005

Copyright 2005 Taylor & Francis Group plc, London, UK


BALKEMA Proceedings and Monographs in
Engineering, Water and Earth Sciences

Copyright 2005 Taylor & Francis Group plc, London, UK


PROCEEDINGS OF THE FIRST INTERNATIONAL SYMPOSIUM ON FRONTIERS IN
OFFSHORE GEOTECHNICS, UNIVERSITY OF WESTERN AUSTRALIA, PERTH,
1921 SEPTEMBER 2005

Frontiers in Offshore
Geotechnics
ISFOG 2005

Susan Gourvenec & Mark Cassidy


Centre for Offshore Foundation Systems, University of Western Australia

LONDON/LEIDEN/NEW YORK/PHILADELPHIA/SINGAPORE

Copyright 2005 Taylor & Francis Group plc, London, UK


Front Cover: ENSCO 104, a KFELS B Class jackup constructed by Keppel FELS, Singapore (Source: Keppel
Offshore & Marine).

Back Cover: Arial shot of coastline of Australias North West Shelf (top right) and calcareous sand (middle).

ISFOG logo designed by Hilit Einav.

Copyright 2005 Taylor & Francis Group plc, London, UK

All rights reserved. No part of this publication or the information contained herein may be reproduced, stored
in a retrieval system, or transmitted in any form or by any means, electronic, mechanical, by photocopying,
recording or otherwise, without written prior permission from the publisher.

Although all care is taken to ensure the integrity and quality of this publication and the information herein, no
responsibility is assumed by the publishers nor the author for any damage to property or persons as a result of
operation or use of this publication and/or the information contained herein.

Published by: Taylor & Francis/Balkema


P.O. Box 447, 2300 AK Leiden, The Netherlands
e-mail: Pub.NL@tandf.co.uk
www.balkema.nl, www.tandf.co.uk, www.crcpress.com

ISBN Set (book  cd-rom): 0 415 39063 X


ISBN Book: 0 415 39064 8
ISBN CD-ROM: 0 415 39065 6

Printed in Great-Britain

Copyright 2005 Taylor & Francis Group plc, London, UK


Table of Contents

Preface XV

Committees XVII

Reviewers XIX

Keynote papers
Suction anchors for deepwater applications 3
K.H. Andersen, J.D. Murff, M.F. Randolph, E.C. Clukey, C.T. Erbrich, H.P. Jostad,
B. Hansen, C. Aubeny, P. Sharma & C. Supachawarote

Vertically loaded plate anchors for deepwater applications 31


J.D. Murff, M.F. Randolph, S. Elkhatib, H.J. Kolk, R.M. Ruinen, P.J. Strom & C.P. Thorne

Australian frontiers spudcans on the edge 49


C.T. Erbrich

Suction caissons for wind turbines 75


G.T. Houlsby, L. Bo Ibsen & B.W. Byrne

Pipeline geotechnics state-of-the-art 95


D.N. Cathie, C. Jaeck, J.-C. Ballard & J.-F. Wintgens

An operators perspective on offshore risk assessment and geotechnical design in


geohazard-prone areas 115
P. Jeanjean, K. Hampson, T. Evans, E. Liedtke & E.C. Clukey

Offshore site investigations: new frontiers 145


H.J. Kolk & J. Wegerif

Deepwater developments: drag and plate anchors


Influence of anchor geometry and soil properties on numerical modeling of
drag anchor behavior in soft clay 165
R.M. Ruinen

The effect of interface friction on the performance of drag-in plate anchors 171
S. Elkhatib & M.F. Randolph

Proposed upper bound analysis for drag embedment anchors in soft clay 179
C.P. Aubeny, B.M. Kim & J.D. Murff

Ultimate pullout capacity of SBMs VErtically Loaded Plate Anchor (VELPA) in


deep sea sediments 185
P.Y. Foray, S. Alhayari, E. Pons, L. Thorel, N. Thetiot, S. Bale & E. Flavigny

Copyright 2005 Taylor & Francis Group plc, London, UK


Effects of long term loading on storm capacity of vertically loaded anchors 191
H.A. Taiebat, C.P. Thorne & J.P. Carter

An experimental and numerical study of rate effects for plate anchors in clay 197
M.J. Rattley, B.M. Lehane, D.J. Richards & C. Gaudin

Vertical pullout behavior of plate anchors in uniform clay 205


Z. Song & Y. Hu

Pullout capacity of circular plate anchors in double-layered clays 213


J. Liu, Y. Hu & L. Wu

The use of inflatable anchors in offshore sandy soils 221


T.A. Newson, F.W. Smith & P. Brunning

Deepwater developments: suction caissons


Suction caisson soil displacement during installation 229
E.C. Clukey

Lessons learned from several suction caisson installation projects in clay 235
Y.C. Lee, J.M.E. Audibert & K.-M. Tjok

Centrifuge tests on axial capacity of suction caissons in clay 243


W. Chen & M.F. Randolph

Evaluation of recovery of wall friction after penetration of skirts with laboratory and field tests 251
Y. Yoshida, N. Masui & M. Ito

Study of sand heave formation in suction caissons using Particle Image Velocimetry (PIV) 259
M.N. Tran, M.F. Randolph & D.W. Airey

Electrokinetic and electrochemical stabilization of caissons in calcareous sand 267


A. Rittirong, J.Q. Shang, M.A. Ismail & M.F. Randolph

Vertical uplift capacity of suction caisson in clay 273


L. Thorel, J. Garnier, G. Rault & A. Bisson

Capacity of suction caissons under inclined loading in normally consolidated clay 281
R.M. El-Sherbiny, R.E. Olson, R.B. Gilbert & S.K. Vanka

A failure surface for caisson foundations in undrained soils 289


H.A. Taiebat & J.P. Carter

Parametric finite element analyses of suction anchors 297


L. Zdravkovic & D.M. Potts

Centrifuge modelling of suction piles in clay 303


R.D. Raines, O. Ugaz & J. Garnier

Establishing a model testing capability for deep water foundation systems 309
K.S. Prakasha, H.A. Joer & M.F. Randolph

Reliability-based design considerations for deepwater mooring system foundations 317


R.B. Gilbert, Y.J. Choi, S. Dangayach & S.S. Najjar

VI

Copyright 2005 Taylor & Francis Group plc, London, UK


Validation of the use of beam-column method for suction caisson design 325
J. Cao, Y. Li, K.-M. Tjok & J.M.E. Audibert

Validation of the use of finite element method for suction caisson design 333
J. Cao, J.M.E. Audibert, K.-M. Tjok & M.K. Hossain

Developments in the Australian frontiers


Geotechnical interpretation for the Yolla A Platform 343
P.G. Watson & C. Humpheson

Preloading of drag anchors in carbonate sediments 351


S.R. Neubecker, M.P. ONeill & C.T. Erbrich

The geotechnical performance of Deep Penetrating Anchors in calcareous sand 357


M.D. Richardson, C.D. OLoughlin & M.F. Randolph

Seabed geotechnical characterisation with a ball penetrometer deployed from the


Portable Remotely Operated Drill 365
P.J. Kelleher & M.F. Randolph

Static and cyclic behavior of laterally loaded piles in calcareous sand 373
W.D. Guo & B.T. Zhu

Foundation solutions for offshore wind turbines


Design aspects of monopiles in German offshore wind farms 383
K. Lesny & J. Wiemann

Finite element modelling of horizontally loaded monopile foundations for


offshore wind energy converters in Germany 391
K. Abdel-Rahman & M. Achmus

Tripods with suction caissons as foundations for offshore wind turbines on sand 397
M. Senders

The tensile capacity of suction caissons in sand under rapid loading 405
G.T. Houlsby, R.B. Kelly & B.W. Byrne

Moment loading of caissons installed in saturated sand 411


F.A. Villalobos, B.W. Byrne & G.T. Houlsby

The theoretical modelling of a suction caisson foundation using hyperplasticity theory 417
L. Nguyen-Sy & G.T. Houlsby

Shallow foundations: vertical bearing capacity


Bearing capacity of parallel strip footings on non-homogeneous clay 427
C.M. Martin & E.C.J. Hazell

Study on bearing behaviors of foundations on multi-layer subsoil 435


F.F. Yuan, M.T. Luan & S.W. Yan

Numerical study of shallow foundations on calcareous sand 443


N. Yamamoto, M.F. Randolph & I. Einav

VII

Copyright 2005 Taylor & Francis Group plc, London, UK


Undrained bearing capacity of shallow foundations on structured soils 451
D.S. Liyanapathirana & J.P. Carter

An investigation into the vertical bearing capacity of perforated mudmats 459


D.J. White, A.J. Maconochie, C.Y. Cheuk, M.D. Bolton, D. Joray & S.M. Springman

Shallow foundations: combined loading


Yield loci for shallow foundations by swipe testing 469
G. Gottardi, L. Govoni & R. Butterfield

Investigating 6 degree-of-freedom loading on shallow foundations 477


B.W. Byrne & G.T. Houlsby

Single surface hardening model a system law to describe the foundation-soil interaction 483
A. Kisse & K. Lesny

Bearing capacity of strip footings subjected to complex loading 491


L. Thorel, A.-H. Soubra, J. Garnier & R. Assaf

Numerical analysis of bearing capacity of foundation under combined loading 499


S.F. Zhao, M. Luan & A-Z. Lu

Centrifuge tests on improving offshore foundation systems 507


H.G.B. Allersma

Shallow foundations: mobile jack-up units


Extraction of jackup spudcan foundations 517
O.A. Purwana, C.F. Leung, Y.K. Chow & K.S. Foo

Numerical simulation of the breakout process of an object at the ocean bottom 523
X.X. Zhou, Y.K. Chow & C.F. Leung

Spudcan penetration in sand overlying clay 529


K.L. Teh, C.F. Leung & Y.K. Chow

Punch-through of spudcan foundations in two-layer clay 535


M.S. Hossain, Y. Hu, M.F. Randolph & D.J. White

Influence of jack-up operation adjacent to a piled structure 543


D.P. Stewart

FE modelling of spudcan pipeline interaction 551


L. Kellezi, G. Kudsk & P.B. Hansen

Jack-up footing penetration and fixity analyses 559


L. Kellezi, H.W.L. Hofstede & P.B. Hansen

Breakwater caissons and liquifaction


Performance of caisson breakwater subjected to breaking wave loads 569
X.Y. Zhang, C.F. Leung & F.H. Lee

An interaction model for seismic stability analysis of caisson type structure 577
H. Hazarika

VIII

Copyright 2005 Taylor & Francis Group plc, London, UK


Progressive ocean wave modelling in drum centrifuge 583
F.P. Gao & M.F. Randolph

Numerical analysis of dynamic response of seabed under random wave loading 589
Z. Wang, M. Luan, Z. Liu & D. Wang

Pipelines
The performance of pipeline ploughs in layered soils 597
M.F. Bransby, G.J. Yun, D.R. Morrow & P. Brunning

Physical and numerical modelling of lateral buckling of a pipeline in very soft clay 607
J.R.M.S. Oliveira, M.S.S. Almeida, M.C.F. Almeida, R.G. Borges, C.S. Amaral & A.M. Costa

Bearing capacity and large penetration of a cylindrical object at shallow embedment 615
E.R. Barbosa-Cruz & M.F. Randolph

Stability design of untrenched pipelines geotechnical aspects 623


J. Zhang & C.T. Erbrich

Pipeline-seabed interaction analysis subjected to horizontal cyclic loading 629


T. Takatani

A numerical model of onset of scour below offshore pipelines subject to steady currents 637
D. Liang & L. Cheng

Arctic seabed ice gouging and large sub-gouge deformations 645


A.C. Palmer, I. Konuk, A.W. Niedoroda, K. Been & K.R. Croasdale

Model tests to simulate riser-soil interaction in touchdown point region 651


E.C. Clukey, L. Haustermans & R. Dyvik

Capacity of piles in sand


Results from axial load tests on pipe piles in very dense sands: the EURIPIDES JIP 661
H.J. Kolk, A.E. Baaijens & P. Vergobbi

CPT-based design method for steel pipe piles driven in very dense silica
sands compared to the Euripides pile load test results 669
P.Y. Foray & J.-L. Colliat

Bearing capacity of driven piles in sand, the NGI approach 677


C.J.F. Clausen, P.M. Aas & K. Karlsrud

The UWA-05 method for prediction of axial capacity of driven piles in sand 683
B.M. Lehane, J.A. Schneider & X. Xu

An updated assessment of the ICP pile capacity procedures 691


R.J. Jardine, F.C. Chow, J.R. Standing, R.F. Overy, E. Saldivar-Moguel,
C. Strick van Linschoten & A. Ridgway

A general framework for shaft resistance on displacement piles in sand 697


D.J. White

Field research into the effects of time on the shaft capacity of piles driven in sand 705
R.J. Jardine, J.R. Standing & F.C. Chow

IX

Copyright 2005 Taylor & Francis Group plc, London, UK


Design criteria for pipe piles in silica sands 711
H.J. Kolk, A.E. Baaijens & M. Senders

Estimating the end bearing resistance of pipe piles in sand using the final filling ratio 717
K. Gavin & B.M. Lehane

Evaluation of end-bearing capacity of open-ended piles driven in sand from CPT data 725
X. Xu, B.M. Lehane & J.A. Schneider

Evaluation of end-bearing capacity of closed-ended pile in sand from cone penetration data 733
X. Xu & B.M. Lehane

The influence of effective area ratio on shaft friction of displacement piles in sand 741
D.J. White, J.A. Schneider & B.M. Lehane

A centrifuge study of the monotonic and cyclic resistance of piles and pile groups in sand 749
C. Gaudin, B.M. Lehane & P.F. Wallis

Correlations for shaft capacity of offshore piles in sand 757


J.A. Schneider & B.M. Lehane

Axial load tests on pipe piles in very dense sands at Ras Tanajib 765
H.J. Kolk, A.E. Baaijens, K.A. Shafei & O.A. Dakhil

Piles
Bearing capacity of driven piles in clay, the NGI approach 775
K. Karlsrud, C.J.F. Clausen & P.M. Aas

Lateral pile design of the Ursa tension leg platform 783


E.H. Doyle, E.T. Richard Dean & J.A. Newlin

Case study on soil plugging of open-ended steel pipe piles in Tokyo Bay 791
T. Matsumoto & P. Kitiyodom

Foundation capacity of piled offshore platforms 799


N. Morgan, I.M.S. Finnie, G. Stewart & J.L. Price

Soil characterization for consistent reliability in the Load and Resistance


Factor Design of pile foundations 807
K.C. Foye & R. Salgado

Buckling considerations in pile design 815


S. Bhattacharya, T.M. Carrington & T.R. Aldridge

Propagation of pile tip damage during installation 823


T.R. Aldridge, T.M. Carrington & N.R. Kee

One dimensional wave propagation analysis of an open-ended pipe pile with


consideration of the excess pore pressure in soil plug 829
T. Matsumoto, T. Wakisaka & A. Numata

Simplified analysis of single pile subjected to dynamic active and passive loadings 837
P. Kitiyodom, R. Sonoda & T. Matsumoto

Simplified dynamic analysis of pile group subjected to horizontal loading 845


R. Sonoda, P. Kitiyodom & T. Matsumoto

Copyright 2005 Taylor & Francis Group plc, London, UK


Reliability analysis on axially loaded pile foundation of offshore platforms 853
Y. Shuwang, Z. Hongjie & L. Run

Numerical analysis of pile axial loading test on Ryukyu calcareous sediments 859
M. Ohuchi, M. Kiyosumi, N. Umeda, F.L. Peng & O. Kusakabe

A preliminary investigation into the effect of axial load on piles subjected to


lateral soil movement 865
W.D. Guo & E.H. Ghee

Modelling combined loading of piles with local interacting yield surfaces 873
N.H. Levy, I. Einav & M.F. Randolph

Assessing geohazards
Assessment of the hydrate geohazard 883
A.J. Digby

Gas hydrates and their potential effects on deep water exploration activities 889
J.A. Priest, C.R.I. Clayton & A.I. Best

OpenSees modeling of the 3D plastic behavior of underwater slopes:


achievements and limitations 897
H.G. Brandes & S. Wang

3DSTAB: a history of 3D stability analysis applied in offshore geotechnics 903


H.J. Luger, J.L. Bijnagte & J.A.M. Teunissen

A study of ice as an analog of methane hydrate on the basis of static shear strength 909
Y. Nabeshima & Y. Takai

Tackling geohazards a case study from the Turkmenistan shelf, Caspian Sea 913
J. Wegerif, M. Galavazi, I. Hamilton & Z.B.A. Razak

Site investigation techniques


Deepwater geotechnical site investigation practice in the Gulf of Guinea 921
D. Borel, A. Puech, H. Dendani & J.L. Colliat

The origin of near-seafloor crust zones in deepwater 927


C.J. Ehlers, J. Chen, H.H. Roberts & Y.C. Lee

Casing mounted method used for geotechnical drilling in Lavan 935


A. Fakher & A. Cheshomi

Self/barge Installing Platform, SIP II: the Calder experience 939


R.J. van den Heuvel & M.E. Riemers

Assessment of sand quality using seismic techniques at Fisherman Islands, Brisbane 945
R.J. Whiteley, J. Ameratunga & P.J. Boyle

Site characterization of Bootlegger Cove clay for Port of Anchorage 951


P.W. Mayne & R.A. Pearce

The geotechnical diving bell equipment used in Brazil to perform nearshore and
offshore geotechnical investigations 957
F. Bogossian, A. Muxfeldt & A.B. Dutra

XI

Copyright 2005 Taylor & Francis Group plc, London, UK


Conquering new frontiers in underwater cone penetration testing 961
K. van den Berg, A. Walta & T. de Wolff

Cyclic friction piezocone tests for offshore applications 967


G.L. Hebeler, J.D. Frost, J.A. Schneider & B.M. Lehane

Practice notes on push-in penetrometers for offshore geotechnical investigation 973


J. Peuchen, J. Adrichem & P.A. Hefer

Comparison of cone and T-bar factors in two onshore and


one offshore clay sediments 981
T. Lunne, M.F. Randolph, S.F. Chung, K.H. Andersen & M. Sjursen

Considerations in evaluating the remoulded undrained shear strength from


full flow penetrometer cycling 991
N.J. Yafrate & J.T. DeJong

Well deformations at West Azeri, Caspian Sea 999


J.D. Allen, K. Hampson, C.J.F. Clausen & C. Vermeijden

Borehole squeezing in soft clays 1005


S. Kay

Soil characterization
Modelling the effects of structure in deep-ocean sediments 1013
B.A. Baudet & E.W.L. Ho

Simulating the mechanical behaviour of some calcareous soils using


the Structured Cam Clay model 1019
M.D. Liu & J.P. Carter

Degradation of cementation for artificially cemented carbonate sands 1027


A.K.M. Mohsin & D.W. Airey

Detection of slight cementation in offshore carbonate deposits from laboratory testing 1033
M.A. Ismail, S.S. Sharma & M. Fahey

Experimental study on cyclic pore water pressure and volumetric changes of


saturated loose sands under complex stress condition 1039
M. Luan, Y. He, C. Xu, Y. Guo & M. Li

Some geotechnical specificities of Gulf of Guinea deepwater sediments 1047


A. Puech, J.L. Colliat, J.-F. Nauroy & J. Meunier

Mineralogical characteristics of the Gulf of Guinea deep water sediments 1055


F. Thomas, B. Rebours, J.-F. Nauroy & J. Meunier

On the compressibility of deepwater sediments of the Gulf of Guinea 1063


V. De Gennaro, P. Delage & A. Puech

On the classification of Bangkok clay deposits and their compressibility 1071


S. Horpibulsuk & R. Rachan

Long-term consolidation behavior of Pleistocene clays in Osaka Bay and applicability


of elasto-viscoplastic one-dimensional consolidation model 1079
K. Oda, K. Tokida & T. Matsui

XII

Copyright 2005 Taylor & Francis Group plc, London, UK


Influence of salinity on soil properties 1087
S. Kay, S.S. Goedemoed & C.A. Vermeijden

The determination of undrained shear strength and plasticity properties from


quasi-static penetration tests 1095
K.J.L. Stone & B.S. Kyambadde

Influence of height and boundary conditions in simple shear tests 1101


A.J. Reyno, D.W. Airey & H.A. Taiebat

XIII

Copyright 2005 Taylor & Francis Group plc, London, UK


Preface

Professor Mark Randolph


Chair of Scientific Committee, International Symposium on Frontiers in Offshore Geotechnics (ISFOG)

The idea for this symposium was first mooted in April 2003, as one of the possible activities of TC1,
the Technical Committee of the International Society of Soil Mechanics and Geotechnical Engineering,
responsible for Offshore and Nearshore Geotechnical Engineering. Harry Kolk, the Chair of TC1, was in
agreement that the number of conferences with a high offshore geotechnical content had diminished signifi-
cantly in recent years, and there had been sufficient developments in offshore geotechnics to justify a specialist
conference.
With some diffidence, I raised the idea of hosting an international conference with colleagues in the Centre
for Offshore Foundation Systems (COFS) at the University of Western Australia, aware of the inordinate
amount of work involved in such ventures. However, the challenge was accepted enthusiastically, spearheaded
by Dr Susan Gourvenec who agreed to chair the organising committee. Some fairly fast decisions were made,
fixing the overall framework of the conference for the 3 days immediately following the Osaka International
Conference on Soil Mechanics and Geotechnical Engineering, with a single-session format rather than parallel
sessions, and aiming for a relatively small gathering of international specialists. The viability of the conference
was assured in the early days of planning by major sponsorship support from the international company, Fugro,
and local companies Woodside and Advanced Geomechanics.
An obvious frontier worldwide in recent years has been the trend towards deep water, and the resulting
emphasis on geohazards, anchoring systems, pipelines and risers, together with the need for improved methods
of quantifying the seabed strength in the upper 20 to 30 m, and in particular the upper 0.5 to 1 m that is critical
for pipeline design. The proposed timing of the symposium fitted well with the conclusion of a major project
sponsored by the American Petroleum Institute, joined later by the Deepstar Project, to review analysis and
design of deep-water anchoring systems. The culmination of that project has formed the basis of the first two
keynote papers, led respectively by Don Murff (Offshore Technology Research Center) and Knut Andersen
(Norwegian Geotechnical Institute). Other keynote papers with a strong focus on deep water include geohazard
assessment (led by Philippe Jeanjean of BP), geotechnical and geophysical site investigation (led by Harry Kolk
of Fugro) and geotechnical aspects of pipeline design (led by David Cathie of Cathie Associates).
While geotechnical design in deep water is primarily concerned with soft fine-grained sediments, the car-
bonate silts and sands that comprise seabed sediments around Australia continue to provide challenges, partic-
ularly in respect of their layering and intermediate consolidation characteristics. A keynote paper from
Carl Erbrich (Advanced Geomechanics) describes entertainingly the difficulties in extrapolating penetration
resistance in such sediments from the scale of a T-bar or ball penetrometer to an 18 m diameter spudcan.
Another timely coincidence for ISFOG has been the imminent revision of the API design code for fixed plat-
forms, and the determination to update the guidelines for driven piles in sand. This triggered a major effort led
by Barry Lehane (of UWA) to review a range of design methods based on cone resistance, resulting in a signif-
icant body of papers addressing this topic in the proceedings; these formed the basis of a workshop during the
symposium itself.
A final keynote paper and topic for the symposium focuses on geotechnical issues associated with offshore
renewable sources of energy. Foundations for offshore wind and wave generators are in important frontier and
the keynote paper by Guy Houlsby (Oxford University) and Lars Bo Ibsen (Alborg University) summarises
recent research results on caisson foundations for wind turbines.
There was a gratifyingly warm response to our call for papers, resulting in 134 papers originating from 26
different countries. These papers were carefully reviewed, with at least 2 reviews per paper and involving 122
reviewers in total; these were drawn from academic institutions and industry within Australia, with support
from 78 colleagues around the world. We are indebted to the efforts of all the reviewers, who have undoubtedly
raised the quality of the proceedings, and also to the editorial support provided by Nina Levy.

XV

Copyright 2005 Taylor & Francis Group plc, London, UK


I would like to express my heartfelt thanks to the local organising committee, particularly Susie Gourvenec
as Chair for her indefatigable efforts and to Monica Mackman, secretary of the committee, who took responsi-
bility for so much of the detailed correspondence and organisation. The final proceedings have been a joint
effort between Susie Gourvenec and Mark Cassidy, not forgetting our publishers Taylor & Francis/Balkema. I
am confident that this set of papers reflect well the current state-of-the-art in terms of practice and new research
ideas, and will prove a milestone in the development of offshore geotechnics.

Mark Randolph
June 2005

XVI

Copyright 2005 Taylor & Francis Group plc, London, UK


Committee

Organising Committee
Susan Gourvenec (Chair) Centre for Offshore Foundation Systems (COFS), UWA
Monica Mackman (Secretary) COFS, UWA
Mark Cassidy COFS, UWA
Diane Christensen COFS, UWA
Geoff Cole Woodside Energy Ltd
Itai Einav COFS, UWA
Martin Fahey COFS, UWA
Christophe Gaudin COFS, UWA
Yuxia Hu Curtin University
Mostafa Ismail COFS, UWA
Barry Lehane School of Civil and Resource Engineering, UWA
Rob Male Woodside Energy Ltd
Conleth O`Loughlin COFS, UWA
Mike O`Neill Advanced Geomechanics
Mark Randolph COFS, UWA
Marc Senders COFS, UWA
Phil Watson Arup Energy
International Scientific Committee
Mark Randolph (Chair) Australia
Knut Andersen Norway
Malcolm Bolton United Kingdom
Fraser Bransby United Kingdom
William Bryant United States of America
John Carter Australia
Mark Cassidy Australia
David Cathie Belgium
Ed Clukey United States of America
Jean-Louis Colliat France
Gijs Degenkamp The Netherlands
Jayme Mello Brazil
Earl Doyle United States of America
Carl Erbrich Australia
Martin Fahey Australia
Ian Finnie United Kingdom
Pierre Foray France
Jacques Garnier France
Jim Hooper United States of America
Guy Houlsby United Kingdom
Harry Kolk The Netherlands
Fook Hou Lee Singapore
Maotian Luan China
Don Murff United States of America
Steiner Nordal Norway
Derek Pennington Australia
Kuppalli Prakasha India
Dick Raines United States of America
Juddith Whittick Canada

XVII

Copyright 2005 Taylor & Francis Group plc, London, UK


Reviewers

Tony Abbs Martin Fahey Jayme Mello


Khalid Abdel-Rahman Sue Feller Richard Merifield
David Airey Ian Finnie Eltayeb Mohamed-Elhassan
Tom Aldridge Pierre Foray Don Murff
Junius Allen Sam Frydman Yasuyuki Nabeshima
Henderikus Allersma Jacques Garnier Steve Neubecker
Jay Ameratunga Christophe Gaudin Tim Newson
Knut Andersen Ken Gavin Conleth OLoughlin
Sina Aragh Guido Gottardi Lee ONeill
Charles Aubeny Susan Gourvenec Mike ONeill
Jean Audibert Wei Dong Guo Joek Peuchen
Aukje Baaijens Hans Hanse David Potts
Christophe Ballard Hemanta Hazarika Harry Poulos
Beatrice Baudet Paul Hefer Jeffrey Priest
Britta Bienen Jim Hooper Alain Puech
Lars Bo Ibsen Guy Houlsby Mark Randolph
Malcolm Bolton Yuxia Hu Mike Rattley
Horst Brandes Buddhima Indraratna David Richards
Andrew Brennan Roger Ingersoll Mark Richardson
Roy Butterfield Mohammad Islam Shibuya Saturo
Byron Byrne Mostafa Ismail James Schneider
John Carter Richard Jardine Robert Semple
Mark Cassidy Dong-Sheng Jeng Marc Senders
David Cathie Hackmet Joer Shambhu Sharma
Liang Cheng Steven Kay Abdul-Hamid Soubra
Fiona Chow Lindita Kellezi Jamie Standing
Shin Fun Chung Richard Kelly Robert Stevens
Chris Clayton Mohamed Khorshid Doug Stewart
Ed Clukey Harry Kolk Ken Stokoe
Geoff Cole Fook Hou Lee Kevin Stone
Jean-Louis Colliat Barry Lehane Chairat Supachawarote
Andrew Deeks Kerstin Lesny Hossein Taiebat
Gijs Degenkamp Colin Leung Tomiya Takatani
Don DeGroot Nina Levy Luc Thorel
Jason DeJong Jon Tor Lieng George Vlahos
Earl Doyle Martin Liu Phil Watson
Clarence Ehlers Samanthika Liyanapathirana David White
Itai Einav Moatian Luan Judith Whittick
Sarah Elkhatib Tom Lunne Xiangtao Xu
Rami El-Sherbiny Chris Martin Lidija Zdravkovic
Carl Erbrich Paul Mayne

XIX

Copyright 2005 Taylor & Francis Group plc, London, UK


Keynote papers

Copyright 2005 Taylor & Francis Group plc, London, UK


Suction anchors for deepwater applications

K.H. Andersen
Norwegian Geotechnical Institute (NGI), Oslo, Norway

J.D. Murff
Offshore Technology Research Center (OTRC), College Station, Texas, USA

M.F. Randolph
Centre for Offshore Foundation Systems (COFS), UWA, Perth, Australia

E.C. Clukey
BP America Inc., Houston, Texas

C.T. Erbrich
Advanced Geomechanics (AG), Perth, Australia

H.P. Jostad & B. Hansen (NGI)


C. Aubeny & P. Sharma (OTRC)
C. Supachawarote (COFS)

ABSTRACT: This paper summarizes the results of an industry sponsored study on the design and analyses of
suction anchors in soft clays. References on suction anchors (200) were collected and a number of prediction
methods and data related to installation performance and holding capacity were identified and summarized. The
practices for predicting the installation performance and capacity of suction anchors were evaluated, including an
assessment of their simplicity, completeness, sensitivity, practicality, and generality. Research topics with the
potential for improving current practice were identified. The basis of the evaluation was a comparison of predic-
tions of hypothetical cases of various simplified methods as well as a comparison of predictions using these meth-
ods with ground truth data from either rigorous 3D finite element analyses or prototype data where available.

1 INTRODUCTION 2 SUCTION ANCHORS

An industry sponsored study on the design and analy- A suction anchor is a large diameter cylinder, open-
sis of deepwater anchors in soft clay was completed ended at the bottom and closed at the top (Figure 1).
in 2003. The overall objective was to provide the API Mooring loads are applied by an anchor line usually
Geotechnical Workgroup (RG7) and the Deepstar Joint attached to the side of the caisson. The length to
Industry Project VI with background, data and other diameter ratio of the caisson is typically six or less.
information needed to develop a widely applicable rec- Once installed, the caisson acts much like a short rigid
ommended practice for the design and installation of pile and is capable of resisting both lateral and axial
deepwater anchors. loads. The maximum holding capacity is obtained if the
This paper summarizes the part of the work related chain is attached at a depth where the anchor failure
to the design and analysis of suction anchors. The part mode is large translational displacements with min-
of the work related to vertically loaded drag anchors is imal rotation (optimum load attachment point).
summarized in the accompanying paper by Murff et al. The suction caisson gets its name from the fact
(2005). that it is usually installed by applying under-pressure

Copyright 2005 Taylor & Francis Group plc, London, UK


uncertainty in predicting the anchors installation per-
formance and holding capacity.
In Phase II the current practices for predicting
the installation performance and capacity of suction
anchors identified in Phase I were evaluated, including
an assessment of their simplicity, completeness, sensi-
tivity, practicality, and generality. The various prediction
methods were also compared and evaluated by using
them to predict penetration behaviour and holding
capacity of hypothetical cases, and to predict ground
truth data from either rigorous numerical analyses
(finite element analyses) or prototype data where avail-
able. Predictions were performed by the Offshore
Technology Research Center (OTRC), the Centre for
Offshore Foundation Systems (COFS), the Norwegian
Geotechnical Institute (NGI) and one industry predic-
tor. COFS co-operated with Advanced Geomechanics
(AG) on the installation predictions. Research topics
with the potential for improving current practice were
also identified.
The objective of the hypothetical case prediction
Figure 1. Suction caissons for the Horn Mountain field in exercise was to have various companies using prac-
the Gulf of Mexico. (Photo: E.C. Clukey, BP). tical design methods to predict installation behaviour
and capacity for well defined suction anchor cases.
Comparisons of the results from the various pre-
(suction) to its interior after it is allowed to penetrate dictors were used to give an assessment of the vari-
under its own weight. The difference between the ability in the design methods for well defined input,
hydrostatic water pressure outside the cylinder and where the input interpretation such as soil strength
the reduced water pressure inside provides a differen- and geometry simplifications, were removed from the
tial pressure that acts as a penetration force in addition assessment.
to the weight. After installation the caissons interior is As ground truth data, installation data for 6 well
sealed off and vertical loading creates an internal defined prototype suction anchor cases in normally
underpressure which in turn mobilizes the end bearing consolidated and lightly overconsolidated clays were
resistance of the soil at the caisson tip. used. For capacity, the cases defined for the hypothetical
Proof loading to check the anchor holding capacity capacity cases were also analyzed by 3D finite element
after installation is not required for suction anchors analyses. Independent 3D finite element analyses were
because (1) their positioning is well controlled; (2) the performed by NGI, COFS and OTRC to ensure the
foundation design is based on prediction methods cali- quality of the 3D finite element results.
brated against model test data and detailed numerical
analyses, and (3) the soil conditions are normally well
documented by in situ and laboratory testing. As suc- 4 COLLECTED INFORMATION
tion anchors are relatively shallow structures, deep
soil borings are not needed, but more detailed soil 4.1 Data from prototypes and experimental studies
data are needed at shallow depths than for piles.
Thanks to the willingness of the industry to provide
proprietary data, the project has compiled a good set
3 DESCRIPTION OF THE STUDY of detailed prototype and experimental data on suc-
tion anchor applications and experiments. More than
The study had two phases. Phase I focused on collec- 485 suction caissons installed at more than 50 locations
tion of references, prediction methods, and data from in water depths to nearly 2000 m were identified by the
actual applications, field tests, and experimental stud- end of the project (Table 1). Suction anchors have been
ies. It was an attempt to establish a baseline of data and applied in most of the worlds deepwater oil produ-
prediction methods for suction anchors. This baseline, cing areas (North Sea, Gulf of Mexico, Offshore West
in turn, served as a jumping off point for Phase II Africa, Offshore Brazil, West of Shetland, South China
which was aimed at evaluating the general methods in Sea, Adriatic Sea, and Timor Sea).
use by industry for analysis and design of these anchors. Detailed installation data have been compiled for
Of particular interest was an assessment of the bias and 16 of the more than 50 locations where suction

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 1. List of installed suction anchors.

Depth DL
Year Field Floater (m) (m  m) No. Operator

1981 Gorm1) FSO 40 3.5  8.5 12 Shell


1991 Snorre1) TLP 335 30  13 4 Saga
1993 Tordis1) Well Head 200 9  6.5 1 Saga
1994 Heidrun1) TLP 375 44  4.5 4 Conoco
1995 Nkossa2) Barge 170 4.55  12 12 Elf
1995 Yme1) Loading buoy 100 57 8 Statoil
1995 Harding1) Loading buoy 110 5  810 8 BP
1996 Norne1) FPSO 375 5  10 12 Statoil
1997 Njord1) Semi 330 Norsk
FPU\FSU 5  810 12 Hydro
5  710 8
1997 Curlew1) FPSO 90 57  912 9 Shell
1997 Marlim Semi 7701000 4.7  13 32 Petrobras
P19P263) FPU
1997 Schiehallion4) FPSO 400 6.5  12 14 BP
1997 Visund1) Semi FPU 345 5  11 16 Norsk
Hydro
1997 Lufeng5) FPSO 30 5  10 8 Statoil
1997 Aquila6) FPSO 850 4.55  16 8 Agip
1998 Laminaria7) FPSO 400 5.5  13 12 Woodside
1998 Marlim P333) FPSO 740840 4.7  20 6 Petrobras
1998 Marlim P18 3) Riser 900 18  16.2 2 Petrobras
support
1998 Siri1) Loading 60 4.25  4.6 1 Statoil
buoy (4 cells)
1)
Snorre B 335 Norsk
Hydro
1998 Banff1) 4 cells 90 4.3  11 1 Conoco
1998 Aasgard1) Mid water arch 303 9.5  7.57.7 3 Statoil
1998 Aasgard A1) FPSO 350 5  11 12 Statoil
1999 Aasgard A1) Tieback 350 Statoil
1999 Aasgard B & C1) Semi- 350 5  10 16 Statoil
FPU & FSO 5  12 9
1999 Marlim P35 3)
FPSO 810910 4.8  17 6 Petrobras
1999 Troll C1) Semi- 350 5  15 12 Norsk
FPU Hydro
1999 Kuito2) FPSO 400 3.5  14 12 Chevron
3.5  11
1999 Diana8) SPAR 1500 6.5  30 12 Exxon-
Mobil
1999 Preset anchors8) 2.8  16 1 BP
2.8  13 1
Green Canyon 8548) Semi-Rig 1650 3.7  18.5 8
1999 MODU8) ? Shell
2000 Hanze1) 40 6.5  6.1 1 VEBA
2001 Girassol2) Riser 1350 8  20 3 TFE
tower
2001 Girassol2) FPSO 1350 4.5  17 16 TFE
2001 Girassol2) Loading 1350 5  18 6 TFE
buoy 5  16.1 3
2002 NaKika8) FDS 1920 4.3  23.8 16 Shell/BP
2002 Horn Mountain8) SPAR 1650 5.5  27.4 6 BP
5.5  29 3
2002 Wenchang5) FPSO 120 5.5  12.1 9 CNOOC
5.5  12.8
2003 Barracuda3) FPSO 825 5  16.5 18 Petrobras
2003 Caratinga3) FPSO 1030 5  16.5 18 Petrobras
(Continued)

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 1. (Continued)

Depth DL
Year Field Floater (m) (m  m) No. Operator

2003 Bonga9) FPSO 980 5  17.5 12 Shell


5  16 Nigeria
2003 Bonga9) SPM 943 3.5  18 9 Shell
3.5  16 Nigeria
2003 Red 1600 5.5  22.9 8 Kerr
Hawk8) McGee
2003 Devils 1700 5.8  34.8 9 Dominion
Tower8)
2003 Ardmore1) 4 cells 78 6.5  3 1 Tuscon
6.5  5.5 1
2003 Panyu5) 105 5  11.7 9 CNOOC
6  12.7
2003 Holstein8) 1280 5.5  36.3/38.4 16 BP
2004 Thunder Semi 1830 5.5  27.5 16 BP
Horse8) FPU
2004 Thunder Manifold 1830 6.4  23.8 4 BP
Horse8)
2004 Thunder Horse8) PLET 1830 5.5  26 4 BP
2004 Thunder Water injection 1830 3.4  19 3 BP
Horse8) 3.4  20 2
2004 Mad 1600 5.5  11 BP
Dog8) 7.6 
1)
North Sea 2) West Africa 3) Offshore Brazil 4) West of Shetlands 5) South China Sea 6) Adriatic Sea 7) Timor Sea 8) Gulf of
Mexico 9) Offshore Nigeria.

anchors were identified. There are no reports of mis- and 8 computer programs to calculate holding
behaviour during operation, and thus no data on hold- capacity.
ing capacity. Of the installation programs, three are EXCEL
Experimental studies that were identified by the end spreadsheets, one is a MathCad document and one is
of the study (2003) are listed in Table 2. The list con- a FORTRAN code. One program is based on limit
tains 19 cases, of which there are 4 full scale field tests, equilibrium of forces, but includes minimization of
3 large scale field model tests, 10 centrifuge tests, and plastic work in assessing flow around ring stiffeners.
21 g laboratory model tests. Thirteen of the cases con- The other programs are based on limit equilibrium of
tain installation, 4 cases contain removal (extraction), forces.
and 14 cases contain loading to failure (capacity). Of the holding capacity programs, seven use limit
Installation data for six of the prototype cases are equilibrium methods, one has the option to use a spe-
presented in Figure 2 and Table 3. The cases have rea- cially formulated equivalent 2D finite element code,
sonably well defined soil conditions and good quality and one program is a 3D finite element code that can
measurements during skirt penetration. They are all in be used both for undrained situations and drained or
soft normally consolidated or lightly overconsolidated partly drained long term load situations. The programs
clay, but cover various clay plasticity and strength have different capabilities with respect to failure mech-
profiles with different strength increase with depth anisms, coupling between vertical and horizontal loads,
and near surface strength. One case (Laminaria) is in and anisotropic shear strength modelling.
calcareous soil. The 6 cases also cover various types of Interpretation of the collected information about
stiffener arrangement and two cases (Girassol FPSO parameters needed for installation and holding cap-
and Girassol Offshore Loading Buoy) have a partly acity predictions indicated that for installation it
painted outside skirt wall. seems that:
A list of more than 200 published references was
established. There is agreement about general principles (e.g.
use of remoulded shear strength to calculate skirt
friction, and critical underpressure with respect to
4.2 Prediction methods
soil heave inside the anchor governed by inverse
The prediction method collection contains detailed bearing capacity of clay plug at skirt tip level and
information of 5 computer programs for installation inside skirt friction).

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 2. Experimental suction anchor studies.

Year Test site Test type Test description Reference

1985 Gullfaks Full scale field Installation and extraction of 2 large diam. Tjelta et al. (1986)
North Sea test (6.5  22 m) concrete cylinders
198? Univ of Col., Centrifuge Monotonic lateral load tests on Unpublished
Boulder model cylinders
1989 NGI/Lysaker Large scale field Monotonic and cyclic TLP loads Dyvik et al. (1993)
model tests 10 from vertical. Andersen et al. (1993)
1991 Focomorto Large scale field Installation of concrete skirt pile. ONeill et al. (1991)
model test
1991 ISMES Centrifuge Installation and monotonic & cyclic Renzi et al. (1991)
vertical load.
1991 DGI Centrifuge Installation and uplift tests, two Fuglsang et al. (1991)
uniform shear strength profiles. Steensen-Bach (1992)
1991 NGI/Lysaker Large scale field Monotonic and cyclic lateral loads Keaveny et al. (1994)
model tests 10 from horizontal.
199093 LCPC Centrifuge Monotonic and cyclic uplift tests on two Clukey et al. (1993/95)
different size caissons. One lateral test. Morrison et al. (1994)
1994 Stavanger Field test Installation & extraction of suction Unpubl.
anchor (5  8.5 m) (Statoil/APL/NGI)
1996 Tordis Field test Installation and removal of skirted anchor Offshore Engr. (1996a)
(5  8 m). Incl. 3 months testing of fiber rope.
1996 Marlin Field test Installation, testing and removal of 3.6 m Offshore Engr. (1996b)
diameter, 18 m long skirted anchor.
1998 MIT 1 g lab model Installation & capacity. Miniature Whittle et al. (1998)
test caisson. Clay.
1998 U W Aust Centrifuge Monotonic and cyclic lateral loads Randolph et al. (1998)
199799 GeoDelft Centrifuge Installation, monotonic & cyclic capacity Andersen et al. (2003)
1999 U W Aust. Centrifuge Installation and undrained uplift McNamara (2000)
2000 CCore Centrifuge Installation and undrained uplift Cao et al. (2002)
199804 Univ. of Texas, 1-g lab models Installation and monotonic Olson et al. (2003)
Austin capacity. Kaolin Rauch et al. (2004)
2001 U W Aust. Centrifuge Installation and undrained uplift House & Randolph (2001)
2002 C-Core Centrifuge Uplift capacity for installation with Clukey & Phillips (2002)
or without suction. Kaolin

For capacity, the interpretation of the collected infor-


The sensitivity is measured by different methods by
mation indicated that there seems to be:
the different companies. The sensitivity may depend
on the apparatus used, and this should be given Uncertainty in prediction of skirt friction. Difficulties
attention. include uncertainty over thixotropic and consolida-
Stiffeners may cause a gap along the wall. Determi- tion strength gains.
nation of gap formation is uncertain, especially above Some controversy as to whether conventional bear-
outside stiffeners. ing capacity factor, allowing for embedment ratio of
Internal ring stiffeners may cause some uncertainty, the caisson, can be used to calculate uplift capacity
as soil may be trapped between them. at skirt tip level.
There is some controversy as to the appropriate Agreement that coupling of vertical, horizontal and
bearing capacity factor to be used for uplift to cal- moment capacities is necessary.
culate plug failure. Agreement that an open crack along the active side
There are differences in safety factor philosophy may be important. More work is needed to establish
and numerical safety factor requirement. when a crack may occur.

Copyright 2005 Taylor & Francis Group plc, London, UK


Required Underpressure (kPa) Required Underpressure (kPa) Required Underpressure (kPa)
0 50 100 150 0 25 50 75 100 0 50 100 150
0 0 0

2 2
5
4 4
10
6 6

Depth (m)
Depth (m)
Depth (m)

15 8 8

20 10 10
12 12
25
14 14
30
16 16
35 18 18
(a) Diana (b) Marlin offshore test (c) Girassol FPSO
Required Underpressure (kPa) Required Underpressure (kPa) Required Underpressure (kPa)
0 25 50 75 100 0 25 50 75 100 0 50 100 150
0 0 0
2
2 2
4
4 4
6
Depth (m)

Depth (m)
Depth (m)

8 6 6

10 8 8
12
10 10
14
12 12
16
18 14 14
(d) Girassol Offload Buoy (e) Laminaria (f) Nkossa Type 1

Figure 2. Installation data for prototype suction anchors at the 6 sites in Table 3.

Differences in manner in which large long term loads approach for both the hypothetical and prototype instal-
are addressed. lation cases, while Predictor 2 also used an effective
Differences in safety factor philosophy and numer- stress approach for the prototype cases.
ical safety factor requirement. In the total stress approach, the remoulded shear
strength is determined either from (1) direct measure-
ments of the strength of remoulded samples, or (2)
5 INSTALLATION
the intact shear strength divided by the sensitivity (i.e.
as   su, with  equal to the inverse of the sensitivity).
5.1 Calculation procedures
Additional correction factors are applied in cases where
5.1.1 Penetration resistance and underpressure the interface between the anchor and the soil is expected
The penetration resistance of suction caisson anchors to have lower strength than the clay, like in cases with
is calculated as the sum of the integrated interface painted walls. The correction factor can be determined
shear strength along the outer and inner skirt walls from ring shear tests.
and along potential plate and ring stiffeners, and the The remoulded shear strength can be determined
end bearing resistance of skirt tips, plate stiffeners, by various methods, but fall cone and miniature vane
ring stiffeners and potential anchor diameter changes. are presently the most common. In cases where the full
In the case of penetration by underpressure, the shear strength may not be mobilized along the skirt
required underpressure is calculated as the penetration wall, ring shear tests may be used to measure the actual
resistance minus the submerged anchor weight, divided interface shear strength. When the interface strength
by the inside cross section area beneath the top lid. is calculated as   su, the undrained direct simple shear
The clay along the outside skirt wall is assumed to strength is normally used as the reference intact
be remoulded, and the remoulded shear strength is cal- shear strength. The advantage of using   su rather
culated either by a total stress or an effective stress than the remoulded shear strength directly is that the
approach. All four predictors used a total stress intact shear strength profile often is better defined

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 3. Key data for suction anchors in Figure 2.

Dimensions and weight Soil

Field Location D (m) L (m) L/D ttip (mm) W(kN) suDSS (kPa) Ip (%) St Stiffeners Comments

Diana Gulf of 6.4 30.5 4.8 51 2150 2.25 + 0.79  z 7030 34 3.05 m high plate stiffener Offshore Engineer
Mexico (014.2 m) 13.5  at pad eye (1999).
0.65  (z-14.2)
(14.2 m)
Marlin Gulf of 3.7 18.3 5 46 625 1.45  z 5535 2.25 1.8 m high 0.0254 m web Offshore Engineer
Offsh. Test Mexico plate at pad eye (1999b).
Girassol Offshore 4.5 17.3 3.8 20 540 6 (03 m) 6  80130 2.75 0.4 m wide ring stiffeners 20% of outside
FPSO Angola 1.26  (z-3) 3.9 & 5 m from tip. 0.07 m skirt painted.
(3 m) wide ring stiffeners 8.82, 11.7 & Dendani & Colliat
14.5 m from tip (2002).
9

Girassol Offshore 5 18 3.6 20 490 6 (03 m) 6  80130 2.75 0.2 m wide ring stiffeners, 20% of outside
Offload Angola 1.26  (z-3) 1.2 m spacing, starting 1.3 m skirt painted. Colliat &
Buoy (3 m) above tip. 0.8 m high 0.045 m Dendani (2002).
plate 5.4 m above tip
Laminaria Offshore 5.5 12.2 2.2 20 380 7.5  1.68  z 3035 2.8 8  0.165 m wide ring stiffeners. Calcareous soft soil.
Australia 2 1.6 m high 0.045 m plate Erbrich & Hefer
stiffeners at padeye (2002).
Nkossa Gulf of 4.5 7.5 2.7 15 355 2.5  1.5  z 2535 3.3 0.1 m wide ring stiffeners,
Type 1 Guinea upper upper spacing 1.8 to 2 m Variable Diameter.
4 4.8 3.1 Colliat et al. (1996).
lower lower

Copyright 2005 Taylor & Francis Group plc, London, UK


than the remoulded strength profile. The disadvan- (1) Equilibrium is checked between the shear stresses
tage is that the sensitivity is an uncertain parameter, from the weight of the clay plug and a certain propor-
and the sensitivity may be too low if the intact shear tion of the mean triaxial compression strength over
strength is measured on disturbed samples. When the the height of the clay plug (Andersen & Jostad 2004).
design remoulded shear strength is established, both The proportion of the triaxial compression strength
approaches should be applied and used as input to can be determined from results of special triaxial tests
establish the final profile. deformed in extension followed by compression to
In the effective stress approach, the external friction, simulate the strain history of a soil element that enters
fse, is determined as: into the anchor, passes a ring stiffener and expands
laterally after having passed the ring stiffener.
(1) If there is only one ring stiffener, the shear strength
along the inside skirt wall is set to zero to the depth
where ri is the radial effective stress during installa- where the clay plug deforms back to the wall. Below
tion, accounting for remoulding of the soil adjacent to this depth, the shear strength along the inside skirt
the pile wall, and interface is the interface friction angle wall is calculated as for the external wall.
between the pile and soil. In cases with more than one ring stiffener, the shear
Two methods are used to determine the radial effect- strength along the inside skirt wall above the first stiff-
ive stress with the higher value obtained from these ener is set equal to either zero or the remoulded shear
two methods used in each case. strength of clay from the upper part of the profile,
In the first method, the mean effective stress when assuming that either water or clay from the upper part of
the soil is fully remoulded is determined, based on the the profile will be trapped in the compartments between
remoulded shear strength and the friction angle of the stiffeners. It is assumed that water is trapped if the
the soil. Then the excess pore pressure caused by plug can stand to a depth that is larger than the distance
remoulding is calculated from the difference between between the stiffeners.
the initial mean effective stress and the mean effective The bearing capacity of the lowest ring stiffener is
stress when the soil is fully remoulded. This excess pore calculated by conventional bearing capacity formula,
pressure is then subtracted from the initial horizontal with a bearing capacity factor reduced for the effect of
effective stress, after making an adjustment to the total remoulded clay along the skirt wall and above the stiff-
stress based on cavity expansion theory. ener, and for the possibility of squeezing of the clay
In some cases this approach does not lead to sens- plug instead of a local bearing capacity failure around
ible values for the radial effective stress, and hence the stiffener.
there is another check just based on the individual (2) Full flow-round is assumed to occur only if
effective stress components determined for the fully su/vo  0.4 at any depth (Erbrich & Hefer 2002).
remoulded soil. In this assessment the minor principal Where full flow-round does not occur, the internal
effective stress is used as the radial effective stress friction behind the stiffener is taken as the lesser of the
but this is determined using assumptions which implic- remoulded strength or 1 kPa in order to represent a
itly ensure that this errs on the high side of what is soil/water mixture. Where full flow-round is assumed
likely to be the true minor principal effective stress to occur, the internal friction is taken as the remoulded
(as it is not possible to be sure of the actual individual shear strength, but with due consideration for entrap-
effective stress components). ment of weak material from near the surface between
The effective stress approach is partly based on the successive stiffeners. End bearing on internal ring stiff-
undrained remoulded shear strength or soil sensitivity, eners is determined using a rigorous upper bound plas-
but there are also uncertainties in determining the radial ticity solution including flow around stiffeners and
effective stress. The approach can incorporate appro- extrusion of the soil plug. The assessment of internal
priate values of the interface friction angle, interface stiffeners is the same for the total and the effective
for different pile surfaces, for example as determined stress models.
from ring shear tests. The bearing capacity of skirt tip and plate stiffen-
In cases without ring stiffeners, and below the first ers is calculated by conventional bearing capacity for-
ring stiffener in cases with ring stiffeners, the shear mulas using the average intact shear strength (average
strength along the inside skirt wall is calculated as for of compression, direct simple shear (DSS) and exten-
the outside skirt wall, both for the total stress and the sion) and including the overburden (  z) above the
effective stress models. bearing area. The bearing capacity factor may be influ-
In cases with internal ring stiffeners, the practice enced by interaction from the wall friction, but this is
varies. Generally, it is necessary to check whether the often neglected in tip resistance calculations. This is
upper part of the clay plug will deform back to the skirt normally acceptable, because the bearing capacity of
wall after passing the 1st ring stiffener. Two approaches these elements is small compared to the total penetra-
used in this project are described below. tion resistance.

10

Copyright 2005 Taylor & Francis Group plc, London, UK


In special cases where the bearing capacity is a more greater than unity (representing a worst case estimate
significant contribution, however, the coupling effect of plug base resistance) is applied to the soil below the
may be taken into account. Different bearing capacity caisson tip.
factors in the range Nc  7.5 to 11 were applied by The maximum recommended penetration depth is
the predictors in this project. The high number takes determined as the depth where the safety factor with
into account the influence of the wall friction above respect to critical underpressure is 1.5 (two predictors)
the bearing area. Reductions in Nc were made for depth or 1.3 (one predictor), or where the material coeffi-
effects in cases with shallow penetration. For inside ring cient with respect to the shear strength in the inverse
stiffeners, bearing capacity factor in the range Nc  5 bearing capacity term is 1.5 (one predictor).
to 13.5 were applied, depending on the penetration Careful field monitoring of the soil plug position
depth and the strength of the soil above and below the during installation can mitigate possible plug heave
ring stiffener. and, if required, justify lower safety factors.

5.1.2 Critical underpressure 5.1.4 Plug heave


The critical underpressure with respect to soil heave The plug heave is calculated as the sum of the volume
inside the anchor is calculated as the inverse bearing of the soil displaced into the anchor at the skirt tip level
capacity of the clay plug at skirt tip level, plus internal and the volume of the inside stiffeners. The volume of
resistance from the inside skirt wall and inside stiffeners. soil displaced into the anchor at skirt tip level depends
The bearing capacity factor applied by the various on the skirt wall thickness at skirt tip level and whether
predictors varied from Nc  7 to 9, reduced as appro- the anchor penetrates under self weight or by under-
priate at shallow penetrations. The low value of Nc  7 pressure. All predictors assume that half the displaced
was applied by one of the predictors to account for a soil goes into the anchor when the anchor penetrates
possible lower capacity in uplift than in compression. under self weight. When the anchor is penetrated by
The shear strength to calculate the inverse bearing underpressure, the assumptions by the various pre-
capacity was taken as the average shear strength (aver- dictors vary between 50% and 100% of the displaced
age of compression, DSS and extension) at skirt tip soil going into the anchor.
level for all but one of the predictors, who used an If there are inside ring stiffeners, the upper part of
average of this average shear strength and the extension the clay plug may not deform back to the wall, and the
shear strength to account for a possible lower capacity extension of the clay plug and trapped water may add
in uplift than in compression due to anisotropy. to the plug heave. The depth to which the clay plug can
stand without deforming back to the skirt wall and the
5.1.3 Safety factor and maximum penetration depth possibility of trapped water can be calculated as
If the critical underpressure is exceeded it may not be described for calculation of penetration resistance in
possible to penetrate the anchor deeper. Underpressures Section 5.1.1.
exceeding the critical underpressure will cause the soil
plug inside the anchor to move upwards without fur-
5.2 Hypothetical cases
ther penetration of the skirts.
The safety factors with respect to large plug heave Four different hypothetical installation cases were
are calculated in different ways by the predictors: one defined, as specified in Table 4. Two different clay
is as the ratio between the critical and the required profiles (one normally consolidated and one lightly
underpressures (essentially applying a material factor overconsolidated) were specified. The clay data are
to all shear strengths or equivalent in an effective stress defined in Table 5. Two of the cases (I1 and I3) did not
approach); the other is based on the material coeffi-
cient for the shear strength used to calculate the uplift Table 4. Hypothetical installation cases.
capacity of the clay plug at skirt tip level, assuming a
material coefficient of unity on the remaining compon- Case: I1 I2 I3 I4
ents of internal plug resistance (Andersen & Jostad
1999). The logic for the second approach is that the Diameter (m) 5 5 5 5
internal plug resistance contributes equally to the Penetr. depth (m) 30 30 30 30
required underpressure and the plug resistance, and so Depth/diameter 6 6 6 6
uncertainty in the internal resistance should not affect Skirt wall 35 mm steel skirts
Weight (kN) 1100 1100 1100 1100
the estimation of safety.
Inside stiffeners No Yes * No Yes *
Taking this approach further suggests an approach Overconsolidation 1.0 1.0 1.6 1.6
whereby the external resistance is estimated using a ratio, OCR
material coefficient smaller than unity (representing a
worst case estimate of caisson resistance, and hence of * 0.2 m wide, 2.5 m spacing ring stiffeners. 1st stiffener 2.5 m
required underpressure) while a material coefficient above skirt tip.

11

Copyright 2005 Taylor & Francis Group plc, London, UK


have inside stiffeners, whereas the two others (I2 and as the difference between the anchor height and the
I4) had 0.2 m wide ring stiffeners with 2.5 m vertical penetration depth at the end of penetration. One uncer-
spacing. The lowest stiffener was located 2.5 m above tainty in this evaluation is the effect of stiffeners and
the skirt tip. other geometrical effects underneath the top plate. For
The predictors were asked to calculate the follow- Laminaria, a reduction in diameter in the upper part of
ing for each of the 4 cases: (1) penetration resistance the anchor also led to some uncertainty. The Diana
with depth; (2) self weight penetration depth; (3) required anchors were not penetrated to refusal, and the differ-
underpressure with depth; (4) critical underpressure as ence between the anchor height and the penetration
function of depth; (5) soil heave as function of depth; depth only gives an upper limit of the soil heave. The
(6) maximum recommended penetration depth, includ- Marlin offshore test anchor was not penetrated to
ing recommended safety factor. The calculation pro- refusal either, but the plug heave could in this case be
cedure and all input parameters were documented by estimated from the height difference of clay attached to
each predictor. the outside and inside skirts after retrieval of the test
anchor. In the Nkossa case, echo sounders showed that
5.3 Prototype cases the clay plug was generally 1 to 1.5 m lower than the
outside clay surface at the end of penetration (Colliat
The predictors were asked to make their best predic-
et al. 1996). The reason for the clay plug being lower
tions with their normal design method for a set of
than the outside surface at Nkossa is that the anchor
well-documented prototype installation cases. If the
diameter is larger in the upper 7.5 m than in the lower
first calculations did not match the measured behav-
4.8 m, and therefore the clay plug will fill out the extra
iour, the predictors were encouraged to make correc-
space and sink down when it enters the upper part.
tions or calibrations that they find appropriate to get a
better agreement. This could be corrections to the input
data or to the calculation method. It was emphasized, 5.4 Findings and recommendations; hypothetical
however, that modifications should be identified and installation cases
that the relevance of the modifications be discussed. Calculated required and critical underpressures, plug
The 6 cases specified in Table 3 were selected for heave and safety factors are presented in Figure 3 for
back-calculation. As mentioned in Section 4.1 these cases I1 and I3. The effective stress approach was not
cases cover various clay types, stiffener arrangements applied for the hypothetical cases. The comparison of
and skirt wall treatment. The underpressures required the results from the 4 predictors for these and the two
to penetrate the anchors are presented in Figure 2. The other hypothetical cases shows various similarities and
soil heave observed inside the six anchors at the end differences between the predictors, as discussed below.
of penetration is summarized in Table 6.
The Girassol and the Laminaria anchors were pene- 5.4.1 Similarities
trated to refusal, and the plug heave was estimated Penetration resistance of skirts with no stiffeners. The
predicted penetration resistance, required underpres-
Table 5. Soil data for hypothetical cases. sure and self weight penetration are all very similar for
skirts with no stiffeners, both in normally consolidated
Overcons. ratio, OCR: 1.0 1.6 and lightly overconsolidated clays (e.g. Figure 3a, top
diagram). The predictors use very similar methods for
sDSS
u (kPa) 1.25  z 10 (z 5 m); 2  these items. One deviation is Predictor 4, who calcu-
z (z 5 m) lates higher resistance in normally consolidated clay
sCu (kPa) 1.2  sDSS
u 1.2  sDSS
u due to use of a higher -value than the inverse of the
suE (kPa) 0.8  sDSS
u 0.8  sDSS
u sensitivity   1/St.
0 0
c/  0/31 0/31
su along outside skirt 0.65  sDSS
u 0.65  sDSS
u 5.4.2 Differences
wall (kPa) Critical underpressure. Predictor 1 calculates a crit-
vc (kPa) 6z 7.2  z
ical underpressure with respect to soil heave inside the
K0 0.55 1.0 (z 5 m);
0.65 (z 5 m) anchor that is 16% lower than Predictors 2 and 3 in
the case with no stiffeners (e.g. Figure 3). The reason

Table 6. Observed soil heave inside anchor at end of penetration.

Marlin Girassol Girassol


Case: Diana test FPSO buoy Laminaria Nkossa

Soil heave (m) 1.0 0.45 0.8 1.6 0.9 to 1.5


1 to
1.5

12

Copyright 2005 Taylor & Francis Group plc, London, UK


is that Predictor 1 uses a reduced bearing capacity 2 and 3 due to the use of a higher -value than the
factor (Nc  7 at large depths) to calculate the uplift inverse of the sensitivity   1/St.
capacity of the clay plug at skirt tip level. The other Safety against plug heave. Predictors 1, 2 and 4 cal-
predictors use the same factor for uplift as for bearing culate significantly lower safety factors than Predictor
capacity (Nc  9 at large depths). Predictor 4 calcu- 3. The reason is that the safety factor is defined dif-
lates 13% higher critical underpressure than Predictors ferently by the predictors. Predictors 1, 2 and 4 define

Required Underpressure (kPa) Required Underpressure (kPa)


0 100 200 300 400 0 100 200 300 400 500
0 0

5 P1 5 P1

10 10 P2
P2

Depth (m)
Depth (m)

P3
15 P3 15
P4
20 P4 20

25 25

30 30
35 35

Allowable Underpressure (kPa) Allowable Underpressure (kPa)


0 100 200 300 400 500 600 0 200 400 600 800 1000
0 0

5 5

10 10
Depth (m)
Depth (m)

15 15

20 20

25 25

30 30
35 35

Plug Heave (m) Plug Heave (m)


0.0 0.2 0.4 0.6 0.8 0.0 0.5 1.0 1.5
0 0

5 5

10 10
Depth (m)
Depth (m)

15 15

20 20

25 25
30 30
35 35

Safety factor Safety factor


0 2 4 6 8 10 0 2 4 6 8 10
0 0

5 5

10 10
Depth (m)
Depth (m)

15 15

20 20

25 25

30 30
35 35
(a) Hypothetical installation case I1. OCR=1. No stiffeners. (b) Hypothetical installation case I4. OCR=1.6. With stiffeners.

Figure 3. Calculated behaviour of 2 of the hypothetical installation cases.

13

Copyright 2005 Taylor & Francis Group plc, London, UK


the safety factor as the ratio between the critical and the uses a reduced bearing capacity factor to account for
required underpressures, whereas Predictor 3 defines effect of remoulded clay and squeezing of the clay plug.
the safety factor as a material coefficient on the shear Predictor 4 assumes full inside friction with shear
strength used to calculate the uplift capacity of the strength higher than the remoulded strength in normally
clay plug at skirt tip level. There is also some differ- consolidated clay and no shear strength along the inside
ence between Predictors 1, 2 and 4. This is due to the skirt wall above the first stiffener in lightly overcon-
differences in the calculation of the required and crit- solidated clay.
ical underpressures, as discussed above. Self weight penetration depth. The different methods
Recommended safety factor against plug heave. to account for ring stiffeners also give some differ-
Predictor 1 recommends a safety factor of 1.3, whereas ences in the self weight penetration depth, with a
Predictors 2 and 3 recommend a safety factor of 1.5. range of 10.2 m to 13.2 m in the normally consolidated
As mentioned above, Predictors 1, 2 and 4 define the clay, and 6.6 m to 9.2 m in the lightly overconsolidated
safety differently from Predictor 3. clay (Figure 3b, top diagram).
Recommended maximum penetration depth. The Plug heave for skirts with no stiffeners. Predictor 1
recommended maximum penetration depth varies quite calculates 12%19% smaller plug heave than Predictors
significantly due to the differences in the definition 2, 3 and 4, because Predictor 1 assumes that 80% of the
of safety, the recommended value of the safety factor, soil displaced by the wall goes into the caisson, whereas
and the critical underpressure. The largest difference Predictors 2, 3 and 4 assume 100% during penetration
is for Case I1 (anchor with no stiffeners in normally by underpressure.
consolidated clay), where the recommended maximum Plug heave for skirts with stiffeners in lightly over-
penetration depth varies from 7.5 times the diameter consolidated clay. The soil plug heave is in this case
to 9.8 times the diameter. higher for Predictor 2 (1.16 m) than for Predictor 3
Effect of ring stiffeners. The predictors calculate (0.97 m) and Predictor 4 (0.85 m) (Figure 3b). Predictor
significantly different behaviour in the cases with inside 2 assumes the upper part of the soil plug stands under
ring stiffeners. its own weight to a larger depth than estimated by
Predictor 2 predicts penetration resistance and Predictor 3. Predictor 3 assumes some strength reduc-
required underpressure that are 2225% higher, and a tion in the clay plug due to the stress history of the clay
critical underpressure that is 1115% higher than as the plug enters the caisson and passes the first stiff-
Predictor 3. The lower numbers are for the normally ener. Predictor 4 estimates that the clay above the first
consolidated clay, and the upper numbers are for the stiffener expands half the distance from the edge of
lightly overconsolidated clay (Figure 3b). Predictor 1 the stiffener to the anchor wall.
predicts penetration resistance and required underpres- There is essentially no difference in predicted plug
sure between the other two for the normally consoli- heave for anchors with stiffeners in normally consoli-
dated clay and close to the lower numbers for the lightly dated clay, where Predictors 2, 3 and 4 all assume that
overconsolidated clay. Predictor 1s critical underpres- the clay plug will not stand.
sure is, however, lower than the two other predictors,
because Predictor 1 uses a lower Nc factor for uplift
5.5 Findings and recommendations: prototype
capacity, as mentioned above.
installation cases
Predictor 4 calculates significantly higher penetra-
tion resistance, required underpressure and critical Calculated and measured required underpressures for
underpressure than the three other predictors for the the 6 prototype cases are compared in Figure 4. The
case with ring stiffeners in normally consolidated clay. calculated underpressures are based on the initial
For the case with ring stiffeners in lightly overconsoli- assumptions. The total stress approach was used by
dated clay, however, Predictor 4 calculates lower pene- all 4 predictors, and one predictor also used the effect-
tration resistance than the other predictors. ive stress approach.
The reason for the differences is that the predictors
treat the ring stiffeners differently. Predictor 1 sums the 5.5.1 Penetration resistance and required
bearing capacity of each stiffener in remoulded soil and underpressure
neglects the shear strength along the inside skirt wall The comparison between calculated and measured
between the stiffeners. Predictors 2 and 3 both assume required underpressures shows that calculations with
that remoulded clay from the upper part of the clay the initial assumptions may give quite poor predictions
profile is trapped between the stiffeners for these clay in some cases. In most of these cases, good agreement
profiles and stiffener arrangement. However, Predictor with measurement could be achieved by making rea-
2 calculates the bearing capacity of the first stiffener sonable modifications to the initial assumptions. It is
with a gradually increasing bearing capacity factor that not obvious, however, whether the correct modifying
accounts for either plug extrusion or bearing resist- assumptions would be made in any given new case,
ance from subsequent stiffeners, whereas Predictor 3 especially with respect to the soil parameters.

14

Copyright 2005 Taylor & Francis Group plc, London, UK


Required Underpressure (kPa) Required Underpressure (kPa) Required Underpressure (kPa)
0 50 100 150 0 25 50 75 100 0 50 100 150
0 0 0
Measured
P1 2 2
5
P2 Total
4 4
10 P2 Effective
P3 6 6

Depth (m)
Depth (m)
Depth (m)

15 P4 8 8

20 10 10

12 12
25
14 14
30
16 16

35 18 18

(a) Diana (b) Marlin offshore test (c) Girassol FPSO

Required Underpressure (kPa) Required Underpressure (kPa)


Required Underpressure (kPa)
0 25 50 75 100 0 25 50 75 100
0 50 100 150
0 0
0
2
2
2
4
4 4
6
Depth (m)

Depth (m)

Depth (m)
8 6 6

10 8 8
12
10 10
14
12 12
16

18 14 14
(d) Girassol Offload Buoy (e) Laminaria (f) Nkossa Type 1

Figure 4. Comparison of calculated and measured required underpressures for installation in the 6 prototype cases in Table
3. Calculations with initial assumptions.

The total stress method gives required underpres- could be due to creep in the period between self weight
sures in good agreement with the measured ones, pro- penetration and penetration by underpressure, together
vided that (1) trapped water or trapped clay is accounted with set-up leading to higher resistance on re-starting
for in cases with more than one ring stiffener, and (2) penetration, but it is difficult to understand why these
that the soil parameters assumed originally are adjusted, effects should be more pronounced in this case than in
while remaining within the scatter of the data, for 3 of the others. The Marlin case showed similar (although
the 6 cases. If conditions (1) and (2) are not fulfilled, less abrupt) indications of set-up effects, and here the
the scatter in required underpressures at the final pene- trend of underpressure with depth predicted using the
tration depth varies between 50% and 200% of those total stress methods did not agree well with the meas-
measured, considering all the six cases and all the ured trend.
four predictors. The effective stress method for estimating frictional
The calculated self weight penetration depths also resistance mostly gave good agreement between pre-
agreed well with those measured once the above adjust- dicted and measured underpressures provided that
ments were made, except for the Diana case where the trapped water or clay is accounted for, except in two
calculated self weight penetration is slightly smaller cases where the predictions at the final penetration
than measured. It is uncertain why the agreement is depth were 55% and 150% of the measured data.
less good in this case, although the measured under- Increasing the assumed interface friction angle from
pressure with depth in this case deviates from the trend 12 to 17, which is within the range typically assumed
in the other 5 cases. Possibly, the deviation between for Gulf of Mexico soils, would correct the 45%
calculated and measured self weight penetration depth underprediction.

15

Copyright 2005 Taylor & Francis Group plc, London, UK


However, this case is also the one in which none of is neglected in the calculations, the soil strength or the
the total stress methods could match the self weight sensitivity would have to be adjusted by the same per-
penetration and hence other undetermined factors may centage. The results thus seem to indicate that paint
also be at play (see above discussion). For the over has an effect on the penetration resistance and that the
predicted case (in the carbonate rich Laminaria soils), effect is properly accounted for by reducing the -fac-
the interface friction angle had to be reduced from 18 tor or interface friction angle, , based on ring shear
to 9 to obtain good agreement, and this is outside the tests. However, the effect is too small to draw definite
expected range. This highlights the need for improved conclusions with respect to the effect of paint, since
understanding of installation friction values in car- this 1015% effect could be within the uncertainties
bonate clays. in the soil data.
Inside ring stiffeners. For the caissons where ring Overall, these cases illustrate that uncertainty in soil
stiffeners were used (Figures 4cf ) it was important data may be a key uncertainty in prediction of pene-
to estimate the extent to which the soil plug would tration resistance, self weight penetration and required
flow back against the internal wall, or would extrude underpressure, although it is also clear that the assump-
within the inner ring stiffener diameter, trapping tion of how the soil flows around and between ring
water (or extremely soft soil) between the stiffeners. stiffeners, and whether water or clay is trapped between
The latter assumption leads to lower predicted resist- the stiffeners, has a strong influence on the predicted
ance and gave better agreement with the measured installation resistance.
data. One predictor calculated the depth to which
the clay plug can stand freely, using approach 1 in 5.5.2 Soil heave inside anchors
Section 5.1.1 with a reduced shear strength deter- The observed data seem to support the soil heave cal-
mined from special triaxial tests where the strain path culations for anchors without ring stiffeners (Table 7),
of the clay plug is followed, giving strengths typically but the data is not sufficiently accurate to decide what
50% of the peak value in cases with relatively wide proportion of the soil displaced by the caisson wall
ring stiffeners. will go into the anchor when it is penetrated by under-
Bearing capacity factor. The predictors used a range pressure.
of bearing capacity factors, Nc, between 5 and 13.5 to Comparison between calculated and observed soil
calculate tip resistances of skirt walls, ring stiffeners, heave shows that it is likely that a mixture of water
plate stiffeners and changes in anchor diameter. The and clay can be trapped between internal ring stiffen-
breadth of the range in assumed Nc values suggests ers, and that trapped water can give a significant con-
further study is required for different geometries of tribution to the soil heave. However, assuming that
internal stiffener or external protuberances. Of particu- the compartment will be filled only with water over-
lar note are the divergent assumptions for Nc for the predicts the plug heave, and it is likely that some clay
internal ring stiffeners adopted for the Girassol Off- will also fall into the compartment.
loading Buoy case with one predictor using Nc increas-
ing from around 5.8 at shallow embedment to a 5.5.3 Critical underpressure
maximum of 13.5 and another predictor using Nc of Since none of the anchors experienced soil heave fail-
5 or 6.5. ure, the differences in the bearing capacity factors and
The effect of paint on the outside skirt wall of the the shear strength used to calculate the critical under-
Girassol FPSO and Offloading Buoy anchors was pressure could not be evaluated. It is therefore proposed
taken into account in all the calculations. The adjust- that the critical underpressure is calculated as proposed
ment in -factor or interface friction angle, , for the for the inverse bearing capacity in the anchor capacity
painted sections has an effect of 1015% on the cal- calculation, where the method is checked against 3D
culated required underpressure. If the effect of paint finite element analyses.

Table 7. Calculated and observed soil heave (in m) inside anchors at end of penetration.

Predictor Diana Marlin test Girassol FPSO Girassol buoy Laminaria Nkossa

Observed 1.0 0.45 0.8 1.6 0.9 to 1.5


1.0 to
1.5
P1 0.53 0.48 0.25 0.28 0.32 0.17
P2 0.6 0.5 0.8 to 0.9 0.4 to 0.5 1.3 to 1.6
0.7 to
0.9
P3 0.52 0.55 0.3* to 1.9** 0.3* to 2.6** 0.3* to 1.4**
1.4
P4 0.65 0.57 0.29 0.29 0.27
1.5

* Assumes trapped clay between ring stiffeners.


** Assumes trapped water between ring stiffeners.

16

Copyright 2005 Taylor & Francis Group plc, London, UK


5.5.4 Safety factor versions. One example of an existing problem-
The calculated safety factors at the final penetration specific finite element program is the BIFURC2D-
depth are in excess of 2.1 for all the anchors, using version incorporated in HVMCap (NGI, 2000), which
any of the definitions in 5.1.3. This agrees with none is tailor made to determine the capacity of suction
of the anchors experiencing plug heave failure. anchors in clay. The input to this program is simple
The calculated safety factors are lowest when cal- and the same as the input to a limit equilibrium pro-
culated as the ratio between the critical and required gram. Parameters that are not important for the capac-
underpressures (in the range 2.1 to 4.4), and highest ity, like sub-failure properties, can be set by the
when calculated as the material coefficient on the program. The user friendliness is at the sacrifice of
undrained shear strength below skirt tip level (in the flexibility with respect to more general use for prob-
range 7 to higher than 10). lems other than for suction caisson anchors.

6.1.2 Limit equilibrium and plastic limit analysis


6 CAPACITY methods
These models are more approximate than finite elem-
6.1 Calculation procedures ent models but are generally easier to use than general
Analysis tools to determine holding capacity of suc- finite element programs. The methods involve esti-
tion caissons can be classified as one of three general mating the ultimate capacity of plastic systems using
methods. These are (in order of detail) the finite assumed failure mechanisms. These mechanisms are
element method (advanced numerical analysis), limit typically based on a combination of experimental
equilibrium or plastic limit analysis methods (models observation, more rigorous numerical or analytical
involving soil failure mechanisms), and semi-empirical studies, and engineering judgment. These methods may
methods (highly simplified models of soil resistance also include the ability to incorporate complex geom-
including beam column models). This project has etry and soil strength variability and do not require
focused on the first two methods, since semi-empirical characterizing sub-failure behaviour. Disadvantages
methods are not industry practice and are not recom- are the approximate nature of the analysis and the dif-
mended to be used for important suction anchor ficulty of generalizing results, i.e. the need to cali-
applications. brate the models to experiment or more rigorous
The soil-caisson system is very stiff compared to analysis for specific structural configurations and soil
the mooring line (even for taut moorings) in the serv- profiles. The argument of calibration to experiment
iceability limit state, and its compliance is relatively is, however, also valid for the more advanced finite
insignificant with regard to the mooring system element models.
response. At failure, however, the anchor displacements In general there are two approaches that can be
may become large (of the order of 1 m or more) and taken using assumed mechanisms: the limit equilib-
this could have an effect on the anchor capacity, as the rium method and the plastic limit analysis method. In
displacements may reduce the effective penetration the limit equilibrium method, a failure mechanism is
depth and the depth of the load attachment point. On assumed, usually described in terms of one or more
the other hand, anchor displacements may lead to a geometric parameters (e.g. Andersen & Jostad 1999).
reduction in the load that could compensate some of The body force distribution, stress boundary condi-
the potential reduction in capacity. tions, and the stress or force distribution on failure sur-
faces are estimated, and a search is conducted to find
6.1.1 Finite element method the geometry that is closest to equilibrium conditions.
The finite element method will find the critical failure The plastic limit analysis method, such as incorp-
mechanism without prior user assumptions, provided orated within the software AGSPANC (AG 2002,
an appropriate constitutive model and a sufficiently Randolph & House 2002), uses an assumed failure
fine mesh are used. The finite element method also has 3D mechanism based on that for laterally loaded piles
many advantages including the ability to include com- (Murff & Hamilton 1993). The mechanism must satisfy
plex geometries, spatially varying soil properties, non- various kinematic constraints. For example, for a purely
linear constitutive behaviour with failure criterion, and cohesive undrained material the mechanism must sat-
partial consolidation under long term loading, to name isfy incompressibility; that is, the material must not
a few. Major disadvantages include the required spe- expand or contract, displacements normal to slip sur-
ciality knowledge of advanced numerical analysis, the faces must be continuous across the slip surface, dis-
large time investment to set up a model and the diffi- placement boundary conditions must be satisfied, etc.
culties of nursing a solution through initial loading to With the plastic limit analysis method, it is straight-
ultimate capacity. forward to include deforming regions as well as slip
The disadvantages can be eliminated by develop- surfaces. The plastic limit analysis calculations are car-
ing user friendly, robust, problem specific program ried out by equating the external work (or work rate)

17

Copyright 2005 Taylor & Francis Group plc, London, UK


done by boundary and body forces to the energy dis- any point along the length of the caisson. The depth of
sipation (or dissipation rate) expended in plastically the assumed axis of rotation and the relative vertical
deforming the soil during an assumed virtual displace- velocity to horizontal velocity at the mud line are
ment of the mechanism. This provides an equation treated as optimization parameters. The interaction
used for solving for the unknown force driving the equations are empirical fits to results from FEM stud-
failure (Calladine 1969). ies for idealized conditions.
In general both limit equilibrium and limit analysis The lateral bearing capacity factor (Nc) and the axial
methods give upper bound estimates of collapse load strength factor (Na) along the caisson (which are multi-
such that minimizing the collapse load with respect plied by local soil strength to obtain local horizontal
to the geometric parameters gives the best answer and vertical resistance) are calculated using an empir-
for the particular mechanism. However, the best ical equation based on a series of analyses using the
answer may or may not be close to the exact answer complete upper bound failure mechanism (see Murff
depending on the assumed mechanism. In the limit & Hamilton 1993, for details) and modifications by
equilibrium method the result will not be a true upper Han (2002). The factors are functions of soil depth to
bound if the mechanism does not satisfy kinematic diameter ratio, a non-dimensional shear strength
constraints. However, the limit equilibrium method parameter, and the soil-caisson roughness. Soil unit
has the advantage of being more natural to most weight is also included in the resistance calculation.
engineers because it is based on equilibrium solu- The vertical capacity is calculated by the -method,
tions. A discussion of these methods is provided in using the specified side shear strength for the side shear.
Chen (1975). The end bearing is calculated with an Nc value of 6 at
the mud line, increasing to 10 at three diameters
6.1.3 Semi-empirical methods depth, with the shear strength taken at the tip of the
These models are the most approximate. They are caisson.
labelled semi-empirical to suggest that they incorpor- The soil is modelled by a simple rigid-plastic,
ate the basic mechanics of a suction caisson loaded to undrained cohesive soil model (von Mises or Tresca).
failure, but depend on a set of empirical rules to rep- The strength is based on a linear profile with a finite
resent the soil resistance. These rules are typically strength at the mud line. Strengths for Cases C3 and
less general (or rigorous) than the methods discussed C4 were therefore approximated as:
above. For example, they do not explicitly incorporate Case C3: along caisson su  8  1.44 z kPa;
soil failure mechanisms, but instead represent the soil at tip su  50  2  (z
25) kPa
resistance as a load distribution varying along the Case C4: along caisson su  9  0.533 z kPa;
boundary at the soil-caisson interface. It is difficult to at tip su  15  2  (z
7.5) kPa.
generalize such a load distribution for a wide range of
soil profile types so a particular solution may apply, 6.2.2 Predictor P2
say, only to a normally consolidated strength profile. The calculations were made using AGSPANC (AG
Rules for constructing these distributions are typically 2002) with a limit equilibrium model based on the
based on a combination of experimental and analyt- (asymmetric) conical mechanism at shallow depths
ical results. In the so-called load distribution model, (Murff & Hamilton 1993) and either limiting hori-
the soil resistance is represented as a stress distribu- zontal pressure at depth (Randolph & Houlsby 1984)
tion on the caisson boundary. In the so-called beam- or a spherical base mechanism. The conical mech-
column model the soil is represented by uncoupled, anism incorporates a cos variation of soil displace-
non-linear, soil springs along the caisson boundary. ments about the vertical axis of symmetry. Base
capacity under vertical motion is estimated using a
conventional bearing capacity factor. The end-bearing
6.2 Calculation procedures used by the predictors
capacity is based on Nc  9, the shear strength at
6.2.1 Predictor P1 skirt tip level, and a characteristic shear strength equal
The calculations were made with an upper bound to the geometric mean of the triaxial extension and
plastic limit analysis that finds the location of the crit- average values. Plastic work within deforming
ical axis of rotation for the failure mechanism that regions of the soil is based on an anisotropic strength
gives the minimum upper bound capacity. The failure version of the von Mises model (Randolph 2000).
mechanism is modelled by a passive wedge near the Coupling between horizontal and vertical loads is
surface and with an active wedge for a no crack con- incorporated by calculating plastic work for each mode
dition; flow around (plane strain) below the wedge; (in terms of local resultant velocity at each point):
and a spherical failure surface at the tip. along the shaft in the flow region, based on finite elem-
The coupling between horizontal and vertical load is ent results of Aubeny et al. (2001) (reducing the lat-
incorporated by interaction equations that depend on eral limit pressure according to mobilization of vertical
the relative vertical to horizontal virtual velocities at shaft friction); at the base by a simple interaction

18

Copyright 2005 Taylor & Francis Group plc, London, UK


approach (root mean square of separate capacities), towards the active side and close the crack. If it is found
provided base failure due to horizontal loading is exter- that the crack will close, the capacity for the intact
nal to anchor. (i.e. no crack) case is used.
The version of AGSPANC used by Predictor P2 is
a later version of that used for some of the predictions 6.2.4 Predictor P4
by Predictor 4. In particular the later version allows An early version of the program AGSPANC was basic-
both forward and backward rotation of the caisson; any ally used for all load inclinations except the uplift
profile of soil strength (layered etc); and anisotropic case. The program is based on the plasticity model
shear strength. It also allows for interaction between initially proposed by Murff & Hamilton (1993) for
vertical and horizontal failure modes, partly through laterally loaded piles. A horizontal bearing capacity
appropriate summing of velocity components (in the factor of 11.7 (Bransby & Randolph 1998) was used
conical wedge region) and partly through modifying in the analyses to determine the horizontal pressure for
the separate horizontal or vertical limiting resistances the portion of the caisson where the soil flows hori-
in a manner consistent with results from 3D finite zontally around the caisson. To estimate the external
element analyses. skin friction on the caisson, the direct simple shear
Since AGSPANC is an upper bound approach, it will (DSS) shear strength was used along with the speci-
tend to over predict the actual capacity. Experience fied  coefficient of 0.65. Based on centrifuge tests
from finite element analysis suggests that this over pre- (Clukey & Phillips 2002) the reverse end bearing
diction is seldom more than 10%, and this would be capacity factor was estimated to be 10 for long and 8
allowed for in real designs. However, no such allowance for short caissons. This factor was used along with the
has been included in the present prediction exercise. DSS shear strength at the bottom (tip) of the caisson
to determine the reverse end bearing in uplift. The soil
6.2.3 Predictor P3 anisotropy factor used to make the predictions for the
The capacities of the anchors with optimal loading were lateral passive pressure was 80% of the DSS shear
calculated by a limiting equilibrium model, HVCap strength.
(Andersen and Jostad, 1999). The capacities of the For Case C1 the capacity in the interaction region
anchors with a specified load attachment point were was defined by centrifuge model tests and a plasticity
calculated by a finite element program, BIFURC-2D based solution calibrated to finite element results
(part of HVMCap, NGI, 2000), specially made for (Clukey et al. 2003). For non-optimal loading, the
analyses of suction anchors and incorporating auto- capacity was modified based on results in Aubeny et al.
matic mesh generation and default values for secondary (2003).
input (e.g. stress-strain relationship, solution algorithm
parameters).
Both models are plane models with the circular 6.3 Hypothetical cases
geometry transformed to a rectangle with the width
equal to the diameter and the same cross sectional area Four hypothetical cases with two different depth/
as the circular geometry. The 3D effect in the upper part diameter ratios (D/B  5 and D/B  1.5) in two dif-
with active and passive earth pressures is modelled by ferent soil profiles (one normally consolidated and one
side shear factors calibrated from full 3D finite elem- lightly overconsolidated clay) were defined. The geom-
ent analyses. Side shear factors of 0.5 between soil etry and loads are specified in Table 8, and the soil
and structure and 0.6 in the soil were used in the pre- properties are defined in Table 5.
sent calculations. In the deeper part, where the soil
flows around the anchor, the 3D effect is modelled by Table 8. Hypothetical capacity cases.
coupling between vertical shear stress and roughness
found from 3D finite element analyses. The finite elem- CASE C1 C2 C3 C4
ent model has interface elements for modelling
reduced interface strength. Both models handle Diameter (m): 5 5 5 5
anisotropic shear strengths. Penetr. depth (m): 25 7.5 25 7.5
The coupling between horizontal and vertical loads Depth/diameter: 5 1.5 5 1.5
is automatically achieved in the finite element model. Weight (kN): 1100 300 1100 300
In the limiting equilibrium model, coupling is mod- Load inclinations**:   0, 22.5, 30, 45, 60 and 90
Specif d load 12.5&20 2.5&6 12.5&20 2.5&6
elled by optimization of mobilized interface shear
point* (m):
strengths until maximum capacity is achieved. Crack at active side: No No With & without
In cases with specified load attachment point and OCR: 1.0 1.0 1.6 1.6
the potential for an open crack at the active side, the
displacement pattern from the finite element analysis * Cases with 30 load inclination.
is used to determine whether the anchor top will move ** Cases with optimal load point.

19

Copyright 2005 Taylor & Francis Group plc, London, UK


The following analyses were made for each of the would open and be assigned zero shear strength if the
4 cases in Table 1: calculated total normal stress reduced to zero.
OTRC modelled the anchor as a solid rigid cylinder.
capacity and depth of load attachment point for
The reduced soil strength along the skirt wall was mod-
optimally loaded anchors with load inclinations at
elled by a narrow (approximately 0.2 m thick) zone of
pad eye ranging from 0 to 90;
elements adjacent to the anchor wall having a strength
capacity for anchors with 30 load inclination from
equal to 0.65 times the undrained DSS strength.
the horizontal at pad eye and specified load attach-
Variations in soil strength with depth were modelled by
ment point beneath seabed at anchor wall of 12.5 m
assigning the estimated strength at the centre of the ele-
and 20 m for anchor with D/B  5 and 2.5 m and
ment to the entire element. Failure was defined as the
6 m for anchor with D/B  1.5.
load where the slope of the load-displacement curve
The predictions also included an evaluation of whether was zero.
a crack will form at the windward (active) side of the
anchors in the lightly overconsolidated clay, and 6.4.2 Soil models
the effect such a crack will have on the capacity and NGI used an elasto-plastic soil model with an approxi-
the depth of the optimal load attachment point. mated Tresca yield criterion and anisotropic undrained
The cases were analysed by 3D finite element shear strength.
analyses by the 3 groups NGI, COFS and OTRC. The COFS and OTRC used a linearly elastic-perfectly
results of the 3D finite element analyses were used to plastic soil model with a von Mises failure criterion,
check the quality of the predictions with the simpli- with the von Mises parameter taken as the undrained
fied calculation procedures used in more conven- DSS shear strength. The shear strength under triaxial
tional practical design analyses. conditions, where the intermediate principle stress is
equal to either the major or the minor principal stress,
will then be 0.866 times the undrained DSS shear
6.4 3D finite element analyses
strength. This strength model thus leads to an average
6.4.1 Finite element models shear strength (average of triaxial compression, DSS
NGI used the program BIFURC-3D (NGI 1999) for and triaxial extension) of 0.91 times the DSS shear
their analyses. COFS and OTRC used the program strength.
ABAQUS (HKS 2002). All used small displacement
formulation, with no mesh updating. 6.4.3 Results of 3D finite element analyses
NGI used 20 node isoparametric brick elements with The results of the 3D finite element analyses from the
(2  2  2) reduced Gaussian integration. The cais- 3 groups are presented by the curves in Figure 5af.
son is modelled by a cylindrical body with stiff elas- The capacity and the location of the optimal load
tic elements (a shear modulus of about 500 times the attachment point for anchors with optimal loading are
elastic shear modulus of the surrounding soil). 16 node given in the two figures to the left. The capacity for
isoparametric zero thickness interface elements are tests with specified load attachment point and 30 load
included between the brick elements at a radius of 2.5 m inclination is given in the figure to the right. The results
and horizontally at skirt tip level. The interface elem- show that there is generally an excellent agreement
ents along the skirt wall are given a shear strength of between the capacities calculated by the three groups.
65% of the DSS soil strength. Failure was defined as This is true for vertical, horizontal and intermediate
the load where the slope of the load displacement (coupling) capacities for optimally loaded anchors, for
curve was zero. depth of optimal load attachment point, and for anchors
COFS and OTRC used 3D 8 node hybrid brick with load attachment point above and below the opti-
elements (C3D8H) with full integration. mal. The difference in capacity is generally less than
COFS modelled the anchor as a rigid cylindrical 3%, with a tendency for COFS to be lower than NGI. A
shell and a top cap. The reduced shear strength along difference greater than this occurs for Case C3 without
the skirt wall was modelled by a 50 mm thin zone of ele- crack. The vertical capacity calculated by OTRC is in
ments adjacent to the outside anchor wall. Failure was this case about 11% higher than by NGI and COFS.
defined as a displacement of 1 m, by which stage all The good agreement between the ABAQUS (used by
load-displacement curves had reached a plateau except COFS and OTRC) and the BIFURC 3D (used by NGI)
those for the longer caissons failing vertically; the end- analyses is somewhat surprising, since the ABAQUS
bearing capacity of the longer caissons requires greater analyses use a soil model that gives an average shear
displacement to mobilize fully, as it involves radially strength that is 0.91 of the actual one. Reasons why
inward soil movement rather than a rigid-plastic mech- this does not give a larger difference may be that the
anism. In the analyses with a crack at the active side, the capacity is governed more by the DSS shear strength
crack was modelled by introducing a contact surface than by the average shear strength, and that the lower
with Mohr-Coulomb properties. The contact surface order elements used by COFS and OTRC compared

20

Copyright 2005 Taylor & Francis Group plc, London, UK


with NGI, together with the different modelling of the The 3D finite element analyses of cases with a crack
interface between the soil and the anchor wall, may at the active side were only performed by COFS, and
give more overshoot in the COFS and OTRC analyses comparisons cannot therefore be made. Modelling of
than in the NGI analyses. The reason why OTRC cal- the crack was somewhat crude, with no attempt to allow
culates about 11% higher vertical capacity than COFS for potential suction. Thus, when the anchor was
and NGI for Case C3 is not clear, but may possibly allowed to rotate backwards, cracks could form near
have to do with the finite element discretization, since the skirt tips, even though there was no direct commu-
OTRC use fewer elements below the caisson than NGI nication with the soil surface. While the failure under
and COFS in Case C3. purely vertical loading was unaffected by the poten-

Optimal load attachment point Optimal load attachment point Specified load attachment point.
30 load inclination.
Load Attachm. Point at Centerl., zcl/D

12000 0.8

Relative Capacity, Pf / Pf, optimal


1.0
10000 0.7
Vertical Load (kN)

0.6 0.8
8000
0.5
0.6
6000 0.4

4000 0.3 0.4


0.2
2000 0.2
0.1
0 0.0 0.0
0 5000 10000 15000 20000 25000 0.0 22.5 45.0 67.5 90.0 0.5 0.6 0.7 0.8 0.9 1.0
Horizontal Load (kN) Load Inclination (degrees) Load Attachm. Point at Centerl., zcl/D

(a) Case C1. Depth/diameter =5. Normally consolidated clay.

3000 0.8
Load Attachm. Point at Centerl.,zcl /D

1.0
Relative Capacity, Pf / Pf, optimal

0.7
2500
0.6
Vertical Load (kN)

0.8
2000
P1 0.5
0.6
1500 P2 0.4
P3
0.3 0.4
1000 P4
NGI 3DFE 0.2
500 OTRC 3D FE 0.2
0.1
COFS 3D FE
0 0.0 0.0
0 500 1000 1500 2000 0 22.5 45 67.5 90 0.5 0.6 0.7 0.8 0.9 1.0
Horizontal Load (kN) Load Inclination (degrees) Load Attachm. Point at Centerl., zcl/D

(b) Case C2. Depth/diameter =1.5. Normally consolidated clay.


Load Attachm. Point at Centerl., zcl/D

20000 0.8
Relative Capacity, Pf / Pf, optimal

1.0
0.7
15000 0.6
Vertical Load (kN)

0.8
0.5
0.6
10000 0.4
0.3 0.4

5000 0.2
0.2
0.1
0 0.0 0.0
0 10000 20000 30000 40000 0 22.5 45 67.5 90 0.5 0.6 0.7 0.8 0.9 1.0
Horizontal Load (kN) Load Inclination (degrees) Load Attachm. Point at Centerl., zcl/D

(c) Case C3. No crack. Depth/diameter =5. Lightly overconsolidated clay, OCR=1.6.

Figure 5ac. Capacity calculated by 3D finite element analyses and simplified prediction methods.

21

Copyright 2005 Taylor & Francis Group plc, London, UK


Optimal load attachment point Optimal load attachment point Specified load attachment point.
30 load inclination.

Load Attachm. Point at Centerl., zcl/D


18000 0.8

Relative Capacity, Pf / Pf, optimal


16000 0.7 1.0

14000 0.6 0.8


Vertical Load (kN)

12000 0.5
10000 0.6
0.4
8000
0.3 0.4
6000
0.2
4000 0.2
0.1
2000
0.0 0.0
0
0.0 22.5 45.0 67.5 90.0 0.5 0.6 0.7 0.8 0.9 1.0
0 10000 20000 30000 40000
Load Inclination (degrees) Load Attachm. Point at Centerl., zcl/D
Horizontal Load (kN)

(d) Case C3. With crack. Depth/diameter =5. Lightly overconsolidated clay. OCR =1.6.
Load Attachm. Point at Centerl., zcl/D

5000 0.9

Relative Capacity, Pf / Pf, optimal


0.8 1.0
4000 0.7
Vertical Load (kN)

0.8
0.6
3000
0.5 0.6
0.4
2000
0.3 0.4

1000 0.2
0.2
0.1
0 0.0 0.0
0 1000 2000 3000 4000 5000 0.0 22.5 45.0 67.5 90.0 0.5 0.6 0.7 0.8 0.9 1.0
Horizontal Load (kN) Load Inclination (degrees) Load Attachm. Point at Centerl., zcl/D

(e) Case 4. No crack. Depth/diameter =1.5. Lightly overconsolidated clay. OCR=1.6.


Load Attachm. Point at Centerl., zcl/D

4000 0.8
Relative Capacity, Pf / Pf, optimal

0.7 1.0
3500
3000 0.6
Vertical Load (kN)

0.8
2500 0.5
0.6
2000 0.4

1500 0.3 0.4


1000 0.2
0.2
500 0.1

0 0.0 0.0
0 1000 2000 3000 4000 0 22.5 45 67.5 90 0.5 0.6 0.7 0.8 0.9 1.0
Horizontal Load (kN) Load Inclination (degrees) Load Attachm. Point at Centerl., zcl/D

(f) Case 4. With crack. Depth/diameter =1.5. Lightly overconsolidated clay. OCR=1.6.

Figure 5df. Capacity calculated by 3D finite element analyses and simplified prediction methods.

6.5 Comparison between simplified prediction


tial crack, it was found that if a crack was first opened
methods and 3D finite element analyses
by horizontal movement, the vertical capacity could
be reduced. The actual calculation results for the cases The simplified calculation procedures used by the pre-
with a crack gave reductions between 10 to 17%, with dictors were checked by comparing the results with the
the exception of pure vertical loading where the 3D finite element analyses. The comparison is shown
reduction would be less. A conservative failure envel- in Figure 5af and the differences are quantified in
ope in the failure interaction diagram has been sug- Table 9. The results from the simplified prediction
gested using a reduction factor of 0.83 on the failure methods are shown with open symbols in Figure 5.
envelope for the corresponding case with no crack. The results are summarized below.

22

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 9. Ratio between capacities calculated by simplified methods and 3D finite element analyses.

C3 no C3 with C4 no C4 with
CASE Predictor C1 C2 crack crack crack crack

Anchor D/B: 5 1.5 5 5 1.5 1.5


OCR: 1.0 1.0 1.6 1.6 1.6 1.6
Vertical capacity P1 0.98 0.83 0.98 1.20 0.86 0.99
P2 0.88 0.85 0.87 0.85 0.88 0.90
P3 0.94 0.93 0.93 1.00 0.97 1.03
P4 0.98 0.82 0.89 0.89 0.67 0.68
Horizontal capacity P1 0.95 1.03 0.95 1.09 1.03 0.99
P2 0.98 1.19 0.97 1.11 1.14 0.99
P3 0.95 0.89 0.95 1.03 0.89 0.87
P4 0.97 0.98 0.98 1.03 0.85 0.85
Capacity at intermediate : 22.5/30 45/60 22.5/30 22.5/30 45/60 45/60
load inclination,  P1
/0.97 1.01/0.98
/1.02
/1.18 1.01/0.89 1.06/1.06
P2
/0.95 1.14/1.02 0.96/0.95 1.03/0.91 1.07/0.91 1.01/0.95
P3
/0.97 0.91/1.00 0.97/1.00 1.05/0.95 0.98/0.98 0.96/1.03
P4
/0.88 1.01/0.94 0.98/0.98 1.03/0.85 0.77/0.69 0.84/0.72
Depth of optimum load P1 1.01 1.01 1.02 1.02 1.00 1.02
attachment point P2 1.06 1.01 1.02 1.00 1.00 1.02
P3 1.01 1.01 1.02 1.00 1.00 1.00
P4 1.03 1.03 1.02 1.00 1.08 1.05
Attachment point P1 0.95 1.00 0.93 (1.09) 0.96 0.89
below optimum P2 0.91 1.33 0.99 (1.09) 1.06 1.13
P3 0.98 1.00 1.00 (1.20) 0.95 1.04
P4 0.99 1.55 1.16 (1.00) 1.48 1.38
Attachment point P1 1.04 0.92 1.10 1.18 0.96 0.99
above optimum P2 1.06 1.10 1.08 1.01 1.14 0.99
P3 0.90 0.95 0.98 0.91 0.92 1.01
P4 0.78 0.89 1.03 1.05 0.75 0.80

The shadings indicate the following: :1020% difference : 20% difference.

6.5.1 Vertical capacity sion, DSS and extension), and it is not recommended
The best and most consistent agreement for the vari- to weight the end bearing capacity towards the exten-
ous cases is obtained when the inverse bearing cap- sion strength.
acity below skirt tip is calculated with a bearing capacity
factor varying from Nc  6.2 at the surface to Nc  9 6.5.2 Horizontal capacity
at depths greater than 4.5 times the diameter (e.g. The calculations with the plane models where 3D
Brinch Hansen, 1970) and a shear strength determined effects are taken into account by side shear factors give
at a depth of 0.25 times the diameter below the skirt capacities 5% to 10% on the low side (figures to
tip elevation. The resulting capacities tend to be slightly the left in Figure 5 and Table 9). There may thus be a
on the low side. This empirical combination of bear- potential for some increase in the side shear factors.
ing capacity factor and reference depth includes the The calculations based on limit analyses using
effect that the outside skirt wall force may have on the Murff and Hamilton (1993) give good results for the
bearing capacity. The effect of the skirt wall friction long anchor, and also for the short anchor using the
may be one reason why the bearing capacity factor is fitted function for the lateral bearing capacity factor,
higher than the one from Houlsby & Martin (2003), Np. For the short anchor, the rigorous upper bound
which was derived for smooth-sided caissons. If the capacity may be up to 19% too high, of which 5% is
shear strength profile deviates from a linear variation due to allowance for shear strength anisotropy.
with depth, one should be careful about taking the
shear strength at a depth of 0.25 times the diameter 6.5.3 Inclined loading capacity
and use a more conservative strength at a different Good agreement in the shape of the failure envelope
reference depth. was generally obtained by the simplified methods,
The end bearing capacity seems to be best related with most discrepancies arising from errors in predic-
to the average shear strength (average of compres- tion of the uniaxial vertical or horizontal capacities

23

Copyright 2005 Taylor & Francis Group plc, London, UK


(figures to the left in Figure 5 and Table 9). There was most likely cases would be anchors with 0 and 22.5
a slight tendency for the plastic limit analysis load inclination.
approaches to underpredict the reduction in vertical
capacity due to increasing horizontal load for the
6.5.8 Discussion
shorter caissons, where the 3D finite element results
The 3D finite element analyses have given an excel-
showed greater curvature in the failure envelope.
lent check of the simplified capacity prediction meth-
ods. However, there are factors that have been
6.5.4 Depth of optimum load attachment point
assumed and specified in the 3D finite element analy-
The calculated depth of the optimal load attachment
ses, and that are thus not checked by this exercise.
point typically deviates from the 3D finite element
These factors are:
analyses by less than 10% for load angles of 45
(middle figures in Figure 5 and Table 9). A deviation Strain softening. If the clay along the skirt wall
of 10% may in a few cases result in 6 to 8% over exhibits strain softening, the peak strength along the
estimation of capacity. skirt wall may be mobilized before the capacity
below the skirt tip level, and the resulting anchor
6.5.5 Attachment point below optimum capacity may be smaller than the sum of the two
The calculations with a plane finite element model components. It is believed, however, that the clay
where 3D effects are taken into account by side shear along the skirt wall exhibits relatively little strain
factors give capacities in good agreement with the 3D softening, at least soon after installation, since it
finite element capacities (within less than 5%) (fig- has been remoulded during installation.
ures to the right in Figure 5 and Table 9). One exception Strain softening may also influence the capacity
is the deep anchor with a crack in lightly overconsol- of the clay below skirt tip level, since the compres-
idated clay, but there is also some uncertainty in the sion, DSS and extension strengths may be mobil-
reference 3D finite element analyses for that case. ized at different shear strains. In cases with strain
The calculations based on Murff & Hamilton (1993) softening clays, the shear strengths may be deter-
using the fitted function for the lateral bearing cap- mined as the shear stress at the shear strain where
acity factor, Np, also give good agreement with the finite the average of the shear stresses from the 3 test
element results. Rigorous use of the upper bound types has its maximum.
approach for the shallow anchor in normally consoli- Set-up. The shear strength along the outside skirt
dated soil tended to underestimate the reduction in wall after installation and subsequent set-up was
capacity for loading below the optimum point, and in specified as 0.65 times the initial DSS shear strength
the worst case gave a significant (33%) overpredic- in this exercise. This exercise did not provide any
tion of capacity. check of this assumption.
Material model. In the 3D finite element analyses,
6.5.6 Attachment point above optimum the soil was modelled by an elasto-plastic material
The calculations with a plane finite element model model based on an anisotropic Tresca yield criter-
where 3D effects are taken into account by side shear ion (NGI) and a von Mises yield criterion (COFS
factors give capacities in reasonably good agreement and OTRC). The strengths for the anisotropic Tresca
with the 3D finite element capacities (within less than model were based on results from undrained tri-

10%) (figures to the right in Figure 5 and Table 9). axial compression and extension tests and DSS tests.
The agreement of the capacities calculated based The anisotropic undrained shear strength for a gen-
on Murff & Hamilton (1993) mostly gave reasonably eral 3D situation is obtained by extrapolating from
good agreement with the finite element results, these strengths. The actual shear strength may there-
although with a tendency to overpredict capacities fore not be modelled correctly for some stress paths.
(underpredicting the reduction due to forward rota- This difficulty is even more pronounced using the
tion of the anchor). von Mises criterion with a shear strength based on
the DSS strength. The shear strengths in compres-
6.5.7 Potential for crack at the active side sion and extension stress paths will in this case be
The predictors are generally uncertain about whether underestimated by 39 and
8%, respectively. For
a crack will form at the active (windward) side of the situations with combined static and cyclic loading
anchor in the lightly overconsolidated clay. Predictors the uncertainties related to the soil model may be
1 and 3 point out that there are large uncertainties in even larger.
estimates with respect to potential for cracking, and Design shear strength profile. The interpretation of
would use conservative assumptions in the capacity in situ and laboratory test data to establish a design
calculations. Predictor 4 believes that cracking will not shear strength profile may often be one of the
occur for the relatively low overconsolidation ratios major uncertainties in anchor design. The shear
in the cases herein, and if cracking should occur, the strength profile was specified in this exercise,

24

Copyright 2005 Taylor & Francis Group plc, London, UK


which therefore did not provide a check of the develop or not along the active (windward) side of a
determination of the design strength profile. suction anchor. The consequence of assuming a crack
Effect of large vertical displacements at failure. or not can in many cases be important for the calcu-
The analyses in this project do not account for lated anchor capacity. It is recommended that theoret-
effect of geometry changes due to large displace- ical analyses are performed to better understand the
ments. At failure the anchor displacements may mechanism for cracking and to develop a model to
become large (of the order of 1 m or more), which predict whether a crack will form or not. This theoret-
could have an effect on the anchor capacity, as the ical work should be checked against existing and new
displacements may reduce the effective penetration model tests with focus on cracking.
depth and the depth of the load attachment point. Set-up. The set-up factor is based on theoretical
On the other hand, anchor displacements may lead analyses, and it is recommended to collect relevant
to a reduction in the mooring load that could com- existing experimental data (model tests and prototype
pensate some of the potential reduction in capacity. data) and to consider new model tests on natural clay
with focus on set-up. Further studies should also assess
the time-scale of consolidation and to what extent the
7 RESEARCH NEEDS set-up factor is affected by suction installation, com-
pared with self-weight penetration.
During the project several items that may deserve fur- Effect of large displacements at failure. The import-
ther study were identified. These are discussed below ance of this topic can be studied by exploring the
for the two main design aspects: installation and effect that large displacements can have on the anchor
operational capacity. capacity by reducing the effective penetration depth
and the depth of the load attachment point. The study
should be accompanied by a study to see if large anchor
7.1 Installation
displacements may lead to some compensating reduc-
Safety against soil plug failure during installation. tion in the mooring line load.
An agreed basis should be decided for expressing the Strain softening. The effect of strain softening
safety factor against plug failure during suction on the anchor capacity may be studied by finite elem-
installation, whether defined as the ratio between the ent analyses with a strain softening soil model.
critical and required underpressures or as material Laboratory DSS tests on consolidated, remoulded
coefficients on the undrained shear strength below clay may be performed to see how significant the strain
skirt tip level and along the outside skirt wall. softening can be for elements along the anchor wall.
Acceptable design values of the safety factor should Safety factor and characteristic shear strength.
also be determined, taking account of whether the Since there is no consensus on the safety factors to be
characteristic shear strength is determined as a best used, this should be evaluated. The numerical value
estimate mean strength or upper and lower bounds. should depend on the definition of the characteristic
Remoulded shear strength. The penetration resist- shear strength (average versus upper or lower bound),
ance and the required underpressure depend strongly and this should also be considered in the evaluation.
on the remoulded shear strength, and it should be
investigated whether an appropriate remoulded shear
strength can be determined directly from the CPT 8 ADDITIONAL DESIGN CONSIDERATIONS
sleeve friction measurement or by other in situ devices
(e.g. cyclic T-bar tests). In addition to the design methods and the findings
Bearing capacity factors, Nc. Bearing capacity fac- discussed above, there are a number of considerations
tors should be determined for different geometries and that the anchor designer needs to make. Some of these
boundary conditions of internal stiffeners and exter- are discussed below.
nal protuberances to resolve the differences in bear- Shear strength for penetration. For penetration
ing capacity factors used by the various predictors. analyses, it is recommended to determine the strength
Installation friction in carbonate soils. Further stud- along the skirt wall by considering both the directly
ies should be made to understand the interface shear measured remoulded shear strength and the intact
strength during installation in carbonate rich soils, both shear strength divided by the sensitivity. If available,
in terms of the radial effective stress and the appro- backfigured values of the interface strength from
priate interface friction angle. actual field installations is another, and perhaps the
most accurate, way to determine the interface strength.
Tilt and misorientation. An anchor will normally
7.2 Capacity
have some tilt and mis-orientation after installation.
Potential for crack at the active side. There are con- Tilt and mis-orientation will reduce the anchor cap-
siderable uncertainties related to whether a crack will acity and tolerances must be specified and accounted

25

Copyright 2005 Taylor & Francis Group plc, London, UK


for in design. The installation costs may depend on at the interface between the clay and the skirt wall,
the tolerances that are specified, and too strict limits both during installation and operation. If the skirt
may not be economical. wall for some reason should not be rusty or is painted
Plug heave. The soil plug inside the anchor may or treated in other ways that reduce the wall friction, a
rise due to soil displaced at the anchor tip moving into reduced interface strength must be considered. The
the anchor during skirt penetration. Inside stiffeners, interface strength for a specific surface can be deter-
diameter changes and trapped water will also influ- mined by ring shear tests with the actual wall surface
ence the soil heave. Shear strains in the soil below and modelled in the test. The Girassol anchors are one
outside the anchor due to applied underpressure dur- case where the interface strength was influenced by
ing penetration may also contribute, depending on the paint (e.g. Colliat & Dendani 2002).
degree of mobilized inverse bearing capacity during Set-up after installation. The reduced interface
penetration. In cases with sloping or uneven seafloor, strength along the outside skirt wall was in the exam-
the potential for contact at a point higher than at the ples in this project specified as   suDSS, with   0.65.
centre of the anchor should also be considered. The The set-up value, , is not a constant, however, but may
plug heave due to the displaced soil at skirt tip level depend on factors like clay plasticity, overconsolidation
can possibly be reduced by tapering the skirt tips. ratio, sensitivity and time after anchor installation (e.g.
Outside diameter variations. The outside diameter Andersen & Jostad 2002). The interface shear strength
may vary due to variations in wall thickness or exter- may also vary between the zone where penetration was
nal ring stiffeners. It is uncertain how the clay will by self-weight and the deeper zone where penetration
deform along the outside wall above an extended diam- was by suction, and it will also vary with time.
eter, but the outside set-up factor may be reduced, or Inside skirt interface strength. The strength along
the clay may not even deform back to the wall above the the inside skirt wall is not mobilized under undrained
extension. To avoid this uncertainty it is recommended loading if the top cap is sealed. If the top cap is not
to avoid variations in outside anchor diameter. sealed, however, the shear strength along the inside
Chain configuration. The load and the load inclin- skirt wall can be an important contribution to the
ation are normally given at seabed elevation, and it is capacity. As for the outside skirt wall, there is disturb-
necessary to account for the inverse catenary shape of ance during installation that may reduce the strength
the mooring line in the soil between the seabed and along the wall after installation, especially if there are
the attachment point at the anchor wall. The load inclin- inside stiffeners or diameter variations (e.g. Andersen
ation from horizontal will be higher and the load will & Jostad 2004). Compartments with trapped water
be smaller at the anchor wall than at the seabed. between ring stiffeners will also be potential drainage
Optimal loading. Highest anchor capacity is channels that need to be considered in cases with sig-
obtained if the failure mode is translational displace- nificant long term loads.
ments with no rotation. A translational failure mode Retrieval. Anchors may have to be extracted at the
can be achieved by applying the load at a depth that end of their service life or due to installation prob-
gives no resultant moment loading. The depth of the lems. Set-up factors must then be established for both
optimal load attachment point at the anchor wall will outside and inside interface skirt strength for times
depend on the shear strength profile, the shear strength different from the time used in the capacity analyses.
at the outside skirt wall, the load inclination, and the It should also be noted that set-up factors that are con-
depth to diameter ratio of the anchor. The imaginary servative (low values) for capacity will not be conser-
optimal load attachment point at the centreline of the vative for retrieval.
anchor is typically about 0.67 to 0.7 of the anchor Cyclic loading. The anchors must be designed for the
penetration depth in clay with shear strength increas- expected load history, which is likely to include cyclic
ing linearly with depth from zero at the clay surface. loading. The capacity under cyclic loading may be
Crack at active side. A crack may form at the active higher or lower than the monotonic capacity, depending
(windward) side of the anchor. Since there are uncer- on the cyclic load history composition and the cyclic
tainties about how to predict whether a crack will form load period. Long period cyclic loading will give more
or not unless the clay is soft and with essentially zero degradation than 10s load periods that are often used in
strength intercept at the seabed, consideration should laboratory tests, and the laboratory test program should
be given to placing the load attachment point far be designed to model the actual load period.
enough below the optimal attachment point for the The capacity analyses in this project have not
anchor top to move backwards (away from the load included cyclic loading per see. Cyclic capacity ana-
direction) during loading to prevent a gap forming. lyses may be performed by the procedure proposed by
Note, however, that for short anchors this may lead to Andersen & Lauritzsen (1988).
a significant reduction in capacity. Long term loads. In the cases of long term loads
Skirt wall roughness. The skirt wall is normally the undrained shear strengths should be reduced to
rusty with a roughness high enough to prevent sliding account for creep effects.

26

Copyright 2005 Taylor & Francis Group plc, London, UK


In cases where significant loads act for months, it should be greater than unity for calculating the inverse
is also necessary to consider whether drained condi- bearing capacity of the soil plug at skirt tip level, and
tions may develop, and the extent to which full base smaller than unity for calculating the interface fric-
suction can be maintained. The possibility of drainage tion and other resistances (for example at the padeye)
channels between inside skirt stiffeners, above inside along the outside skirt.
stiffeners and extended skirt wall thicknesses, and Monitoring during installation. The installation of
along open cracks outside the anchor at the active side a suction anchor must be monitored to make sure that
of the anchor must be considered. the installation proceeds as expected and that the
Where suction caissons are subjected to long-term anchor is installed as designed. The measurements
tensile loading, the effect of pore pressure redistribu- should include penetration depth, applied underpres-
tion and swelling should be considered, since this may sure, penetration rate, plug heave, tilt, and orientation.
lead to reduction in effective stresses and undrained The lowering speed at time of touch-down at the sea
shear strengths, and hence the capacity under transient bottom can be critical, in particular for large caissons
wave loading. with relatively limited top venting for water evacu-
Clukey et al. (2004) show how the capacity of a ation. It is strongly recommended that underpressure
typical suction caisson under sustained loading may during installation should be monitored inside the
be only 70% of the undrained capacity. anchor, rather than within the pumping unit, in order
Closed vs. open top. The capacity examples in this to avoid errors due to pipe losses and Venturi effects.
project assumed that the anchor is closed at the top. Simple preliminary capacity estimate. A simple
For anchors with essentially horizontal loading, a estimate of the capacity of anchors with predomin-
sealed top is not essential for the capacity; the top antly vertical loading may be obtained by calculat-
part can therefore be removed after installation and ing the vertical anchor capacity as the sum of the
reused for subsequent anchors, as was the case for the integrated interface shear strength along the external
Nkossa anchors (Colliat et al. 1995). For anchors with skirt wall, the inverse bearing capacity at skirt tip
higher load inclinations from the horizontal, the cap- level, and the submerged caisson weight. The result-
acity will be limited by the vertical failure load, ant capacity can be calculated as Pv/sin , where Pv is
which, for open-topped caissons, will be limited to the vertical anchor capacity and  is the load inclina-
the contribution from the shaft friction along the out- tion from the horizontal. These calculations can be
side and inside skirt walls and the submerged caisson done by hand, and may provide an early ball-
weight. park capacity. However, the calculation of the result-
Sand layers. The present study has focused on ant capacity neglects the coupling between vertical
anchors in clay. Sand layers will increase the skirt tip and horizontal capacities, and will overestimate the
resistance during penetration, and the underpressure resultant capacity except for load inclinations close to
required for penetration will increase. If the sand vertical.
layer is continuous the anchor must be penetrated at a For the cases studied in this project, the simple
rate high enough to limit the seepage flow through the estimate gives excellent agreement for the deep anchors
sand layer below the anchor, as excessive seepage for load angles greater than about 45 from the hori-
would allow the clay plug above the sand layer to move zontal, acceptable agreement for load angles of 30
up inside the anchor and cause unacceptable plug (5% too high), and significantly too high capacities
heave. Sand layers beneath skirt tip level may also sig- for load angles of 22.5 (20% too high). For the
nificantly reduce the inverse bearing capacity of the shallow anchors the simple estimate significantly
anchor, as drainage or pore pressure redistribution overpredicts the capacity already at load angles of 60
within the sand layer may allow dissipation of the from the horizontal.
suction.
Safety factors. Safety factors for capacity should
be specified with due consideration to the definition
9 SUMMARY AND CONCLUSIONS
of the characteristic shear strength (average versus
upper or lower bound) and failure (limiting displace-
9.1 Collected information
ments to a percentage of diameter versus ultimate
failure with excessive displacements). The conse- Suction anchors have been used at about 50 locations
quence of a failure should also be considered. in water depths to nearly 2000 m during the last decade.
Safety factor for installation should be related to Thanks to the willingness of the industry to pro-
avoiding large plug heave inside the anchor. Rather vide proprietary data, a good set of data on suction
than defining the safety factor as the ratio between anchor applications, experiments and prediction meth-
critical and required underpressure, consideration ods has been compiled. This includes detailed instal-
should be given to defining it in terms of material lation data for 16 of the 50 locations where suction
coefficients on the shear strength. Such coefficients anchors have been identified.

27

Copyright 2005 Taylor & Francis Group plc, London, UK


9.2 Installation plug at skirt tip level, plus internal resistance from
the inside skirt wall and inside stiffeners. Different
The penetration resistance of suction anchors is cal-
approaches are possible in calculating the safety fac-
culated as the sum of the integrated interface shear
tor with respect to internal plug heave. However,
strength along outer and inner skirt walls and any
since internal shaft resistance contributes both to
internal plate and ring stiffeners, and the end bearing
the required underpressure for penetration and also
resistance of skirt tips, plate stiffeners, ring stiffeners
internal plug stability, a suggested design approach is
and any changes in anchor diameter.
to evaluate external resistance using a material coeffi-
In the case of penetration by underpressure, the
cient less than unity (representing worst case estimate
required underpressure is calculated as the penetra-
of caisson resistance, and hence of required under-
tion resistance minus the submerged anchor weight,
pressure) together with a material coefficient greater
divided by the inside cross section area beneath the
than unity for the soil strength at the caisson tip
top lid.
(representing worst case estimate of plug base resist-
As for pile design, both total and effective stress
ance). Maximum recommended penetration depth
methods may be used to estimate the shear strength
would then be determined as the depth where the
along the inside and outside of the anchor. Both
material (safety) coefficient is around 1.5 (or 0.67 as
approaches require knowledge of the remoulded
appropriate).
strength of the soil. The effective stress approach also
Excessive plug heave may prevent full installation
requires the interface friction angle between caisson
of a suction anchor. Provided soil plug failure is not
and clay and the effective normal stress along the skirt
approached, plug heave may be calculated on the
wall. If non-standard wall surfaces are used, such as
basis of the soil displaced by some proportion of
painted walls, potential reductions in the interface
the caisson wall (between 50 and 100% assumed by
strength or friction angle must be taken into account.
the various predictors) and by internal stiffeners. The
Bearing resistance on the skirt tip, external protu-
largest degree of uncertainty among the predictors
berances, or internal stiffeners is based on bearing
was the assumed free-standing height of soil plug
capacity factors, Nc multiplied by the local shear
above internal ring stiffeners, and the extent to which
strength.
water may be trapped between the soil plug and the
All four predictors used a total stress approach for
caisson wall.
estimating penetration resistance, while only one pre-
dictor also used an effective stress approach for com-
parison. Either method could be adjusted to give 9.3 Capacity
reasonable agreement with measured data, although The capacities of suction anchors were predicted with
such agreement involved hindsight that would not plane limiting equilibrium methods, plastic limit ana-
necessarily be available during a design. Significant lyses methods, and tailor made plane finite element
variations among the predictors were associated with: methods. More approximate semi-empirical methods
were not applied since they are not industry practice
handling of soil flow around internal stiffeners,
and are not recommended to be used for future import-
particularly in respect of internal friction above
ring stiffeners; ant suction anchor applications.
The results of reference 3D finite element analyses
bearing capacity factors appropriate for internal
from three organizations generally gave excellent
ring stiffeners and external diameter changes;
agreement, and provided a good basis for checking
evaluation of interface friction.
the quality of the more simplified capacity prediction
The first of these contributed to the highest degree methods.
of variability in prediction of performance, particu- The comparison of results from simplified predic-
larly for lightly overconsolidated soil. A consensus tion methods and 3D finite element analyses showed
view is that the internal soil plug may not flow back that reverse end bearing can be calculated with the same
around internal ring stiffeners for a significant plug bearing capacity factor, Nc, as for downward loading.
height, and that when flow-back occurs it may trap an The reverse end bearing seems best related to the aver-
interface zone of water or high water content soft clay age shear strength (average of compression, DSS and
leading to low internal friction. extension), and it is not recommended to weight the
Characteristic soil data was also identified as a key end bearing capacity towards the extension strength.
uncertainty, even in cases with reasonably good qual- The plane limiting equilibrium method (optimal load
ity soil investigations. attachment point) and the plane finite element ana-
Excessive underpressure may lead to the soil plug lysis method (non-optimal load attachment point) where
rising up within the caisson, without further penetra- 3D effects were accounted for by side shear, generally
tion of the caisson. The critical underpressure is cal- gave good agreement with the 3D finite element anal-
culated from the inverse bearing capacity of the clay yses. The plastic limit analysis method using a function

28

Copyright 2005 Taylor & Francis Group plc, London, UK


fitted to approximate upper bound results, Murff & REFERENCES
Hamilton (1993), also gave good results. Available
mechanisms that rigorously satisfied upper bound AG. 2002. Suction pile analysis code: AGSPANC, Version
constraints indicated significant errors for shallow cais- 4.1, Advanced Geomechanics, Perth.
sons, but gave good agreement for the longer caissons. Andersen, K.H. & Lauritzsen, R. 1988. Bearing capacity for
foundation with cyclic loads. American Society of Civil
The capacity at intermediate load angles where there Engineers, Journal of Geotechnical Engineering, Vol.
is coupling between vertical and horizontal failure 114, No. GT5, pp. 540555.
mechanisms is well predicted when the interaction is Andersen, K.H., Dyvik, R., Schroder, K., Hansteen, O.E. &
determined by optimizing the failure mechanism in Bysveen, S. 1993. Field tests of anchors in clay. II:
plane limiting equilibrium analyses. If the interaction Predictions and interpretation. Journal of Geotechnical
is based on results from previous finite element Engineering, ASCE, Vol. 119, No. 10, pp. 15321549.
analyses and model tests, one should be cautious if Andersen, K.H. & Jostad, H.P. 1999. Foundation Design
the conditions differ from those in previous analyses of Skirted Foundations and Anchors in Clay. Proc.
or model tests. Offshore Technology Conference: OTC Paper No. 10824,
Houston.
There is some uncertainty over conditions for a crack Andersen, K.H. & Jostad, H.P. 2002. Shear Strength Along
to form at the windward side of an anchor in lightly Outside Wall of Suction Anchors in Clay After Installation.
overconsolidated clay. In the case that a crack does Proceedings, 12th ISOPE Conference, ISOPE 10824.
occur, adjustments of the simple solutions to allow for Andersen, K.H., Jeanjean, P., Luger, D. & Jostad, H.P. 2003.
a crack gave good agreement with finite element cal- Centrifuge tests on installation of suction anchors in soft
culations undertaken using ABAQUS with zero ten- clay. Int. Symp. on Deepwater Mooring Systems, Houston,
sile criterion on the total normal stress along the wall. Texas. Oct. 2003. ASCE Proc., pp. 1327. Also to be
There is no industry consensus on the safety factor publ. in ASCE J. of Ocean Engrg.
to use for capacity of suction anchor. The safety factor Andersen, K.H. & Jostad, H.P. 2004. Shear strength along
inside of suction anchor skirt wall in clay. Proc. Paper
varies with certifying agency and may also be client 16844, Offshore Technology Conference, Houston.
dependent. It should depend on the consequence of a Aubeny, C.P., Murff, J.D. & Moon S.K. 2001. Lateral
failure. Undrained Resistance of Suction Caisson Anchors.
Some factors that have been assumed and speci- International Journal of Offshore and Polar
fied in the 3D finite element analyses, and that are Engineering, Vol. 11, No.3, September 2001.
thus not checked by the comparison exercise are (1) Aubeny, C.P, Han, S.W. & Murff, J.D. 2003. Inclined load
strain softening and resulting progressive failure, (2) capacity of suction caissons. Accepted for publication in
set-up along the outside skirt wall, (3) interpretation the International Journal for Numerical and Analytical
of in situ and laboratory test data to establish a design Methods in Geomechanics.
Bransby, M.F. & Randolph, M.F., 1998. Combined loading
shear strength profile, and (4) effect of large vertical of skirted foundations. Geotechnique, Vol. 48 (5),
displacements at failure. pp. 637655.
Brinch Hansen, J. 1970. A Revised and extended formula for
bearing capacity. Geoteknisk Institutt, Bulletin No. 28,
ACKNOWLEDGEMENTS pp. 511. Copenhagen.
Calladine, C. R. 1969. Engineering Plasticity. Pergamon
The authors would like to acknowledge the help of Press, Oxford.
many colleagues from around the world who took time Cao, J., Phillips, R., Popescu, R., Al-Khafaji, Z. &
out of busy schedules to help with this study. We are Audibert, J.M.E. 2002. Penetration resistance of suction
particularly grateful to those who supplied data and caissons in clay. Proc., ISOPE 2002.
appreciate their willingness to share. In this regard we Chen, W.F. 1975. Limit Analysis and Soil Plasticity, Elsevier
Publishing Co., Amsterdam, The Netherlands.
would particularly like to extend our thanks to mem- Clukey, E.C. & Morrison, M.J. 1993. A centrifuge and ana-
bers of the API Advisory Committee and its chairman lytical study to evaluate suction caissons for TLP appli-
Philippe Jeanjean (BP), the Deepstar Joint Industry cations in the Gulf of Mexico. ASCE Spec. Publ. in
Project, BP, Chevron, ExxonMobil, Norsk Hydro, Design & Perform. of Deep Found.: Piles & Piers in Soil
Petrobras, Shell, Statoil and Total. We would also like to & Soft Rock.
thank Jean Audibert, Fugro McClelland, for contribut- Clukey, E.C., Morrison, M.J., Garnier, J. & Corte, J.F. 1995.
ing in the first phase of the project. In addition, the The response of suction caissons in normally consoli-
OTRC would like to acknowledge the U. S. Minerals dated clays to cyclic TLP loading conditions. Proc. 27th
Management Service for their support of the research Offshore Techn. Conference. Houston. Paper 7796.
Clukey, E.C. & Phillips, R. 2002. Centrifuge model tests to
program on deepwater anchors, the Centre for Offshore verify suction caisson capacities for taut and semi-taut
Foundation Systems would like to acknowledge fund- legged mooring systems. Proc. Deep Offshore
ing from the Australian Research Councils Research Technology Conference, New Orleans.
Centres Program, and the NGI would like to acknow- Clukey, E.C., Aubeny, C.P. & Murff, J.D. 2003. Comparison
ledge funding from the Research Council of Norway. of analytical and centrifuge model tests for suction caissons

29

Copyright 2005 Taylor & Francis Group plc, London, UK


subjected to combined loads. OMAE03. 22nd Intern. Murff, J.D., Randolph, M.F., Elkhatib, S., Kolk, H.J.,
Conf. on Offsh. Mech. & Arctic Engrg, Cancun, Mexico. Ruinen, R., Strom, P.J. & Thorne, C. 2005. Vertically
Clukey, E.C., Templeton, J.S., Randolph, M.F. & Phillips, R.A. loaded plate anchors for deepwater applications. Proc.
2004. Suction caisson response under sustained loop-cur- Int. Symp. On Frontiers in Offshore Geotechnics, ISFOG.
rent loads. Proc. Offshore Tech. Conf., Houston, Paper Perth, Western Australia, 1921 September 2005.
OTC 16843. Norwegian Geotechnical Institute. 1999. BIFURC-3D. A
Colliat, J-L., Boisard, P., Andersen, K.H. & Schroeder, K. finite element program for 3 dimensional geotechnical
1995. Caisson foundations as alternative anchors for per- problems. Report 514065-1, 31 December 1999.
manent mooring of a process barge offshore Congo. Norwegian Geotechnical Institute. 2000. Windows Program
Proceedings 27th Annual Offshore Technology Conference, HVMCap. Version 2.0. Theory, user manual and certifi-
Houston, Texas, OTC 7797. cation. Report 524096-7, Rev. 1, 30 June 2000. Conf.
Colliat, J-L., Boissard, P., Gramet, J-C. & Sparrevik, P. 1996. Offshore Engineer. 1996. Taut leg tested in rig role. November
Design and installation of suction anchor piles at a soft 1996, pp. 1517.
clay site in the Gulf of Guinea. 28th Offshore Technology Offshore Engineer. 1996b. Suction success for Shell.
Conference. Houston 1996. Proc., Vol. 3, pp. 325337, November 1996, pp. 26.
Paper 8150. Offshore Engineer. 1999. Record Setting Deep Draft
Colliat, J-L. & Dendani, H. 2002. Girassol: Geotechnical Caisson Set for Installation. September 1999.
design analyses and installation of the suction anchors. Olson, R.E., Rauch, A.F., Luke, A.M., Maniar, D.R.,
SUT 2002 Intern. Conf. on Offshore Site Investigation Tassoulas, J.L. & Mecham, E.C. 2003. Soil reconsolida-
and Geotechnics, London, November 2002. tion following the installation of suction caissons. Proc.,
Dendani, H. & Colliat, J-L. 2002. Girassol: Design Analyses Offshore Technology Conf., Paper 15263. Houston.
and Installation of the Suction Anchors. Proc. Offshore ONeill, D., Pezzetti, G. & Manes, V. 1991. In Situ
Technology Conference, Paper 14209. Houston. Penetration of a Large Scale Instrumented Model Skirt
Dyvik, R., Andersen, K.H., Hansen, S.B. & Pile. Field Measurements in Geotechnics, 1991 Balkema,
Christophersen, H.P. 1993. Field Tests of Anchors in Rotterdam. ISBN 90 5410 0257.
Clay, I: Description. J. of Geotechnical Engrg, ASCE, Randolph, M.F. & Houlsby, G.T. 1984. The Limiting
Vol. 119, No. 10, Oct. 1993, pp. 15151531. Pressure on a Circular Pile Loaded Laterally in Cohesive
Erbrich, C. & Hefer, P. 2002. Installation of the Laminaria Soil. Geotechnique, London, England, 34(4), 613623.
Suction piles A Case History. OTC paper 14240. Proc. Randolph, M.F., ONeill, M.P., Stewart, D.P. & Erbrich, C.T.
Offshore Technology Conference, Houston, May 2002. 1998. Performance of suction anchors in fine-grained
Fuglsang, L.D. & Steensen-Bach, J.O. 1991. Breakout calcareous soils. Paper 8831, Offshore Technology
resistance of suction piles in clay. Int. Conf. Centrifuge Conference, Houston. Proc. Vol. IV, pp. 521594.
91, Boulder Colorado, Proceedings, pp. 153159. Randolph, M.F. 2000. Effect of strength anisotropy on
Han, S. 2002. Inclined Load Capacity of Suction Caissons. capacity of foundations. Proc. John Booker Memorial
Dissert. subm. to Texas A&M Univ. in partial fulfillment Symposium, Sydney, 313328.
of requirements for the degree of Doctor of Philosophy. Randolph, M.F. & House, A.R. 2002. Analysis of Suction
HKS. 2002. ABAQUS Users Manual, Version 6.3-1, Hibbit, Caisson Capacity in Clay. OTC Paper 14236, Proc.
Karlsson & Sorensen, Inc. Offshore Technology Conference, Houston, Texas, May
House, A.R. & Randolph, M.F. 2001. Installation and pull- 69.
out capacity of stiffened suction caissons in cohesive sedi- Rauch, A.F., Olson, R.E., Coffman, R.A. & El-Sherbiny, R.M.
ments. Proc., 11th Int. Offshore & Polar Engrg. Conf., 2004. Measured horizontal capacity of suction caissons.
Stavanger, Norway, June 2001. Proc., Offshore Technology Conf., Houston, Paper 16161.
Houlsby, G.T. & Martin, C.M. 2003. Undrained bearing Renzi, R., Maggioni, W., Smits, F. & Manes, V. 1991. A cen-
capacity factors for conical footings on clay. Geotechnique trifugal study on the behavior of suction piles. Proc. Int.
53(5), 513520. Conf. Centrifuge 91, Boulder, Colorado, pp. 169176.
Keaveny, J.M., Hansen, S.B., Madshus, C. & Dyvik, R. Steensen-Bach, J.O. 1992. Recent model tests with suction
1994. Horizontal capacity of large scale model anchors. piles in clay and sand. Proc., 24th Offshore Technology
Proc. XIII ICSMFE. New Delhi, Vol. 2, pp. 677680. Conference. Houston. pp. 323330.
McNamara, A. P. 2000. Behavior of soil around the internal Tjelta, T.I., Guttormsen, T.R., & Hermstad, J. 1986. Large-
stiffeners of suction caissons and the effect on installa- scale penetration test at a deepwater site. Proc., 18th
tion and pullout resistance. Final Year Honours Thesis. Offshore Technology Conference. Houston, Texas.
University of Western Australia. Whittle, A.J., Germaine, J.T. & Cauble, D.F. 1998. Behavior
Morrison, M.J., Clukey, E.C. and Garnier, J. 1994. Behavior of Miniature Suction Caissons in Clay. Offshore Site Inv.
of suction caissons under static uplift loading. Proc. Found. Behavior 98, SUT 1998, pp. 279300.
International Conference Centrifuge 94, Singapore.
Murff, J. D. & Hamilton, J. M. 1993. P-ultimate for undrained
analysis of laterally loaded piles. J. of Geotechnical
Engrg, ASCE, Vol. 119, No. 1, pp. 91107.

30

Copyright 2005 Taylor & Francis Group plc, London, UK


Vertically loaded plate anchors for deepwater applications

J.D. Murff
Offshore Technology Research Center, Texas A&M University, College Station Texas, USA

M.F. Randolph & S.Elkhatib


Centre for Offshore Foundation Systems, University of Western Australia, Perth, Australia

H.J. Kolk
Fugro Engineers, Leidschendam, The Netherlands

R.M. Ruinen
Vryhof Anchors, Krimpen ad IJssel, The Netherlands

P.J. Strom
det Norske Veritas, Oslo, Norway

C.P. Thorne
University of Sydney, Sydney, Australia

ABSTRACT: This paper summarizes the results of an industry sponsored study on the design and analysis of
vertically loaded plate anchors (VLAs) embedded in soft clays. Phase I of the study focused on collection of ref-
erences, prediction methods, and data on actual applications, field tests, and experimental studies to establish a
baseline of experience and understanding. In Phase II, the current practices for predicting the installation per-
formance and capacity of VLAs were evaluated, including an assessment of their simplicity, completeness, sen-
sitivity, practicality, and generality. Research topics with the potential for improving current practice were
identified. The basis of the evaluation was a comparison of predictions of hypothetical cases among various
simplified methods and predictions using these methods with ground truth data from either rigorous numer-
ical analyses (FEM results) or field/experimental tests where available.

1 INTRODUCTION geometry and configuration, the mechanics of drag


embedment are very similar whether an anchor is a
In the first phase of this study an attempt was made to conventional drag anchor or a drag embedded VLA.
establish a baseline of data and prediction methods Further, the most challenging aspect of a drag embed-
for VLAs which could serve as a jumping off point ded VLA is its installation and the associated uncer-
for the second phase of research to be aimed at clarify- tainties of anchor placement.
ing the uncertainties involved in analysis and design In the second phase of the study the prediction
and identifying ways to reduce these uncertainties. In methods were evaluated to the extent practicable using
particular Phase I focused on (1) identifying and col- several techniques. The primary means of evaluation
lecting applicable references, (2) developing a collec- is a comparison of predictions of hypothetical cases
tion of summaries of prediction methods relevant to among various prediction methods and predictions
VLA installation assessment and holding capacity esti- using the methods with ground truth data from
mation, and (3) developing a database of actual VLA either field or experimental tests or more rigorous
applications as well as relevant field tests and experi- numerical analyses (finite element modeling (FEM)
ments. The focus of the field test studies is the behav- results). Specifically we carried out the following:
ior of vertically loaded plate anchors. However we have
included data on more conventional drag anchors as 1 Selected simplified methods were used to predict the
well. This was done since, despite differences in anchor behavior of hypothetical cases for both installation

31

Copyright 2005 Taylor & Francis Group plc, London, UK


and capacity and these predictions were compared
to assess variability inherent among the methods
for well defined input.
2 Results of simplified prediction methods were
compared with selected field/experimental meas-
urements taken during installations to assess the
qualitative agreement as well as the bias and uncer-
tainties in predictions and the calibrations required
to improve agreement.
3 A series of FEM analyses was conducted and the
results were used to develop a method for estimat-
ing anchor holding capacity under multiaxial load-
ing. Results of simplified prediction methods were
compared with the interpreted numerical analyses Figure 1a. Bruce Dennla anchor.
(FEM) to assess bias and uncertainties in predic-
tions as well as sensitivity to method assumptions.
While the research proposed herein is not intended to
include the development of new prediction methods,
it does identify areas where new developments might
serve to improve and extend current procedures.

2 PHASE I

2.1 Background of vertically loaded anchors


2.1.1 Principles of VLA operation
A vertically loaded anchor typically derives its hold-
ing capacity from a large bearing plate (called a
fluke). During installation of a VLA, load is applied
to the plate through an attached anchor line by various Figure 1b. Schematic of Vryhof Stevmanta.
means such as through a connecting rigid bar (shank)
or through a harness or bridle. The anchor is placed
on the seafloor such that, as the anchor is pulled along Seabed Mudline force
the bottom, it penetrates the soil. Initially, the anchor
dives more or less parallel to the fluke, eventually
rotating such that the target penetration depth is
achieved. Various methods are then used to activate the
anchor, i.e. orient the anchor fluke so that it becomes
perpendicular to the anchor line force. For example,
Figure 1a is a photograph of a VLA called the Bruce
Dennla anchor, while Figure 1b shows a schematic of
a Vryhof Stevmanta anchor. For both these anchors, Figure 2. Schematic of drag anchor installation.
the shank or bridle arrangement is designed to rotate
after placement so that the anchor line load is approxi-
mately perpendicular to the fluke. Their design has evolved by a systematic trial and
Conventional mooring lines are catenaries between error or iterative approach.
the vessel and the seafloor such that the anchor line
enters the soil horizontally. For deepwater applica- 2.1.2 Impetus for the study
tions (300 m), this can result in the anchor being over Simplicity and relatively low cost serve as strong
1000 m from the vessel. Owing to soil resistance, the incentives to use drag anchors for permanent deepwa-
anchor line takes on a reverse curvature below the ter production facilities. However, the large uncer-
mudline, such that the line imposes some vertical load tainties involved in anchor placement and capacity
component on the anchor, as shown in Figure 2. and the requisite large horizontal anchor spreads have
Historically, drag anchors have been used for temporary dissuaded many operators from their use for perman-
mooring systems although a few permanent installa- ent systems. In spite of recent innovations that have
tions for FPSOs, FSOs, and CALMs have used them. improved the reliability of drag anchor systems, such

32

Copyright 2005 Taylor & Francis Group plc, London, UK


as the development of anchors specially designed to forces, the soil exerts bearing pressure normal to the
maximize vertical capacity (e.g. the Vryhof Stevmanta line and shear resistance tangent to the line. The gov-
anchor (Agnevall 1997) and the Bruce Denla anchor erning differential equations for this system of forces
(an earlier version of the Dennla discussed by Leite are nonlinear and require an iterative, numerical solu-
et al. 2000), there remains a degree of uneasiness with tion such as the finite difference approach described
the use of drag anchors. This is primarily because of by Vivatrat et al. (1982). Neubecker & Randolph
lack of confidence during the design stage (before (1996a, b) have proposed simplifying approximations
placement) that the anchors will achieve adequate (assuming small angles and weightless line) that lin-
depth and capacity. However, if the anchor penetra- earize the equations and provide a surprisingly robust
tion depth and orientation is known or measured to be solution. The latter solution can be coupled with anchor
within expectations, the uncertainty in holding capacity behavior models to estimate both installation and hold-
is probably commensurate with typical bearing capacity ing capacity performance.
predictions. The capacity of a drag anchor depends strongly on
its final orientation and depth below the seabed, hence
2.1.3 Summary of documentation and references prediction of the anchor trajectory during installation
A bibliography of over eighty references relevant to is a critical issue in VLA design. As discussed above,
VLA applications were identified in the study. These anchor capacity is only a special case of the installa-
references are primarily published papers, although tion sequence and, hence, the methods underlying
we were able to obtain certain organization publica- installation prediction are directly applicable. Methods
tions, brochures, and reports and, in a few cases, we for predicting trajectory generally fall into four groups:
were provided with research reports such as JIP (Joint empirical methods, limit equilibrium methods, plastic
Industry Project) studies. Unfortunately, some of the limit analysis methods, and advanced numerical
most important documents, such as those detailing methods.
major field applications, remain proprietary. In these Empirical prediction methods for installation are
cases, we had to rely on open file papers which are typically based on correlations with observed anchor
necessarily brief and, therefore, it is likely that con- performance (NCEL 1987, Vryhof Anchors 1999).
siderable valuable information is not yet accessible. These methods generally involve the prediction of
anchor depth and capacity as a function of the anchor
weight and, at least, a crude measure of soil strength.
2.2 Overview of prediction methods
Limit equilibrium methods generally take into
There are three aspects of VLA behavior for which account a more detailed description of the soil and
prediction methods are needed: (1) anchor line mechan- the anchor (Stewart Technology Associates 1995,
ics, (2) installation performance, and (3) holding Neubecker & Randolph 1996b, Dahlberg 1998). The
capacity. Our literature search and industry survey anchor line mechanics are combined with the model,
identified several prediction methods in each of these at least in a simplified way. Limit equilibrium meth-
areas. It should be pointed out, however, that holding ods are typically incremental methods based on an
capacity prediction is usually implicit in the installa- estimated distribution of soil forces on the anchor at
tion prediction methods. The converse is not neces- its failure condition.
sarily true. Therefore, discussion of prediction methods Plastic limit analysis is in many ways similar to
for installation performance and holding capacity are limit equilibrium methods but also fundamentally dif-
lumped together herein. Each of these topics is dis- ferent. A solution is determined using an assumed
cussed below. failure mechanism along with virtual work principles.
Figure 2 is a schematic of a deployed anchor show- The calculated failure load is minimized with respect
ing the reverse curvature of the anchor line as it cuts to the geometric parameters defining the mechanism.
through the soil, with increasing vertical component The only method that we identified in this category is
of load on the anchor as the embedment increases. that of Bransby & ONeill (1999). This is a simple,
For a given embedment depth and orientation, as the elegant idea and achieves results similar to those that
load in the anchor line increases, the inclination of the the limit equilibrium methods purport to achieve,
line with the horizontal at the anchor attachment point except that the force distributions on the anchor are
decreases giving rise to an interaction between the implicit in the assumed mechanism.
anchor line and the installation/holding capacity of Advanced numerical methods (usually the finite
the anchor, the latter being dependent on the direction element method) have the potential to obtain a rigor-
of the resultant force. ous solution for all aspects of anchor behavior. In
In general, the anchor line problem is approached practice, however, they have considerable practical
in the same manner as that for predicting the displaced limitations because a simple anchor trajectory predic-
shape of a catenary, fixed at both ends, and deformed tion would require a prodigious effort. On the other
only by its own weight. In addition to the usual catenary hand, numerical analysis can be used to advantage to

33

Copyright 2005 Taylor & Francis Group plc, London, UK


check calculations for a specific snapshot in the Marine 1997, Omega Marine 1990) and are reason-
anchor trajectory and to therefore enhance the other ably well documented, although project reports
prediction methods described above, as shown by remain proprietary for the most part. Most of the
Rowe & Davis (1982) and Bransby & ONeill (1999). other tests were carried out by operators at or near a
possible application site (e.g. Foxton 1997, Ruinen &
Degenkamp 1999b).
2.3 Database of VLA applications and experiments
2.3.1 General 2.3.4 Model experiments
A search of the literature for data relevant to verti- The model test data are the most comprehensive of
cally loaded drag embedded or plate anchors resulted the data sets since, in these cases, investigators were
in the identification of a number of data sources. able to exercise considerably more control than is
These sources have been grouped into three cate- possible for an offshore test. Included in the data sets
gories: (1) full scale applications, (2) offshore field are centrifuge programs (ONeill et al. 1997, ONeill
experiments, and (3) onshore field or laboratory model et al. 2003), small scale 1-g lab tests (Das et al. 1985,
experiments. In each case, one or more aspects of per- Das & Puri 1989, Rowe & Davis 1982), and relatively
formance (anchor line mechanics, installation or hold- large scale 1-g onshore field tests (Dahlberg & Strom
ing capacity) were tested. In some cases, data from 1999, Heyerdahl & Eklund 2001).
conventional drag embedded anchors (not VLAs) are
included because of the relevance to VLAs. Each of
the three categories of data is briefly discussed below. 3 PHASE II HYPOTHETICAL STUDIES
2.3.2 Field applications 3.1 Objective of hypothetical case predictions
Most of the field applications are floating production
systems located offshore Brazil with Petrobras as The hypothetical case prediction exercise involves
operator. Table 1 provides a summary of field appli- having various individuals use practical design methods
cations for VLAs used to moor permanent facilities. to predict the installation behavior and capacity
Unfortunately, the data available on these applica- for well defined anchor cases. The objective of this
tions at this time are relatively sketchy. Anchor types exercise is to assess the variability among typical pre-
employed include the Vryhof Stevmanta (Ruinen & diction methods for well defined input, where the
Degenkamp 1999a, b, Agnevall 1997), the Bruce input interpretation, such as soil strength and anchor
Denla (Leite et al. 2000, Barusco 1999, Del Vecchio geometry simplifications, are removed from the
et al. 1999), and the Bruce FFTS Mk4 (Foxton 1996, assessment.
1997). The latter is not a VLA per se nor is the SBM
Mag but they provide reasonable data to test instal-
3.2 Definition of hypothetical cases
lation and capacity prediction methods.
Seven separate hypothetical cases were defined in
2.3.3 Field experiments which design parameters are varied one at a time. A
The field tests include tests ranging from very small brief discussion of these cases is provided below.
scale to full scale. Several of these tests were carried Predictors were asked to provide their estimates of the
out as JIPs (Ruinen & Degenkamp 1999a, Aker trajectory, orientation, and loads on the anchor in

Table 1. Application of VLAs for deepwater floaters.

Water Fluke
Year Field & Type Location depth (m) Anchor type area (m2) Operator

1995 Nkossa FSO Gulf of Guinea 1125 SBM Mag Elf


1996 Liuhua 11-1 South China Sea 310 Bruce 16.4 Amoco
FFTS Mk4
1998 Voador P27 Semi-FPU Offshore Brazil 530 Stevmanta 11 Petrobras
1999 Marlim South EPS FPSO-II Offshore Brazil 1215 Bruce Dennla 10 Petrobras
1999 Roncador P36 Semi-FPU Offshore Brazil 1350 Stevmanta 13 Petrobras
2000 Marlim P40 Semi-FPU Offshore Brazil 1080 Stevmanta 13 Petrobras
2002 Roncador FPSO Offshore Brazil 1150 to Stevmanta 14 Petrobras
1475
2003 Fluminese FPSO Offshore Brazil 700 Stevmanta 11 Shell
2004 Marlim FPSO Offshore Brazil 1210 Stevmanta 13 Petrobras

34

Copyright 2005 Taylor & Francis Group plc, London, UK


each of the seven cases given the following assumed vary considerably but the general nature of these cal-
conditions: culations is similar.
In each method the anchor trajectory is computed in
1 The anchor is initially embedded one meter in the a series of steps (e.g. Dahlberg 1998, Neubecker &
soil with the shank in a horizontal plane. Randolph 1996a, Bransby & ONeill 1999, Thorne
2 The anchor line is a catenary so that it is essentially 1997). At each step the anchor is assumed to move a
horizontal at the mudline and remains so during small increment based on certain assumptions or rules
the installation process. about anchor behavior. For four of the methods
The seven cases studied here are as follows: (Predictors 1, 2, 4 and 5) equilibrium is then checked
Case 1 Base case where the forces include the soil resistance and its
Case 2 Vary anchor weight (weight  2) assumed location and direction for various anchor
Case 3 Vary anchor line diameter (diameter  2) components, the anchor weight, and the anchor line
Case 4 Vary fluke shank angle (decrease from 50 tension at the attachment point. The anchor is rotated
to 35 degrees) (iteratively) so that at the final orientation at the end of
Case 5 Vary shank cross-section (increase cross- the step, equilibrium is achieved consistent with the
section area  2.25) anchor line tension and orientation. Soil resistances are
Case 6 Vary fluke aspect ratio (change fluke area based on theoretical models or empirical data sources.
from a 2:1 rectangle to a square) In the method used by Predictor 3 the anchor tra-
Case 7 Vary soil profile (change from linear jectory is computed in a series of steps using macro-
increase with depth to uniform strength). scopic plasticity concepts (Bransby & ONeill 1999).
Deepwater anchors have evolved by trial and error The method uses an interaction diagram among
and typically have rather complex geometries.
Modeling the many details of these anchors would
obviously require considerable interpretation by a Table 2. Anchor and soil parameters for hypothetical
predictor. For this reason it was decided to provide the cases.
predictors with an interpreted geometry which roughly
represents real anchors (i.e. a fluke, shank, and anchor Anchor Case 1 Variation case
line) but with a highly simplified shape as shown in parameter base case number and value
Figure 3. The geometry and weight were selected
such that the anchor has properties in the same range A 2.98 m Case 4, A  2.23 m
as typical real anchors. Table 2 provides details of the B 3.0 m Case 6, B  2.12 m
anchor and soil properties for the seven cases studied. C 1.5 m Case 6, C  2.12 m
D 3.89 m None
D1 0.7 m Case 4, D1  0.81 m
3.3 Installation calculation procedures Case 5, D1  0.89 m
Case 6, D1  0.58 m
Five organizations or individuals contributed to the D2 0.49 m Case 4, D2  0.37 m
prediction exercise. Details of each prediction method Case 5, D2  0.41 m
Case 6, D2  0.35 m
E 0.2 m Case 5, E  0.3 m
F 0.2 m None
G 0.2 m Case 5, G  0.3 m
l
D H 0m None
Fluke-shank 50 degrees Case 4, f  35 degrees
G angle, f
D1 D2 A Anchor 15 kN Case 2, W  30 kN
f weight, W
CG Anchor Line Wire rope Case 3 Wire rope
H 50 mm 100 mm diameter
F
diameter
Soil strength A  0 kPa Case 7
E su  B  1.5 kPa/m A  20 kPa
B A  B*depth B0
Soil unit 18 kN/m3 None
weight, kN/m3
Remolding ** None
factor, 1/St
C
** Predictors selected their own remolding factors which
Figure 3. Schematic of anchor for hypothetical cases. ranged from approximately 0.3 to 0.5.

35

Copyright 2005 Taylor & Francis Group plc, London, UK


Drag Distance, m Drag Distance, m
0 100 200 300 400 500 600 0 100 200 300 400 500
0 0

10 10 Case 1
Case 2

Shackle Depth, m
Shackle Depth, m

Predictor 1 20
20 Predictor 2 Case 3
30 Case 4
30 Predictor 3
Predictor 4 40 Case 5
40 Predictor 5 Case 6
50 Case 7
50
60
60 70
70 80

Figure 4a. Comparison for base case: depth vs. drag.


Figure 5a. Predictor 3 Case comparisons: shackle depth
vs. drag.
3000
4000
2500
3500
Shackle Load, KN

Case 1

Shackle Load, KN
2000 3000
Predictor 1 Case 2
2500 Case 3
1500 Predictor 2
Predictor 3 2000 Case 4
1000 Predictor 4 Case 5
1500 Case 6
Predictor 5
50
0 1000 Case 7

0 500
0 100 200 300 400 500 600 0
Drag Distance, m 0 100 200 300 400 500
Drag Distance, m
Figure 4b. Comparison for base case: shackle load vs. drag.
Figure 5b. Predictor 3 Case comparisons: shackle load
vs. drag.
forces parallel and normal to the fluke and moment as
a plastic potential surface. A similar surface using anchors studied is qualitatively similar among partici-
only parallel and normal forces (ignoring moment) is pants. For example, the penetration depth and anchor
used for the shank (Aubeny et al. 2003). Interaction capacities generally increase with drag distance,
surfaces are developed using limit analysis solutions asymptotically approaching a limiting value. There
and FEM analyses. The upper bound theorem of plas- are, however, significant differences in the quantita-
ticity is used to find the displacement/rotation incre- tive predictions and in some cases in the directional
ment that minimizes the resistance and is compatible effects of the various parameters relative to the base
with the anchor line tension and orientation. Resist- case. In the following we will discuss these compari-
ances acting normal to the components are based on sons in more detail.
local undrained shear strengths and those acting par- For these purposes we will first discuss the predicted
allel to the components are based on actual or implicit effects of the various parameter variations. Due to the
remolded strength values. large number of results we will primarily consider
them in a statistical sense and will focus on the final
anchor depths and corresponding loads.
3.4 Discussion of predicted results
Figures 6a, b depict statistics on predicted depth
For brevity only two types of example plots are pro- and shackle load for each of the cases studied. The
vided here. The first plot type (Fig. 4a, b) shows com- vertical bars represent the range of predicted values
parisons among predictors for the base case; the and the horizontal tick marks represent the mean val-
second type (Fig. 5a, b) shows comparisons for the ues. For example, for the base case, the mean pre-
different cases for a typical predictor. The parameters dicted maximum or ultimate depth among the five
plotted are shackle depth (a measure of anchor pene- predictions is 50 m with a range of approximately
tration, with the shackle being the connection point 23 m. Table 3 addresses the directional consistency of
between anchor chain and shank) vs. horizontal drag the predicted parameter effects of each participant
distance and shackle load vs. horizontal drag distance. relative to that participants base case. For example,
A review of the participants results indicates that, the effect of each parameter relative to the base case
for the most part, the predicted performance of the is represented by normalizing each prediction with

36

Copyright 2005 Taylor & Francis Group plc, London, UK


Case Case Case Case Case Case Case the respective predictors base case results. Thus, a
0
1 2 3 4 5 6 7 number larger than 1 indicates that parameter
increased the predicted quantity and a number less
10
than 1 reduced it relative to the base case. The base
20 case results for each predictor therefore have a nor-
Predicted Depth, m

30 malized value of 1 and a COV (coefficient of vari-


40 ation) of 0 by definition. The following are some brief
50 observations based on these results.
60
3.4.1 Anchor weight
70
The participants predicted that doubling the anchor
80 weight on average would have a negligible effect on
90 the penetration depth. It should be mentioned that
most empirical plots (e.g. API 1995) of anchor depth
Figure 6a. Mean and range of ultimate predicted depths. or capacity are shown as linear log-log functions of
anchor weight whereas the numerical models show no
effect of weight. The reason for this is that weight is
usually highly correlated with anchor size (fluke area,
4000 etc.) which generally controls the depth and capacity
3500 but weight itself is relatively unimportant for fluke-
type anchors.
3000
Shackle Load, KN

2500 3.4.2 Anchor line diameter


2000 The results show that increasing anchor line diameter
1500
significantly reduces predicted penetration depth and
hence holding capacity of the anchor. Participants used
1000 various algorithms (Degenkamp & Dutta 1989,
500 Neubecker & Randolph 1996b, Vivatrat et al. 1982) to
0
represent the anchor line behavior but they are based on
Case Case Case Case Case Case Case the same principles. The thicker anchor line provides
1 2 3 4 5 6 7 more resistance during dragging causing the anchor
fluke to rotate toward a horizontal position more
Figure 6b. Mean and range of ultimate predicted shackle quickly thus limiting the depth and the anchor capacity.
loads. Sizing the anchor line (or chain) is clearly an important
design issue.

3.4.3 Fluke-shank angle


Table 3. Summary of relative prediction statistics for pre-
dicted ultimate values. Current rules of thumb dictate that a fluke-shank angle
of approximately 50 degrees be used to maximize
Shank anchor holding capacity in soft clay. A smaller angle
Depth force Mudline of approximately 35 degrees is used for sand and stiff
ratio ratio force ratio clay. This value (35) was investigated for the soft clay
Case Parameter z/zcase1 T/Tcase1 T/Tcase1 case to measure the predicted effects. As indicated in
the figures and tables, the predicted results are direc-
1 Base Mean 1 1 1 tionally consistent with this conventional wisdom.
COV 0 0 0
2 Vary Mean 1.00 1.01 1.01
weight COV 0.006 0.007 0.008
3.4.4 Shank cross-section
3 Vary line Mean 0.54 0.57 0.58 The shank dimensions were increased to study the
diameter COV 0.036 0.040 0.033 effects of increased soil resistance parallel and nor-
4 Vary fluke Mean 0.38 0.29 0.28 mal to the shank relative to the forces on the fluke.
shank angle COV 0.26 0.31 0.32 Predictions indicate that increasing the shank cross-
5 Vary shank Mean 1.06 1.20 1.19 section by a factor of 2.25 increases the penetration
cross-section COV 0.035 0.065 0.062 depth of the anchor less than 10% and increases the
6 Vary fluke Mean 1.13 1.15 1.17 capacity on the order of 20%. As shown in Table 2,
aspect ratio COV 0.27 0.27 0.28 these trends were relatively consistent among partici-
7 Uniform Mean 0.49 0.26 0.27
soil strength COV 0.049 0.204 0.193
pants with depth COVs of about 0.03 and capacity
COVs on the order of 0.06.

37

Copyright 2005 Taylor & Francis Group plc, London, UK


3.4.5 Fluke aspect ratio of area. For the Vryhof Stevpris the weight (in kN) is
The fluke aspect ratio was varied from 2:1 to 1:1 with approximately,
the fluke area being held constant. The prediction
models for this case are not directionally consistent, (3)
with normalized depth predictions ranging from 0.77
(significantly shallower penetration) to 1.42 (signifi- hence
cantly deeper penetration) with a mean value of 1.13.
The predicted effect has a relatively large COV of 0.27. (4)
This is clearly an area that warrants more investigation.

3.4.6 Soil strength profile where A  fluke area in m2. For the base case (Case 1)
The final case was a variation in the soil strength pro- in this exercise the fluke area is 4.5 m2 so the ultimate
file from the base case normally consolidated clay to a holding capacity according to Equations 24 is 864 kN.
uniform strength clay. Although the anchor is consist- For this case the predictors simulation methods gave
ently predicted to penetrate in a trajectory similar to the a mean capacity of 2985 kN with a range of 1963 kN to
base case, the maximum installation load of the anchor 3811 kN, several times the capacity given by Equation 4.
tends to develop quickly and thereafter remains rela- It should be pointed out that, because of the
tively constant. This seems reasonable as, once the decreasing rate in anchor depth and load at large drag
anchor achieves a depth where it is no longer influ- distances, an anchor will typically only be pulled to
enced by the proximity of the seabed surface, then the 60 or 70% of its estimated capacity or a drag distance
resistance becomes independent of the anchor orienta- of 40 to 50% of the maximum (Vryhof Manual 1999).
tion the soil is the same in all directions. The general This fact along with the wire forerunner assumed in
nature of the penetration is consistent among predictors the hypothetical studies would significantly improve
with a depth ratio of 0.49 and a relatively small COV of agreement.
0.049. The depth is of course dependent on the particu- Another result of interest is the ultimate uplift
lar value of uniform strength selected so the ratio itself capacity of an anchor subjected to normal (or near
is not particularly meaningful but the depth COV is. normal) loading of the fluke such as a VLA. Based on
results by Martin & Randolph (2001) the capacity of
a plate loaded normal to its surface is approximately,
3.5 Design chart predictions
(5)
As a final note in this exercise it is of interest to con-
sider predictions for a hypothetical case using con-
In the base case the soil strength su equals 1.5z kPa
ventional empirical design methods similar to the
where z  depth in meters. For the mean simulation
NCEL Chart (1987) included in the API Recommended
depth the capacity for the 4.5 m2 fluke is
Design Practice
RP2SK (1995). In general the
Tnhc  87.8z kN. At 20 m depth this gives 1755 kN
design curves have the form,
and for 50 m it gives 4390 kN. The ratio of ultimate
(1) uplift (or normal) capacity to installation load in these
two cases is then 1755/864  2.0 (based on design
where Thc  anchor holding capacity (at mudline); chart) and 4390/2985  1.5 (based on simulations).
C  dimensional constant; W  anchor weight; and These numbers are indicative of the efficiencies (per-
n  dimensionless exponent. The anchor weight is formance ratios) of the anchor. For comparison, the
only relevant as it is correlated with the anchor geom- Vryhof Manual (1999) gives performance ratios of
etry, especially the fluke area. For the purposes here approximately 3 for the Stevmanta VLA.
we will use the charts for the Vryhof Stevpris Mk 5
anchor provided in the Vryhof Manual (1999).
However, the charts are based on chain forerunners 4 PHASE II CASE HISTORIES
which will tend to cause the charts to underpredict
anchor depth and hence holding capacity for anchors 4.1 General
with wire forerunners. This should be borne in mind This section describes the field test cases that were
when comparing predictions. For very soft clay the selected and the results of the predictions by different
holding capacity of the Mk 5 is given as: methods. Four organizations participated in the pre-
diction exercise; the Centre for Offshore Foundation
(2) Systems (COFS), Fugro, University of Sydney, and
Vryhof Anchors. The installation and capacity predic-
where the weight and holding capacity are expressed in tions are interdependent and are therefore presented
kN. It is convenient to express the weight as a function together in the same sections.

38

Copyright 2005 Taylor & Francis Group plc, London, UK


4.2 Objective of field test predictions Table 4. Summary of field test case histories.
The field test prediction exercise involved having var- Case 1 Case 2 Case 3 Case 4
ious companies or individuals use practical design/
analysis methods to hindcast the installation behavior Location Gulf of Gulf of Gulf of Centrifuge
and capacity for some of the best documented field Mexico Mexico Mexico UWA (80 g)
tests available. The objective of this exercise was to Operator Omega Aker Aker COFS
demonstrate the capability among typical prediction Marine Maritime Maritime
methods to assess the performance of actual case his- JIP JIP
Water Unknown 91 m 91 m NA
tories. The detailed input data for the various models
depth
were left to the individual predictors to develop from Anchor Vryhof Bruce Vryhof Based on
the case descriptions. type Stevpris Dennla Stevmanta Stevpris
In addition to the test descriptions, predictors were Weight 68.6 kN 12.7 kN 31.6 kN 373 kN
provided with the results of each test as these results (prototype)
are already publicly available. This being the case, the Fluke 9 m2 5 m2 5 m2 21 m2
predictors were encouraged to make adjustments to area
their models to improve agreement with measure-
ments where they saw fit and to document such adjust-
ments. Because of the approach taken, no attempt is pose considerable practical measurement difficulties.
made to judge results among predictors as this is not Onshore tests and small scale tests allow for consid-
considered to be meaningful. The main benefit from erable improvement in data quality but introduce
this exercise is to provide the reader insight into the other issues such as scale effects and unrepresentative
application of simplified methods to the prediction of soils. In short, there are no ideal data sets and con-
anchor performance including the potential of these siderable compromise in the acceptance criteria was
methods as well as their limitations. This will hopefully necessary to make a selection.
lead to a better understanding of how these methods Four test sets were selected that were deemed
can be used in design and analysis and in identifying adequate as summarized in Table 4. These test sets
critical areas requiring further study. were documented to the extent possible. This docu-
mentation and a list of instructions to predictors were
4.3 Selection of field cases included in a prediction package which was distrib-
uted to a wide range of potential participants.
The criteria for selection of the field cases to be hind- Participants were asked to develop and document
cast were as follows: the input data for use in their prediction model, mak-
1 Cases with the highest quality information avail- ing assumptions based on judgment as needed. They
able including anchor geometry details and soil were then asked to predict the particular (i.e. either
properties. mudline load or shackle load, whichever was pro-
2 Cases that include commonly used prototype vided) anchor depth and load vs. horizontal pull dis-
anchors including both drag anchors and vertically tance. Finally, predictors were encouraged to adjust
loaded anchors. their models to improve agreement with measure-
3 Cases that include reliable observations or meas- ments as they saw fit, documenting adjustments made.
urements. These include initial anchor position,
drag distance, anchor load (both mudline and 4.4 Predictor results
shackle where possible) vs. drag distance, and
This data represented some of the best test data avail-
anchor depth vs. drag distance.
able but, as mentioned above, was not ideal in several
4 Cases where the anchor was tested to failure after
respects including uncertainties in soil properties and
installation.
questions regarding initial anchor orientation as well
The tests selected for study were taken from the as depth and load measurements. Each predictor used
database assembled in Phase I of the project. A careful a different model with its own set of assumptions and
review indicated that all available tests have significant methodology. With sufficient fine tuning of the
shortcomings. Perhaps most notable is the minimal soil input data all models were capable of reasonably
data available for most of the tests. Secondly, measure- replicating the installation trajectories and loads for
ments of drag distance, anchor load, and anchor depth the majority of the tests. It is also important to point
have a high degree of variability and uncertainty. Since out that while this involved adjusting parameters in
the most desirable tests were actual prototype anchors the model inputs, adjustments were well within the
tested in the ocean environment, the relatively low range of the uncertainties in the data. The degree of
data quality is understandable. Tests from floating fine tuning to use was left to the predictor to decide.
vessels in a wave and current environment obviously The predicted trajectories of the more conventional

39

Copyright 2005 Taylor & Francis Group plc, London, UK


Drag Distance, m the predictors choice of where to fit the data. Since the
0 50 100 150 200 250 data set available is quite small, the generality of the
0
measurements should be viewed with some circum-
5 Predictor 2 spection until additional experimental confirmation
10 Predictor 3 can be made.
Shackle Depth, m

Predictor 4
15 It was observed that mudline load was, for the cases
Predictor 5
20 Measured
considered, a linear function of penetration depth in all
cases except possibly in close proximity to the mud-
25
line. Likewise the models all seem to predict this
30 trend. This emphasizes a very important point it
35 appears to be the dependence of penetration and load
40 on drag distance that introduces the largest uncer-
tainty. If the anchor depth and soil strength profile are
(a) Shackle depth vs. drag distance known, the anchor capacity can be estimated with rea-
1600 sonable accuracy (uncertainty commensurate with the
1400 bearing capacity of a footing, for example) including
the effects of interaction of the anchor line with the
1200
Mudline Load, KN

anchor. Predicting the ultimate anchor penetration


1000 depth and the drag distance to achieve this result a pri-
Predictor 2
800 Predictor 3
ori seems to be a much more formidable challenge.
600 Predictor 4
Predictor 5
400 Measured 4.5 Design chart predictions
200 As in the hypothetical case studies it is of interest to
0 consider predictions for the field cases using conven-
0 50 100 150 200 250 tional empirical design methods similar to the NCEL
Drag Distance, m
(b) Mudline capacity vs. drag distance Chart (1987) included in the API Recommended
Design Practice RP2SK (1995). As discussed for the
1600 hypothetical studies, an estimate of the ultimate mud-
1400 line load during the drag-in phase is given by
1200 Equations 14. In addition the anchor depth and drag
Mudline Load, KN

distance (both in m, with A in m2) may be expressed as:


1000
800 (6)
Predictor 2
600 Predictor 3

400
Predictor 4 and
Predictor 5
200 Measured
(7)
0
0 10 20 30 40
Depth, m For the purposes here we will use the above men-
(c) Shackle depth vs. drag distance tioned relationships in Equations 14 for the field case
predictions, noting that they may underpredict anchor
Figure 7. Field test predictions for Case 2. capacity and embedment depth since they are based
on chain forerunners rather than wire. By expressing
drag anchors (in this case the Stevpris anchors, Cases the parameters as a function of fluke area, the equa-
1 and 4) were curved upward (higher penetration rates tions have a similar form for the installation phase
earlier in the trajectory) and generally in good quali- (drag-in) of fluke anchors in general although they are
tative agreement with the measurements even before strictly intended for the Stevpris MK 5 anchor. Case 4
adjustments were made. With adjustments, the is a Stevpris anchor with a 32 degree fluke-shank
models were able to fit the data extremely well. For angle (vs. the 50 degree angle used as a basis in the
the two plate anchors (Cases 2 and 3) the measured charts) and is thus not included in this comparison.
depths appear to be more linear with drag distance Table 5 provides a comparison between the meas-
whereas the predictions were again curved upward. ured field data and the chart predictions for Cases 13,
As an example, for Case 2 the adjusted predictions Case 1 being the most relevant to these specific equa-
among the four models are virtually identical qualita- tions. It is interesting that the comparisons here are
tively and quite close quantitatively as shown in Figs. quite reasonable whereas the chart predictions seri-
7a, b, c. The quantitative differences are mainly due to ously underestimated the anchor depth and capacity

40

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 5. Comparison of field test results and chart
predictions.

Drag Penetration Mudline


distance depth load
Case Quantity (m) (m) (kN)

1 Measured 75 21.3 1983


Chart 141 27.7 2260
2 Measured 77 21 1445
Chart 102 20.1 886
3 Measured 77 24.0 1053
Chart 110 21.6 1089

simulated in the hypothetical case studies. Based on


the results in this study, the reasons for these discrep-
ancies cannot be determined with certainty. One
possibility is that the simulations simply tend to over-
predict depth and capacity. However, another possibil-
ity is that the anchors in the field tests were not
dragged far enough to develop their ultimate capacity.
Resolving this discrepancy has the potential to signifi-
cantly improve anchor design technology and seems to
be a promising area for future research.

5 PHASE II FINITE ELEMENT STUDIES

5.1 General: Idealized anchors for FEM analysis


Modeling the full details of anchors would require a Figure 8. Finite element mesh for L/t  20 (detail below).
very complex three-dimensional numerical analysis.
Such a model would be extremely time consuming to Establish solutions for use in calibrating and/or
develop and analyze and would be of limited value in validating simplified models.
the development of insight a principal purpose of
this study. For similar reasons we do not attempt to 5.3 Two-dimensional finite element model and
simulate the anchor installation process per se using analysis results
FEM. Instead we have selected highly simplified The two-dimensional FEM analyses undertaken here
cases for analysis, i.e. assessing anchor fluke capacity have employed the commercial program, ABAQUS
at specific depths and orientations. (HKS, 2003) to assess the capacity of plates subjected
In the base case we consider a two-dimensional to multi-axial loading. Conventional small strain
plate i.e. a slice of unit width for an infinitely wide analyses were undertaken, with the anchor movement
plate, embedded in a soil of uniform strength. This limited to 0.1 times the anchor length.
approximation ignores so-called end effects. By con- The soil has been modeled as homogeneous and
sidering a few examples of three-dimensional plates linearly elastic, with failure determined using a Tresca
we will estimate these effects and discuss methods to criterion (shear strength, su). The Youngs modulus, E,
compensate for this error. Likewise we will discuss is given by a modulus ratio of E/su  500, and
examples of varying soil strength with depth as well as Poissons ratio was taken as 0.49. The anchor has been
effects of soil-fluke interface strength. modeled essentially as a rigid body, with Youngs
modulus 107 times that of the soil, and Poissons ratio
5.2 Objective of finite element studies of 0.15. A range of length to thickness ratios, L/t, have
been analyzed. An example of the finite element
The finite element studies involve predictions of
meshes (for a plate with aspect ratio L/t  20) is
capacities of rigid plates under multi-axial loading for
shown in Figure 8. The ABAQUS interface model has
a range of plate geometries and soil conditions. The
been used to model interface friction ratios, , vary-
objectives of this study are twofold:
ing from 0 (fully smooth) to 1 (fully rough), where 
Establish a suite of ground truth solutions for is the ratio of limiting friction to the soil shear
direct use in assessing plate anchor capacity under strength. Analyses have also been undertaken without
multi-axial loading. using an interface, where the anchor is fully bonded

41

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 6. Non-dimensionalized ultimate capacities. Table 7. Theoretical capacity factors for fully rough
plates.
Aspect ratio Value of  Nn Ns Nm
Aspect ratio, L/t Nn Ns Nm
L/t  7 0 11.15 2.10 1.48
0.2 11.24 2.58 1.56 5 12.39 5.00 1.63
0.4 11.32 3.04 1.61 6 12.23 4.50 1.61
0.6 11.39 3.50 1.63 7 12.11 4.14 1.60
0.8 11.45 3.95 1.65 8 12.03 3.88 1.60
1 11.49 4.36 1.67 9 11.96 3.67 1.59
Bonded 11.58 4.49 1.74 10 11.91 3.50 1.59
L/t  20 Bonded 11.33 3.21 1.71 15 11.75 3.00 1.58
20 11.67 2.75 1.57

to the adjacent soil elements. These analyses always 5


L/t = 7, fully rough
give slightly higher capacities (by up to 4%), presum-

Parellel, Ns, or Moment, Nm, factor


ably due to forcing of the failure surface away from Normal-parallel
4
the edge of the anchor.
Displacement controlled analyses were undertaken
whereby all the nodes defining the anchor-soil inter- 3
face were forced to move together, either parallel to
the anchor plate (sliding), normal to the anchor plate, 2
or in a path corresponding to rotation of the anchor
Normal-moment
plate about its centre. All results are presented here in 1
non-dimensional terms, using factors defined as:
0
(8) 0 2 4 6 8 10 12
Normal capacity factor, Nn

where L is the length of the two-dimensional anchor Figure 9. Interaction curves in normal-parallel and
plate. In addition, ultimate values of these non- normal-moment space.
dimensional quantities will be written as Nn, Ns and
Nm, representing the different forms of bearing cap- Results for varying L/t values are given in Table 7.
acity factors. Comparing these results with those in Table 6 for
L/t  7 (fully rough) and 20 (bonded) generally
5.4 Capacities under uniaxial loading shows reasonable agreement, although the finite elem-
ent results for parallel and rotational motion are
A summary of results for the non-dimensional capaci- respectively 17% and 9% greater than the theoretical
ties under normal, parallel and rotational load is given values. It is possible that mesh refinement has led to
in Table 6 for L/t  7 and varying interface friction some loss in accuracy, as it is nearly impossible to
ratios, and also for the fully bonded case with have sufficient mesh density for the very thin plate.
L/t  20.
The ultimate capacities given in the table may be
compared with theoretical limit analysis predictions 5.5 Interaction diagrams
developed by ONeill et al. (2003). Finite element analyses have been undertaken for
fully rough plate conditions in the primary planes
M  0, Fs  0 and Fn  0. The resulting interaction
(9) curves in the Fn:Fs and Fn:M planes are shown in
Figure 9. A key difference between the two curves is
the gradient close to zero Fs or M. The Fn:Fs curve
rises steeply, implying no sliding motion for small
ratios of sliding to normal force, while the Fn:M curve
(10) shows a much smaller negative gradient, implying
significant rotation at low values of moment.
Interaction curves were developed for parallel-
(11) moment loading, for a range of normal loads represent-
ing fractions of 0, 0.2, 0.4, 0.6, 0.8 to 0.9 of the

42

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 8. Exponents for the Bransby-ONeill equation. 2
1.8

Moment Capacity Factor, Nm


Parameter m n p q 1.6
1.4
Value 1.37 3.74 1.22 3.68
1.2
1
5.00 0.8
4.50 FEM-COFS, L/t=7
0.6
Parallel Capacity Factor, Ns

Eqs. 2-5, L/t=7


4.00 0.4
FEM- OTRC, L/t=6
3.50 0.2 Eqs. 2-5, L/t=6
3.00 0
0 5 10 15
2.50 Normal Capacity Factor, Nn
2.00
1.50 FEM-COFS,L/t=7 Figure 11. Interaction curves in moment-normal space.
Eqs. 2-5, L/t=7
1.00
FEM-OTRC, L/t=10
0.50 Eqs. 2-5, L/t=10 5
0.00 4.5

Parallel Capacity Factor, Ns


0 5 10 15 4
Normal Capacity Factor, Nn
3.5
3
Figure 10. Interaction curves in parallel-normal space.
2.5
2
normal capacity. As expected, the size of the interac- 1.5
Eqs. 2-5,L/t=7

tion curves reduce as the normal load increases, but 1


FEM-COFS, L/t=7
Eqs. 2-5, L/t=6
the shapes remain similar. 0.5 FEM-OTRC, L/t=6
0
0 0.5 1 1.5 2
5.6 A generalized model Moment Capacity Factor, Nm
While the FEM results are very useful in themselves
it is convenient to put them in a generalized form Figure 12. Interaction curves in parallel-moment space.
representing the range of parameters of interest. Here
we use the general form of an equation proposed by anchor using the upper bound method of plasticity as
Bransby & ONeill (1999). described by ONeill et al. (2003). In this approach
Equation 12 takes the role of a generalized yield sur-
face where the loads and moments are generalized
stresses and the displacements and rotations are gen-
eralized strains as discussed by Prager (1959).
Predictor 3 used this approach in the hypothetical
(12) studies and field test hindcasts described herein.
Perhaps a more obvious application is to simply
In the application of this equation the values of assess the capacity of a plate anchor at a known loca-
Fnmax, Fsmax, and Mmax are determined from equations tion and orientation in a given soil profile subjected to
presented by ONeill et al. (2003). The exponents in a generalized force. As an example of this application
the equation were determined by a least squares fit of consider an anchor plate with Length L, width W and
Equation 12 to the finite element results and are given thickness t. Take the aspect ratio L/t  7 with a fully
in Table 8. rough surface so that   1 and the undrained
Figures 1012 are plots showing various cross sec- strength at the centroid  su. The uniaxial capacities,
tions of the surface described by Equation 12 compared Fnmax, Fsmax, and Mmax can then be taken directly from
to finite element results. Note that in addition to the pri- Table 6. To account for the finite width of the footing
mary finite element studies some additional, independ- the forces and moments for the plane strain cases
ent analyses were carried out to validate the results. (given per unit width) are simply multiplied by the
Equation 12 agrees well with the FEM results although plate width W. This conservatively ignores the end
the parallel and moment capacity FEM results are effects of a footing of finite width.
somewhat high. Overpredictions of 510% are fairly Equation 12 can be used to develop a range of non-
common for capacity estimates using finite elements. dimensional solutions for varying load inclination
Equation 12 has several possible applications. For (parallel load) and offset from the centroid (moment).
example, it can be used to predict the trajectory of an The uniaxial capacities from Table 6 are determined

43

Copyright 2005 Taylor & Francis Group plc, London, UK


F soil profiles. Fortunately some analytical solutions
exist for more general geometries. Martin & Randolph
(2001) published bearing (or uplift) capacity factors
Centroid
for infinitely thin, smooth and rough, circular foot-
t ings loaded in a normal direction of 12.42 and 13.11
b respectively. Merifield et al. (2003) published a factor
of 11.9 for a square footing using a novel lower bound
L
FEM approach. This is believed to be very close to an
Figure 13. Schematic of loaded anchor plate. exact solution. These solutions compare with the fac-
tor for an infinitely thin, two-dimensional plate of
11.42 (rough and smooth solutions are the same for
14 this case). For this case the three-dimensional effect
12 b/L=0 for a square plate is then about 4% a fairly minor
effect. The three-dimensional effects for the parallel
Resultant Force/LWSu

b/L=0.1
10 b/L=0.2 and moment loading are also relatively small.
8
b/L=0.3 One approach to incorporate the three-dimensional
b/L=0.4
effects is to simply adjust the uniaxial capacities for
6 b/L=0.5
normal, parallel, and moment loading required for
4 input into Equation 12 as discussed above. This
assumes the exponents in Equation 12 would not be
2
significantly affected by three dimensional effects.
0 Alternatively, the plane strain solutions (per unit
0 20 40 60 80 width) can be determined directly and multiplied by
Load Inclination from Vertical, Degrees
the plate width, which always leads to conservative
results as previously mentioned.
Figure 14. Resultant capacity of a plate under multiaxial
loading based on Equation 12. Since the soil strength profile typically varies with
depth we consider here the effect of that variation on
the anchor capacity. Consider a typical soft clay profile
and substituted into Equation 12 giving a single non- with a strength gradient of 1.25 kPa/m. A high capacity
linear equation where the resultant load F is the only anchor in this profile might have an ultimate depth of
unknown as shown in Figure 13, a schematic of the 25 m where the soil strength would be 31.25 kPa and a
loaded plate. The equation is well behaved and straight- minimum fluke dimension of the order of 2 m. The fail-
forward to solve with most standard techniques. Figure ure zone around the anchor might extend 2 m above
14 contains plots of the normalized value of F as a func- and below the anchor in which case the soil strength
tion of inclination, , and non-dimensional offset, b. would vary by 5 kPa over the vertical extent of the fail-
Consider the case of zero load inclination in Figure ure zone or about 8%. In our view such a variation is
13. For b  0, the capacity is the normal capacity from very unlikely to significantly change the anchor cap-
Table 6, i.e. 11.49. As b increases beyond this point, F acity from that estimated by using a homogeneous
will monotonically decrease but the product solution with the average strength at the fluke centroid.
b/L  F/suWL will approach the normalized moment This assumption could lead to larger errors in other
capacity from Table 6, i.e. a value of 1.67. As the load profiles such as a layered soil or more severe strength
inclination approaches 90 degrees (pure parallel load- gradients. This also may not be appropriate in predict-
ing), all solutions approach the parallel capacity, 4.36. ing behavior at shallow depths during the anchor tra-
Note also that for smaller offsets the resultant load jectory, in particular in estimating the drag length
capacity is largest at zero inclination and monotonically required to achieve a certain depth.
decreases thereafter. At large offsets the peak capacity The above issues of three-dimensional geometry and
occurs at intermediate inclinations. This is because the soil strength variation effects could be addressed in fol-
normal load decreases with inclination and this effect is low-up studies to develop more definitive guidelines
exaggerated due to the relatively large exponent on the and ultimately to improve anchor prediction models.
normal load terms. Thus Equation 12 provides a simple,
practical method for evaluating the multi-axial capacity
5.8 Interpretation of results from hypothetical
of a plate anchor in an undrained clay soil.
studies
It is of interest to compare results from various sim-
5.7 Extension of solutions
plified methods with more rigorous finite element
The solutions in the above section are derived for results. In order to do this it is essential to establish a
deeply embedded two-dimensional plates with uniform common basis for comparison. The FEM results are

44

Copyright 2005 Taylor & Francis Group plc, London, UK


F St  soil sensitivity. It can further be shown that for
X soil strength increasing linearly with depth.
lab
fbn
lbc a
Shackle (18)
fbs Shank
b
fdn
c where
md
Y d fds
e (19)
Fluke
General direction
of travel
In this part of the study a limited number of results
from the hypothetical studies were selected for inter-
Figure 15. Schematic of soil loads on anchor components. pretation. In particular, specific snapshots from Cases
1, 4 and 7 were interpreted for all five predictors to
obtain the implied forces and moment acting on the
represented using Equation 12, which provides a rela- fluke centroid. For a consistent comparison we
tionship among the failure interaction loads applied at selected a depth (during the anchor trajectory) to the
the midpoint of a rectangular plate. The simplified fluke midpoint of 13.33 m such that the soil strength
methods used in the hypothetical studies are based on at the fluke midpoint is the same in all three cases
the anchor line load applied to the shackle. Since the (13.33 * 1.5  20 kPa). These results were compared
fluke in the hypothetical studies is also a rectangular with Equation 12 based on the FEM studies. Since the
plate, one approach is to resolve the loads from the simplified methods all use different assumptions and
simplified analyses to the fluke midpoint. approximations, the anchor orientations and load
The results from each of the simplified methods angles vary somewhat among the various predictions
include the anchor shackle depth, shackle load and at the same fluke midpoint depth and thus the com-
load angle, and the anchor orientation in a series of parisons are not completely consistent. However, we
snapshots along the predicted anchor trajectory. From estimated the forces and moment at the fluke mid-
the anchor geometry and estimates of soil loads on point for all cases and used the interaction equation,
the shank the shackle loads are resolved to the fluke Equation 12, from FEM studies to provide a compari-
mid-point. Figure 15 is a schematic diagram showing son with simplified method results.
the forces acting on the anchor components. Equilib- Two key results in the analyses by simplified methods
rium considerations result in the following equations: are the resultant force and the angle the resultant
force makes with the shank axis. For the cases con-
(13) sidered, the resultant forces range from about 500 kN
to 700 kN and the force angles range from about 10
degrees to
10 degrees. Figure 16 is a plot showing
the inferred values of the failure function, Equation
12 (actually  f  1) for the hypothetical case snap-
(14) shots. Two sets of data points are shown one set with
the moment included, the other set with the moment
omitted. In all cases with positive load angles, the
(15) fluke-soil system is below yield as defined by
Equation 12 even with moment included. On the
other hand, points with negative load angles exceed
where lf is the fluke length.
yield in most cases. This result points out the import-
For the case of linearly varying soil strength with
ance of moment on the failure function and the sensi-
depth, the forces on the shank are taken as:
tivity to load angle, especially for negative values.
This is a topic that may deserve further study in any
(16) attempt to improve prediction tools for installation
and capacity estimates.
While the above results are hopefully useful, their
(17) importance should be kept in perspective in assessing
present prediction methods. These methods have gen-
where sua  strength at point a; su1  strength gradi- erally been developed using a system of assumptions
ent; Np  bearing capacity factor; ls  shank length; that tend to be self-correcting. All the simplified
ws  shank width; Cs  shank circumference; and methods provide qualitatively realistic predictions

45

Copyright 2005 Taylor & Francis Group plc, London, UK


4 (assuming a chain forerunner and correcting for dif-
Inferred from Hypothetical
3.5 Data ferences in drag distance) and the average mudline
Inferred from Hypothetical ultimate capacity was approximately 3.5 times the
3 Data, No moment conventional chart prediction. Some of this difference
Failure Function

2.5 is due to the larger depths and capacities associated


with wire rope vs. chain. The predicted ratio of uplift
2
capacity (load normal to fluke) to installation load
1.5 ranges from 1.5 (simulations) to 2.0 (charts).
1
6.3 Phase II: Case histories
0.5
Failure Line To provide further insight into simplified industry
0 anchor installation models, a hindcast exercise was
-10 -5 0 5 10
Force Angle with Shank, Degrees conducted using available field and centrifuge data.
Four different sets of drag anchor or plate anchor
Figure 16. Deviation from Equation 12 for inferred forces installation data, including measured anchor response,
from hypothetical cases. were provided to volunteer predictors. This data
represented some of the best test data available but
was not ideal in several respects.
and can be calibrated to give accurate quantitative
Each predictor used a different model with its own
results for specific situations. However, as in any
set of assumptions and methodology. With sufficient
semi-empirical method, one should not attempt to
fine tuning of the input data all models were capable
improve any part of the method without due regard to
of reasonably replicating the installation trajectories
the total recipe.
and loads for the majority of the tests. The predicted
trajectories of the more conventional drag anchors
6 SUMMARY AND CONCLUSIONS were curved upward and generally in good qualitative
agreement with the measurements. For the two plate
6.1 Phase I: Compilation of database anchors, the measured data showed a much more lin-
ear variation of depth with drag distance whereas the
This phase of the study focused on compilation of predictions were again curved upward.
information related to vertically loaded anchors. The It was observed that mudline load was a linear func-
results include a listing of references, descriptions of tion of penetration depth in all cases except possibly
various prediction methods, and data from field in close proximity to the mudline. Likewise the
applications and tests. The information collected pro- models all seem to predict this trend. This emphasizes a
vided a good baseline for Phase II studies, which very important point it appears to be the dependence
included assessing current design and analysis of penetration and load on drag distance that intro-
methods in an effort to improve prediction methods duces the largest uncertainty.
and to better understand the uncertainties involved. An additional study was carried out using conven-
tional chart prediction methods, similar to the NCEL
6.2 Phase II: Hypothetical studies charts (1985) included in API RP2SK (1995), to hind-
cast the four field tests. The chart predictions pro-
This study was carried out to assess the variability in
vided reasonable estimates of the field tests for drag
predictions of drag anchor installation performance
distance, depth, and mudline load except for Case 4
for currently available numerical models. These
which had significantly different anchor line and
models typically employ limit equilibrium methods
fluke-shank angle than those on which the chart pre-
using simplifying assumptions.
diction was based. This agreement is contrary to the
All models predicted qualitatively similar trajector-
significant differences between chart predictions and
ies with the anchor initially penetrating parallel to the
simulation predictions for the hypothetical cases.
fluke. The anchor gradually rotates and ultimately
reaches a steady state where the anchor translates
6.4 Phase II: FEM studies
horizontally.
Based on predictors results the most important In order to supplement the few experimental results
effects were found to be anchor line diameter, fluke- on plate anchor capacity and to develop a better under-
shank angle, fluke aspect ratio, and soil strength char- standing of the behavior of plate anchors under multi-
acteristics. axial loading, a series of finite element calculations was
For the base case, the average simulation depth pre- conducted. Failure surface diagrams were developed
diction (assuming a wire rope forerunner) was approxi- for multi-axial loading comprising normal, parallel
mately 2.5 times the conventional chart prediction and moment loads. Although coupling is evident

46

Copyright 2005 Taylor & Francis Group plc, London, UK


among all three load components, it is strongest Andersen, K.A., Murff, J.D. & Randolph, M.F. 2001.
between normal load and moment. For example, at Deepwater Anchor Design Practice- Vertically Loaded
high normal loads the anchor undergoes significant Drag Anchors, First Year Report to API Volume III.
rotation at small moments. Andersen, K.A., Murff, J.D. & Randolph, M.F. 2004.
Deepwater Anchor Design Practice- Vertically Loaded
It was found that the FEM results were in good Drag Anchors, Phase II Report to API/Deepstar JIP,
agreement with available analytical results with a ten- Volume III.
dency to slightly overpredict. The largest discrepancy Aubeny, C.P., Han, S. & Murff, J.D. 2003. Inclined load
was found to be in the parallel capacity. The analytical capacity of suction caissons, Int. J. for Numerical and
results were combined with the FEM results into a Analytical Methods in Geomechanics, 27: 12351254.
single empirical interaction equation of a form origin- Barusco, P. 1999. Mooring and anchoring systems
ally suggested by Bransby & ONeill (1999). The developed in Marlin field, Proc. 31st Annual Offshore
equation was fit to the FEM results using the least Technology Conference, Houston, Texas, OTC 10720.
squares method and was found to be an excellent fit. Barusco Filho, P.J. & Ferrari, J.A. 1997. Case study exam-
ining the use of taut leg, fibre rope mooring lines for
It was pointed out that relatively minor adjustments in deepwater platforms. Proc. FPS97 Conference.
the interaction equation may be sufficient to extend Bransby, M.F. & ONeill, M.P. 1999. Drag anchor fluke-soil
these results to three-dimensional plate anchors and interaction in clays. Proc. Int. Symp. on Numerical
to typical normally consolidated, soft clay profiles. Models in Geomechanics (NUMOG VII), Graz, Austria,
The FEM results were also used to gain insight into 489494.
the simplified methods used to estimate anchor tra- Dahlberg, R. 1998. Design procedures for deepwater
jectories. The FEM results were compared to the anchors in clay, Proc. 30th Annual Offshore Technology
inferred load capacities of selected snapshots from Conference, Houston, Texas, OTC 8837.
the trajectories predicted in the hypothetical studies. Dahlberg, R. & Strm, P. 1999. Unique onshore tests of
deepwater drag-in plate anchors, Proc. 31st Annual
These comparisons point out that prediction of Offshore Technology Conference, Houston, Texas, 1:
anchor trajectories is a difficult task and the simpli- 713724.
fied methods should be calibrated for specific appli- Das, B.M., Moreno, R. & Dallo, K.F. 1985. Ultimate pullout
cations in order to provide the most reliable results. capacity of shallow vertical anchors in clay. Soils and
Foundations, 25(2): 148152.
Das, B.M. & Puri, V.K. 1989. Holding capacity of inclined
ACKNOWLEDGEMENTS square plate anchors in clay. Soils and Foundations,
29(3): 138144.
The authors would like to acknowledge the help of Degenkamp, G. & Dutta, A. 1989. Soil resistances to
many colleagues from around the world who took embedded anchor chain in soft clay, Journal of
Geotechnical Engineering, ASCE, 115(10): 14201438.
time out of busy schedules to help with this study. We Del Vecchio, C.J.M. & Costa, L.C.S. 1999. Recent advances
are particularly grateful to those who supplied data in deepwater mooring systems off Brazil. Proc. 14th IBC
and appreciate their willingness to share. In this Conference FPS99.
regard we would particularly like to extend our thanks DnV 1999. Design and installation of drag-in plate anchors
to members of the API Advisory Committee and its in clay, Draft version, June 1999. Recommended Practice
chairman Philippe Jeanjean (BP), and to the Deepstar RP-E302, Det Norske Veritas, 132.
Joint Industry Project. In addition, the OTRC would Dunnavant, T.W. & Kwan, C.-T.T. 1993. Centrifuge model-
like to acknowledge the U. S. Minerals Management ling and parametric analyses of drag anchor behaviour.
Service for their support of the research program on Proc. 25th Annual Offshore Technology Conference,
Houston, Texas, OTC 7202.
deepwater anchors and the Centre for Offshore Dutta, A. & Degenkamp, G. 1989. Behaviour of embedded
Foundation Systems would like to acknowledge fund- mooring chains in clay during chain tensioning. Proc.
ing from the Australian Research Councils Research 21st Annual Offshore Technology Conference, Houston,
Centres Program. Texas, OTC 6031.
Foxton, P. 1996. Deepwater moorings: The Amoco Liuhua
experience & beyond. Proc. IBC Conference on Mooring
REFERENCES & Anchoring, Aberdeen.
Foxton, P. 1997. Latest development for vertically loaded
Agnevall, T. 1997. Installation and performance of P27 anchors. Proc. 2nd Annual Conference on Mooring &
Stevmanta-VLA anchors. Proc. 2nd Annual Conf. on Anchoring, Aberdeen.
Mooring & Anchoring, Aberdeen. Heyerdahl, H. & Eklund, T. 2001. Testing of plate anchors.
Aker Maritime Contractors 1997. The Deepstar Project Proc. 33rd Annual Offshore Technology Conference,
CTR 3405: Final Report on VLA Tests, Test Report, Oslo, Houston, Texas, OTC 13273.
Norway. HKS. 2003. ABAQUS Users Manual, Version 6.4, Hibbit,
American Petroleum Institute 1995. Recommended practice Karlsson and Sorensen, Inc.
for design and analysis of station keeping systems for Leite, A.J.P., Costa, L.C.S., Skusa, W., Schuurmans, S.T. &
floating structures, API-RP-2SK, Washington. Zanutto, J.C. 2000. Marlim South early production

47

Copyright 2005 Taylor & Francis Group plc, London, UK


system (FPSO II), The first taut-leg moored FPSO in the Omega Marine Services International 1990. Joint industry
world. Proc. 12th DOT Conference. project: Gulf of Mexico large scale anchor tests test
Martin, C.M. & Randolph, M.F. 2001. Applications of the report, Omega Marine Services International, Houston,
lower and upper bound theorems of plasticity to collapse Texas.
of circular foundations. Proc. 10th Int. Conf. on ONeill, M.P., Bransby, M.F. & Randolph, M.F. 2003. Drag
Computer Methods and Advances in Geomechanics, anchor fluke-soil interaction in clays. Canadian
Tucson, 2: 14171428. Rotterdam: Balkema. Geotechnical Journal, 40: 7894.
Martin, C.M. 2001. Vertical bearing capacity of skirted cir- ONeill, M.P., Randolph, M.F. & Neubecker, S.R. 1997. A
cular foundations on Tresca soil. Proc. 15th Int. Conf. on novel procedure for testing model drag anchors, Proc. 7th
Soil Mechanics and Geotechnical Eng., Istanbul, 1: International Offshore and Polar Engineering
743746. Conference, Honolulu, Hawaii, 1: 939945.
Merifield, R.S., Sloan, S.W. & Yu, H.S. 1999. Stability of Prager, W. 1959. An Introduction to Plasticity, Addison-
plate anchors in undrained clay, Research Report No. Wesley.
174.02.1999, Department of Civil, Surveying and Environ- Rowe, R.K. & Davis, E.H. 1982. The behaviour of anchor
mental Engineering, The University of Newcastle. plates in clay, Geotechnique, 32(1): 923.
Merifield, R.S., Lyamin, A.V., Sloan, S.W. & Yu, H.S. 2003. Ruinen, R. & Degenkamp, G. 1999a. First Application of 12
Three-dimensional lower bound solutions for stability of Stevmanta Anchors (VLA) in the P27 Taut Leg Mooring
plate anchors in clay. Journal of Geotechnical System. Proc. 11th DOT Conference (Deep Offshore
Engineering, ASCE, 129(3): 243253. Technology), Stavanger.
NCEL 1987. Drag embedment anchors for navy moorings. Ruinen, R. & Degenkamp, G. 1999b. Advances in the devel-
Techdata Sheet 83-08R, Naval Civil Engineering opment and operational experience with Stevmanta
Laboratory, Port Hueneme, California. VLAs in deepwater environments. Proc. 4th IBC confer-
Neubecker, S.R. & Randolph, M.F. 1995. Profile and fric- ence Mooring & Anchors, Aberdeen.
tional capacity of embedded anchor chain. J. Stewart Technology Assoc. 1995. STA ANCHOR User
Geotechnical Engineering. Eng. Div., ASCE, 121(11): Manual and Technical Documentation, Houston.
787803. Thorne, C.P. 1998. Penetration and load capacity of
Neubecker, S.R. & Randolph, M.F. 1996a. The performance marine drag anchors in soft clay. Journal of Geotechnical
of drag anchor and chain systems in cohesive soil. and Geoenvironmental Engineering, ASCE, 124(10):
Marine Georesources and Geotechnology, 14: 7796. 945953.
Neubecker, S.R. & Randolph, M.F. 1996b. Performance of Vivatrat, V., Valent, P.J. & Ponterio, A. 1982. The influence
embedded anchor chains and consequences for anchor of chain friction on anchor pile design, Proc. 14th Annual
design, Proc. 28th Annual Offshore Technology Offshore Technology Conference, Houston, Texas, OTC
Conference, Houston, Texas, OTC 7712. 4178.
Offshore Engineer 1999. Floating production Brazil stays Vryhof Anchors 1999. Anchor Manual 2000, Krimpen ad
at the cutting edge, (January 1999). YJssel, The Netherlands.
Offshore 2000. Trial results from deepwater anchors, poly- Vryhof Anchors 2000. Company Presentation, Private
ester moorings, (March 2000). Communication.

48

Copyright 2005 Taylor & Francis Group plc, London, UK


Australian frontiers spudcans on the edge

C.T. Erbrich
Advanced Geomechanics

ABSTRACT: This paper presents results and interpretation from two extraordinary case histories of jackup oper-
ation in the Yolla field and at the Trefoil prospect in the Bass Strait, offshore Australia. A variety of events occurred
during installation that appear unprecedented in the literature, including static punch-through at depths exceeding
any of four a-priori predictions of maximum spudcan penetration and the use of cyclic preloading to increase
spudcan penetration. Many lessons can and should be learnt from these examples and incorporated into stan-
dard practice in future cases where difficult soils such as carbonate (or some non-carbonate) silty sands and
sandy silts are encountered. These include the need to reconsider the methods used to assess spudcan penetra-
tion where high sensitivity soils are encountered and where partial drainage effects are significantly different
during spudcan penetration and conventional SI probing (PCPT or T-bar). New methods are proposed to predict
spudcan penetration directly from modern SI tools such as T-bar and ball penetrometers and to assess the effect
of cyclic strength degradation when calculating the jackup rig foundation stability under design storm events.

1 INTRODUCTION 18.2 m

The Yolla field and the Trefoil prospect in the Bass


Strait, offshore Australia, are frontier areas, which until
recently had been unexplored and undeveloped. The 5.49 m
Yolla field is located due east of King Island and due
North of Burnie, Tasmania, while the Trefoil prospect is
located a further 38 km to the west of the Yolla devel-
opment. An unmanned platform (Yolla A), founded
on a large skirted raft was recently installed in the Yolla Figure 1. Ensco 102 spudcan geometry.
field to produce gas for export by pipeline to Kilcunda
in Victoria. The production wells for this platform were
spudded in mid 2004 using the Ensco 102 jackup rig, clay layers are also interbedded amongst the carbon-
which was bought on station shortly after the platform ate soils.
installation. Subsequent to completion of these pro- Due to the development of the Yolla A platform, a
duction wells, the Ensco 102 moved to spud an explo- comprehensive site investigation, focused on the per-
ration well at Trefoil. formance of shallow foundations, was performed at the
The Ensco 102 is a Keppel Fels Mod V A class platform location. A more limited but nevertheless
jackup unit. The spudcans each have an area of 256 m2 invaluable site investigation was also performed at
which gives an effective diameter of about 18.2 m Trefoil prior to the arrival of the Ensco 102. These
(Fig. 1), and the jackup unit can apply preloads up to investigations revealed unusual soil characteristics; in
98 MN. A feature of this modern class of jackup rig is particular very high sensitivities and a marked suscep-
the ability to apply full preload on one leg at a time tibility to degradation under cyclic loading. However,
while the hull remains submerged in the water. As will as will become apparent, even this information proved
become apparent, this feature repeatedly prevented insufficient to enable accurate a-priori predictions of
catastrophe in the conditions that were encountered. the spudcan penetration at Yolla. With the benefit of
The soil conditions found at Yolla and Trefoil are learnt experience from Yolla, the situation at Trefoil
quite different to those found in previously developed was forewarned, but even despite this, the sequence of
areas of the Bass Strait. The profile at both locations events that unfolded appears unprecedented. MSL
mostly comprises carbonate sandy silt/silty sand or (2004) presents a comprehensive review of the State of
sand. Over the depth range of interest, two calcareous the Art on jackup foundation integrity, including the

49

Copyright 2005 Taylor & Francis Group plc, London, UK


compiliation of over 50 incidents from 84 different Records were also kept of the ongoing settlement
references. However, none of the quoted examples of the jackup rig, which amounted to between 250 mm
appear to have much similarity to Yolla or Trefoil. and 300 mm in a two month period after installation.
Nevertheless, it is suspected that some of the reported
incidents may be due (at least in part) to similar
mechanisms to those encountered at Yolla and Trefoil,
but these were not identified at the time due to insuf- 3 A-PRIORI PREDICTIONS OF SPUDCAN
ficient data and inadequate understanding of the basic PENETRATION RESPONSE AT YOLLA
mechanics.
This paper will present the facts of the Yolla and A number of parties made various a-priori predictions
Trefoil case histories and a geotechnical interpret- as to the likely spudcan penetration response of the
ation of these events. New ideas are introduced, which Ensco 102 spudcans when jacked down at the Yolla A
have specific application to these cases, but also reveal platform location. The predicted penetration depth to
a deeper understanding of a number of aspects of the tip of the actual spudcan (Fig. 1) is given below
jackup spudcan behaviour in a more general sense. for each of these:

1 Predictor 1 estimated penetration depths of 11.6 m


2 OBSERVED SPUDCAN PENETRATION to 16.6 m for a range of different jackups. The Ensco
AT YOLLA 102 is a Keppel Fels Mod V class jackup which was
not explicitly addressed by this predictor although
The Ensco 102 was bought on station next to the Yolla they did consider the Hitachi Giant class of jackup,
A platform at the beginning of June 2004. During which imposes a slightly higher bearing stress on
preloading of each of the spudcans (with the hull in the soil. Hence it may be inferred that Predictor 1
the water), unexpectedly large penetrations occurred would have estimated that the Keppel Fels Mod V
after about 85% of the maximum preload had been would penetrate to slightly less than 15 m.
applied. For all three spudcans, a static punch- 2 Predictor 2 estimated initial penetration depths of
through occurred after the tip of the spudcans had between 13.1 m to 18.6 m for the Keppel Fels Mod
penetrated to a depth of about 20 m even though this V class jackup, increasing to between 13.6 m and
depth already exceeded the expected maximum pene- 20.1 m over time due to consolidation of the under-
tration depth. As demonstrated later, piezocone pene- lying soil.
tration tests (PCPT) and T-bar penetrometer data 3 Predictor 3 estimated penetration depths of 15.4 m
suggested an increase in strength at this depth, which to 19.0 m for the Keppel Fels Mod V class jackup,
is the last thing usually expected to be associated with assuming either no backflow or including backflow
a punch-through risk. The punch-through involved a of soil behind the spudcan respectively. However, it
free-fall of about 4 m, before the bottom of the spud- was suggested that the best estimate would be the
cans finally came to rest at depths of about 23 m to no backflow case (ie. 15.4 m).
24 m below mudline. These punch-throughs led to the 4 Predictor 4 estimated penetration depths of between
jackup tilting about 4, and in the worst case, the hull 10.3 m and 20 m although it was suggested that the
of the jackup came within 3 m of contact with the most likely result would be a penetration depth
Yolla A platform. Each spudcan was then successfully towards the deeper end of the range.
loaded to the maximum preload of 98 MN with min-
imal extra penetration. Hence the final spudcan pene- Hence it can be seen that the upper bound esti-
trations were between depths of 23 m and 24 m in all mates made by Predictors 2, 3 and 4 are all very simi-
cases. All three spudcans performed in much the lar, with the tip of the spudcan penetrating about 19 m
same manner and hence little lateral variability was to 20 m. However, the actual observed penetration of
apparent at the site. the spudcans was about 20% higher than even these
An ROV (remotely operated vehicle) survey was upper bound estimates. Hence the question must be
performed of the seabed around the spudcans after the asked; what caused the spudcan penetration to be so
final penetration had been achieved, which revealed much higher than all the a-priori initial estimates?
that the area inside the legs appeared to be filled with In the following sections the available soil data at
sea floor material, with no evidence of a significant the location is discussed. Then the various a-priori
crater around the spudcan entry points into the seabed. predictions are considered in more detail and the limi-
However, an ROV survey undertaken just before the tations of these methods when applied to the Yolla soil
spudcans were extracted revealed circular features are discussed. Finally a back-analysis, is presented
with a diameter of approximately 1.5 times that of the along with appropriate revisions to both the soil
spudcan. Anecdotal evidence suggests that these were parameters and theory in order to enable the observed
the edges of seabed craters with depths of around 1 m. response to be predicted.

50

Copyright 2005 Taylor & Francis Group plc, London, UK


4 SOIL DATA AT YOLLA Consolidation Coefficient, ch (m2/yr)

4.1 General 1 10 100 1000 10000 100000


0
To enable the detailed design of the Yolla A platform
foundation, comprehensive programmes of site inves-
5 CPT 2 - Ir = 500
tigation activities and subsequent laboratory testing
were undertaken at the platform location. An overview CPT 2 - Ir = 50
of these investigations and the data obtained is pre- 10 CPT 3 - Ir = 500
sented in Watson & Humpheson (2005). It is perhaps CPT 3 - Ir = 50
ironic that the large amount of data that has been col-

Depth, z (m)
15
lected compared to that which would normally be avail-
able for a typical spudcan penetration analysis did not
enable better a-priori predictions of spudcan penetra- 20
tion to be made, but has highlighted the severity of the
discrepancy between the predictions and the measured
25
data. With a more typical quantity of SI data it is likely
that the aberrant behaviour could (and would) have
been more readily written off as due to some lack of 30
knowledge brought about by the limited quantity and
quality of the data. On the other hand, the high quantity
35
and quality of available data means that it is now pos-
sible (with the benefit of hindsight) to obtain a robust Figure 2. PCPT coefficient of consolidation at Yolla.
interpretation of what actually did happen, which prob-
ably would not have been possible otherwise.
The soil profile has been described by Watson &
Humpheson (2005) as mostly comprising a mixture rem N95
0 0.1 0.2 0.3 0.4 0 1 2 3 4 5
of interbedded carbonate sandy silts and very sandy 0 0
silts. However, several calcareous clay layers were also 5 5
Depth (m)

Depth (m)
identified, at various depths. The most significant for 10 10
the current evaluation is that found between approx- 15 15
imately 15 m and 19.5 m below the mudline. The other 20 20
25 25
clay layer that needs to be noted when interpreting the
30 30
data given herein, but which is not significant for the
spudcan interpretation, is that found between about Figure 3. Parameters derived from cyclic T-bar tests at Yolla.
1.6 m and 3.2 m below the mudline.

An extensive programme of T-bar testing was also


4.2 In situ data
performed during the final site investigation, includ-
Three separate site investigation programmes have been ing both static and cyclic tests. Up until recently,
undertaken at the Yolla A platform location, each of cyclic T-bar tests have only rarely been performed and
which involved measurement of in situ soil properties. previously these have been thought of as providing
All of the programmes included some PCPT prob- qualitative information only. However, recent work
ing, albeit each with a cone of different diameter. The (Einav & Randolph 2004) has revolutionised our under-
data collected show that the site exhibits little lateral standing of this data and, as will be shown later, this
variability. Generally, it was found that the measured provides considerable insight into the current problem.
PCPT resistance reduced with increasing cone diam- Of particular importance are two parameters; rem, the
eter, which is an effect that is now believed to be fully remoulded shear strength ratio obtained after
mostly attributable to some degree of partial drainage many penetration/extraction cycles of the T-bar, and
during the testing (see Section 6.2 for more discus- N95, the number of cycles to achieve 95% of the total
sion on this subject). degradation that occurs during cycling. These data
The final site investigation performed by Fugro are summarised on Figure 3. In fully undrained soils
also included an extensive programme of PCPT dissi- rem is the inverse of the sensitivity, St. However, where
pation tests. These results are considered to be of partial drainage occurs during the testing, there has
great importance in understanding the spudcan pene- been some doubt as to whether rem remains a true
tration behaviour that occurred. An interpretation of measure of the sensitivity. Unfortunately, there is no
these results is presented on Figure 2. other test data (eg. laboratory fall cone tests) from

51

Copyright 2005 Taylor & Francis Group plc, London, UK


Penetrometer Resistance (MPa) require further consideration or reconsideration in
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6
light of the observed spudcan penetration:
0
Laboratory strengths exceeding field derived values.
Viscous rate effects during loading.
5 During detailed design of the Yolla A platform
there was much discussion on the subject of why the
monotonic undrained shear strengths measured in
Average CPT
10 the laboratory appeared, in many cases, to be signifi-
Depth (m)

cantly higher than the strength deduced from the


T-bar. This discrepancy is apparently compounded by
Average T-bar
15
the observed spudcan response, which seems to imply
(at face value) that even the strength measured using
the T-bar is excessive when applied to the spudcan
response. Fortunately, it will be shown later that this
20 latter presumption is not really correct, but a similar
discrepancy to that obtained previously still remains
between the laboratory test strengths and the T-bar
25 strength. It will transpire that the main argument used
by the platform designers (ie. strain softening of the
Figure 4. T-bar and PCPT resistance at Yolla. soil during T-bar penetration) is certainly applicable.
However, the strain softening effect is offset by
another factor (ie. partial drainage) that had not previ-
Yolla to confirm or deny this belief; but from another ously been accounted for. Hence the author is now
site that was recently studied, where the coefficient of convinced that sample disturbance was the main
consolidation derived from PCPT tests was very similar explanation for the observed strength difference (it is
to that obtained at Yolla, fall cone tests have con- worth noting at this stage that unlike in most soils,
firmed that the sensitivity derived from the cyclic sample disturbance of carbonate sandy silts tends to
T-bar was a reasonable estimate of the true value. Hence create a denser soil matrix, leading to a reduced sen-
it is believed that the same conclusion should apply at sitivity and a greater tendency to dilation in the dis-
Yolla, although in the analyses that are presented later, turbed sample compared to the in situ soil).
the adopted value for rem is higher than the most Another important factor in the spudcan penetra-
extreme low values that were actually measured. tion story is the influence of the rate of loading on the
Figure 4 presents a comparison between the net undrained shear strength; most soils exhibit some vis-
bearing resistance (averaged for all individual profiles) cous rate effect, which leads to higher undrained
obtained with the T-bar and cone during the final site strengths for higher applied strain rates. To assess this
investigation. One unusual and striking feature of this factor, the available static and cyclic simple shear
data is that the average cone resistance is much higher tests were reprocessed and the measured stressstrain
than the average T-bar resistance. In the sandy silt and responses obtained during monotonic loading were
very sandy silt, this deviation may be explained by compared with the first cycle of cyclic tests for sam-
the higher degree of drainage expected for the PCPT ples from similar depths. From these results the strain
compared to the T-bar (see Section 6.2). However, rate parameter  was determined, which is the rate of
this explanation cannot apply in the clay layers, where increase in shear strength for a 10 fold increase in
both tools would certainly have exhibited a fully strain rate:
undrained response. The deviation between the PCPT
and T-bar in these layers remains something that can-
not be readily explained. (1)

4.3 Laboratory data


where: .
A comprehensive programme of laboratory tests was su  undrained shear strength at a strain rate of 
undertaken on samples recovered during the two main su(ref)  undrained
. shear strength at a strain rate of
site investigation programmes (performed by Benthic ref
Geotech and Fugro). The data obtained is comprehen- A typical result from this assessment is presented
sive and is summarised in Watson & Humpheson on Figure 5 while the results obtained from each batch
(2005). It is believed that there are two aspects that of tests are summarised in Table 1.

52

Copyright 2005 Taylor & Francis Group plc, London, UK


20 50 Monotonic su (kPa)
Test Depth = 6.3 m to 6.7 m
0 25 50 75 100 125 150
Shear stress (kPa)

15 40
0

Time (sec)
30
10
20 5
5
10
10
0 0
0 1 2 3 0 1 2 3
Shear Strain (%) Shear Strain (%) 15

Depth (m)
Figure 5. Stress-strain response for different rates; Yolla soil. 20

Table 1. Summary of viscous rate effect 25


parameter.
30 Predictor 2; Lower Bound
Depth Range (m) Predictor 2; Upper Bound
35 Predictor 3
Core from to  Predictor 4
Average T-bar (Nbar=10.5)
Benthic 6.3 6.7 0 40
Fugro 6.9 7.2 0.03
Fugro 9.1 9.2
0.03 Figure 6. Undrained strength used by spudcan penetration
Benthic 9.1 9.3 0.10 predictors.
Fugro 10.7 10.9 0
Benthic 11.5 11.6 0.02
Benthic 13.7 13.9 0.08 the inclusion of a stronger layer at about 24 m in the
Benthic 23.0 23.2 0.06 profile of Predictor 4. These both provide a good aver-
Average 0.03 age fit to the T-bar strength data over most of the depth
range of interest. The two profiles used by Predictor 2
were developed at an earlier stage based on only
There is a good deal of scatter from batch to batch
limited preliminary site investigation information,
(and between individual tests in the same batch),
but nevertheless generally bound the profiles used by
which reflects the inherent uncertainties associated
Predictors 3 and 4.
with this sort of exercise. In one case,  was shown to
The spudcan penetrations estimated by Predictor 3
be negative, which is physically unlikely, but this
are based on the approach outlined in the recom-
value has been maintained to ensure that the statistical
mended practice, SNAME (2002):
average is correctly calculated. Note that all the test
results apply to samples of sandy silt and very sandy
silt; no data is available for the clay layers.
(2)

5 COMMENTARY ON A-PRIORI YOLLA where VLo is the applied preload (pressure), Nc is a


SPUDCAN PENETRATION PREDICTIONS bearing capacity factor and su is the shear strength at
the bottom of the widest section of the spudcan.
All the a-priori estimates of spudcan penetration con- The bearing capacity factors proposed in the
sidered undrained soil behaviour. Predictor 4 also SNAME practice apply to undrained strength profiles
made an assessment assuming drained soil behaviour, that vary linearly with depth. These factors explicitly
but since this led to significantly less penetration than account for the depth of embedment, the spudcan base
obtained using their undrained assumption, and since roughness and the spudcan shape. As discussed by
it will become clear that fully undrained conditions Houlsby & Martin (2003), these factors are rigorous
pertained, this case will not be considered any further. lower bounds (ie. the true values must be equal to or
Figure 6 presents the various undrained strength higher than the given values). Indeed it has been sub-
profiles used by Predictors 2, 3 and 4 along with an sequently shown (Martin 2001) that for a rough and
interpreted undrained strength assuming that a stan- flat bottom spudcan, the true Nc factors are typically
dard T-bar factor of 10.5 is applied to the T-bar data. around 15% higher than the SNAME values. Adopting
It may be seen that the strength profiles of Predictors these higher values would, of course, have reduced the
3 and 4 are virtually identical to each other, excepting predicted penetration and increased the divergence

53

Copyright 2005 Taylor & Francis Group plc, London, UK


with the measured data. The soil weight term and the data has shown their method to be quite reliable for
backflow term largely cancel out in the above equa- predicting spudcan penetrations in various soft clays
tion when full backflow occurs (ie. where no crater is found in South East Asia.
left behind the spudcan after penetration) this scen- The method used by Predictor 2 is a hybrid of
ario leads to the largest penetrations. In soils where a methodologies. The bearing capacity factors used to
stable hole can be formed, there may be no infilling assess the spudcan penetration were the SNAME fac-
behind the penetrating spudcan, which in turn leads to tors for a rough flat base, but excluding the embed-
significantly reduced spudcan penetrations. Based on ment factor. It was argued by Predictor 2 that since
their assumed in situ soil strength, Predictor 3 assessed penetrating spudcans remould the soil above them,
that no backflow was the most likely scenario; however, leading to a significant reduction in the soil strength,
the spudcan installation shows that this was not a good the embedment factor would be much lower than the-
assumption. Predictor 3 assessed backflow using the ory would suggest. In addition, it was argued that the
wall failure criteria currently specified by SNAME. underside of most spudcans is not flat or fully rough,
However, recent research (Hossain et al. 2005) has resulting in lower bearing capacity factors than would
shown that the wall failure criterion significantly be obtained for flat, rough bases. However, a soil
overestimates the stable depth and that a flow failure weight term was included in the analysis of Predictor
criterion should be used for this assessment instead. 2 in such a way as to imply a no-backflow scenario.
The Yolla observations are consistent with this find- As will become clear in the next section, it is now evi-
ing but probably also reflect an additional factor that dent that these assumptions about the significance of
was not explicitly considered by Hossain et al. soil remoulding and spudcan shape were generally
(2005); the soil sensitivity. Kaolin with a sensitivity appropriate, but the implied assumption of no-back-
of only about 2 was used by Hossain et al. (2005) but, flow offset this effect, resulting in equivalent Nc fac-
as explained above, the Yolla carbonate soils have a tors similar to the SNAME factors for full backflow,
much higher sensitivity. Hence both wall and flow but where a depth effect is explicitly included. If the
failure mechanisms are likely to be initiated much Nc factors assumed by Predictor 2 had been used
more readily in the partially remoulded soils above along with a full backflow assumption, significantly
the spudcan in Yolla type soils compared to kaolin larger penetrations would have been predicted; per-
with the same in situ peak strength. haps as large as actually measured in the field.
When full backflow is included Predictor 3 esti- However, while this is convenient, it does not provide
mated a spudcan tip penetration of 19 m, which is still an adequate explanation of everything that occurred
20% to 25% less than observed. in the field; the reality is considerably more complex.
It is understood that Predictor 4 used the same Before closing this discussion, it should also be
basic equation as that adopted by Predictor 3 (assum- appreciated that the under-prediction of spudcan pene-
ing full backflow) but with a different form for Nc: tration is more severe than evident from the raw stat-
istics presented above since the soil strength profiles
used by both Predictors 3 and 4, which are apparently
(3) supported by the T-bar data, include an approximate
25% jump in the soil strength at a depth of about
20 m. In order to advance the spudcan beyond this
where d is the spudcan embedment depth and D is the depth, a proportional increase in the penetration force
diameter. should have been required.
This equation generally gives lower values of Nc In summary, the methods used by all the predictors
than would be obtained from the more rigorous include a variety of theoretical and empirical assump-
SNAME practice. In addition, the undrained strength tions with differing levels of conservatism. Despite this,
assumed in the analysis of Predictor 4 was an aver- all of the upper bound predictions of spudcan penetra-
aged value over a depth equal to one quarter of the tion were significantly less than actually encountered.
spudcan diameter below the tip, which is also a less In order to explain the actual field behaviour a more
rigorous approach than the SNAME practice (for cases fundamental understanding of the basic mechanics is
where the strength varies linearly with depth). It is required, which is the subject of the next section.
believed that these factors have led to Predictor 4s
slightly higher calculated penetration for full backflow
compared to Predictor 3. It is noteworthy that while 6 YOLLA SPUDCAN PENETRATION
the method used by Predictor 4 appears less rigorous REVISED INTERPRETATION
than that used by Predictor 3, it has led to slightly
more accurate predictions of the actual spudcan pene- A sensible starting point in a review of the mechanics
tration (15% to 20% under-predicted). Based on dis- of spudcan penetration is to consider the behaviour
cussion with Predictor 4, it is understood that field of another type of penetrating object; the T-bar

54

Copyright 2005 Taylor & Francis Group plc, London, UK


penetrometer. As discussed previously, this tool was In contrast, Cassidy et al. (2004) present data from
used extensively during the second Yolla site investiga- centrifuge tests where model T-bar and spudcans were
tion and at face value it might be considered to give a penetrated into a pure kaolin clay. In this case the
direct reading of the bearing resistance of other kinds of T-bar resistance is much closer to the spudcan resist-
penetrating objects, such as spudcans. This is because ance albeit the spudcan resistance is still about 15%
the geometry of the T-bar allows full flow-around of less than obtained with the T-bar (Fig. 8). A series of
soil as it penetrates and hence this might be con- model spudcan tests performed in reconstituted car-
sidered to properly include the effect of soil remould- bonate silt from the North West Shelf (Cassidy, pers.
ing; in this respect it is a more appropriate tool for comm.) also provides useful data. These results seem
assessing spudcan penetration than the piezocone. to indicate that the spudcan and T-bar penetration
Figure 7 presents the equivalent penetration resist- resistance are very similar to one another for
ance of the Ensco 102 spudcans, assuming that the undrained conditions (Fig. 9).
measured T-bar resistance may be applied directly to
the spudcan geometry:
Bearing Pressure (kPa)
(4) -40 -20 0 20 40 60
0
Spudcan; 0.1 mm/s
where: Spudcan; 0.2 mm/s
Qspudcan  force required to penetrate spudcan T-bar 1 mm/s
1 T-bar 1 mm/s
qTbar  T-bar penetration resistance (pressure) T-bar 1 mm/s
Aspudcan  base area of spudcan
Penetration (m)

It must be remembered that the T-bar will see 2


every small change in the soil layering whereas the
much larger spudcan will tend to integrate the behav- 3
iour of all the different soil layers for some distance
around the spudcan; hence the true spudcan resist-
ance would be something of an averaged version of 4
the presented resistance profile over perhaps 4 m to Prototype Dimensions of Centrifuge Models
5 m below the spudcan tip. Also shown on Figure 7 Spudcan Diameter = 6 m
are the actual recorded spudcan penetration responses T-bar Diameter = 0.5 m
as supplied by Ensco; it is clear that the observed
resistance is much less than implied using the Figure 8. Centrifuge model test data; kaolin.
T-bar  spudcan hypothesis.

Ave. T-bar Resistance x Spudcan Area (MN) Bearing Pressure (kPa)


0 50 100 150 200 250 300 -20 -10 0 10 20 30 40
0 0
Spudcan from raw T-bar
Measured (Bow Leg)
Measured (Port Leg)
Measured (Starboard Leg) 1
5
2
Penetration (m)

10
Depth (m)

4
15
5
Spudcan; 3 mm/s
20 Spudcan; 1.0 mm/s
Spudcan; 0.33 mm/s
T-bar; 1 mm/s
25 T-bar; 1.5 mm/s

Figure 7. Raw T-bar prediction of spudcan penetration. Figure 9. Centrifuge model test data; carbonate silt.

55

Copyright 2005 Taylor & Francis Group plc, London, UK


For the in situ carbonate silts at Yolla the situation is where:
clearly more complex than the simple T-bar  spud-  and rem are as defined earlier
can hypothesis allows, even though other model test 95  the cumulative plastic shear strain required
data have shown that this does not appear to be an to cause 95% reduction of shear strength
unreasonable assumption. It is therefore necessary to (from peak to remoulded)
consider the fundamental differences between the p  the theoretical average shear strain per pas-
T-bar and spudcan and develop models that address sage of the penetrometer
these factors and which can also explain the differ- Nideal  theoretical T-bar factor derived for an ideal
ence between the Yolla field scenario and the other rigid-plastic material.
model test cases. These differences are as follows: 95 may be determined from cyclic T-bar tests
using the following equation:
1 The T-bar is obviously much smaller than the spud-
can; 40 mm diameter compared to 18 m diameter
respectively. The averaging of different strength (6)
layers is one effect of this which has already been
discussed, but in addition it is necessary to con- where N95 is as defined earlier.
sider the influence of other scale factors such as Values for Nideal and p are presented in Table 2 as
the effect of different consolidation times and a function of the surface roughness () of the T-bar.
embedment depth to diameter ratios. Einav & Randolph suggest that for consistency 
2 The rate of strain imposed on the soil during T-bar should be set equal to rem.
penetration is several orders of magnitude faster Equation 5 is intended to correlate the field
than during spudcan penetration; the T-bar is pene- T-bar resistance to an average laboratory monotonic
trated at approximately half a diameter per second strength determined at a strain rate of 1% per hour.
while the spudcan is penetrated 15,000 to 20,000 However, the standard strain rate adopted in mono-
times slower (in a normalised sense) than this. tonic simple shear tests for offshore projects is
Therefore the potential effect of different loading approximately 20% per hour and hence the strain rate
rates on the soil strength also needs to be considered. term requires some modification:
3 The T-bar comprises a relatively long cylindrical
bar, whereas the spudcan is a relatively flat bot-
tomed, approximately circular footing. The influ-
ence of the different bottom geometries and the (7)
difference between axisymmetric and plane strain
geometries need to be considered. This equation has been applied to several locations
where all the necessary data is available and with one
Fortunately, a good body of recent research has exception (see Section 7) it appears to give sensible
shed considerable light on the influence of the various results. For the Yolla sandy silt and very sandy silt, 
factors that govern the resistance of penetrating was assessed earlier as 0.03, rem as 0.05 and N95 as 2,
objects of different types, and this work is considered whereas in the clay layers rem was assessed as 0.3 and
in the following sections. N95 as 4. No information is available for  in the clay
layers but a value of 0.1 was estimated based on
knowledge of other clays.
6.1 Theoretical modelling of T-bar penetration Using these parameters with equation 7, an Nbar
Einav & Randolph (2004) present a comprehensive the- value of 7.65 is estimated for the sandy silt and very
oretical evaluation of the basic mechanics of penetro- sandy silt while 13.9 is deduced for the clay layers. It
meter behaviour. Traditionally (Stewart & Randolph is interesting to note that the standard cone factors, Nk
1994) the T-bar factor (Nbar) used to assess the
undrained shear strength (su  qbar/Nbar) has been
selected as 10.5. However, Einav & Randolph have Table 2. Parameters for T-bar analysis.
developed models that specifically account for the
Nideal
effect of soil remoulding and for viscous rate effects in
the soil. These authors now propose the following  Lower bound Upper bound Average p
approximate analytical expression to evaluate the bar
factor: 0 9.14 9.20 9.17 4.42
0.25 10.06 10.09 10.08 3.93
0.5 10.82 10.83 10.83 3.60
0.75 11.45 11.46 11.46 3.38
1 11.94 11.94 11.94 3.27
(5)

56

Copyright 2005 Taylor & Francis Group plc, London, UK


( qnet/su), implied by these Nbar factors are about 12 of Western Australia, which have addressed the pene-
for the sandy silt and 23 for the clay layers; the latter tration response of T-bars, PCPTs and spudcans when
is surprisingly high, which reflects the earlier obser- penetrated into different soils. In all of the interpret-
vation on the large disparity between the raw penetra- ations, a good degree of judgement was applied in
tion resistance obtained from the T-bar and PCPT. selecting various parameters, in particular the spe-
Further discussion on the clay strength is provided in cific value of cv to use in the normalisation. Slightly
Section 14. different conclusions may therefore be drawn from
different data sets and hence several different sets of
results are considered here.
6.2 Consolidation properties; spudcan The first set of tests was reported by Finnie (1993),
versus T-bar and considered a model spudcan penetrated into car-
Prior to the occurrence of the unexpected spudcan bonate sands and silts. His results for the silt case are
penetration at Yolla, it had been assumed that the shown on Figure 10, along with the interpreted values
PCPT and T-bar penetration tests were performed of vD/cv determined for the Yolla sandy silt and very
under essentially fully undrained conditions in all sandy silt layers given in Table 3. More recently, Watson
layers, except for the thin sand layer at a depth of performed a series of PCPT and T-bar penetrometer
about 3.5 m. The appropriateness of this assumption tests in a different carbonate silt recovered from the
appeared to be reinforced by the fact that the labora- North West Shelf, while Cassidy also undertook T-bar
tory undrained strengths were significantly higher tests in the same material (Cassidy, pers. comm.). The
than the backfigured value obtained from the T-bar latter results are presented on Figure 11 along with
using the standard bar factor of 10.5, as discussed the Yolla parameters.
earlier. However, the unexpected spudcan behaviour Finally, Randolph & Hope (2004) summarise a
gave cause for this assumption to be reassessed and it number of studies of PCPT and T-bar penetration in
is now believed that the T-bar and PCPT tests were
both affected by drainage to some degree in the sandy
silt and very sandy silt layers. The critical parameter 40
Finnie Silt
that determines the degree of drainage is the ratio Fugro Tbar (Yolla)
vD/cv, where v and D are the velocity and diameter of 30 Fugro CPT (Yolla)
the penetrating object respectively and cv (or alter- Ensco 102 Spudcans(Yolla)
natively ch) is the coefficient of consolidation.
q/'vo

Consolidation properties for the various soil layers at 20


the Yolla A location were discussed earlier. Based on
the interpretation of the data shown on Figure 2, it is 10
believed that the coefficient of consolidation can be
assigned an average value of about 7000 m2/yr for all 0
of the sandy silt and very sandy silt layers while in the 0.001 0.1 10 1000
various clay layers, lower values of between about vD/cv
20 m2/yr and 100 m2/yr seem appropriate.
The parameters required to assess the ratio vD/cv, Figure 10. Drainage during penetration; after Finnie 1993.
and the calculation of this ratio, are summarised in
Table 3.
Various tests have been performed at the Centre for
Offshore Foundation Systems (COFS) at the University 11
T-bar Twitch Test
Fugro Tbar (Yolla)
Table 3. Consolidation parameters; PCPT, T-bar and spudcan. 9 Fugro CPT (Yolla)
Ensco 102 Spudcans (Yolla)
VSS/ SS* Clay 7
q/qu

D (m) v (m/s) vD/cv vD/cv 5

Thales PCPT 0.0159 0.02 1.44 100 3


Benthic PCPT 0.0357 0.02 3.22 225
Fugro PCPT 0.0437 0.02 3.94 276 1
Fugro T-bar 0.04 0.02 3.60 252 0.01 0.1 1 10 100
Ensco 102 18.2 5.5  10
4 45.0 3150 vD/cv
Spudcan
Figure 11. Drainage during penetration; after Cassidy,
* VSS/ SS  Very sandy silt/ sandy silt. pers. comm.

57

Copyright 2005 Taylor & Francis Group plc, London, UK


4 consolidation for the upper part of the sample would
Randolph and Hope (CPT) be lower than the reference value, which was obtained
3.5 Randolph and Hope (T bar)
Fugro Tbar (Yolla) from Rowe cell testing of a perfect sample. The effect
Fugro CPT (Yolla) of this would be to shift the normalised response to
3 Ensco 102 Spudcans (Yolla) the right, leading to a bigger enhancement factor for
q/qu

2.5 any given value of vD/cv.


Considering all of the different results presented
2 above, it is believed that there is a strong case to
1.5
assume that the penetration resistance measured in
the Yolla PCPT and T-bar tests has been enhanced to
1 some degree by drainage effects. Despite the different
0.1 1 10 100 1000 responses for different soils, the data appears to sup-
vD/cv port a minimum degree of enhancement of about 30%
(ie. the true undrained resistance would have been
Figure 12. Drainage during penetration; after Randolph & 0.77 times the measured value). It will be shown in
Hope 2004. Section 6.5 that this may be insufficient to explain the
observed spudcan behaviour, but that an enhance-
ment factor of no more than 1.55 (ie. a true undrained
kaolin clay and suggest new best fit lines for PCPT resistance of 0.65 times the measured value) is suffi-
and T-bar tests using their most recent data; this is cient to match the observed spudcan behaviour,
shown on Figure 12 along with the Yolla parameters. which also appears quite plausible.
The most notable difference between the various For the PCPT and T-bar in the clay soils, and for
tests is the large variation in the ratio of penetration the spudcan penetrating through any of the Yolla soils,
resistance between the undrained and fully drained it may be seen by inspection of the vD/cv ratios given
tests; the silts have a much bigger ratio than obtained in Table 3 that fully undrained conditions are certain
for the clay. There are also differences between the to have been induced. Hence for assessing the spud-
value of vD/cv at which the penetration resistance starts can response the fully undrained resistance must be
to increase significantly above the fully undrained used in all layers.
resistance. Finally, it should be appreciated that it is
theoretically expected that T-bar and PCPT/ spudcan
results should be offset from each other (with the 6.3 Rate Effects: Spudcan versus T-bar
PCPT and spudcan plotting to the right of the T-bar);
It has previously been mentioned that T-bar penetra-
this is due to the difference between axisymmetric
tion induces a much higher strain rate in the soil than
and plane-strain geometries. This theoretical observa-
spudcan penetration. Hence it is to be expected that
tion is supported from the empirical data for the kaolin
the shear strength mobilised during the latter will be
clay but is less obvious in Watsons carbonate silt data.
less than during the former due to viscous rate effects.
For assessing the conditions pertaining during
As discussed earlier, the viscous rate effect for the
PCPT and T-bar testing at Yolla, considerable weight
sandy silt and very sandy silt is quite low, whereas for
has been given to the kaolin test data even though this
the clay it is expected to be much higher.
is the most dissimilar to the sandy silt and very sandy
Assuming that the spudcan penetration rate is
silt found at Yolla. It is considered that the kaolin tests
about 2 m per hour, the normalised penetration rate
were the most controlled of all those that have been
(penetration rate divided by diameter) is 0.11 diam-
undertaken and it is believed that these have the clear-
eters per hour as compared to 0.5 diameters per second
est definition of all the necessary parameters. From
for the T-bar. The ratio of strain rates for these two
the kaolin tests, the enhancement factors are about
cases is therefore 16,200 (ie. 4.2 log cycles). Using
1.3 from the T-bar data and 1.5 from the PCPT data.
the  values given earlier, the penetration resistance
The various tests in the silt show a good deal of scatter,
for the T-bar penetration is therefore about 13% higher
with the Finnie data suggesting a substantial enhance-
than would be expected for a spudcan in the sandy silt
ment factor (about 4). The best average fit to the
and very sandy silt, and about 42% higher in the clay.
Watson data gave an enhancement factor of between
about 1.3 from the T-bar data and 1.5 from the PCPT
data, whereas Cassidys T-bar data suggests an enhance-
6.4 Geometry Effects: Spudcan versus T-bar
ment factor of about 1.25. One problem with inter-
preting the silt tests is that there appears to be some The Ensco 102 spudcan is a relatively flat bottomed
evidence of sample segregation, with finer grained and axisymmetric footing, while a T-bar is a horizontal
material at the top, grading to coarser below. Under cylindrical bar. These geometry differences may give
such conditions, it is likely that the true coefficient of rise to a different bearing response in the two cases.

58

Copyright 2005 Taylor & Francis Group plc, London, UK


T-bar Ring Stiffener means that the initial T-bar roughness is higher, leading
to a bigger effect from a trapped wedge; a Nc value
of about 0.4 has been estimated for this case.

Trapped
Wedge 6.4.2 Axisymetric vs plane strain
Another major geometric difference between the
Remoulded strength T-bar and spudcan is that the former is oeffectively a
Ensco 102 Spudcan plane strain problem while the latter is axisymmetric.
Theoretical work reported in Einav & Randolph (2004)
suggests that the bearing resistance of a ball pen-
etrometer (literally a spherical ball pushed into the
ground in the same manner as a T-bar or PCPT) should
be higher than obtained with a T-bar. However, most
Trapped empirical data available to date indicates a similar
Wedge resistance for the T-bar and ball, or in some cases, a
lower resistance for the ball. Further work is ongoing
Remoulded strength in this area but it is believed (Randolph, pers. comm.)
that the poor theoretical results are due to the forma-
Figure 13. Trapped wedge mechanisms. tion of a highly inhomogeneous strain field during
ball penetration, compared to a fairly homogenous
strain field for the T-bar. It is expected that an improved
6.4.1 Trapped wedge
theoretical model that properly accounts for the inhomo-
Experience with ring stiffeners in suction piles (eg.
geneous strain field will resolve this discrepancy. It is
Erbrich & Hefer 2002) has demonstrated that a flat
also believed that an improved model will demon-
strip may give a lower bearing capacity than a cylin-
strate that the ball resistance will be similar to the
drical T-bar due to the ability of the former to trap a
T-bar for soils with low sensitivity, but lower than the
wedge of weak soil from the surface below the base
T-bar where the sensitivity is high, which appears
(Fig. 13).
consistent with current empirical data. The Yolla sandy
The interface between this trapped wedge of weak
silt and very sandy silt are highly sensitive and hence
soil and the surrounding soil that it displaces is sub-
it is anticipated that a ball penetrometer would pene-
ject to very high strains, leading to full remoulding
trate with lower resistance than the T-bar (assuming
and hence a very low shear strength. Erbrich & Hefer
the same drainage conditions in both cases). To model
(2002) present an analysis of the effect of a trapped
this effect a factor FA ( the ratio of ball to T-bar
wedge, which suggests that the bearing capacity fac-
resistance) has been included in the analysis but since
tor Nstrip could be only 80% of Nbar. However, their
at this stage there is no firm basis to assess what level
analysis also includes various other factors that are
of reduction might occur, a value of 1 has been
not pertinent to the current case and hence this result
adopted. However, the authors best guess of this fac-
is not directly applicable here.
tor would be around 0.9.
For the current scenario it is believed that a useful
approach is to assume that the trapped wedge can be
modelled as equivalent to a cone protruding from the 6.4.3 Embedment
base of the spudcan. The influence of the very weak Finally, there is the difference in relative penetration
remoulded interface between this wedge and the sur- depths for the two foundations; the very small T-bar is
rounding soil may be treated as an effectively smooth penetrated many hundreds of diameters into the soil
interface in this case. SNAME (2002) and Houlsby & whereas the much larger spudcan only penetrated
Martin (2003) present bearing capacity factors for about 1.3 diameters into the soil. However, Mehryar
different roughness assumptions and these have been et al. (2002), show that for a spudcan penetrated into
used to assess the effect of a trapped wedge. In the normally consolidated soil, this level of embedment is
carbonate sandy silt and very sandy silt, the low strength sufficient to ensure that the limiting resistance for deep
assigned to the T-bar/ soil interface means that the failure is virtually obtained. In addition, these analyses
absolute effect of the trapped wedge must be rather were for soils with a sensitivity of 1 whereas for soils
small; the roughness of the T-bar is assumed to be with a higher sensitivity it is to be expected that the
0.05 in the carbonate silt while on the trapped wedge embedment effect will be further suppressed leading
it is about 0.005. Hence subtracting an incremental to a limiting resistance being obtained at even lower
bearing capacity factor (Nc) of 0.1 is considered to normalised penetration depths. A factor FE is there-
be an appropriate allowance for the trapped wedge fore included in the analysis to address the embed-
beneath the spudcan. The lower sensitivity in the clay ment effect, but it has been assigned a value of 1.

59

Copyright 2005 Taylor & Francis Group plc, London, UK


6.5 Revised prediction of spudcan penetration conclusion, or whether modifying some other factor
resistance (such as adopting an FA or FE less than 1.0) is more
realistic; however, both scenarios appear credible.
The foregoing sections have outlined the various fac-
As an interesting aside it is worth considering the
tors that are believed to differentiate penetration of a
bearing capacity factor, Nc, required to predict the
spudcan from a T-bar. Combining these various factors
final spudcan penetration depth at Yolla if a simple con-
allows a revised prediction to be made of the spudcan
ventional bearing capacity approach is followed (ie.
penetration at Yolla using the T-bar data as the basis.
q  Nc su) and this is applied to the revised T-bar
This is achieved using the following equation:
interpretation of the undrained shear strength presented

Adjusted Ave. T-bar Resistance x Spudcan


Area (MN)
0 50 100 150 200 250 300
0
(8) Spudcan from adjusted T-bar
Spudcan from raw T-bar
where: Measured (Bow Leg)
5 Measured (Port Leg)
Qspudcan  force required to penetrate spudcan Measured (Starboard Leg)
qTbar  T-bar penetration resistance (pressure)
Fd  drainage correction factor Depth (m) 10
Nbar  theoretical bar factor
Nc  trapped wedge adjustment factor 15
FR  T-bar to spudcan viscous rate effect adjust-
ment factor
FA  axisymmetric geometry correction factor 20
FE  embedment correction factor
Aspudcan  base area of spudcan
25
The values adopted in the analysis for each of these
various factors are summarised in Table 4.
Figure 14 and Figure 15 present the revised predic- 30
tions of spudcan penetration resistance, using the for-
Figure 14. Revised T-bar prediction of spudcan penetra-
mula and parameters outlined above, along with the tion Fd  0.77.
observed penetration behaviour. Also shown for com-
parison are the original T-bar estimates of penetration
resistance. Remember that the spudcan will average
Adjusted Ave. T-bar Resistance x spudcan
the predicted penetration resistance over about 4 m to
Area (MN)
5 m below the spudcan tip.
Figure 14 applies to the case where Fd is 0.77 in the 0 50 100 150 200 250 300
0 Spudcan from adjusted T-bar
sandy silt and very sandy silt, and it can be seen that Spudcan from raw T-bar
the average penetration resistance still appears some- Measured (Bow Leg)
what higher than the observed behaviour. However, 5 Measured (Port Leg)
Measured (Starboard Leg)
adopting an Fd of 0.62 seems to fit the measured data
well, as shown on Figure 15. With the available soil 10
data it is not really possible to determine whether
Depth (m)

adopting an Fd as low as 0.62 is the most appropriate


15
Table 4. Revised spudcan penetration analysis parameters.
20
Parameter Sandy Silt/Very Sandy Silt Clay

Nbar 7.65 13.9 25


Fd 0.77 or 0.62 1.0
FR 1.13 1.42
FA 1 1 30
FE 1 1
Nc 0.1 0.4 Figure 15. Revised T-bar prediction of spudcan penetra-
tion Fd  0.62.

60

Copyright 2005 Taylor & Francis Group plc, London, UK


above (ie. su  Fd qTbar /Nbar). Using an Fd of 0.77, the and spudcan penetration resistance and therefore it
required Nc is determined to be about 5.2, while if an would be predicted that they should be similar, as
Fd as low as 0.62 is used then Nc is estimated as 6.4. indeed has been observed.
These Nc values are extremely low and could not ordin- In summary, it appears that the new predicative
arily be anticipated with any conventional theory. model for spudcan penetration presented in this section
Before closing this section it is useful to consider is capable of providing an explanation of the behaviour
what the revised model for spudcan penetration would observed in two sets of centrifuge model tests in dif-
predict for the model test results presented by Cassidy ferent soils, as well as the field observations at Yolla.
et al. (2004) and Cassidy (pers. comm.), that were
discussed earlier. With respect to the model tests per-
formed in the kaolin clay, the revised model does 7 CAUTIONARY NOTE
capture the observed behaviour, where the spudcan
resistance was about 15% less than the T-bar. For that The analyses presented above are based on new ideas
case the FR factor can be computed as about 20% that stem from the latest research and experience.
since the difference between the normalised penetra- When applied to the Yolla situation these lead to a
tion rates of the model T-bar and the model spudcan is consistent model, strongly supported by the available
about 2 log cycles while  for the kaolin is about 0.1. data, which provides a neat and convincing explan-
All the kaolin tests were completely undrained and ation of the observed phenomena. In addition the avail-
hence Fd  1. In addition it is believed that a trapped able data does not readily allow any other alternative
wedge could not form with the spudcan geometry explanation (except by a matter of degree).
used for the model tests and hence Nc is zero. The Nevertheless, the author is aware of one case else-
kaolin also has a relatively low sensitivity (about 2 to where offshore Australia (referred to as Site A) where
2.5) and hence it is not anticipated that FA would dif- the new T-bar theory cannot provide a convincing
fer significantly from 1. Assuming also that FE  1, explanation of the observed difference between labora-
the only divergence between the T-bar and the spud- tory strengths and T-bar strengths. This concern is
can for this case is due to the viscous rate effect, and accentuated by the fact that the soil at Yolla and at Site
the computed difference of about 20% compares well A appear to have a number of striking similarities; both
with the observed 15% lower penetration resistance for are carbonate silts and have similar grading, similar
the model spudcan. consolidation parameters, similar net PCPT and T-bar
Slightly less information is available to define the resistance and similar sensitivities. In addition, both
critical soil parameters for the model tests performed soils give poor correlation between undrained strengths
in the North West Shelf carbonate silt, particularly the determined from the T-bar and from laboratory strength
viscous rate effect. However, it is understood that the tests. However, they differ in the two following
viscous rate effect is probably small or non-existent important regards, and it is these that lead to an appar-
and hence FR is likely to be around 1. One problem ently irreconcilable difference for Site A as compared
that has proven consistently difficult to solve with to Yolla soil, where the theory seems to work nicely:
model testing in this material has been that of obtain-
ing a realistic in situ void ratio for the reconstituted 1 For the soil at Site A, a strong viscous rate effect
sediment; generally it is found that centrifuge sam- has been observed in laboratory simple shear tests,
ples have much lower void ratios than the in situ soils, albeit there is some doubt as to the veracity of
and it is believed that this leads to a significantly these results (Goodison 2005). This high rate effect
reduced sensitivity. Strong supporting evidence for offsets the competing effect of the high sensitivity.
this assumption comes from comparing the ratio of If the theory presented in this paper is applied to
first extraction to penetration resistance of the model Site A soils, a T-bar factor of about 12.5 is esti-
T-bars, which gives a value close to 1 (as compared to mated due to the high rate effect which is higher
about 0.2 to 0.3 for the in situ Yolla silts). It is there- than the default value of 10.5. This compares to an
fore believed that FA is unlikely to significantly devi- average T-bar factor of nearer 5 for Site A, derived
ate from 1 in this case either. The same spudcan was from comparing the laboratory strength tests with
used in these tests as for the kaolin clay tests and the T-bar resistance (and this is before any drainage
hence Nc is also zero. Model tests were performed at correction is made, which would make the com-
a variety of different loading rates from fully undrained parison even worse).
to partially drained. The conclusion reached earlier 2 The obvious explanation for this discrepancy (and
that similar T-bar and spudcan penetration resistances one that the author believes in for Yolla) is that
were obtained during these tests was for cases where the laboratory samples have been substantially dis-
both the T-bar and spudcan were fully undrained; turbed at some stage, leading to re-consolidation at
hence Fd  1. For such cases there appear to be no lower void ratio. However, for the soils at Site A
factors which lead to divergence between the T-bar there is very little evidence for this; as compared to

61

Copyright 2005 Taylor & Francis Group plc, London, UK


the Yolla soil the other soil had no shells (which the Yolla A raft supported foundation, the spudcan
act like an extra thick sampling tube wall) and the extraction operation also proved to be of great interest.
samples exhibited no evidence of compaction (in Generally the literature is rather sparse in terms of
the form of water filled voids at the ends of the what is to be expected after spudcan extraction and
sample) when removed from the sample tubes. In the impact that this may have on adjacent structures.
addition, none of the X-rays of the Site A sample For example, while there have been many experi-
tubes showed any visible evidence of disturbance. mental studies on spudcan penetration, few have con-
It is not impossible that the Site A samples were sidered spudcan extraction (one notable exception
actually disturbed. However, if they were, then all being Stewart & Finnie (2001). Even where extrac-
reconsolidation must have occurred before the tion has been addressed during model tests, this has
samples were recovered to the deck of the drilling generally just involved pulling the spudcan out as a
vessel and before the samples were cut and end straight reversal of the installation. However, in prac-
waxed. Furthermore all of this must have occurred tice, real jackups are mostly unable to apply large ten-
with no visible signs on the X-rays. Overall this sile loads to the spudcans and hence field procedures
scenario does not appear very plausible. generally involve jetting and working of the spudcan
to destroy the soil above the spudcan and to break
The fact that the soil at Site A does not fit the the-
down the suction between the underside of the base
ory suggests that there is still something fundamen-
and the soil. Intuitively it seems obvious that there is
tally missing; either some of the data is just incorrect
potential for the formation of a much deeper crater
(for an unknown reason) or there is a hidden factor
under these conditions compared to what would be
which has not been accounted for.
expected where extraction is through direct pulling of
The fact that the theory and the data provide an
the spudcan. Over time the steep side walls of such a
internally consistent argument for Yolla provides strong
crater are likely to collapse, potentially affecting a
support for the veracity of the explanation provided in
wide area.
this paper. But the inability to also explain the behav-
The Yolla extraction data confirms these hypoth-
iour at Site A suggests that an unknown factor at Yolla
eses. The key features of the extraction operation were:
may not have been identified. With the available data
and the current state of the art, there appears to be lit-
1 The operation proved to be very time consuming
tle prospect of reconciling this difficulty. To achieve
(approximately 7 days) due to the considerable time
this, significant extra research will be required and
required to break-down the soil resistance on each
this will inevitably need to include the collection of
spudcan. During the first three days, the Port and
more field data at various sites.
Bow spudcans were released but little impact was
made on the Starboard spudcan despite injecting
8 50-YEAR STORM STABILITY AT YOLLA up to 1200 gallons per minute of water above and
below this spudcan. The weather then deteriorated
Even after the spudcans had achieved their final pene- to the point where the rig had to jack back down
tration depth and had successfully sustained the max- again, where it remained for another three days. As
imum preload, it was not immediately obvious that a the weather started to improve, the third leg was
safe penetration depth had been reached with respect finally released through a combination of jetting
to stability under the design 50 year return period and load cycling acting on the jackup hull from the
storm. This is because the PCPT and T-bar profiles both decreasing seas. Despite having being jacked back
exhibited a strong cut-back in resistance below about down after their initial release, the other two legs
25 m and because the strength testing performed dur- also broke free quite quickly.
ing design of the Yolla A platform indicated that the 2 As hinted above, there was a clear correlation
soil had a high susceptibility to degrade under cyclic between the rate of release of all three spudcans and
loading. However, through additional analyses it was the seastate. When the hull was in the water and the
possible to demonstrate that the achieved penetrations swell was relatively high, the swaying motion of the
were in fact acceptable. These analyses were similar rig resulted in the spudcans being cycled. Conversely,
to those described later in this paper for the Trefoil periods of calm seas and low swells were often asso-
location, where the effect of cyclic loading was even ciated with little or no progress.
more important than at Yolla. 3 After initial release the spudcans were generally
pulled out of the soil with little further resistance.
However, at a number of locations, a temporary
9 SPUDCAN EXTRACTION AT YOLLA increase in the pullout resistance occurred, neces-
sitated further jetting of the soil. These incidents
Due to the substantial penetration depths achieved by coincided with the spudcans passing through
the spudcans at Yolla and the very close proximity of layers which exhibited a high T-bar resistance, and

62

Copyright 2005 Taylor & Francis Group plc, London, UK


are therefore suggestive of a strongly dilative soil that at least two back-scarps associated with slump
response in these materials. failures were observed a major one with a radius of
4 After retraction of the legs to the seabed an approxi- about 14 m and a minor one at a radius of about 21 m.
mately 10 m high soil plug was found on top of the Unfortunately, no reliable estimate of the crater depths
Port and Starboard spudcans (Fig. 16). A similar could be made, but the current best estimate for the
plug was also noted on top of the Bow spudcan, but Port spudcan is about 4 m from the outer edge to the
the height of this plug was not recorded. centre, albeit this is qualified with a large degree of
uncertainty. Broadly similar craters were found at the
The relative ineffectiveness of the jetting in this
other two locations, although the Starboard crater
case appears at least partly attributable to the layout
appeared deeper, but with only the inner main scarp
of the underbase jets, which comprised only a single
evident, whilst a larger diameter (35 m) but shal-
ring of jets near the outer perimeter of the spudcan. This
lower main crater was found at the Bow spudcan.
design precluded the injection of any water directly
Given the height of soil found on top of the recovered
beneath the centre of the baseplate which should be a
spudcans it is evident that the initial crater depths
much more effective method of breaking the hydraulic
were much deeper, possibly around 12 m, albeit this is
seal between the soil and the spudcan base.
once again a qualified estimate.
Almost 2 months after the spudcans were extracted
Empirical methods have been developed for
an ROV survey was performed of the seabed sur-
assessing the interaction between a spudcan and
rounding the Yolla A platform, with particular focus
pile foundation after spudcan penetration (eg.
on the three craters resulting from the Ensco 102
Siciliano, et al. 1990). Based on such results it is gen-
jackup operations. Substantial craters were found at
erally assumed that the significant zone of influence
each spudcan location; Figure 17 presents a schematic
is no more than about one spudcan radius from the
cross-section through the Port crater. It may be noted
edge of the spudcan. However, the craters observed at
Yolla show a much larger zone of influence after
spudcan extraction and after allowing for the result-
ing craters to collapse and infill. Analyses were there-
fore performed, using the computer program FLAC
(Itasca 1996), to assess the likely consequences of an
infilling crater on the adjacent Yolla A platform. For
simplicity, an axisymmetric model was used for this
assessment and hence the raft foundation wraps
around the spudcan crater; clearly this is more oner-
ous than the real situation. For these analyses three
different soil models were considered:
1 Unconsolidated undrained conditions throughout
the mesh, except for the sand layer at around 3.5 m
depth which was assumed to be fully drained.
2 Drained conditions throughout the entire mesh
(soil friction angle  40), except for undrained
Figure 16. Soil plug on spudcan after extraction at Yolla.
clay layers between 1.5 m and 3.5 m and between
15 m and 19.5 m.
3 Drained conditions throughout the whole mesh
Yolla A Platform Raft (soil friction angle  40).
0 10 20 30
Distance in metres These represent (simplistically) the different stages
0

in the collapse process, where the soil gradually tran-


sitions from a fully undrained to a fully drained state
(assuming that complete collapse and infilling has
10

Estimate of potential not occurred in a previous, less drained, step).


depth backfilled at
time of spudcan extraction
The failure mechanisms that develop in the three dif-
20

Initial spudcan hole


ferent soil models are illustrated on Figure 18. It may
be seen that even with an assumed 24 m deep initial
Diameter = 18.2 m hole, collapse of the crater was predicted to be confined
30

Depth in metres to soil outside the raft skirt perimeter in all cases. The
analyses also indicated very small raft settlements
Figure 17. Estimated crater geometry after extraction (Port and only a minor redistribution of the vertical stress
spudcan at Yolla). away from the raft edge as the crater collapsed.

63

Copyright 2005 Taylor & Francis Group plc, London, UK


Raft Foundation Raft Foundation Yolla soil and a review of the Trefoil SI data, we
advised that this was a highly unsatisfactory state of
affairs and that it was improbable that the two shallow
legs of the jackup could sustain a 50 year storm with-
out initiating a cyclic punch-through failure, whereby
the foundation would progressively settle once the
storm loading exceeded a certain magnitude. We rec-
ommended that the spudcan penetration should be
Step 1: All undrained Step 2: All drained except increased significantly in order to attain a safe condi-
clay layers
tion. After some discussion it was agreed by all par-
Raft Foundation ties that the two high legs would be worked through
a combination of jetting and cycling of the preload
between zero and 62 MN (ie. just under 65% of the
previously applied maximum preload) in order to
attempt to advance the legs, with the legs pulled free
of the seabed during the low end of each cycle.
Cycling and jetting was first initiated on the Port
leg, but the Starboard leg was also subject to simultan-
Step 3: All drained
eous cyclic loading. After around 3 hours of cycling
at a rate of about 1 to 2 cycles per hour the weather
Figure 18. Mechanisms of hole collapse after spudcan
extraction at Yolla.
deteriorated and it was necessary to stop working and
to jackup in order to clear the rising seas. At this stage
there was no evidence that either of the two high legs
The ROV survey evidence supports the FLAC had increased their penetrations. As the weather started
analysis, with no evidence of slump failures propa- to improve, the jackup hull was lowered into the water
gating beneath the raft skirts. again, and this was combined with jetting of the Port
leg in order to try and pull this leg free of the base of
its hole; wave loading on the hull from the subsiding
10 OBSERVED SPUDCAN PENETRATION sea was believed to be significant during this oper-
AT TREFOIL ation. After the Port leg was pulled free, the rig was
jacked back up out of the water (leg load approxi-
After completing operations at Yolla the Ensco 102 mately 53 MN to 62 MN) and during this operation
moved 38 km to the west to a new site, Trefoil. The the penetration of the Starboard leg increased to 9.8 m
penetration and preloading operation at Trefoil proved and the Port leg increased to 8.5 m, while the Bow leg
to be unusual, complex and time consuming. This was reported to be at 13.4 m of penetration. The full
operation is summarised in Table 5, and discussed in preload was slowly applied to the Port leg and then
more detail below. held for several hours, but no further increase in leg
On 17 October 2004 the rig arrived on location. All penetration occurred. Under the 100% preload, jet-
the legs were then jacked down under a maximum ting was then performed under the Port leg for about
load of 62 MN and attained 4.9 m of penetration (to half an hour, but again no increase in leg penetration
the base of the spudcan) under all legs. Preloading could be achieved and therefore some of the preload
was then commenced under the Bow leg which then was dumped and the rig was jacked down into the
penetrated to 6.4 m under a preload of 83 MN, which water while holding a leg load of 62 MN.
was 85% of the maximum preload. This leg then After further discussion, it was agreed that jetting
punched through to 12 m penetration, causing a 5 tilt was probably hindering rather than helping the leg
of the rig towards the Bow leg. The preload was penetration process. While jetting has been used
dumped, the rig levelled and a second stage of pre- successfully to advance spudcans through thin sand
loading was then commenced on the Bow leg, which layers in the past (eg. Baglioni et al. 1982), at Trefoil
led to a final penetration of 12.8 m under the max- the location of the jets and the nature of the soil pre-
imum preload of 98 MN. cluded significant erosion instead it is more likely
During the following 24 hours the Port and that a downward flow gradient was set up through this
Starboard legs were both preloaded, but neither could material leading to a reinforcing rather than degrad-
be advanced beyond a penetration depth of 6.1 m, ing effect. It was therefore decided to use only cyclic
even under the full preload of 98 MN. Origin Energy loading to try and advance the spudcans further. The
(the field operators) then sought advice from the Starboard and Port legs were therefore cycled in turn
authors company on the acceptability or otherwise of between loads of nominally zero (maybe a small ten-
this situation. Based on previous experience with the sion) and 62 MN, with the spudcan pulled free during

64

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 5. Spudcan penetration and preloading at Trefoil.

Penetration

Action Port (m) Star. (m) Bow (m)

17 October 2004 Ensco 102 Jackup rig arrived on location


All legs jacked down; 62 MN load 4.9 4.9 4.9
Preloading of Bow leg up to 83 MN 4.9 4.9 6.4
Punch-through of Bow leg at 83 MN 4.9 4.9 12.0
Dump preload then reapply to Bow leg up to 98 MN 4.9 4.9 12.8
Preload port and starboard leg to 98 MN 6.1 6.1 12.8
Jetting and cycling (0 to 62 MN) of Port leg for 3 hours 6.1 6.1 12.8
Wait on weather; hull jacked clear 6.1 6.1 12.8
Weather improves; hull jacked into water; Port leg jetted and pulled free 6.1 6.1 12.8
Hull jacked out of water; leg load  53 MN to 62 MN 8.5 9.8 13.4
Full preload (98 MN) applied to Port leg; jetting at 98 MN preload 8.5 9.8 13.4
Cycling of both Port and Starboard spudcans for 16 hours (0 to 62 MN) 9.1 12.5 13.4
Static preloading of Port leg at 98 MN 9.1 12.5 13.4
Cycling of Port leg for 11 hours (0 to 53 MN) plus wave load on hull 10.1 14.3 14.0
Wait on weather 10.1 14.3 14.0
Further cycling of Port leg for 6 hours 10.1 16.5 14.0
Alternative cycling of Port leg (62 MN to 84 MN) for 11 hours 10.1 16.5 14.0
Ensco elect to reposition rig pivot about Bow leg 14.0
Port leg into old Starboard hole; preload to 62 MN 12.2 14.0
Virgin location for Starboard leg; preload to 98 MN 12.2 7.9 14.0
Preload immediately dumped; leg pulled free; preload back to 98 MN. 12.2 11.6 14.0
Preload immediately dumped; hull jacked clear of water 14.0 13.7 14.0
Maximum preload applied to Starboard leg (98 MN) 14.0 14.0 14.0
Final preloading on all legs (48 hours) combined with severe weather 14.0 14.0 14.0
Ensco 102 declared ready for operation on 28 October 2004 14.0 14.0 14.0

the low part of each cycle. This process was repeated return period storm, which was based on an assumption
6 times in succession over a 6 hour period, which led that the soil would be fully consolidated under the
to the Starboard leg penetration increasing by 1.2 m static load imposed by the jackup rig on the seabed
but the Port leg only advanced another 150 mm. This (these calculations will be discussed in Section 13).
process was continued several more times over the After considering this advice it was agreed that the
next 10 hours, albeit broken up with some operational static preloading should be aborted and the leg cycling
delays, and this gradually increased the Starboard leg recommenced.
penetration to 12.5 m and the Port leg to 9.1 m. Further cycling of the Port leg was undertaken over
Static preloading of the Port leg was then reat- an 11 hour period between loads of zero and about
tempted but during discussions with all interested par- 53 MN (ie. no preload) but also using wave action on
ties, it was agreed that this was unlikely to help the the hull to increase the cyclic action acting on the foun-
situation since the soil was consolidating and strength- dations. This increased the Port leg to 10.1 m penetra-
ening under the static load, thereby reducing the likeli- tion, but the Starboard leg increased by a much more
hood that the leg could be worked down further under significant degree to 14.3 m penetration, despite only
the cyclic preloading regime that could be physically receiving incidental cyclic loading (albeit this was
imposed. In addition, preliminary analyses had been believed to be of similar magnitude to the Port leg
performed that suggested that the spudcan penetrations cyclic loading). The Bow leg was reported to be at
needed to be in the order of 12 m or more in order to 14.0 m of penetration. After about 5 hours of waiting
achieve an adequate factor of safety for the 50 year on weather, with the hull jacked out of the water,

65

Copyright 2005 Taylor & Francis Group plc, London, UK


static preloading of the Starboard leg was commenced, The behaviour of the spudcans at Trefoil shocked
but after further discussion, it was decided to con- most of the people involved who had never encoun-
tinue working the Port leg instead, in order to try and tered such conditions before; where the application of
advance it to a safe working depth. However, a further cyclic pre-loading could more than double the initial
six hours of cyclic loading only managed to advance penetration of the spudcans. Similar behaviour has
the Starboard leg (to 16.5 m penetration), but left the been seen in model tests (e.g. Finnie 1993) but, to the
Port leg unchanged at 10.1 m penetration and the Bow authors knowledge, has not been reported before in
leg at 14.0 m. the field. It is particularly noteworthy that once a
After further discussions, it was decided to attempt cyclic failure had been initiated it was not readily
an alternative preloading strategy for the Port leg stopped (consider the Starboard leg in particular),
whereby the leg would be cycled between preloads of even where the amplitude of the cyclic loading was
62 MN and 84 MN, where the latter load was a bit subsequently reduced. This suggests a trigger mech-
higher than the maximum load-factored vertical load anism, very similar to that observed in model tests of
that would be imposed on the spudcan during the 50 bucket foundations on silica sand (Bye et al. 1995);
year storm. Over an 11 hour period, 7 cycles of such only through curtailing all cyclic loading and allow-
loading were applied, but (due to the slow cyclic rate) ing time for at least partial dissipation of accumulated
this was unable to advance any of the spudcans further. pore pressures was it possible to halt the development
Having exhausted all practical options to advance of progressive settlements.
the Port leg to the preliminary recommended depth, Fortunately (and to the due credit of all concerned)
Ensco elected to reposition the rig. This would be the geotechnical advice was heeded despite the sig-
achieved by pivoting the rig around the Bow leg, nificant extra costs incurred due to the time consum-
respudding the Port leg in the existing Starboard ing spudcan installation. The site investigation data
leg hole and making a new hole for the Starboard leg. from Trefoil combined with that from Yolla enabled
Due to adverse weather, a delay of 1.5 days was the observed mechanism to be successfully predicted
incurred before this operation could be commenced. in advance. However, the key question then arises; at
After lifting all the legs and rotating the rig, the Port what penetration depth may the spudcan foundations
leg was jacked down into the old Starboard hole and be deemed to have attained a safe depth in order to
then penetrated to 12.2 m under about 62 MN preload. satisfactorily resist the design environmental loads
Preloading was then commenced on the Starboard leg from a 50 year return period storm, without triggering
which attained a penetration of 7.9 m under the max- a cyclic failure? In the following sections the soil
imum preload of 98 MN. The preload was immedi- conditions at Trefoil are discussed and compared to
ately dumped and the Starboard spudcan pulled free the conditions found at Yolla. The reasons for the
before recommencing preloading. This caused the unusual installation behaviour will then become appar-
Starboard leg penetration to increase to 11.6 m after ent. The analyses performed to justify the 50 year
the maximum preload of 98 MN had been reapplied. storm stability for the Ensco 102 at the Trefoil loca-
The preload was dumped again and then the rig was tion will then be discussed. A number of important
jacked up to a 2 m air gap which caused the Port leg adjustments to the standard SNAME procedure are
penetration to increase to 14.0 m and the Starboard necessary in order to address the behaviour of soils
leg to increase to 13.7 m. Maximum preload was then such as those found at Trefoil, which are subject to
applied to the Starboard leg which caused the pene- severe degradation of strength under cyclic loading.
tration to increase to 14.0 m.
Over the next 48 hours, the maximum preload of
98 MN was first held on each leg for at least six
11 SOIL DATA AT TREFOIL
hours, in accordance with the requirements of the
Warranty Surveyor. During the latter part of this oper-
11.1 Site investigation
ation, some fairly severe weather occurred with reported
seas of 6 m to 7.5 m significant wave height and A site investigation was performed by Benthic Geotech
winds of 25 knots, gusting to 35 or 45 knots. During at the Trefoil location during September 2004 (Kelleher
this latter period a maximum leg load of up to 77 MN & Randolph 2005). This investigation comprised a
was imposed on each spudcan leg, which comprised comprehensive programme of static penetrometer
both preload and environmental load components; the tests (PCPT), and ball penetrometer, (BPT), cyclic
preload was adjusted to the prevailing weather during BPT tests, and PCPT dissipation tests. In addition, 2
that time based on wind and wave leg load charts sampling boreholes were also undertaken to a max-
prepared by Ensco. The spudcan penetration remained imum depth of 27.4 m below the seabed. All four of
unchanged at 14.0 m on all legs and the rig was the PCPT probes, including one beneath each spud-
therefore declared ready for operation on the 28 can location, revealed that the Trefoil site is very uni-
October 2004. form with minimal lateral variability from hole to hole.

66

Copyright 2005 Taylor & Francis Group plc, London, UK


qnet (MPa) qnet (MPa) ch (m2/yr)
0 5 10 0 1 2 3 4 5
0 0 1 100 10000 1000000
nd 0
Trefoil Silty sa Yolla
5 Clay 5
Sand 5

Depth (m)
Depth (m)

10 10

Depth (m)
10
15 t 15
dy sil
/ San 15
sand
Silty
20 Clay 20
20
25 25 TRE-01 Ir = 200
CPT TRE-01
Figure 19. Comparison of soil types Yolla and Trefoil. 25
0 2 4 6 8 10 12
qnet (kPa)
However, a number of different material types were
encountered over the depth range examined. Figure 20. PCPT Dissipation test results at Trefoil.
At face value the soil at Trefoil appears signifi-
cantly different (and stronger) compared to that at rem
the Yolla location 38 km to the East. However, closer 0 0.1 0.2 0.3 0.4
inspection suggests that the soil stratigraphy is actu- 0
ally very similar. This is indicated on Figure 19 which
shows that the two clay layers at Yolla can be traced
through to Trefoil, but are located typically 3 m to 4 m 5
deeper. In addition the carbonate sand layer just below
the upper clay layer at Yolla is also present just below
Depth (m)

10
the same layer at Trefoil. The material below this sand
layer and above the lower clay comprised carbonate
sandy silt at Yolla, but is more like carbonate silty sand 15
at Trefoil. This gradational difference is the reason for
the significantly higher PCPT resistance obtained in
this layer at Trefoil compared to Yolla. At Yolla, we saw 20
earlier in this paper that the PCPT resistance in the
carbonate sandy silt exhibits a small degree of partial 25
drainage but at Trefoil the degree of partial drainage is
much greater due to the slightly coarser soil grading Figure 21. Soil sensitivity from BPT at Trefoil.
and this leads to a substantially higher PCPT resist-
ance. It is reasonable to assume that this difference in One of the main features of the non-clay soils at
soil grading at the two sites has come about due to a Yolla were their very high susceptibility to severe
slightly higher energy depositional environment in the strength degradation during cyclic T-bar tests. At
shallower water at Trefoil. Trefoil the cyclic BPT tests suggest that the silty sand
The enhanced degree of drainage at Trefoil is between the two clay layers is similarly susceptible to
clearly evident on Figure 20, which presents the cyclic degradation. This is evident on Figure 21, which
deduced coefficient of consolidation (ch) from the summarises rem recorded during these tests at Trefoil,
PCPT dissipation test results at Trefoil. This may be which may be compared with the equivalent results
compared with the equivalent data from Yolla (Fig. 2). obtained with the T-bar at Yolla (Fig. 3). It may be
It may be seen that in the silty sand at Trefoil, the seen that very low values (ie. 0.05) were recorded at
average value for ch is around 20,000 m2/yr whereas both sites and hence both the Yolla and Trefoil non-
in the sandy silt at Yolla it is more like 7000 m2/yr. In clay soils would be defined as highly sensitive.
addition there is at least one sandier layer at Trefoil
(between depths of 11 m and 13 m) where the PCPT
11.2 Undrained shear strength
exhibited a fully drained response, which indicates
that ch is at least 250,000 m2/yr. In the clay layers at 11.2.1 In situ monotonic strength
both Yolla and Trefoil, ch is several orders of magni- No laboratory strength tests were performed on sam-
tude smaller at around 20 m2/yr. ples of the Trefoil soil. However, a direct assessment

67

Copyright 2005 Taylor & Francis Group plc, London, UK


of the in situ monotonic undrained strength (su-mono) experience with other carbonate silty sands and sandy
can be made from the BPT data, by dividing the silts the monotonic undrained strength was defined as
measured penetration resistance by an appropriate su-mono  4z, where z is the depth below mudline in
factor (Nball): metres, and su-mono is in kPa.
The monotonic undrained strength derived from
this relationship is typically about double that deter-
(9) mined for the Yolla sandy silt. However, much experi-
mental data from various carbonate soils shows that
the monotonic undrained strength tends to be strongly
Due to the similarity between the Yolla and Trefoil
dependent on the grading of the soil with lower mono-
soils, it was considered reasonable to apply the same
tonic undrained strengths obtained for soils with
factors as were found to be applicable at Yolla. At
higher fines content. Hence the average undrained
Yolla, the T-bar was used instead of the BPT, but as
strength for the Trefoil silty sand is expected to be
discussed earlier, most experimental data indicates
higher than at Yolla, although it should also be appre-
that the ball and T-bar generally exhibit similar resist-
ciated that some thin layers may be weaker and some
ance. Hence it was considered reasonable to apply the
stronger than defined by this simple profile.
Yolla T-bar factors (Nbar) to the BPT results at Trefoil
For the clay layers, the average trendline deter-
(ie. Nbar  Nball  7.65 for the carbonate sandy silt/
mined using Nball  13.9 is best fitted with an aver-
silty sand and Nbar  Nball  13.9 for the non-
age trendline defined as su-mono  2.2z, and this
carbonate clay layers).
strength profile was adopted for the storm stability
Figure 22 presents the undrained strength derived
analysis described later. Further discussion on the
by applying the Yolla derived factors to the BPT
appropriateness of this strength profile is presented in
results from Trefoil. However, it should be remem-
Section 14.
bered that throughout all of the carbonate silty sand
the BPT exhibited drained or partially drained behav-
iour and hence the undrained strength derived in this 11.2.2 Normalised cyclic strength
manner is not realistic. Nevertheless, troughs in An extensive programme of cyclic soil strength tests
the data occur at a number of depths, which are was performed to enable design of the Yolla A plat-
believed to be associated with siltier and hence close form foundations. Due to the inherently similar soil
to undrained layers. Considering these troughs, types at Yolla and Trefoil, it was also considered
the inherent degree of drainage and the authors appropriate to assess cyclic strength properties for
Trefoil using appropriately normalised Yolla data.
su (kPa)
Some of the interpreted data from the Yolla tests is
presented in Watson & Humpheson (2005). However,
0 50 100 150 200
0 the author has generally found that it is very useful to
present the test data in a different manner; the nor-
malised cyclic strength (su-cyc/su-mono) as a function of
5 the normalised monotonic strength (su-mono/vo). The
Clay (Nball = 13.9) cyclic strength (su-cyc) may be defined for any pre-
determined equivalent number of cycles of the design
10 peak load at a given shear strain level (in this case a
single amplitude shear strain of 10%). Presentation of
Silty Sand (Nball = 7.65)
the test results in this way directly addresses the effect
of changing density and confining stress level for any
Depth (m)

15
soil; high density and/or low confining stress will
generally lead to high values of su-mono/vo. The gen-
20 eral form of the relationship shown is defined as:
Clay (Nball = 13.9)

25 (10)

30
k = 2.2 kPa/m In this case it is necessary to assess the cyclic
strength for the soil supporting the spudcan founda-
k = 4 kPa/m tions during the 50 year return period storm. As a
35
basis for this assessment the strengths obtained from
Figure 22. Monotonic undrained strength from BPT. simple shear tests with full 1-way cyclic loading were

68

Copyright 2005 Taylor & Francis Group plc, London, UK


considered. Thirteen equivalent cycles of the maxi- soil strength may be modelled using the following
mum design load was found to be an appropriate rep- general relationship:
resentation of the entire storm. The derived 1-way
strength for 13 equivalent cycles is presented in a nor- (11)
malised form on Figure 23.
It may be noted from Figure 23 that the soil types
where D is an empirical coefficient, and n is a curve
most susceptible to cyclic degradation (ie. those with
fitting parameter that fits the laboratory data. Testing
the lower ratios of cyclic strength to monotonic
of a wide variety of such soils has revealed that n typ-
strength) are the stronger materials (ie. those with
ically lies between
0.4 and
0.6 provided that the
higher ratios of monotonic strength to initial vertical
in-situ value of su-mono/v is less than about 2.5 (which
effective stress). This is part of the reason why the
is applicable in this case). No specific tests were per-
spudcan foundations were in a meta-stable and quite
formed to determine n for either the Yolla or Trefoil
unsafe position after penetrating to 6.1 m, despite
carbonate soil and hence the more conservative value
holding the full static preload at this depth.
of
0.6 was adopted for the analyses.
Full 1-way cyclic loading is essentially the type of
D is derived from known values of su-mono and v
cyclic loading imposed on the spudcans during the
(i.e. the in situ values) at the depth of interest. The
cyclic preloading operations that were used to work
consolidated value of su-mono at any depth may then be
the spudcans down to 14.0 m penetration at Trefoil.
readily determined by substituting the applied
However, the design loads acting on the jackup during
enhanced stress level back into the equation, along
the 50 year storm show a lesser degree of cycling; the
with the deduced value of D.
ratio of cyclic to average shear stress was deduced to
be only about 0.2 as compared to a ratio of 1 for full
1-way cyclic loading. The normalised cyclic strength
relationship for this degree of cycling was determined 12 MODIFICATIONS TO SNAME PRACTICE;
based on linear interpolation between the 1-way cyclic CYCLICALLY DEGRADABLE SOILS
strength data from Yolla and the monotonic strength
(no degradation). The derived relationship is also The SNAME recommended practice (SNAME 2002)
shown on Figure 23. for assessing the geotechnical performance of spudcan
foundations under storm conditions is based on using
11.2.3 Consolidated strength the static preload as a normalising parameter for a
As discussed above, the coefficient of consolidation geotechnical yield envelope that defines the inter-
for the carbonate soil layers at Trefoil is high. Hence action between vertical, horizontal and moment capac-
these layers consolidate and gain in strength suffi- ities. For most sands and clays, this approach is safe
ciently rapidly under the imposed static weight of the since degradation of the soil strength under cyclic load-
jackup rig to enable the fully consolidated strength to ing is modest. However, as already discussed, carbonate
be adopted in the assessment of the 50 year storm sta- soils like those at Yolla and Trefoil exhibit substantial
bility. However, no significant consolidation of the degradation of strength when subject to cyclic load-
non carbonate clay layers was expected. ing and for such soils the standard SNAME approach
Laboratory testing carried out on various carbon- is unsafe. The SNAME practice does in fact recognise
ate soils indicate that the effect of consolidation on this fact and includes a specific warning with respect
to cyclically degradable soils such as silts and/or car-
bonate materials. However, no further guidance is
1
Fugro-Extrapolated (Full 1-way) given on how to deal with these materials.
Fugro (Full 1-way) Hence, for cyclically degradable soils, the standard
Benthic Sample (Full 1-way)
0.8 SNAME practice needs to be modified and for this
Ensco 102; 50 yr storm; cyc /ave = 0.2
(su-cyc /su-mono)

purpose the approach presented on Figure 24 is pro-


0.6 posed. In the upper part of this figure, the square
symbol indicates the depth at which the in situ mono-
0.4 tonic soil strength first becomes just sufficient to sup-
port the jackup preload and the circular symbol
0.2 indicates the cyclic soil strength that is available at
Full 1-way loading; cyc /ave = 1 this depth. Based on the monotonic strength, in the
0 lower part of Figure 24 the yield envelope under
0 0.5 1 1.5 2 2.5 monotonic loads indicates the limit of the range of
sumono /vo' vertical/horizontal load combinations within which
no further penetration will occur. The corresponding
Figure 23. Normalised cyclic strength data from Yolla. yield envelope for cyclic soil strength is indicated by

69

Copyright 2005 Taylor & Francis Group plc, London, UK


Undrained shear strength The analysis of consolidated pure vertical cyclic
bearing capacity may be performed using various
Average monotonic undrained
strength that just supports approaches. Randolph & Erbrich (2000) present
jackup weight at given depth results for typical cases using a limit analysis program
based on a 3-wedge optimised upper bound plasticity
mechanism. The main limitation of this program is
su-mono Fcyclic = su-cyc/su-mono that it can only deal with one material type. Hence in
su-cyc
this case, the results obtained would only be reliable
up to the point where the clay layer below 20 m depth
starts to influence the results.
The influence of the underlying clay layer can be
Depth
addressed approximately using the methods presented
in Brown & Meyerhof (1969) and Meyerhof & Hanna
Cyclic strength compatible with
required average monotonic strength (1978). Both of these methods apply to 2-layer sys-
Vertical Load
tems with an upper layer comprising a strong homo-
geneous undrained soil and a lower layer comprising
Vertical preload, VP
a weaker homogeneous undrained soil. At Trefoil, the
Yield envelope under assumption of two uniformstrength layers does not
monotonic preload
Equivalent preload =Fcyclic x Vp
apply, since the consolidated undrained strength of the
overlying material (carbonate silty sand) has a com-
Unsafe zone
plex spatial distribution, varying in both the vertical
and lateral directions while the underlying clay has a
Degraded cyclic strength profile that increase linearly with depth. In
yield envelope order to use this approximate method it is therefore
Safe zone (shaded) first necessary to determine equivalent uniform
strengths for the two layers that match the capacities
Material factored
cyclic yield envelope
obtained for each layer individually, using more
accurate methods (eg. the aforementioned limit
Horizontal Load
analysis program for the upper consolidated layer and
Salencon & Matar (1982) for the lower clay layer).
Figure 24. Modified yield envelope approach for cyclically
degradable soils. Due to the approximate nature of the results
obtained using these approaches and the criticality
and sensitivity of the results to the proximity of the
the degraded cyclic yield envelope. The safe zone, clay layer, the results from these simpler models were
bounded by the broken line in the lower part of Figure verified using finite element type analysis. These
24, is determined by applying a material factor to were performed using the program FLAC (Itasca
the cyclic yield envelope. In the standard SNAME 1996). This program can explicitly handle any gener-
approach the material factor would have been applied alised strength profile.
directly to the monotonic yield envelope, leading to a
safe zone that in reality includes unsafe combin-
ations of load. 13.2 Cyclic strength profiles
Hence to summarise, in the standard SNAME Combining the in situ monotonic strength relation-
practice the normalising parameter for the yield envel- ship (su-mono  4z) with the normalised cyclic strength
ope is the applied static preload whereas in the modi- relationship (Equation 10), the cyclic undrained
fied approach the normalising parameter is the strength for in situ conditions may be derived, as
calculated pure vertical cyclic bearing capacity (equiva- shown on Figure 25. Similarly, after consolidation
lent preload). under the static stress imposed by the jackup weight
(219 kPa) the cyclic undrained strength beneath the
centre line of the spudcan may be determined using
13 50-YEAR STORM STABILITY AT TREFOIL
equations 10 and 11, as also shown on Figure 25. At
all other points in the soil, the consolidated cyclic
13.1 Analysis methods
strengths are determined by consideration of the spe-
To assess the cyclic bearing capacity of the Ensco cific imposed stresses caused by the spudcan weight
102 spudcans at Trefoil it is necessary to address both this is performed automatically with the analysis
the strength of the consolidated carbonate silty sand approaches described above.
and the influence of the weaker unconsolidated clay Figure 26 presents the cyclic undrained strength
layer between 20 m and 22.5 m below the seabed. distribution computed with FLAC assuming that the

70

Copyright 2005 Taylor & Francis Group plc, London, UK


su (kPa) su (kPa) Material Factor
0 20 40 60 80 100 0 20 40 60 80 100 0 0.5 1.5 2
5 5
1
(cyc /ave = 0.2) (cyc /ave = 0.2) 0
Method 1; Proprietary spreadsheet program
10 10 Method 2b; Meyerhof & Hanna
Method 2a; Brown & Meyerhof
Depth (m)

Depth (m)
Method 3; FLAC finite element
15 15 5

Depth to Tip of Spudcan (m)


m = 1.18
20 20
m = 1.25
25 25
10
Insitu Monotonic
Insitu Cyclic
100% Consolidated Monotonic
100% Consolidated Cyclic m = 1
Spudcan Tip Embedment
Figure 25. Monotonic and cyclic strength profiles for 50 15
year storm stability analyses at Trefoil.

20
JOB TITLE : (*10^1)
FLAC (Version3.30) -.600

LEGEND Static load = 219 kPa


11/05/2004 13:07 -1.000 25
step 240040
-1.514E+00 <x<2.279E+01
-2.926E+01 <y<-4.955E+00
cohesion
2.00E+01
40 -1.400 Figure 27. Results from cyclic stability analysis.
4.00E+01 80
6.00E+01
8.00E+01 Silty Sand 80 60
1.00E+02
1.20E+02
-1.800
1.40E+02

Contour interval= 2.00E+01


Clay
A summary plot showing the results obtained is
-2.200
presented on Figure 27. It may be seen that at the final
100
-2.600
achieved spudcan tip depth of 14.0 m, virtually the
Silty Sand
same result was obtained irrespective of the analysis
Advanced Geomechanics
Perth, WesternAustralia
.200 .600 1.000 1.400 1.800 2.200
method used, with a material factor of about 1.3 indi-
(*10^1)
cated. The normal SNAME practice requires resist-
ance factors (ie. 1/material factors) of 0.9 or 0.85 (ie.
Figure 26. Undrained strength distribution in FLAC model.
material factors of 1.11 to 1.18), which are lower than
the values that would be adopted for conventional
bottom of the spudcan had penetrated to 14.0 m and geotechnical design (ie. 1.25 to 1.3). However,
allowing for 100% consolidation in the carbonate soil SNAME notes that these lower material factors are
layers under the static preload. predicated on the fact that under normal conditions
the static preload essentially load tests the foundation
to failure. Given that this condition does not apply
13.3 Analysis results when cyclic strength degradation is significant, it can
reasonably be argued that the conventional factors
The SNAME factored 50 year storm design loads act- should be applied rather than the reduced SNAME
ing on a single spudcan at Trefoil were; vertical  values. However, SNAME provides no definitive guid-
80 MN, horizontal  2.57 MN and zero moment (ie. ance on this matter so it is left to the operator and/or
pinned footing assumed). The horizontal load was any certifying authority to make the final assessment.
sufficiently small as to have little influence on the Finally, it should be appreciated that the analysis of
bearing capacity at the base of the spudcan and hence cyclic capacity presented here includes two inherent
was ignored. Material factors were therefore deter- conservatisms, which could be addressed in the future:
mined against cyclic punch-through failure using
the methods outlined above to determine the equiva- 1 The spudcan was treated as a non-embedded sur-
lent vertical preload and dividing this value by the face footing. As demonstrated at Yolla, the net bear-
factored vertical design load. ing capacity factor determined immediately after
It should be appreciated that even finite element penetration was similar to that expected for a non-
type analyses are subject to error and from verifica- embedded footing, due to remoulding of the highly
tion cases using simpler soil profiles it was deter- sensitive soil. However, re-consolidation of this
mined that the model used was likely to over predict remoulded soil would lead to some enhancement
the true capacity by about 3%. Hence a correction of the cyclic capacity over time, even without any
factor of this magnitude was applied to all the com- extra consolidation strength gain arising from the
puted FLAC results. static vertical load acting on the spudcan.

71

Copyright 2005 Taylor & Francis Group plc, London, UK


2 Full backflow was implicitly assumed; while this is layers at Trefoil was a little lower than assumed, but it
probably close to reality it is likely that a shallow is hard to envisage that it could have been as low as
crater existed after penetration, thereby leading to required to justify the observed Yolla spudcan pene-
a small extra overburden component contributing tration. Strain rate effects may provide a partial explan-
to the cyclic bearing capacity. ation for the apparent discrepancy with slower rates
of loading applied during the static penetration of
the Yolla spudcans as compared to the cyclic pene-
14 CLAY LAYER CONTRADICTIONS tration operation performed at Trefoil. During a
storm event, cyclic loading rates are much higher than
Earlier, it was stated that the undrained strength of the during installation and hence even higher strengths
clay layers at both Yolla and Trefoil were determined would be expected to be mobilised in a strongly rate
from the T-bar and ball respectively by assuming that dependent soil. However, whether strain rate effects
Nbar  Nball  13.9. provide a complete explanation remains debatable.
No direct measurements were made of the
undrained strength of the Yolla clay but Kelleher &
Randolph (2005) present results from a number of 15 CONCLUSIONS
torvane tests performed on recovered samples from
the Trefoil clay layers. In the upper clay layer, the pro- This paper has presented results and interpretation
posed undrained strength determined from the ball from two extraordinary case histories of jackup oper-
(su  2.2z) provides a reasonable fit to the data, but ation in the Bass Strait, offshore Australia. While
for the lower clay layer, the torvane tests give a sig- these case histories appear unique in the literature, it
nificantly lower strength. However, the torvane is a is suspected that other similar cases may have gone
rudimentary tool and normally only provides a (con- unreported due to insufficient basic geotechnical data
servative) indication of the true undrained strength. and an inadequate understanding of the underlying
It is also interesting to note that the gradient of the mechanics.
PCPT qnet as a function of depth was only about 8% Many lessons can and should be learnt from these
higher in the clay layers at Trefoil compared to Yolla, examples and incorporated into standard practice in
but that the gradient of the ball resistance at Trefoil was future cases where difficult soils such as carbonate
much higher than the T-bar gradient at Yolla. Hence (or some non-carbonate) silty sands and sandy silts
using a common ball and T-bar factor of 13.5 at both are encountered.
sites implies a 30% lower undrained shear strength The experience at Yolla has shown that there is a
for the Yolla clay (su  1.7z). As discussed earlier, the need to reconsider the methods used to determine spud-
penetration of the Ensco 102 spudcans at Yolla pro- can penetration where high sensitivity soils are encoun-
vides additional evidence for this low interpretation tered. In particular, the typical depth factors applied
of the monotonic undrained strength; if the strength in bearing capacity formulae have been found to be
had been any higher then the spudcans could not have inappropriate. Backflow is also much more likely to
passed through the layer encountered between 16 m occur in such soils, leading to greater penetration than
and 19.5 m. Also, it should be remembered that for the industry standard methods (eg. SNAME) would sug-
analysis of spudcan penetration at Yolla it was assumed gest. A method to predict spudcan penetration directly
that the penetration resistance of a ball-shaped object from modern SI tools such as T-bar and ball pen-
(the spudcan) would be the same as a T-bar shaped etrometers has been proposed. However, even with
object; if a ball-shaped object had in fact exhibited a this method it is necessary to make appropriate modi-
higher resistance then the implied undrained strength fications for a number of factors, of which the most
would need to have been even lower. important are partial drainage and viscous rate effects.
On the other hand, if the strength of the equivalent The penetration rate used in PCPT and T-bar testing
clay layer at Trefoil (between approximately 20 m and will always result in a higher degree of drainage than
23 m below mudline) had a similarly low strength gra- will occur during spudcan penetration. The Yolla expe-
dient to the Yolla clay, then it seems likely that a static rience has demonstrated that these partial drainage
punch-through should have been initiated once (or effects can be sufficient to result in a severe under-
before) the Starboard spudcan had been worked down prediction of final spudcan penetrations and to mask
to 16.5 m below the mudline during the installation serious punch-through risks in soils where the coeffi-
phase. By this time the overlying sandy silt was evi- cient of consolidation of the in situ soil is broadly in a
dently substantially degraded through cyclic action and range between 1000 m2/yr and 250,000 m2/yr; extreme
hence was only offering limited resistance to the spud- caution should be exercised in such conditions.
can. Despite this, a static punch-through did not occur. The importance of cyclic strength degradation in
With the benefit of hindsight, it is possible that the soils such as those found at Yolla and Trefoil has also
in situ static undrained strength gradient of the clay been demonstrated. This can be a useful device to

72

Copyright 2005 Taylor & Francis Group plc, London, UK


increase spudcan penetration in cyclically degradable REFERENCES
soils, but this technique is most effective when per-
formed quickly and without delays, thereby minimis- Baglioni, V.P., Chow, G.S. & Endley, S.N. 1982. Jack-Up
ing consolidation effects. Rig Foundation Stability in Stratified Soil Profiles. Proc.
More problematically, cyclic strength degradation Offshore Technology Conference, OTC 4409, Houston,
Texas.
is an important factor that must be explicitly accounted Brown, J.D. & Meyerhof, G.G. 1969. Experimental Study
for when assessing the jackup rig foundation stability of Bearing Capacity in Layered Clays. Proc. 7th
under design storm events. Modifications to the stand- International Conf. Soil Mech. and Found. Eng., Mexico.
ard SNAME practice are proposed to address situa- Bye, A., Erbrich, C.T., Rognlien, B. & Tjelta, T.I. 1995.
tions of this kind. Most importantly, it should be Geotechnical Design of Bucket Foundations. Proc.
appreciated that the static preload cannot be relied on Offshore Technology Conference, OTC 7793, Houston,
as a load test of the foundation in such cases. As a Texas.
direct consequence, the reduced soil material factors Cassidy, M.J., Byrne, B.W. & Randolph, M.F. 2004. A
allowed in the SNAME practice also need to be Comparison of the Combined Load Behaviour of
Spudcan and Caisson Foundations on Soft Normally
reassessed. The author would recommend that values Consolidated Clay. Geotechnique, Vol 54, No. 2.
consistent with standard geotechnical design of off- Einav, I. & Randolph, M.F. 2004. Combining Upper Bound
shore shallow foundations (ie. between 1.25 and 1.3) and Strain Path Methods for Evaluating Penetration
should be adopted in such cases. Resistance. Submitted for publication.
Valuable insight has also been gained into the Erbrich, C.T. & Hefer, P.A. 2002. Installation of the
spudcan extraction process and the shape and size of Laminaria Suction Piles A Case History. Proc. Offshore
seabed craters that may result. Allowing wave loads to Technology Conference, OTC 14240, Houston, Texas.
apply a cyclic load to the spudcans was found to ease Finnie, I. 1993. Performance of Shallow Foundations in
the extraction process in the soils at Yolla, but the jet- Calcareous Soil. PhD Thesis, University of Western
Australia.
ting system was less effective than it might have been. Goodison, D. 2005 Understanding the Relationship between
The design of the jetting system could be signifi- Field Penetrometers and Laboratory Test Results for
cantly improved. Carbonate Silty Sand. Thesis submitted for BEng degree,
Finally, it should be appreciated that the observed UWA.
behaviour at Trefoil has very important implications Hossain, M.S., Hu, Y., Randolph, M.F. & White, D. 2005.
for other kinds of foundations that rely on preload as Limiting Cavity Depth for Spudcan Foundations
a proof load test, such as drag anchors, for example. Penetrating Clay. Paper submitted to Geotechnique for
In soils susceptible to high degrees of cyclic loading, publication.
normal industry accepted preload requirements Houlsby, G.T. & Martin, C.M. 2003. Undrained Bearing
Capacity Factors for Conical Footings on Clay.
appear inadequate and may need to be significantly Geotechnique, Vol 53, No. 5.
enhanced (Neubecker et al. 2005). Itasca, 1996. FLAC Fast Lagrangian Analysis of Continua
User Manual. Itasca Consulting Group.
Kelleher, P.J. & Randolph, M.F. 2005. Seabed Geotechnical
ACKNOWLEDGEMENTS Characterisation with the Portable Remotely Operated
Drill. Proc. International Symposium on Frontiers in
The author would like to thank the many colleagues Offshore Geotechnics, Perth, Western Australia, Balkema:
who assisted in this work and contributed ideas and/ Rotterdam.
or data. Of particular note are the contributions made Martin, C.M. 2001. Vertical Bearing Capacity of Skirted
Circular Foundations on Tresca Soil. Proc. 15th Int. Conf.
by Mark Cassidy and Itai Einav from COFS and Mark on Soil Mech. and Geotech. Eng., Istanbul, Balkema:
Randolph from AG/COFS. The records of Mike Rotterdam.
ONeill (AG), who acted as the client representative Mehryar, Z., Hu, Y. & Randolph, M.F. 2002. Penetration
during extraction of the Yolla spudcans were also Analysis of Spudcan Foundations in NC Clay. Proc. 12th
invaluable. The assistance given by Ensco is also Intl. Offshore and Polar Engineering Conf., Kitakyushu,
acknowledged, both for reviewing the record of oper- Japan.
ational details and for recovering bulk soil samples Meyerhof, G.G. & Hanna, A.M. 1978. Ultimate Bearing
from the spudcans after extraction at Trefoil, which Capacity of Foundations on Layered Soils under Inclined
will be used for future research purposes. Load. Canadian Geotech. Journal, Vol. 15.
MSL 2004. Guidelines for Jack-up Rigs with Particular
The author acknowledges the co-operation of the Reference to Foundation Integrity. Report prepared for
T/L1 and T/18P Joint Venture Parties (JVPs), and the UK Health and Safety Executive.
thanks them for their permission to publish this paper. Neubecker S.R., O Neill M.P. & Erbrich C.T. 2005.
The JVPs are Origin Energy (and subsidiaries, Preloading of Drag Anchors in Carbonate Sediments,
Operator), Australian Worldwide Exploration (AWE), Proc. International Symposium on Frontiers in Offshore
CalEnergy Gas (Australia) and Wandoo Petroleum (a Geotechnics, Perth, Western Australia, Balkema:
subsidiary of Mitsui). Rotterdam.

73

Copyright 2005 Taylor & Francis Group plc, London, UK


Randolph, M.F. & Erbrich, C.T. 1998. Design of shallow SNAME, 2002. Guidelines for Site Specific Assessments of
foundations for calcareous sediments. Proc. 2nd Intl. Mobile Jack-up Units. New Jersey: The Society of Naval
Conf, Eng. for Calcareous Sediments. Bahrain, Balkema: Architects and Marine Engineers.
Rotterdam. Stewart, D.P. & Finnie, I. 2001 Spudcan-Footprint
Randolph, M.F. & Hope, S. 2004. Effect of Cone Velocity on Interaction During Jack-up Workovers. Proc. 11th Intl.
Cone Resistance and Excess Pore Pressures. Proc. Int. Offshore and Polar Engineering Conf., Stavanger,
Symp. on Eng. Practice and Performance of Soft Deposits, Norway.
Osaka, Japan. Stewart, D.P. & Randolph, M.F. 1994. T-Bar Penetration test-
Salenon, J. & Matar, 1982. Capacit portant des fondations ing in Soft Clay. Journ. of Geotech. Eng. ASCE, Vol. 120,
superficielles circualires. Journal de mchanique theo- No.12.
rique et applique, Vol. 1, No. 2, pp. 237267. Watson, P.G. & Humpheson C. 2005. Geotechnical
Siciliano, R.J., Hamilton J.M. & Murff J.D. 1990. Effect of Interpretation for the Yolla A Platform. Proc. International
Jackup SpudCan on Piles. Proc. Offshore Technology Symposium on Frontiers in Offshore Geotechnics, Perth,
Conference OTC 6467, Houston, Texas. Western Australia, Balkema: Rotterdam.

74

Copyright 2005 Taylor & Francis Group plc, London, UK


Suction caissons for wind turbines

Guy T. Houlsby1, Lars Bo Ibsen2 & Byron W. Byrne1


1
Department of Engineering Science, Oxford University, UK
2
Department of Civil Engineering, Aalborg University, Denmark

ABSTRACT: Suction caissons may be used in the future as the foundations for offshore wind turbines. We
review recent research on the development of design methods for suction caissons for these applications. We
give some attention to installation, but concentrate on design for in-service performance. Whilst much can be
learned from previous offshore experience, the wind turbine problem poses a particularly challenging combin-
ation of a relatively light structure, with large imposed horizontal forces and overturning moments. Monopod
or tripod/tetrapod foundations result in very different loading regimes on the foundations, and we consider both
cases. The results of laboratory studies and field trials are reported. We also outline briefly relevant numerical
and theoretical work. Extensive references are given to sources of further information.

1 INTRODUCTION The dominant device used for large scale wind


power generation is a horizontal axis, 3-bladed tur-
The purpose of this paper is to review recent research bine with the blades upwind of the tower, as shown in
work on the design of suction caisson foundations for Figure 1. The details of the generator, rotational speed
offshore wind turbines. Most of the relevant work has and blade pitch control vary between designs. Most
been conducted at, or in co-operation with, the univer- offshore turbines installed to date generate 2 MW rated
sities of Oxford and Aalborg, so we report here power, and typically have a rotor about 80 m in diam-
mainly the work of our own research groups. eter with a hub about 80 m above mean sea level. The
Suction caissons have been extensively used as size of turbines available is increasing rapidly, and
anchors, principally in clays, and have also been used as prototypes of 5 MW turbines already exist. These
foundations for a small number of offshore platforms involve a rotor of about 128 m diameter at a hub height
in the North Sea. They are currently being considered of about 100 m. The loads on a typical 3.5 MW tur-
as possible foundations for offshore wind turbines. As bine are shown in Figure 2, which is intended to give
discussed by Houlsby and Byrne (2000) and by Byrne no more than a broad indication of the magnitude of
and Houlsby (2003), it is important to realise that the the problem.
loading regimes on offshore turbines differ in several Note that in conditions as might be encountered in
respects from those on structures usually encountered the North Sea, the horizontal load from waves (say
in the offshore oil and gas industry. Firstly the structures 3 MN) is significantly larger than that from the wind
are likely to be founded in much shallower water: 10 m (say 1 MN). However, because the latter acts at a
to 20 m is typical of the early developments, although much higher point (say 90 m above the foundation) it
deeper water applications are already being planned. provides more of the overturning moment than the
Typically the structures are relatively light, with a wave loading, which may only act at say 10 m above
mass of say 600t (vertical deadload 6 MN), but in the foundation. Using these figures the overturning
proportion to the vertical load the horizontal loads moment of 120 MNm would divide as 90 MNm due
and overturning moments are large. For instance the to wind and 30 MNm due to waves.
horizontal load under extreme conditions may be Realistic combinations of loads need to be con-
about 60% of the vertical load. sidered. For instance the maximum thrust on the turbine
An important consideration is that, unlike the oil and occurs when it is generating at the maximum allow-
gas industry where large one-off structures dominate, able wind speed for generation (say 25 m/s). At higher
many relatively small and inexpensive foundations are wind speeds the blades will be feathered and provide
required for a wind farm development, which might much less wind resistance. It is thus unlikely that the
involve anything from 30 to 250 turbines. maximum storm wave loading would occur at the

75

Copyright 2005 Taylor & Francis Group plc, London, UK


m
100

90 m
6 MN

4 MN

h  30 m

Figure 2. Typical loads on a 3.5 MW offshore wind turbine.

Figure 1. Offshore tests in Frederikshavn, Denmark. Front: the rotor at 1P. To the right of the first natural frequency
Vestas V90 3.0 MW turbine. Back: Nordex 2.3 MW turbine. is the 3P frequency. It should be noted that the 1P and
3P frequencies in general cover frequency bands and
not just two particular values, because the Vestas wind
same time as maximum thrust. Turbine designers must turbine is a variable speed device.
also consider important load cases such as emergency To avoid resonances in the structure at the key
braking. It is important to recognise that the design of a excitation frequencies (1P, 3P) the structural designer
turbine foundation is not usually governed by consider- needs to know the stiffness of the foundation with some
ations of ultimate capacity, but is typically dominated confidence, this means that problems of deformation
by (a) considerations of stiffness of the foundation and stiffness are as important as capacity. Furthermore,
and (b) performance under fatigue loading. much of the structural design is dictated by consider-
An operational wind turbine is subjected to har- ations of high cycle fatigue (up to about 108 cycles),
monic excitation from the rotor. The rotors rotational and the foundation too must be designed for these
frequency is the first excitation frequency and is com- conditions.
monly referred to as 1P. The second excitation fre-
quency to consider is the blade passing frequency, often
called 3P (for a three-bladed wind turbine) at three 2 CASES FOR STUDY
times the 1P frequency.
Figure 3 shows a representative frequency plot of a The two main problems that need to be studied in
selection of measured displacements for the Vestas design of a suction caisson as a foundation are:
V90 3.0 MW wind turbine in operational mode. The
foundation is a suction caisson. The measured data, installation;
monitoring system and Output-Only Modal Analysis in service performance.
used to establish the frequency plot are described in In this review we shall discuss installation methods
Ibsen and Liingaard (2005). The first mode of the struc- briefly, but shall concentrate mainly on design for in
ture is estimated, and corresponds to the frequency service performance. The relevant studies involve
observed from idling conditions. The peak to the left of techniques as diverse as laboratory model testing,
the first natural frequency is the forced vibration from centrifuge model testing, field trials at reduced scale,

76

Copyright 2005 Taylor & Francis Group plc, London, UK


Frequency Domain Decomposition - Peak Picking
Average of the Normalized Singular Values of
dB | 1.0 / Hz Spectral Density Matrices of all Data Sets.
20
3P
1P First mode
0

-20

-40

-60

-80

-100
1
Frequency

Figure 3. Frequency plot of measured displacements for a wind turbine in operational mode.

and a full-scale field installation. Complementing these


experiments are numerical studies using finite element
techniques, and the development of plasticity-based
models to represent the foundation behaviour.
Suction caissons may be installed in a variety of
soils, but we shall consider here two somewhat ideal-
ised cases: a caisson installed either in clay, which
may be treated as undrained, or in sand. For typical
sands the combination of permeability value, size of
caisson and loading rates leads to partially drained
conditions, although much of the testing we shall report
is under fully drained conditions. In this paper we report
mainly work on sands.
We shall consider two significantly different loading
regimes, which depend on the nature of the structure
supporting the wind turbine. Most offshore wind tur-
bines to date have been supported on a monopile a L
L D
single large diameter pile, which in effect is a direct s
extension of the tubular steel tower which supports the D
turbine. Some turbines have been supported on circular
(a) (b)
gravity bases. An obvious alternative is to use a single
suction caisson to support the turbine, and we shall call
Figure 4. Caisson foundations for a wind turbine, (a)
this a monopod foundation, Figure 4(a). The mono- monopod, (b) tripod/tetrapod.
pod resists the overturning moment (usually the most
important loading component) directly by its rotational
fixity in the seabed. smaller foundations: a tripod or tetrapod, Figure
As turbines become larger, monopod designs may 4(b). In either of these configurations the overturning
become sufficiently large to be uneconomic, and an moment on the structure is resisted principally by
alternative is a structure founded on three or four pushpull action of opposing vertical loads on the

77

Copyright 2005 Taylor & Francis Group plc, London, UK


upwind and downwind foundations. Alternatives
using asymmetric designs of tripod, and those V
employing jacket type substructures are also under
consideration.
Mudline

3 NORMALISATION PROCEDURES L
h t
A number of studies have been conducted at different Di
scales and it is necessary to compare the results from Do
these various studies. To do this it is appropriate to
normalise all the results so that they can be repre-
sented in non-dimensional form. This procedure also Figure 5. Geometry of a caisson foundation.
allows more confident extrapolation to full scale.
The geometry of a caisson is shown in Figure 5.
The outside radius is R (diameter Do), skirt length is L
and wall thickness t. In practice caissons may also
involve stiffeners on the inside of the caisson, these
being necessary to prevent buckling instability during
suction installation, but we ignore these in a simplified
analysis. Geometric similarity is achieved by requiring
similar values of L/2R and t/2R.
The sign convention for applied loads and dis-
placements is shown in Figure 6.
The rotation of the caisson  is already dimension-
less, and we normalise the displacements simply by
dividing by the caisson diameter, to give w/2R and u/2R.
In sand it is straightforward to show that, for similar
values of dimensionless bearing capacity factor, the
loads at failure would be proportional to  and to R3.
We therefore normalise vertical and horizontal loads as
Figure 6. Loading and displacement conventions for a
V/2 R3
and H/2 R3
, where we have included the caisson foundation (displacements exaggerated).
factor 2 to give the normalisation factor a simple
physical meaning: it is the effective weight of a cylinder
of soil of the same diameter of the caisson, and depth of stiffness, provided that the clays being compared
equal to the diameter. In a similar way we normalise have similar values of Ir  G/su. This condition is
the overturning moment as M/4 R4
. usually satisfied if the clays are of similar compos-
Use of the above normalisation is appropriate for ition and overconsolidation ratio. For sands, however,
comparing tests in sands with similar angles of fric- an extra consideration needs to be taken into account.
tion and dilation. We recognise that these angles both The shear modulus of a sand does not increase in pro-
decrease slightly with pressure and increase rapidly portion to the stress level, but instead can reasonably
with Relative Density (Bolton, 1986). This means that be expressed by:
comparable tests at smaller scales (and therefore
lower stress levels) will need to be at lower Relative
Densities to be comparable with field tests. (1)
In clay the vertical capacity is proportional to a
representative undrained shear strength su and to R2,
so we normalise loads as V/ R2su and H/ R2su, and where g and n are dimensionless constants, and pa is
the moment as M/2 R3su. atmospheric pressure (used as a reference pressure).
In order to be comparable, tests at different scales The value of n is typically about 0.5, so that the stiffness
will need the profile of undrained strength with depth is proportional roughly to the square root of pressure.
to be similar. If the strength profile is fitted by a sim- Comparing rotational stiffnesses on the basis of a
ple straight-line fit su  suo  z, then this requires plot of M/4 R4
against  effectively makes the
similar values of the factor 2R /suo. assumption that the shear stiffness is proportional to
Scaling of results using the above methods should 2R
, which may be regarded as a representative
give satisfactory results in terms of capacity. For clays it stress level. Since in fact the stiffness increases at a
should also lead to satisfactory comparisons in terms lower rate with stress level, this comparison will result

78

Copyright 2005 Taylor & Francis Group plc, London, UK


in larger scale tests giving lower apparent normalised Table 1. Installations in shallow water.
stiffness. This effect can be reduced by multiplying the
 scale by the dimensionless factor ( pa/2R
)1
n, which hw D L
compensates for the stiffness variation with stress Site Soil (m) (m) (m) Ref.
level.
Wilhelmshaven Sand 6.0 16.0 15.0 Installation
Thus we recommend that to compare both stiffness April 2005
and capacity data for sands one should plot M/4 R4

Frederikshavn Sand 1.0 12.0 6.0 30


against (pa/2R
)0.5 (assuming n  0.5) for moment
tests, and V/2 R3
against (w/2R)(pa/2R
)0.5 for Frederikshavn Sand
0.2
2.0 2.0
4.0 4.0
vertical loading tests. A fuller description of these nor-
malisation procedures is given by Kelly et al. (2005a). Sandy Haven Sand 0.5 4.0 2.5 23
Tenby Sand 2.0 2.0 2.0 23
Burry Port Sand 0.5 2.0 2.0
4 INSTALLATION STUDIES
Luce Bay Sand 3.0 1.5
0.2 1.5 1.0 27
The principal difference between installation of a suc-
Bothkennar Clay 3.0 1.5
tion caisson for an offshore wind turbine and for previ- 0.2 26
1.5 1.0
ous applications is that the turbines are likely to be
installed in much shallower water. There is a popular
misconception that suction caissons can only be
installed in deep water, where a very substantial head
difference can be established across the lid of the
caisson. In shallow water the net suction that can be
achieved is indeed much smaller (being limited by the
efficiency of the pumps, as the absolute pressure
approaches zero), but the suctions that can be achieved
are nevertheless sufficient for installation in most cir-
cumstances. Only in stiff clays is it likely that some
possible caisson designs, which might otherwise be
suitable as far as in-service conditions are concerned,
could not be installed by suction in shallow water.
In Table 1 we list the main instances where cais-
sons have been installed in shallow water, as appro-
priate to wind turbine installations. The water depths
hw are approximate only. In addition to the field tests
listed, a large number of small scale model tests of
installation have been carried out at Oxford University
(on caissons of 0.1 m to 0.4 m diameter), the University
of Western Australia (UWA), Aalborg and elsewhere.
The largest completed installation in shallow water is
that of a prototype suction caisson, shown in Figure 7,
installed in the offshore research test facility in
Frederikshavn, Denmark. The prototype has a diameter
of 12 m and a skirt length of 6 m. The operational water
depth is 4 m, and as the site is in a basin, no wave or ice
loads are applied. As seen in Figure 7 the suction cais-
son was installed in only 1 m of water in the basin. The
steel construction has a mass of approximately 140 t,
and the caisson was placed in late October 2002. The Figure 7. Installation of the prototype foundation at the
installation period was about 12 hours, with the soil test site in Frederikshavn: (a) during installation, (b) at the
penetration time being 6 hours. A computer system end of installation.
was used to control the inclination, suction pressure and
penetration rate. Det Norske Veritas (DNV) has certi- described in Ibsen and Brincker (2004). An even larger
fied the design of the prototype in Frederikshavn to B installation is currently in progress at Wilhelmshaven,
level. The Vestas V90 3.0 MW turbine was erected on Denmark.
the foundation in December 2002. The development There are two main ways of predicting firstly the
of the design procedure for the bucket foundation is self-weight penetration of the caisson and secondly

79

Copyright 2005 Taylor & Francis Group plc, London, UK


Suction, s (Pa)
0 500 1000 1500 2000 2500 3000
0

50
Penetration, h (mm)

100

150

200 SCIP Results


Experimental Result

250

Figure 8. Comparison of SCIP with model test.

the suction required to achieve full installation. The Figure 9. Suction required for installation at Frederikshavn.
first method (Houlsby and Byrne, 2005a, b) involves
use of adaptations of pile capacity analysis, in which
where kt is an empirical coefficient relating qt to the
the resistance to penetration is calculated as the sum
tip resistance during static penetration of the caisson,
of an end bearing term on the rim and friction on the
rt is the maximum reduction in tip resistance. ucrit is
inside and outside. In sands the seepage pattern set up
the critical suction resulting in the critical hydraulic
by the suction processes alters the effective stress
gradient icrit  1 along the skirt. t is an empirical
regime in a way that aids installation.
factor.
The calculation has been implemented in a spread-
Kout and Kin are coefficients relating fs to the unit
sheet program SCIP. Figure 8 shows for example a
skin friction on the outside and inside of the skirt. The
comparison between variation of measured suction in
water flow along the skirt changes the skin friction.
a model test installation with tip penetration of the
For the inside skin friction the coefficient reduces the
caisson (Sanham, 2003), and the SCIP calculation.
skin friction when suction is applied, whereas on the
The other approach involves use of CPT data to
outside the skin friction is increased. The coefficients
infer directly the resistance Rd to penetration of the
are established as:
caisson. The required suction ureq to penetrate the
caisson to depth d is calculated as:

(2)
(5a,b)
where G
(d) is the self-weight of the caisson at pene-
tration depth d (reduced for buoyancy), and Asuc is the
area inside the caisson, where the suction is applied.
The penetration resistance is calculated from the where out and in are empirical coefficients relating
following expression, which is based on calibration fs to the unit skin friction during static penetration of
against measured data: the caisson. rout and rin are the maximum changes in
skirt friction. out and in are empirical factors.
The required suction ureq to penetrate the prototype
(3)
in Frederikshavn was predicted using equation (2). The
result of the analysis is shown in Figure 9. The lower
line represents ureq calculated from the CPT tests. The
where qt is the corrected cone resistance and fs the curved line represents the limiting suction upip which
sleeve friction at depth z. Kt is a coefficient relating qt would cause piping to occur. umax is the theoretical
to the unit tip resistance on the rim. This resistance is maximum net suction, limited by the possibility of
adjusted for the reduction due to the applied suction cavitation within the caisson, as the absolute pressure
by the expression: approaches zero, so that umax  100 kPa above water
level and increases linearly with the water depth, as
shown by Figure 9. umax is used to calculate the access-
(4) ible net suction, which is limited by the efficiency of
the pumps, upump. As is seen, the suction in shallow

80

Copyright 2005 Taylor & Francis Group plc, London, UK


Volume, (10-3 m3)
0 5 10 15 20 25 30 35
0
50 Cell 1
Cell 2

Penetration, h (mm)
100
150
200
250
Total
300 Seepage Volume
Volume
350 Volume
Displaced
400

Figure 11. Volumes pumped from 2-cell caisson in sand.


Figure 10. The limiting suction upip has been achieved and
soil failure by piping has occurred. Suction, s (Pa)
0 1000 2000 3000 4000 5000 6000
0
water can be limited either by the suction causing
50 Cell 1
piping or by the accessible net suction available from Cell 2
Penetration, h (mm)
the pumps. 100
The suction upip causing piping has been studied at 150
the test site in Frederikshavn by installation tests on 200
2  2 m and 4  4 m caissons. Figure 10 shows a 250
4  4 m caisson where the limiting suction upip has 300
been achieved, and soil failure by piping has occurred. 350
The soil outside of the skirt is sucked into the caisson 400
and the penetration of the caisson cannot proceed.
If a tripod or tetrapod structure is to be installed, Figure 12. Suctions required for installation of 2-cell cais-
then levelling of the structure can be achieved by sep- son in sand.
arately controlling the suction in each of the caissons.
For a monopod structure, however, an alternative strat-
water pumped represents the volume displaced by the
egy has to be adopted. Experience suggests that for
descending caisson, whilst about 40% represents
installation in either clay or sand, the level of the caisson
seepage beneath the caisson rim. Figure 12 shows that
is rather sensitive to the application of eccentric loads
during the installation the suctions developed in the
(moments), especially in the early stages of installation.
two halves were (as would be expected in a uniform
This offers one possibility for controlling the level of
material) almost equal.
the caisson: by use of an eccentric load that can be
adjusted in position to keep the caisson level.
An alternative strategy, which has proven to be 5 CAISSON PERFORMANCE: MONOPOD
highly successful for installation in sand, is to divide the
rim into sections and to control the pressures at the skirt A large number of tests have been devoted to studying
tip in each section individually. By applying pressure the performance of a caisson under moment loading
over one segment of the caisson rim the upward at relatively small vertical loads, as is relevant to the
hydraulic gradient within the caisson can be enhanced wind turbine design. Some details of the test pro-
locally, thus encouraging additional downward move- grammes are given in Table 2.
ment for that sector. By controlling the pressures at a
number of points the caisson may be maintained level.
5.1 Sand: field tests
This method would not be applicable in clays. One
possibility, as yet untried at large scale, for control- The largest test involves the instrumented Vestas V90
ling level in clays would be to use a segmented cais- 3.0 MW prototype turbine at Frederikshavn, Denmark.
son in which the suctions in the different segments The caisson is installed in a shallow 4 m depth lagoon
could be controlled independently. next to the sea, and the turbine is fully operational.
Some preliminary small scale tests suggest that this The only significant difference between this installa-
approach might be successful in sand too (Coldicott, tion and an offshore one is that the structure is not
2005). Figure 11 shows the volumes of water pumped subjected to wave loading.
from the two halves of a 400 mm diameter caisson The test program involving the prototype (turbine
split by a diametral vertical wall. About 60% of the and caisson) is focusing on long-term deformations,

81

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 2. Moment loading tests.

Site Soil D (m) L (m) Ref.

Frederikshavn Sand 12.0 6.0


Frederikshavn Sand 2.0 2.0
Sandy Haven Sand 4.0 2.5
Burry Port Sand 2.0 2.0 21
Level IV: 89 m
Luce Bay Sand 3.0 1.5 27
0.1 0.00.066 2, 4
0.15 0.05 2, 7
Oxford laboratory Sand 0.15 0.1 42, 43
0.2 0.1 34, 42, 43
0.2 0.2 11, 43
0.3 0.15 11, 42, 43 Level III: 46 m
0.2 0.00.2
Aalborg laboratory Sand 0.3 0.00.3
0.4 0.00.4
Bothkennar Clay 3.0 1.5 26
Level II: 13 m
Oxford laboratory Clay 0.2 0.1 34 Level I: 6 m
0.3 0.15
UWA centrifuge 0.02 12
(100g) Clay 0.06 0.03
0.06

soil structure interaction, stiffness and fatigue. The


prototype has been equipped with: Figure 13. Sensor positions in tower and foundation.

an online monitoring system that measures the


dynamic deformation modes of the foundation and straightforward, with very simple instrumentation, but
the wind turbine, those at Frederikshavn test site and at Luce Bay were
a monitoring system that measures the long-time detailed investigations.
deflection and rotation of the caisson, The large scale tests at Frederikshavn is part of a
a monitoring system that measures the pore pres- research and development program concerning caisson
sure along the inside of the skirt. foundation for offshore wind turbines. The research
program is a co-operation between Aalborg University
The online monitoring system that measures the and MBD offshore power (Ibsen et al., 2003). The large
modes of deformation of the foundation and wind tur- scale tests are complemented by laboratory studies. The
bine involves 15 accelerometers and a real-time data- laboratory and large scale tests are intended to model
acquisition system. The accelerometers are placed at the prototype in Frederikshavn directly. In order to
three different levels in the turbine tower and at one design a caisson foundation for offshore wind turbines
level in compartments inside the caisson foundation. several load combinations have to be investigated. Each
The positions are shown in Figure 13, and the locations load combination is represented by a height of load h
and measuring directions are defined in Figure14. above the foundation and a horizontal force H. The
Output-only Modal Analysis has been used to ana- moment at the seabed is calculated as M hH. Table 3
lyze the structural behaviour of the wind turbine during shows that the resulting loading height varies from
various operational conditions. The modal analysis 10 m (for a wave force in shallow water) to 104.4 m
has shown highly damped mode shapes of the foun- (force from normal production of a 3 MW turbine in
dation/wind turbine system, which the present aero- 20 m of water). Scaling of the tests is achieved by:
elastic codes for wind turbine design cannot model.
Further studies are to be carried out with respect to (6)
soil-structure interaction. A detailed description of
the measuring system and the Output-Only Modal
Analysis is given by Ibsen and Liingaard (2005). where D is the diameter of the caisson and index m
The static moment tests referred to in Table 2 and p are for model and prototype. The values of the
at Sandy Haven and at Burry Port were relatively loading height in the test program are shown in Table 3.

82

Copyright 2005 Taylor & Francis Group plc, London, UK


Figure 15. Caisson for large scale test at Frederikshavn.

Figure 14. Sensor mountings in the tower and foundation


at Frederikshavn.

Table 3. Loading heights in the Aalborg test program.

Field
Prototype LaboratoryModel Model

D p  12 m Dm  0.2 m 0.3 m 0.4 m 2.0 m Figure 16. Setup for combined loading of 2  2 m caisson
at Frederikshavn (Back: prototype 3 MW Vestas wind tur-
hp [m] hm [m] hm [m] bine on the 12  6 m caisson).

104.4 1.74 2.61 3.48 17.40


69.6 1.16 1.74 2.32 11.60 2. Loading phase: An old tower from a wind turbine
38.0 0.63 0.95 1.27 6.33 is mounted on top of the caisson. The caisson is
20.0 0.33 0.50 0.67 3.33 loaded by pulling the tower horizontally with a
10.0 0.17 0.25 0.33 1.67 wire. The combined loading (H, M) is controlled by
changing the height of loading.
3. Dismantling phase: The caisson is removed by
The large scale tests at Frederikshavn employ load- applying overpressure inside the bucket.
ing by applying a horizontal load at a fixed height,
Figure 17 shows the moment rotation curve for a
under constant vertical load. A steel caisson with an
test on the 2  2 m caisson at Frederikshavn. The test
outer diameter of 2 m and a skirt length of 2 m has
is performed with hm  17.4 m and a vertical load on
been used. The skirt is made of 12 mm thick steel
the caisson of 37.3 kN. The fluctuations in the curve
plate. Figure 15 shows the caisson prior to installa-
are caused by wind on the tower.
tion, and Figure 16 the overall test setup. Currently 10
Tests at Luce Bay were designed by Oxford
experiments have been conducted, but the testing pro-
University and conducted by Fugro Ltd.. The moment
gram is ongoing. Each test has three phases:
loading tests were of two types. Firstly small ampli-
1. Installation phase: The caisson is installed by tude (but relatively high frequency) loading was
means of suction. CPT tests are performed before applied by a Structural Eccentric Mass Vibrator
and after installation of the caisson. (SEMV) in which rotating masses are used to apply

83

Copyright 2005 Taylor & Francis Group plc, London, UK


inertial loads at frequencies up to about 12 Hz. amplitude is due to gapping occurring at the sides of
Secondly larger amplitude, but lower frequency, cycles the caisson.
were applied using a hydraulic jack. A diagram of the The secant stiffnesses deduced from both the
loading rig, which allowed both moment and vertical SEMV tests and the hydraulic jacking tests are com-
loading tests, is shown in Figure 18. bined in Figure 21, where they are plotted against the
The SEMV test involve cycles of moment loading amplitude of cyclic rotation. It is clear that the two
at increasing amplitude as the frequency increases.
Figure 19 shows the hysteresis loops obtained from a
series of these cycles at different amplitudes. As the 30
6Hz
cycles become larger the stiffness reduces but hys-
7Hz 20
teresis increases. The tests were interpreted (Houlsby
8Hz
et al., 2005b) using the theory of Wolf (1994), which

Moment (kNm)
10
9Hz
takes account of the dynamic effects in the soil, and the 10H
equivalent secant shear modulus for each amplitude 0
-0.00005 -0.000025 0 0.000025 0.00005
of cycling determined. -10
Figure 20 shows the moment rotation curves for
much larger amplitude cycling applied by the hydraulic -20
jack. Again hysteresis increases and secant stiffness
-30
decreases as the amplitude increases. The unusual Rotation (radians)
waisted shape of the hysteresis loops at very large
Figure 19. Hysteresis loops from SEMV tests on 3 m caisson.

500
400
300
Moment (kNm)

200
100
0
-0.08 -0.06 -0.04 -0.02 -100 0 0.02 0.04 0.06 0.08
-200
-300
-400
-500
-600
Rotation of caisson centre (2R ) (m)

Figure 20. Hysteresis loops from hydraulic jacking tests


Figure 17. Moment-rotation test on 2  2 m caisson. on 3 m caisson.

Figure 18. Field testing equipment, dimensions in mm. Water level and displacement reference frames not shown. (a)
arrangement for jacking tests on 1.5 m and 3.0 m caissons, (b) alternative arrangement during SEMV tests. Labels indicate
(A) A-frame, (B) concrete block, (C) caissons, (H ) hydraulic jacks, (L) load cells, (R) foundations of reaction frame, (V)
SEMV, (W) weight providing offset load for SEMV tests.

84

Copyright 2005 Taylor & Francis Group plc, London, UK


groups of tests give a consistent pattern of reduction The rig is shown in Figure 24, and is capable of apply-
of shear modulus with strain amplitude, similar to ing a wide range of combinations of vertical, horizontal
that obtained for instance from laboratory tests. and moment loading under either displacement or
load control.
Typical moment loading tests involve applying a
5.2 Sand: laboratory tests
fixed vertical load, and then cycling the rotation at
Turning now to model testing, a large number of tests increasing amplitude. An example is given in Figure 25.
have been carried out both at Aalborg and at Oxford. The first interpretation of such tests is to determine
Almost all the model tests have involved in plane the yield surface for a single surface plasticity model
loading (in which the moment is about an axis perpen- (see section 7.2 below, and also Martin and Houlsby
dicular to the horizontal load). However, a test rig cap-
able of applying full 6 degree-of-freedom loading has
recently been developed by Byrne and Houlsby (2005).
The model tests at Aalborg are performed by the
test rig shown in Figure 22. The rig consists of a test
box and loading frame. The test box consists of a steel
frame with an inner width of 1.6 m 1.6 m and an
inner total depth of 0.65 m. The test box is filled with
Aalborg University Sand No 0. After each experiment
the sand in the box is prepared in a systematic way to
ensure homogeneity within the box, and between the
different test boxes. The sand is saturated by the water
reservoir shown in Figure 22. Before each experiment
CPT-tests are performed to verify the density and
strength of the sand. The caisson is then installed and
loaded with a constant vertical load. The vertical load
is kept constant through the experiment, while the
horizontal force is applied to the tower by the loading
device mounted on the loading frame, see Figure 21.
The tower and the loading device are connected by a
wire. The combined loading (H, M) is controlled by
the height of loading h. The loading frame allows the
possibility of changing h from 0.1 m to 4.0 m above
the sand surface (Table 3). The horizontal force H is
measured by a transducer connected to the wire. The
deformation of the foundation and the moment are
measured with the measuring cell mounted on the top
of the caisson, as shown by Figure 23.
Laboratory tests at Oxford University have used a
versatile 3 degree-of-freedom loading rig designed by Figure 22. The caisson test rig at Aalborg University.
Martin (1994) and adapted by Byrne (2000) (see also
Martin and Houlsby (2000) and Gottardi et al. (1999)).

100
90 Jacking
80 SEMV
Hyperbolic curve fit
70
G (MPa)

60
50
40
30
20
10
0
0.000001 0.00001 0.0001 0.001 0.01 0.1
(radians)
Figure 23. The measuring cell connecting the caisson and
Figure 21. Shear modulus against rotation amplitude. the tower.

85

Copyright 2005 Taylor & Francis Group plc, London, UK


100

Moment Load, M/2R (N)


Experiment, M/2RH = 1
Fitted Yield Surface 80

Soil Plug Weight


60

40

20

0
-160 -120 -80 -40 0 40 80 120
Vertical Load, V (N)

Figure 26. Experimentally determined yield surface in


V-M plane.

120

Incremental Rotation, 2Rdtheta (mm)


V = -50 N
80 V= 0N

Moment Load, M/2R(N) 40


V = 50 N

-40

-80

Figure 24. Three degree-of-freedom testing rig at Oxford -120


University. -180 -140 -100 -60 -20 20 60 100 140 180
Horizontal Load, H(N)
Incremental Horizontal Displacement, du (mm)
100
80 Figure 27. Yield surfaces and flow vectors in H-M space.
Moment Load, M/2R (N)

60
40
20
0 stress levels. The flow vectors are also plotted in this
-20 figure, and show that in this plane (unlike the V-M
-40 plane) associated flow is a reasonable approximation
-60 to the behaviour. Feld (2001) has observed similar
-80 shapes of a yield surface for a caisson in sand.
-100
-2.0 -1.5 -1.0 -0.5 0.0 0.5 1.0 1.5 2.0
We now consider the possibility of scaling the results
Rotational Displacement, 2R (mm)
of laboratory tests to the field. The test at Frederikshavn
shown in Figure 17 was on a caisson with a ratio
Figure 25. Moment-rotation test on sand. L/2R  1, at an M/2RH value of approximately 8.7,
and with a value of V/2 R3
of about 0.62. Using the
data from the Oxford laboratory on 0.2  0.2 m cais-
(2001), Houlsby and Cassidy (2002), Houlsby (2003), sons this requires a vertical load of about 60N. In fact
Cassidy et al. (2004)). An example of the yield points a test had been carried out with L/2R  1 and V 
obtained, plotted in the vertical load-moment plane, is 50 N. According to the scaling relationships discussed
given in Figure 26. Of particular importance is the fact in section 3, the moment should be scaled according
that at very low vertical loads there is a significant to R4
(a factor of 6250) and the rotational displace-
moment capacity, and that this extends even into the ment 2R according to  R3 
(a factor of 25). Figures
tensile load range. In these drained tests the ultimate 26 and 27 suggest that for a vertical load of 60 N
load in tension is a significant fraction of the weight of rather than 50 N a moment capacity say 5% higher
the soil plug inside the caisson. might be expected, and that for the higher value of
Sections of the yield surface can also be plotted in M/2RH a further increase of say 15% is appropriate.
H-M space as shown in Figure 27, where the data here We therefore apply a factor of 7500 to the moments and
have been assembled from many tests at different 25 to the rotational displacements. The result is shown

86

Copyright 2005 Taylor & Francis Group plc, London, UK


150 0.3

0.2
100
0.1

M/[su(2R) ]
3
Moment, M (kNm)

50
0
0 -0.1

-0.2
-50
-0.3
-100
-0.4
-0.015 -0.01 -0.005 0 0.005 0.01
-150
-0.06 -0.04 -0.02 0 0.02 0.04 0.06
Rotational Displacement, 2R (m) (a) field test
0.3
Figure 28. Laboratory moment test scaled to field condi- 0.2
tions for comparison with Figure 17.
0.1

M/[su(2R) ]
3
0
in Figure 28. It can be seen that after scaling the
-0.1
moment at a 2R value of 0.04 m is about 120 kNm,
compared to about 280 kNm measured in the field. -0.2
Although there is a factor of about 2 between these -0.3
values, it must be borne in mind that there are a num-
-0.4
ber of possible causes of difference between the tests -0.015 -0.01 -0.005 0 0.005 0.01
(e.g. the sand in the field test may be much denser),
(b) model test
and also that a factor of 7500 has already been applied:
a factor of 2 is relatively small by comparison.
Figure 29. Moment-rotation results presented in non-
dimensional form for laboratory and field tests.
5.3 Clay: field and laboratory tests
Less work has been carried out on clay than on sand. This sort of comparison is vital to establish confidence
The large scale trials at Bothkennar (Houlsby et al. in the use of model testing to develop design guidelines.
2005b) are complemented by laboratory studies
intended to model these trials directly, and therefore
add confidence to the scaling of the results to proto- 6 CAISSON PERFORMANCE: TETRAPOD
type size caissons (Kelly et al., 2005a). OR TRIPOD
At Bothkennar, moment loads were applied to a
3 m 1.5 m caisson by two means. Small amplitude, In the following, in which we consider multiple foot-
but relatively high frequency (10 Hz) loading was ing designs to support the wind turbine, we shall refer
applied by means of the SEMV device described principally to a tetrapod (four footings) rather than a
above, and larger amplitude cycles, but at much lower tripod. As a tripod is perhaps the most obvious mul-
frequency, were applied using a hydraulic jack. In tiple footing design to use, and has the obvious
both cases the loading was 4 m above the caisson, so advantage of simplicity, our preference for the tetra-
that hload/D  1.33. The most important observation pod deserves some explanation.
from these tests was the gradual reduction of secant As is discussed below, prudent design of a multiple
stiffness (and increase in hysteresis) as the amplitude footing structure will avoid tension being applied to
of the load cycles increases. any of the foundations (except under the most extreme
The laboratory tests, specifically modelling the of circumstances). This in effect dictates the separation
field tests, involved just relatively low frequency load- of the foundations for a given overturning moment
ing. After the scaling relationships described in section 3 and weight of structure. Approximate calculations
were applied, there was a satisfactory agreement indicate that the tetrapod structure is usually a more
between laboratory and field data, especially at rela- favourable configuration to avoid tension, as it requires
tively small amplitudes of movement. As an example, somewhat less material. The differences are not large,
Figure 29(a) shows the results (in dimensionless form) and a tripod may be preferred in some circumstances,
for rotation of the 3.0 m diameter caisson in the field, but we shall refer to a tetrapod, as this will probably
and Figure 29(b) the equivalent results, also in dimen- be more efficient. The important mechanism is the
sionless form, from the small scale model test. The pat- same in both cases: the overturning moment is resisted
tern of behaviour is remarkably similar in the two tests. by opposing pushpull action on the foundations.

87

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 4. Vertical loading tests. Normalised Vertical Displacement, w/D
0 0.5 1 1.5 2 2.5 3 3.5
Site Soil D (m) L (m) Ref. 400
350

Vertical Load, V (N)


Luce Bay Sand 1.5 1.0 27 300
0.05 0.00.1 11 250
0.1 0.00.066 2,5 200
0.15 0.05 2,5 150
Oxford Sand 0.15 0.1 34 100
laboratory 0.2 0.133 34 50
0.28 0.18 25, 32, 33, 35 0
0 50 100 150
Bothkennar Clay 1.5 1.0 26
Vertical Displacement, w (mm)
UWA 0.02
centrifuge Clay 0.06 0.03 3 Figure 30. Vertical load-penetration curves for caissons of
(100g) 0.06 different L/D ratios.

In Table 4 we list the tests that have been carried out 1600
on vertical loading of caissons relevant to the wind 1400
turbine problem. In addition to these studies there are 1200
a number of other relevant studies which have been Vertical Stress (kPa) 1000
directed towards vertical loading of caissons for struc- 800
tures in the oil and gas industry or for use as anchors. 600
400
200
6.1 Sand: field and laboratory tests 0
-200
The simplest tests on vertical loading of caissons in -400
sand, which are relevant both to installation and to sub- 200 210 220 230 240 250 260 270
sequent performance, simply involve pushing caissons Vertical Displacement (mm)
vertically into sand to determine the vertical load-dis-
placement response. Figure 30 shows the results of a set Figure 31. Cyclic vertical loading of model caisson.
of such tests on caissons of different L/D ratios, Byrne
et al. (2003). It is clear from the figure that there is a
well-established pattern. While the caisson skirt is The above observations mean that tension must be
penetrating the sand there is relatively low vertical avoided in a prudent design of a tripod or tetrapod foun-
capacity, but as soon as the top plate makes contact with dation for a wind turbine. However, in all but the shal-
the sand there is a sudden increase in capacity. The lowest of water, avoiding this tension means that either
envelope of the ultimate capacities of footings of differ- the foundation must have a large spacing between the
ent initial L/D ratios also forms a single consistent line. footings, or that ballasting must be used. The latter
Of most importance, however, is the performance of may in fact be a cost effective measure in deep water.
the caissons under cyclic vertical loading. Figure 31 Some designers may wish to reduce conservatism by
shows the results of tests on a 300 mm diameter cais- allowing for the possibility of tension under extreme
son subjected to rapid cyclic loading. Smallamplitude circumstances. It is therefore useful to examine the
cycles show a stiff response, with larger cycles show- ultimate tensile capacity under rapid loading. Figure 32
ing both more hysteresis and more accumulated dis- shows the result of three such tests. The slowest test (at
placement per cycle. The most important observation 5 mm/s) is almost drained, and a very low capacity in
is that as soon as the cycles go into tension, a much tension is indicated. The capacity in this case is simply
softer response is observed, and the hysteresis loops the friction on the skirts. The test at 100 mm/s (but zero
acquire a characteristic banana shape. Clearly the ambient water pressure) shows a larger capacity, and it
soft response on achieving tension should be avoided is straightforward to show that this is controlled by
in design. Closer examination of the curves reveals cavitation beneath the foundation. This means that at
that the softening in fact occurs once the drained fric- elevated water pressures (as in the third test) the
tional capacity of the skirts has been exceeded, rather capacity rises approximately in step with to the ambient
than simply the transition into tension. water pressure, as correspondingly larger pressure
Paradoxically, although additional accumulated changes are required to cause cavitation. This problem
displacement is observed once tension is reached, this is studied in more detail by Houlsby et al. (2005a).
accumulated displacement is downwards (not upwards It is important to note, however, that although ambi-
as one might expect because of the tensile loading). ent water pressure increases the ultimate capacity, it

88

Copyright 2005 Taylor & Francis Group plc, London, UK


Vertical Displacement (mm) 0.05
150 160 170 180 190 200 210
0 0.00

[w/(2R)][pa/(2R)]1/2
1 10 100 1000 10000 100000
-50 -0.05
-100
Vertical Stress (kPa)

-150 -0.10

-200
-0.15
-250
Direction of
-0.02 Min
-300 movement
Max
-350 5mm/s, 0kPa
100mm/s, 0kPa -0.25
-400 Number of Cycles
100mm/s, 200kPa
-450
Figure 34. Accumulated displacement during long term
Figure 32. Tensile capacity of model caisson pulled at dif- cyclic vertical loading on sand.
ferent rates and at different ambient pressures.

5 60
1.5m Field

Vertical Stress, V/A (kPa)


4 40
0.15m Suction
0.2m Pushed 20
3
V/['(2R)3]

0.15m Pushed
0
2 0 0.2 0.4 0.6 0.8 1
-20
1
-40
0 Test 1: Post Bearing Capacity
-60
Test 2: Pre Bearing Capacity
-1 -80
0 0.01 0.02 0.03 0.04 0.05 Normalised Displacement, (w + L)/D
[w/(2R)][pa/(2R')]1/2
Figure 35. Tension tests on caisson foundations in clay.
Figure 33. Hysteresis loops from tests at different scales
and rates.
is seen in Figure 34 to increase approximately with
the logarithm of the number of cycles of loading (after
has negligible influence on the tensile load at which a about 1000 cycles). Note that even in this case where
flexible response begins to occur. there is a tensile loading in part of the cycle, the net
Comparison of cyclic loading tests at different scales movement is downwards. The displacement is of course
and at different speeds shows that it is difficult to scale very sensitive also to the amplitude of the cycling.
reliably the accumulated displacements, which reduce
with larger tests and higher loading rates. However,
when the scaling rules described earlier are applied, 6.2 Clay: field and laboratory tests
the shapes of individual hysteresis loops at different Very few vertical loading tests relevant to the wind tur-
scales and at different rates become remarkably simi- bine problem have been completed on caissons in clay,
lar. Figure 33 shows a comparison, for instance, of although there have been a number of studies directed
loops at three different load amplitudes from four dif- towards suction caissons used as tension anchors, e.g.
ferent tests. At each particular load amplitude the loops El-Gharbawy (1998), Watson (1999), House (2002).
from the different tests are very similar. At Bothkennar tests were carried out in which
The accumulation of displacement after very large inclined (but near vertical) loading was applied to a
numbers of cycles is difficult to predict, and so far few 1.5 m diameter caisson (Houlsby et al., 2005b). Diffi-
data are available. Rushton (2005) has carried out verti- culties were encountered with the control of the loads
cal loading tests to about 100 000 cycles on a model using a hydraulic system, and the resulting load paths
caisson in sand, using a simple loading rig which are therefore rather complex, leading to difficulties in
employs a rotating mass and a series of pulleys to apply interpretation. Further work on vertical loading in clay
a cyclic load. A typical result is shown in Figure 34, is required before definitive conclusions can be drawn,
on a caisson 200 mm diameter and 100 mm deep, with and in particular the issue of tensile loading in clay
cycling between 210  260 N. The caisson is there- needs attention. Some preliminary results (Byrne and
fore subjected (at the minimum vertical load) to a small Cassidy, 2002), shown in Figure 35, show that the ten-
tension, but less than the frictional capacity of the skirts. sile response may be sensitive to prior compressive
The dimensionless accumulated vertical displacement loading. Footings loaded in tension immediately after

89

Copyright 2005 Taylor & Francis Group plc, London, UK


installation showed a stiff tensile response, whilst number of other issues which need to be addressed in
those loaded after first applying a compressive load to a caisson design, and we mention them here briefly.
failure showed a more flexible tensile response.
8.1 Scour
7 NUMERICAL STUDIES
Scour is more important for caissons, since they are
relatively shallow, than for piles. The size of caissons,
7.1 Finite element studies
and the fact that part of the caisson inevitably pro-
A number of analyses of suction caissons for offshore trudes above mudline level, creates rather aggressive
wind farms have been carried out as part of commercial conditions for scour. The fact that the caissons may be
investigations for possible projects. A more detailed installed in mobile shallowwater environments means
research project was carried out by Feld (2001). that proper consideration of this problem is essential,
Finite element analysis is particularly appropriate especially in sands.
for establishing the effects of design parameters on the If the scour depth can be determined with suffi-
elastic behaviour of caissons, and has been used by cient confidence (e.g. from comprehensive model test-
Doherty et al. (2004a, b) to determine elastic stiffness ing) then it may be possible to permit the scour to
coefficients for caisson design which take into account occur, and simply allow for this in the design by
the flexibility of the caisson wall as well as coupling ensuring that the caisson is deep enough.
effects between horizontal and moment loading. It is more likely, however, that scour protection
measures such as rock-dumping will need to be
7.2 Plasticity models employed. Practical experience suggests that such pro-
tection must be placed very soon after caisson installa-
An important tool for the analysis of soil-structure tion, as scour can occur very rapidly. In highly mobile
interaction problems, particularly those involving environments, significant scour can, for instance, occur
dynamically sensitive structures are force resultant due to the currents in a single tide. Model testing indi-
models. In these the behaviour of the foundation is cates, however, that scour protection measures can be
represented purely through the force resultants acting effective in preventing further erosion (R. Whitehouse:
upon it, and the resulting displacements (see Figure 4). private communication). For in-service conditions
Details of stresses and deformations within the soil regular monitoring for the possibility of scour would
are ignored. The models are usually framed within the be prudent.
context of work-hardening plasticity theory. Examples
include models for foundations on clay (Martin and
Houlsby, 2001) and on sand (Houlsby and Cassidy, 8.2 Liquefaction
2002). Overviews of the development of these models The transient pore pressures induced in the seabed
are given by Houlsby (2003) and Cassidy et al. (2004). can induce liquefaction, especially if the seabed is
These models have been further developed specif- partially saturated due to the presence of gas (as can
ically for the offshore wind turbine application. The occur in shallow seabeds, largely due to decay of organic
developments include: matter).
Generalisation to full three-dimensional loading The problem is a complex one, but typically, at one
conditions, stage in the wave cycle, the pore pressure in the seabed
Inclusion of special features to represent the cais- can become equal to the overburden stress, and the
son geometry, effective stress falls to zero. This problem is further
Expression of the models within the continuous complicated by the presence of a structure, which
hyperplasticity framework to allow realistic descrip- clearly modifies the pore pressure pattern that would
tion of hysteretic response during cyclic loading. occur in the far field. Although some progress has
been made, the interactions are complex, and theoret-
A model with all these features is described by ical modelling of the problem is not straightforward.
Lam and Houlsby (2005). The fitting of cyclic data to
a continuous hyperplastic model is discussed by
Byrne et al. (2002a). 8.3 Wave-induced forces
A quite different problem from liquefaction is also
8 OTHER CONSIDERATIONS related to the fact that the principal forces on the
structure are wave induced. As a wave passes the col-
We have concentrated here on the design of caisson umn of the structure it exerts large horizontal forces
foundations as far as capacity and stiffness are con- (of the order of a few meganewtons for a large wave),
cerned for in-service conditions. However, there a which also cause overturning moments. However, at

90

Copyright 2005 Taylor & Francis Group plc, London, UK


the same time the wave causes a transient pressure on As the amplitude of vertical loading increases,
the seabed, and on the lid of the caisson. Because the stiffness reduces and hysteresis increases. Once
caissons are in shallow water these pressures are quite tension is reached there is a sudden reduction of
large. The pore water pressure within the caisson is stiffness.
unlikely to change as rapidly as the pressure on the Whilst high ultimate tensile capacities are possible
lid, so there will be pressure differentials across the (especially in deep water) this is at the expense of
lid of the caisson which result in net vertical forces, large movements.
and overturning moments on the caisson. Application of scaling procedures for tests in both
The relative phase of the different sources of load- sand and clay allows model and field tests to be
ing is important. As the crest of the wave just reaches compared successfully as far as stiffness and the
the structure, the wave kinematics are such that the shapes of hysteresis loops is concerned.
horizontal forces are likely to be largest. At this stage Cumulative displacements after very many cycles
the pressure on the upwave side of the caisson is are harder to model.
likely to be larger than on the downwave side. The net The design of caisson foundations also needs to
result is that the moment caused by the pressures on take into consideration issues such as scour and
the caisson lid opposes that caused by the horizontal liquefaction.
loading, so this effect is likely to be beneficial to the
performance of the caisson. Little work has, however, It is hoped that the conclusions above lead in due
yet been completed on the magnitudes of these effects. course to application of suction caissons as founda-
The problem is complicated by the fact that the kin- tions for offshore wind turbines, thereby making an
ematics of large (highly non-linear) shallow water important renewable energy source more econom-
waves is still a matter of research, as is their interaction ically viable.
with structures.
ACKNOWLEDGEMENTS
9 CONCLUSIONS
The work at Oxford University has been supported
In this paper we have provided an overview of the by the Department of Trade and Industry, the Engi-
extensive amount of work that has been carried out on neering and Physical Sciences Research Council and
the design of suction caisson foundations for offshore a consortium of companies: SLP Engineering Ltd,
wind turbines. Further verification of the results pre- Aerolaminates (now Vestas), Fugro Ltd, Garrad Hassan,
sented here is still required, and in due course it is hoped GE Wind and Shell Renewables. An outline of the
that this will come from instrumented caisson foun- project is given by Byrne et al. (2002b). The work of
dations offshore. Our broad conclusions at present are: Richard Kelly, Nguyen-Sy Lam and Felipe Villalobos
on this project is gratefully acknowledged.
Suction caissons could be used as foundations for
offshore wind turbines, either in monopod or tri-
pod/tetrapod layout.
The combination of low vertical load and high REFERENCES
horizontal load and moment is a particular feature
of the wind turbine problem. 1. Bolton, M.D. (1986) The strength and Dilatancy of
Stiffness and fatigue are as important for turbine Sand, Geotechnique, Vol. 36, No. 1, pp 6578
design as ultimate capacity. 2. Byrne, B.W. (2000) Investigations of Suction Caissons
in Dense Sand, D.Phil. Thesis, Oxford University
Monopod foundation design is dominated by
3. Byrne, B.W. and Cassidy, M.J. (2002) Investigating the
moment loading. response of offshore foundations in soft clay soils,
Tripod/tetrapod foundation design is dominated by Proc. OMAE, Oslo, Paper OMAE2002-28057
considerations of tensile loading. 4. Byrne, B.W. and Houlsby, G.T. (1999) Drained Behaviour
The moment-rotation response of caissons in sand of Suction Caisson Foundations on Very Dense Sand,
has been extensively investigated by model tests and Offshore Technology Conference, 36 May, Houston,
field trials, and modelled theoretically by finite Paper 10994
element analyses and force resultant (yield surface) 5. Byrne, B.W. and Houlsby, G.T. (2002) Experimental
models. Investigations of the Response of Suction Caissons to
Transient Vertical Loading, Proc. ASCE, J. of Geot.
As amplitude of moment loading increases, stiff-
Eng., Vol. 128, No. 11, Nov., pp 926939
ness reduces and hysteresis increases. 6. Byrne, B.W. and Houlsby, G.T. (2003) Foundations for
Moment loading in clay has been less extensively Offshore Wind Turbines, Phil. Trans. of the Royal
investigated in the laboratory and field. Society of London, Series A, Vol. 361, Dec., 29092930
Vertical loading in sand has been extensively inves- 7. Byrne, B.W. and Houlsby, G.T. (2004) Experimental
tigated in the laboratory and field. Investigations of the Response of Suction Caissons to

91

Copyright 2005 Taylor & Francis Group plc, London, UK


Transient Combined Loading, Proc. ASCE, J. of 24. Houlsby, G.T. and Cassidy, M.J. (2002) A Plasticity
Geotech. and Geoenvironmental Eng., Vol. 130, No. 3, Model for the Behaviour of Footings on Sand under
pp 240253 Combined Loading, Gotechnique, Vol. 52, No. 2,
8. Byrne, B.W. and Houlsby, G.T. (2005) Investigating 6 Mar., 117129
degree-of-freedom loading on shallow foundations, 25. Houlsby, G.T., Kelly, R.B. and Byrne, B.W. (2005a)
Proc. International Symposium on Frontiers in Offshore The Tensile Capacity of Suction Caissons in Sand under
Geotechnics, Perth, Australia, 1921 September, in press Rapid Loading, Proc. Int. Symp. on Frontiers in Offshore
9. Byrne, B.W., Houlsby, G.T. and Martin, C.M. (2002a) Geotechnics, Perth, Australia, September, in press
Cyclic Loading of Shallow Offshore Foundations on 26. Houlsby, G.T., Kelly, R.B., Huxtable, J. and Byrne, B.W.
Sand, Proc. Int. Conf on Physical Modelling in Geotech., (2005b) Field Trials of Suction Caissons in Clay for
July 1012, St Johns, Newfoundland, 277282 Offshore Wind Turbine Foundations, Gotechnique, in
10. Byrne, B.W., Houlsby, G.T., Martin, C.M. and Fish, P. press
(2002b) Suction Caisson Foundations for Offshore Wind 27. Houlsby, G.T., Kelly, R.B., Huxtable, J. and Byrne, B.W.
Turbines, Wind Engineering, Vol. 26, No. 3, pp 145155 (2005c) Field Trials of Suction Caissons in Sand for
11. Byrne, B.W., Villalobos,, F. Houlsby, G.T. and Martin, Offshore Wind Turbine Foundations, submitted to Go-
C.M. (2003) Laboratory Testing of Shallow Skirted technique
Foundations in Sand, Proc. Int. Conf. on Foundations, 28. House, A. (2002) Suction Caisson Foundations for
Dundee, 25 September, Thomas Telford, pp 161173 Buoyant Offshore Facilities, PhD Thesis, the University
12. Cassidy, M.J., Byrne, B.W. and Randolph, M.F. (2004) of Western Australia
A comparison of the combined load behaviour of spud- 29. Ibsen, L.B., Schakenda, B., Nielsen, S.A. (2003)
can and caisson foundations on soft normally consoli- Development of bucket foundation for offshore wind
dated clay, Gotechnique, Vol. 54, No. 2, pp 91106 turbines, a novel principle. Proc. USA Wind 2003
13. Cassidy, M.J., Martin, C.M. and Houlsby, G.T. (2004) Boston.
Development and Application of Force Resultant 30. Ibsen, L.B. and Brincker, R. (2004) Design of New
Models Describing Jack-up Foundation Behaviour, Foundation for Offshore Wind Turbines, Proceedings
Marine Structures, (special issue on Jack-up Platforms: of The 22nd International Modal Analysis Conference
Papers from 9th Int. Conf. on Jack-Up Platform Design, (IMAC), Detroit, Michigan, 2004.
Construction and Operation, Sept. 2324, 2003, City 31. Ibsen, L.B., Liingaard, M. (2005) Output-Only Modal
Univ., London), Vol. 17, No. 34, MayAug., 165193 Analysis Used on New Foundation Concept for
14. Coldicott, L. (2005) Suction installation of cellular Offshore Wind Turbine, in preparation
skirted foundations, 4th year project report, Dept. of 32. Kelly, R.B., Byrne, B.W., Houlsby, G.T. and Martin,
Engineering Science, Oxford University C.M. (2003) Pressure Chamber Testing of Model
15. Doherty, J.P., Deeks, A.J. and Houlsby, G.T. (2004a) Caisson Foundations in Sand, Proc. Int. Conf. on
Evaluation of Foundation Stiffness Using the Scaled Foundations, Dundee, 25 Sept., Thomas Telford, pp
Boundary Method, Proc. 6th World Congress on 421431
Computational Mechanics, Beijing, 510 Sept., in press 33. Kelly, R.B., Byrne, B.W., Houlsby, G.T. and Martin,
16. Doherty, J.P., Houlsby, G.T. and Deeks, A.J. (2004b) C.M., 2004. Tensile loading of model caisson founda-
Stiffness of Flexible Caisson Foundations Embedded tions for structures on sand, Proc. ISOPE, Toulon,
in Non-Homogeneous Elastic Soil, Submitted to Proc. Vol. 2, 638641
ASCE, Jour. Structural Engineering Division 34. Kelly, R.B., Houlsby, G.T. and Byrne, B.W. (2005a) A
17. El-Gharbawy, S.L. (1998) The Pullout Capacity of Comparison of Field and Laboratory Tests of Caisson
Suction Caisson Foundations, PhD Thesis, University Foundations in Sand and Clay submitted to Go-
of Texas at Austin technique
18. Feld, T. (2001) Suction Buckets, a New Innovative 35. Kelly, R.B., Houlsby, G.T. and Byrne, B.W. (2005b)
Foundation Concept, applied to offshore Wind Transient Vertical Loading of Model Suction Caissons
Turbines Ph.D. Thesis, Aalborg University Geotechnical in a Pressure Chamber, submitted to Gotechnique
Engineering Group, Feb. 36. Lam, N.-S. and Houlsby, G.T. (2005) The Theoretical
19. Gottardi, G., Houlsby, G.T. and Butterfield, R. (1999) Modelling of a Suction Caisson Foundation using Hyper-
The Plastic Response of Circular Footings on Sand plasticity Theory, Proc. Int. Symp. on Frontiers in
under General Planar Loading, Gotechnique, Vol. 49, Offshore Geotechnics, Perth, Australia, Sept., in press
No. 4, pp 453470 37. Martin, C.M. (1994) Physical and Numerical Modelling
20. Houlsby, G.T. (2003) Modelling of Shallow Foundations of Offshore Foundations Under Combined Loads, D.Phil.
for Offshore Structures, Proc. Int. Conf. on Foundations, Thesis, Oxford University
Dundee, 25 Sept., Thomas Telford, pp 1126 38. Martin, C.M. and Houlsby, G.T. (2000) Combined
21. Houlsby, G.T. and Byrne, B.W. (2000) Suction Caisson Loading of Spudcan Foundations on Clay: Laboratory
Foundations for Offshore Wind Turbines and Anemometer Tests, Gotechnique, Vol. 50, No. 4, pp 325338
Masts, Wind Engineering, Vol. 24, No. 4, pp 249255 39. Martin, C.M. and Houlsby, G.T. (2001) Combined
22. Houlsby, G.T. and Byrne, B.W. (2005a) Design Loading of Spudcan Foundations on Clay: Numerical
Procedures for Installation of Suction Caissons in Clay Modelling, Gotechnique, Vol. 51, No. 8, Oct., 687700
and Other Materials, Proc. ICE, Geotechnical Eng., Vol. 40. Rushton, C. (2005) Cyclic testing of model founda-
158 No. GE2, pp 7582 tions for an offshore wind turbine, 4th year project
23. Houlsby, G.T. and Byrne, B.W. (2005b) Design report, Dept. of Engineering Science, Oxford University
Procedures for Installation of Suction Caissons in Sand, 41. Sanham, S.C. (2003) Investigations into the installa-
Proceedings ICE, Geotechnical Eng., in press tion of suction assisted caisson foundations, 4th year

92

Copyright 2005 Taylor & Francis Group plc, London, UK


project report, Dept. of Engineering Science, Oxford Turbines, Proc. 5th Chilean Conference of Geotechnics
University (Congreso Chileno de Geotecnia), Santiago, 2426
42. Villalobos, F.A., Byrne, B.W. and Houlsby, G.T. (2005) November
Moment loading of caissons installed in saturated 44. Watson, P.G. (1999) Performance of Skirted
sand, Proc. Int. Symp. on Frontiers in Offshore Geo- Foundations for Offshore Structures, PhD Thesis, the
technics, Perth, Australia, Sept., in press University of Western Australia
43. Villalobos, F., Houlsby, G.T. and Byrne, B.W. (2004) 45. Wolf, J.P. (1994) Foundation Vibration Analysis Using
Suction Caisson Foundations for Offshore Wind Simple Physical Models, Prentice Hall, New Jersey

93

Copyright 2005 Taylor & Francis Group plc, London, UK


Pipeline geotechnics state-of-the-art

D.N. Cathie & C. Jaeck


Cathie Associates SA/NV, Brussels, Belgium

J.-C. Ballard & J.-F. Wintgens


Fugro Engineers SA/NV, Brussels, Belgium

ABSTRACT: Pipeline geotechnics deals with soil-pipeline interaction. This covers installation issues (pipeline
penetration and short-term lateral stability), axial and lateral response to loads. It then encompasses pipeline
trenching, backfill engineering and pipeline stability when buried. This review provides an overview of all aspects
of pipeline geotechnics except trenching. The focus of the paper has been on the mechanics of each problem,
explaining the issues with a view to developing understanding, rather than providing ready made solutions. The
interested reader can make use of the references for going deeper into particular aspects of the subject.

1 INTRODUCTION geotechnical industry has put together a guidance


document to assist non-specialists to understand the
Subsea pipelines are laid on the seabed and may or basic data required from a survey (OSIF 1999). Char-
may not be buried. Pipeline design needs to account acterisation of the pipeline route with emphasis on
for the ways in which the pipeline interacts with the trenching issues is discussed in Cathie (2001).
soil. This paper is about pipeline-soil interaction and, in Pipeline design and engineering will require geo-
particular, the geotechnical issues that must be faced technical data if major assumptions are to be avoided.
by pipeline designers if soil-pipe interaction is to be For design, all soils require classification tests (par-
captured adequately in the design process. Pipeline ticle size distribution, index tests and basic shear
geotechnics is an emerging specialty that involves strength parameters). The in situ density of granular
applications of geotechnical theory and practice unique soils and the undrained shear strength of cohesive
to the construction of underwater pipelines. soils should be determined.
For this State-of-the-Art review, the material has If the soil is very soft it is not sufficient to collect
been organized broadly in the order in which the phe- drop cores as the disturbance can lead to underestimat-
nomena arise embedment and lateral friction mobil- ing the soil strength. This could affect axial and lateral
isation during pipelay, stability against current and friction assessments, and trenching engineering issues.
wave action when on the seabed, development of axial The box corer coupled with in situ vane testing (OSIF
friction as a result of thermal and pressure loading etc. 1999) is recommended for obtaining good quality data
Since embedment during pipeline affects axial and lat- in very soft soils. A substantial quantity of soil can be
eral friction this appears to be a logical progression. collected using a box corer and this may be necessary if
Some areas that are not covered include self-burial riser-soil interaction needs to be investigated in specific
of pipelines as a result of wave/current action and tests, or if backfill properties need to be investigated.
sediment transport, liquefaction around pipelines, and In granular material, the cone penetration test (CPT)
the whole area of ploughing and trenching (Cathie & remains the most attractive tool to determine the in situ
Wintgens 2001) which would make the paper too density (OSIF 1999). CPTs should always be com-
extensive. bined with sampling techniques in order to determine
the physical properties of the sand, particularly the
2 GEOTECHNICAL PARAMETERS grain size, which may be important for trenchability.

If soil-pipeline interaction is to be understood and 3 PIPELINE EMBEDMENT


modelled then it is clear that geotechnical data along
the pipeline route is required. This basis point is not Pipeline embedment begins during pipe lay and may
always recognized and a group involved in the offshore increase with time due to hydrodynamic forces, pipe

95

Copyright 2005 Taylor & Francis Group plc, London, UK


content load (e.g. hydrotest of the line) and creep. 4.5
Embedment is primarily an issue of penetration -
bearing capacity failure until the soil resistance is suf- 4.0
Very soft

Load Concentration in Soil (-)


ficient to resist the load applied by the pipeline. The Firm
3.5
loads applied by a pipeline during installation will be
discussed first followed by the bearing capacity 3.0
issues.
2.5

3.1 Soil loading during pipeline installation 2.0

Consider a pipeline being laid in perfectly still condi- 1.5


tions. As the pipeline is laid on the seabed the force
applied by the pipe near the touchdown point is much 1.0
greater than the submerged weight of the pipe as a
0.5
result of the seabed interrupting the normal caten-
ary of the pipe. An analysis of soil reaction loads for 0.0
a 12 pipeline in very soft and firm soil conditions is 0 10 20 30 40 50
shown on Figure 1 computed using a finite element Distance from Touch Down Point (m)
analysis which models the pipe lay process and an
elastic-plastic seabed. Figure 1. Example of load concentration at touchdown point.
In the example shown the load concentration fac-
tors (applied to the pipe submerged weight) are 2.0
and 4.2 for the very soft and firm soil respectively. stability), is critically dependent on embedment, par-
The load concentration is also a function of the caten- ticularly for soft clay soils.
ary shape (including the tension in the pipeline) as
well as the soil stiffness. 3.2 Pipe penetration analysis
During real pipelaying operations, vessel heave
occurs which induces an additional vertical motion at Pipe embedment is fundamentally a large deform-
the vessel end of the pipe and cyclic touchdown and ation penetration problem rather than a bearing cap-
lift off cycles at the seabed end. This vessel motion acity problem. As the pipeline touches down, plastic
will inevitably induce some changes in pipeline ten- deformation of the soil occurs until the penetration is
sion on the seabed and therefore additional soil pres- sufficient (i.e. the bearing area is large enough) to
sures. Changes in tension are also introduced as new provide the resistance necessary. If the soil loading
sections of pipe are welded and then over-boarded in during touchdown is greater than the pipeline weight
a stepwise manner. This change in tension causes sec- then the soil experiences unloading as the pipe lay
tions of pipe near the seabed to touchdown and lift off process moves on. Therefore, most pipelines are
more than once. believed to be overpenetrated (terminology of
Wave and current action on the pipeline in the Zhang et al. 1999), i.e. they have experienced a verti-
water column will provide another source of dynamic cal load greater than they are experiencing at present,
loading and this could be in any direction causing the a state that is not unlike overconsolidation in some
pipeline to move horizontally as well as vertically as respects. Of course the degree of overpenetration may
it touches down. Pipeline motion near the seabed vary, for example if the pipe is filled with water for
could also produce a scouring effect from below the hydrotesting, or with oil for production.
line in granular soils. Various approaches have been proposed to analyse
The purpose of describing pipeline behaviour dur- the static penetration problem (Fig. 2a). Early attempts
ing laying is to provide the right framework for under- were based on considering the pipe as a strip footing
standing pipeline embedment. Any analysis that is where the width of the footing is taken as the chord
performed should consider (a) the static pipeline weight, length of the embedded section of the pipe (Small et al.
(b) the amplification of that load at touchdown, and 1971). Recent work on riser-soil interaction has con-
(c) the likely magnitude of cyclic loading (both vertical tinued to use this method (Bridge et al. 2004). Verley
and lateral) that will accompany the laying process. All & Lund (1995) proposed an empirical equation for
of these processes will result in the pipe being embed- penetration in clay, based on an extensive laboratory
ded deeper than a simplified analysis based on the testing programme, expressed as
weight alone would suggest. This is extremely import-
ant because lateral resistance that is needed during
pipelaying to negotiate a turn, or that is needed to resist (1)
hydrodynamic forces during operations (on-bottom

96

Copyright 2005 Taylor & Francis Group plc, London, UK


heave Wagner et al. (1987) and is applicable to remoulded
clay with an undrained shear strength around 1 kPa.
Some of the scatter may arise from the difficulty of
z
measuring strength and its variation with depth.
The plasticity solutions do not account for buoy-
a) ancy effects, soil heave, or any increase in strength
with depth. Murff et al. have shown the potential impact
of heave and increase in strength. The effect of buoy-
ancy would typically be about 510% depending on
the soil strength but would be more important in fluid
muds. Soil heave is potentially even more important.
b) Soil heave around the pipe increases the bearing area
by about 20% at z/D  0.2 and increases the contact
Figure 2. Pipeline embedment (a) initial penetration, (b) area at the same depth by 35%. At a penetration of
lateral movement. about z/D  0.25 almost the full diameter of the pipe
is bearing on the soil when heave is accounted for. It
6
is difficult to see how laboratory and field data can be
properly interpreted without taking account of the
geometric changes that occur with penetration. There
5 is scope for further theoretical work in this respect.
Since pipe penetration involves remoulding the
Normalised resistance (F/SuD)

soil locally to the pipe wall, and since repeated load-


4
ing due to hydrodynamic effects would only accentu-
ate this effect, it seems reasonable to assume that
3 penetration assessment should consider a zone of
remoulded soil below the pipeline during laying. This
would suggest that using a low soil-pipe adhesion for
2
embedment assessment would be appropriate.
In the view of the authors, the issue of initial pene-
1 tration resistance is relatively well understood and
established. Focus should now be on the nature and
magnitude of the loading applied by a pipeline during
0 laying, and the possible magnitude of the cyclic effects.
0.0 0.2 0.4 0.6 0.8 1.0
Normalised penetration (z/r)
Murff et al - rough Murff et al - smooth
Verley linear Verley and Lund 3.3 Penetration due to repeated loads
Data (Murff et al)
As discussed above, as a pipe is laid on the seabed the
movement of the pipe during the laydown process
Figure 3. Lower bound plasticity solutions and empirical will increase the penetration.
approaches for pipeline embedment for cohesive soil. Cyclic vertical loading has been investigated by
Dunlap et al. (1990), Fontaine et al. (2004) and in the
where z  penetration, D  pipeline diameter, STRIDE and CARISIMA joint industry projects
su  undrained shear strength,   soil unit weight, aimed at understanding riser-soil interaction (Bridge
Fz  vertical load per unit length on soil. et al. 2004). The latter authors provide a clear explan-
Verley & Sotberg (1992) propose equivalent equa- ation of repeated penetration and pullout, particularly
tions for penetration in sand but for cyclic loading. focusing on the suction that develops during pullout
Murff et al. (1989) have presented upper and lower in soft cohesive soils. Dunlap et al. indicate that
bound plasticity solutions for rough and smooth pipes limited cyclic loading without break out resulted in
(full adhesion and no adhesion) in cohesive soils. The little additional burial while larger cyclic loads which
lower bound results are shown on Figure 3 and com- broke suction and pulled the pipe free resulted in fur-
pared with experimental data and the Verley & Lund ther penetration. This may be due to the additional
approaches. Upper bound solutions were only slightly remoulding experienced in the full breakout case.
higher. Repeated lateral loading induced by current load-
These solutions are of most importance in very ing and dynamic response of the suspended section of
soft clays where embedment may be significant. Much pipeline can result in a considerable increase in pene-
of the experimental data shown in Figure 3 is from tration (Morris et al. 1988), particularly if the pipe is

97

Copyright 2005 Taylor & Francis Group plc, London, UK


not overpenetrated. Consider a pipe penetrated with a properties also take on more importance for lateral
certain vertical load and let us assume that the load is and axial resistance.
kept constant. Any horizontal load will then result in
both vertical and horizontal movements since the pipe
4.1 Lateral resistance
is on the failure envelope. This is discussed further in
the next section. Three different approaches have evolved for assess-
Morris et al. (1988) carried out laboratory tests to ing lateral resistance:
assess the embedment due to horizontal cyclic load-
ing in very soft clay. The additional penetration was 1 a single friction factor approach where the lat-
largely dependant on the magnitude of the force, or eral resistance is related to the submerged weight
displacement, and the duration during for which it of the pipeline and the soil type;
was applied. Although the rate of penetration was 2 a two component model consisting of a sliding
found to decrease over the duration of each test, this resistance component and a lateral passive pressure
rate did not appear to reach a constant value or the component frictional model supplemented with
pipeline to reach a limiting burial depth. As the pipe passive resistance of the wedge of soil (Nyman
continued to sink, the mounds grew in size, attaining 1984, Wagner et al. 1987, Lieng et al. 1988, Verley &
heights of up to 0.8 D. Morris et al. also considered Sotberg 1992, Verley & Lund 1995);
the accumulated burial effect of variable cyclic load- 3 a plasticity model approach (Zhang et al. 1999,
ing and found that the effect could be modelled with 2002, Cassidy 2004).
sets of uniform cycles. Generally, the two component models are based on
Verley & Lund (1995) proposed the empirical empirically fitting laboratory test data. A summary of
equation some of the proposed equations is given in Table 1.
While being practically useful, these models do not
enlighten the user with the actual mechanics of the prob-
(2) lem. Moreover, they become more and more empirical
when cyclic behaviour needs to be introduced.
The plasticity framework outlined by Zhang et al.
where a is the amplitude of horizontal movement, to (1999, 2002) provides a much more fundamental way
assess the maximum penetration that can be achieved of understanding the mechanisms involved as well as
for a given amplitude of motion and to assess the being much more general. Zhang et al. have developed
development of the penetration as a function of the the plasticity model for calcareous sands but the
work done by the pipe on the soil. The authors suggest application to clays should be straightforward and is
that the applicability should be limited to z/D of 0.3 in progress by the authors. The concepts are now well
due to range investigated in their study. developed for surface footings. As described by
Lund (2000) performed an investigation on the pene- Cassidy (2004) the models contain a yield surface, a
tration depth achieved for a large diameter pipeline strain-hardening expression, elastic behaviour inside
(1.2 m OD). While the calculated embedment was the yield surface, and a flow rule to define the direc-
about 0.05 m, the actual embedment varied between tion of movement during yield.
0.250.4 m. Much of the route was sand. Lund con- Considering pipe behaviour under combined verti-
cluded that much of the additional embedment was cal (V) and horizontal (H) loads in calcareous sand in
due to lateral oscillation at the touchdown point. centrifuge tests, Zhang et al. have confirmed what
was known for surface footings that the shape of the
yield surface is almost parabolic. A simplified repre-
4 LATERAL AND AXIAL RESISTANCE OF sentation was proposed, given by
PARTIALLY EMBEDDED PIPELINES

Lateral and axial resistance of an unburied pipeline (3)


needs to be assessed at the design stage. Lateral resist-
ance is important for the design of the pipeline on-
bottom weight and any weight coating that may be where  is a parameter associated with the friction
required. Axial resistance controls pipeline expansion between pipe and soil for low vertical load, Vmax and
and affects end connections, spool pieces, and upheaval Vmin are the positive and negative intercepts (H  0)
buckling. Since environmental conditions (wave/cur- on the vertical load axis, and
rent) leading to lateral forces and thermal/pressure (4)
effects leading to axial forces are variable and cyclic in
nature, the subject must be considered for both mono- where is determined from calibration tests. This
tonic and cyclic loading. The pipeline outer coating expression gives an intercept on the horizontal load

98

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 1. Examples of lateral resistance models.

Reference Equations Details

Wagner et al. (1987) Sands:  8.6 kN/m3


Fy   (W FL)   A Monotonic
  0.6
 38
Fy  horizontal resistance  9.6 kN/m3
A  0.5xEmbedded area Monotonic
W  submerged pipe wt   0.6
FL  hydrodynamic lift  79
Cyclic load
( static failure)
Embedment x 2
reduced by 50%
Cyclic load
( 5% D)
Embedment x 3
reduced by 8090%
Clays: Monotonic
Fy   (W FL)  Su A/D   0.2
 39.7
Cyclic
( static failure)
Embedment x 2
 31.7
All clays
cyclic load
( static failure)
Embed x 2.5
 15.7
Lieng et al. (1988) Fy   (W FL)  FR   0.6 (sands)
  0.2 (clays)
FR calculated considering
accumulated energy
Verley & Sotberg (1992) Fy  Fc  FR All sands
Fy   (W FL)  FR   0.6
FR  D2 (4.5 0.11D2/Fc) (z/D)1.25
Verley & Lund (1995) Fy  Fc  FR Clays
Fy  (W FL)  FR (Su 70 kPa)
  0.2
FR  4.13 DSu(Su/(D))
0.392 (z/D)1.31

axis when vertical load is zero, and represents a pas- The plastic potential takes a different but similar form
sive soil resistance with a magnitude Vmin. Vmax rep- to the yield surface, and was defined by considering the
resents the maximum vertical load for a given displacement increment vectors in different tests:
penetration (the preload). Figure 4 shows the nor-
malised form of the yield surface for different values (5)
of . The peak horizontal resistance is achieved at
about 40% of the maximum vertical load.
The proposed hardening function is based on the where t is the shape parameter. The exponent m has
monotonic vertical penetration resistance of the pipe the effect of adjusting the value of the vertical load at
(plastic stiffness) and the rebound response gives the which the normal to the plastic potential becomes
elastic stiffness. This enables the increment of vertical parallel to the H axis. Figure 5 shows the plastic
plastic strain to be defined in terms of the elastic and potential for different values of m. The model predicts
plastic stiffnesses and the increment in Vmax. upwards pipe movement and strain softening when

99

Copyright 2005 Taylor & Francis Group plc, London, UK


-0.2 Normalised lateral resistance H/ 'z2
=0.7 0 50 100 150 200
0 0

0.2
0.1
V/Vmax

0.4

Embedment (z/D)
0.6 =0 0.2

=0.1
0.8
=0.2
0.3
1
0.0 0.1 0.2 0.3 0.4 0.5 0.6
H/Vmax 0.4

Figure 4. Yield envelope for sand (Zhang et al. 1999).


0.5
Plasticity model Verley & Sotberg (1994)

-0.2 Lieng et al (1988) Palmer et al, 1988


Brenodden et al, 1989 Wagner et al, 1987
=0.05
Verley and Sotberg (1994) Zhang et al (2001)
0
Figure 6. Lateral resistance models for sands (monotonic
0.2 tests).
V/Vmax

0.4
Normalised lateral resistance H/ 'D2
0.6 m=0.1 0 1 2 3 4
0
m=0.2
0.8 m=0.3
m=0.4
0.1
1
0.0 0.1 0.2 0.3 0.4 0.5 0.6
H/Vmax
Embedment (z/D)

0.2
Figure 5. Plastic potential for sand (Zhang et al. 1999).

0.3
V/Vmax is between 0 and 0.25 depending on the
exponent m.
For the calcareous sand tested by Zhang et al. 0.4

Figure 5 makes it clear that unless the pipe is highly


overpenetrated (i.e. V/Vmax is very low) further pene-
tration of the pipe can be expected under horizontal 0.5

loading. Plasticity model Verley & Sotberg (1994)


The model has been calibrated for the specific cal- Lieng et al (1988) Palmer et al, 1988
careous sand. However, it is of interest to see how the Brenodden et al, 1989 Wagner et al, 1987
Verley and Sotberg (1994) Zhang et al (2001)
model compares with other methods for calculating
lateral resistance. Care must be taken to correctly
interpret laboratory tests in terms of whether they are Figure 7. Lateral resistance model for sands normalised
with pipe diameter.
sideswipe tests (horizontal displacement control at
constant vertical displacement) or probe tests (hori-
zontal displacement at constant vertical load). the data scatter in a good light. Presenting data in
Figure 6 shows experimental data for monotonic terms of D (Fig. 7) shows the wide range of data.
tests with an overpenetration ratio of 1 (current verti- Both plots demonstrate that the Zhang et al. plasticity
cal load is the maximum experienced). Normalisation model is a reasonable representation of all the data
with the penetration depth is logical since vertical despite the fact that it has not been calibrated for
penetration is a key variable but the plot tends to put other sands.

100

Copyright 2005 Taylor & Francis Group plc, London, UK


Friction Factor Table 2. Resistance factor for pipeline axial friction coef-
0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 ficient under fully drained conditions (after Finch et al.
0
2000).

Condition fr
0.1

Granular cohesionless soil and fr  1


Embedment (z/D)

0.2
D50 pipe roughness
Granular cohesionless soil and 0.75 fr 0.9
D50  pipe roughness
0.3 Fine-grained cohesive soil and fr  1
D50 pipe roughness
Clay and D50  pipe roughness fr  0.6
0.4 Silt and D50  pipe roughness fr  0.4

0.5
Strictly, this is only valid for drained conditions,
Plasticity model Verley & Sotberg (1994)
Lieng et al (1988) Palmer et al, 1988 but this may be a reasonable assumption for both
Brenodden et al, 1989 Wagner et al, 1987 sands and clays if the loading rate is slow enough. For
Verley and Sotberg (1994) Zhang et al (2001)
thermal expansion, the temperature increase is likely
to take several hours and this could be taken as justi-
Figure 8. Lateral resistance models for sands in terms of fication for a drained analysis.
friction factor.
Axial friction assessment then reduces to evaluat-
ing the friction coefficient .  depends on the internal
friction angle of the soil and on the properties of the
In order to relate the models back to the traditional soil-pipeline interface. There are various guidelines
friction factor, the horizontal resistance has been nor- used in the offshore industry for both pipelines and
malised with the applied vertical load to give the piles:
friction factor (Fig. 8). The low prediction of the
plasticity model at greater depth reflects the Zhang   tan( 5) (API RP2A WSD 2000)
data but not the other data.   2/3 tan() (Bureau Vritas)
The plasticity model is likely to require develop-   fr tan() (Finch et al. 2000)
ment as the model described above is a single surface The first two formulations assume that the inter-
strain hardening model. Other approaches are likely to face is soil-steel. For pipelines this is rarely the case
be required for modeling cyclic loading. In particular, since the outer coating is generally a corrosion pro-
geometric changes will need to be considered where tection such as polypropylene (PP) or a concrete
large lateral movements of several diameters occur. weight coating. PP coatings may be smooth but spe-
cial materials can be ribbed to improve friction for
4.2 Axial resistance transport and handling. Based on research performed
on a range of coatings (Finch 1999), Finch et al.
Axial loads normally apply some time after installation. (2000) recommend values of fr shown in Table 2 as a
The implications are different for sands and clays. In function of coating roughness and soil grain size:
sands, wave and current action may have induced For fine-grained sediments, and where loading
minor lateral loading which has in turn induced some may be rapid enough to elicit an undrained response
further embedment or densification of the soil. There from the soil, the axial resistance would be a function
could be some build up of sediment against the pipe. of the contact area, and of the undrained shear
In clays, set-up following remoulding will have had strength of the soil, expressed as
time to develop. This could include components of
strength increase arising from thixotropy and consoli- (7)
dation. It is also likely that good adhesion between the
pipe and the soil will have developed in soft clays. where   adhesion factor and L  arc length in
Although the axial resistance is influenced by embedded soil (including heave).
embedment and time dependent factors such as those Appropriate values of the shear strength and adhe-
described above, simple Coulomb friction models are sion factor will depend on whether the peak or residual
often adopted to evaluate the axial resistance of par- axial resistance is required, and how long the pipeline
tially embedded pipelines in all soils, given by has been installed without load. Laboratory shear
tests are recommended for the specific soil and coat-
(6) ing under consideration. In very soft clays, in the

101

Copyright 2005 Taylor & Francis Group plc, London, UK


absence of specific data, and subject to the roughness Table 3. Classification of hydraulic fills (Whitman 1970).
of the interface, the adhesion factor may be taken as 1
for the peak resistance and related to the soil sensitiv- Nature of original material Characteristics of fill
ity, St, for the residual strength (  1/St ).
Fairly clean sand Reasonably uniform fill
The research programmes discussed by Finch
( 15% passing No.200 sieve) of moderate density
(1999) have shown that the peak resistance in sands is Silty or clayey sand Very heterogeneous fill
typically achieved within an axial displacement of of large void ratio
2 mm. This value is consistent with the 0.1 (2.54 mm) Stiff cohesive soil Skeleton of clay balls,
generally considered for the axial friction mobiliza- with matrix of sand and
tion of tz curves of driven piles (API RP2A WSD clay
2000). A linear elastic-perfectly plastic axial resistance Soft cohesive soil Laminated normally
mobilization curve (with a peak resistance reached at consolidated or
around 2 mm) is therefore a reasonable assumption in underconsolidated clay
granular soils.

4.3 Axial creep or walking 5.2 Mechanical backfilling

Observations and analysis have shown that pipelines Mechanical backfilling involves scraping the spoil
can walk or creep axially (Tornes et al. 2000, Carr et al. (previously removed from the trench) back into the
2003) due to internal heating and cooling. The driving trench. It is applicable to all types of soil conditions.
mechanism is the expansion and contraction of the The backfilling process is discussed in detail in Cathie
pipeline and whether there is an effective anchor et al. (1998). Soil in the spoil heaps, and sometimes
point where no movement occurs. The rate of creep some of the in situ seabed soil, is mixed and deposited
will depend not only on the temperature profiles but into the trench very rapidly. Water is believed to be
also on the magnitude of the axial resistance, the entrained with the backfill and the resulting mass is
mobilization distance and the degradation to residual expected to have a higher macro water content than in
conditions. the spoil heaps, particularly if the soil contains a cohe-
sive component.
Mechanical backfills are not unlike hydraulic fills
and a starting point for considering the properties of
5 PROPERTIES OF TRENCH BACKFILLS backfills is to use this work. Whitman (1970) pro-
posed a classification system, which provides a start-
The properties of a trench backfill are necessarily a ing point.
function of how the backfill is placed. Four broad cat- There is no specific published data on the proper-
egories can be considered: ties of mechanically backfilled trenches as far as the
authors are aware.
natural infill (wave/current induced);
For sand backfills, some information about in-situ
mechanical backfilling;
densities after hydraulic filling and slumping is pro-
backfill following jetting;
vided by Stoutjesdijk et al. (1998). They considered
active jet cutting and collapse.
the formation of submarine slopes with hydraulic
sand fill and found that very low densities could
develop during hydraulic filling (1030%) but that
5.1 Natural infill
liquefaction and flow could increase the relative dens-
Wave or current induced natural infill is applicable ity (to 2050%). The rate of filling was apparently not
mainly in sands and in relatively shallow water where important (Bezuijen & Mastbergen 1988). It seems
seabed currents are sufficient to induce transport. reasonable to assume that similar considerations apply
Soil particles are deposited under a relatively high to mechanical backfilling of sands. Therefore, sand
energy environment which results in a structure that is backfills are expected to be in a loose state after back-
typically loose to medium dense. Rates of trench filling. It is not difficult to demonstrate that drainage
infill can be estimated using methods such as Schapp of a 1.5 m trench should be largely complete in a few
(1982) and Niedoroda & Palmer (1986), or by minutes for most sands.
directly modeling the flow regime accounting for Clay backfills or mixed sand/clay materials are
spoil heaps using computational fluid dynamics simu- believed to be heterogeneous after backfilling. Stiff
lations. Other relevant information is given in Van clay is ploughed out of the trench into the spoil heaps
Rijn (1993) and Fredsoe (1978). Note that natural and then left exposed to free water until backfilling
densification can also occur with time as a result of takes place. The surfaces of the lumps will take in water
wave action (Clukey et al. 1989). and soften. During backfilling, further disturbance

102

Copyright 2005 Taylor & Francis Group plc, London, UK


and deformation of lumps will occur and the voids similar to the method of water pluviation to achieve a
between lumps will be filled with water, slurry and minimum density soil (Kolbuszewski 1948) and it is
more mobile sand components. Further details and known that slower rates of deposition/sedimentation
discussion can be found in Karthikeyan et al. (2001, lead to higher densities. Resedimentation tests by
2002), Hartlen & Ingers (1981) and Mendoza & Kvalstad on fine sand have shown that relative dens-
Hartlen (1985). Lumpy clay backfills consisting of ities in the range 1020% have developed. The dens-
stiff clay will generally consolidate much more ity of the backfill is likely to be affected by:
quickly than would be expected for homogeneous in-
rate of trenching and sedimentation, use of back-
situ soil due to the voids and channels available for
flow jets (rate of deposition and water flow through
free water. A model for consolidation of lumpy clay
the sand affects final density (Kolbuszewski
fills has been proposed by Yang et al. (2002).
1948);
Soft clay backfills created by mechanical trench-
particle angularity, grain size distribution (uniform-
ing and backfilling are also believed to be heteroge-
ity) and grain size, mineralogy (Youd 1973, Hight
neous and consist of softened and remoulded material
et al. 1999);
close to the in-situ water content in a slurry of much
depth of backfill (density of very loose sand will
higher water content soil. Their properties are strongly
probably increase with stress level due to the high
time dependent as consolidation and thixotropic regain
contractive potential at very loose states);
takes place. Bruton et al. (1998) studied the consoli-
time, Pipeline movements, wave action, and ageing
dation of clay slurry in the laboratory and showed that
will all tend to cause the loose sand to densify
cavities and channels formed within the mud as it set-
with time.
tled. Such channels tend to form even in a homoge-
neous slurry as it sediments but are further encouraged Some experiences with uplift of pipelines in very
if it is heterogeneous, if there is sufficient adhesion in loose backfills appears to confirm that the sand can
the clay to permit cracks to remain open, and if there be very loose and susceptible to structural collapse
are silt or sand inclusions. This heterogeneity was and static liquefaction. Laboratory testing on soils at
found to result in much more rapid consolidation than very low densities requires special techniques (Lade
would have been measured in an oedometer test on 1992). Logically, samples prepared by water pluvia-
homogeneous soil. Mechanical backfilling probably tion would appear to have the best potential for repre-
does not destructure the soil as completely as jet senting the fabric of a post-jetted sand.
trenching and therefore the quality of the backfill soil Clays. A jetted backfill consisting of clay is
should also be better. This is an area of active work at believed to consist of lumps of semi-intact material in
the present time and much is as yet unpublished. a matrix of unconsolidated slurry. While not unlike
the mechanical backfill of the same material, the
water content is probably higher initially and the
5.3 Backfilling following jetting
remaining intact soil reduced to smaller more dis-
Lowering or burial of a pipeline by jetting results in turbed lumps. Again, very little field data has been
destruction of the in situ soil, local movement of the published but Newson et al. (2004) have demon-
soil behind the trencher and deposition of the soil par- strated by in situ CPT and T-bar tests in a high plas-
ticles or lumps back into the trench or just outside. ticity clay in the Nile delta that about 50% of the in
While clays are cut and broken before being trans- situ strength was regained within 3 months and little
ported, sands and silts are eroded and transported in additional strength gain occurred over the next
suspension. 5 months. This appears to confirm the assumption of
Sand backfills. There is almost no data related to heterogeneous conditions postulated for mechanical
actual soil densities in a real backfill post-jetting. backfilling which allows drainage and consolidation
However, Kvalstad (1999) does indicate that low cone to occur relatively quickly. Unpublished laboratory
resistances have been measured even some time after testing on soft blocky clays does confirm the same
trenching indicating the sand to be loose to very hypothesis but no testing that really simulates the jet
loose. These loose sands had apparently been stable for cutting and deposition process has been attempted to
a long period. We have also seen very loose sand off- date. One important conclusion however from this
shore that have cone resistances equivalent to soft clays. testing is that shear strength of blocky soil for uplift
As discussed for mechanical backfilling, the resistance may be characterised by a frictional model
mechanisms described in Bezuijen and Masterbergen as used for sands (Bruton et al. 1998, Bolton &
(1988) for example for hydraulic filling are likely to Barefoot 1997) although this should be confirmed by
be relevant. However, the mechanism of sedimenta- laboratory testing on project specific soils.
tion of sand in the trench following trenching is quite Finally, the strength regain of a soft clay backfill is
rapid but may occur in a low energy environment. made up of both reconsolidation and thixotropic
Rapid sedimentation of sand after trenching is quite regain. Laboratory testing is beginning to be used to

103

Copyright 2005 Taylor & Francis Group plc, London, UK


characterise both thixotropic regain (without change lower density may need to be held in position, for
in water content) and consolidation strength regain. example, by stitch rockdumping.
Reference is made to Burland (1990), Leroueil &
Hight (2003), Leroueil (2003), Locat et al. (2003),
6.2 Trenching
Schmertmann (1991), Sills (1995), Silva (1974),
Skempton & Northey (1952) and Skempton (1970) for Pipeline lowering and burial by jet trenching requires
further details related to the behaviour and strength of cutting, erosion and fluidization of the soil by the jets
soft clay backfills. and lowering of the pipeline into the area cut or flu-
idized by the jets. Clays are cut and broken by the jets
and transported behind the trencher. Sometimes educ-
6 PIPELINE STABILITY DURING tors are used to clean out the trench since the clay
TRENCHING AND BACKFILLING walls remain stable for some time. In the view of the
author, the problem of pipeline lowering and stability
6.1 Ploughing is mainly dependent on the amount of slurry or lumps
left in the base of the trench. However, Powell et al.
Pipeline stability is generally not an issue during (2002) suggest that the product specific gravity
trenching by ploughing since the pipe is lowered into should be greater than the specific gravity of a lique-
the cut trench as the plough moves forward. However, fied clay at the onset of effective stress. This may not
backfill ploughing can lead to pipeline uplift in cer- be a useful criterion since it is typically around 1.2 for
tain conditions (Cathie et al. 1996, 1998). A pipeline most muds.
lying in a V-shaped trench may be susceptible to uplift For sands, the trench walls are not stable and the
caused by: pipe lowering is dependent on the pipeline being
heavier than the fluidized soil. The sand minimum
transverse flow of the soil down the slopes and high
density could be used as a guide to the minimum pipe
transient hydraulic pressure below the pipe;
weight but the author is unaware of work specifically
turbulence and in-line water flow driven by the
aimed at defining what this minimum pipe weight
backfill plough;
would be. Nevertheless, it is clear that heavier, more
pipe spans and out-of-straightness;
flexible, pipes will be easier to lower than light stiff
low pipe weight;
pipes. Flexure of the pipe below the trencher coupled
soil liquefaction and slow drainage.
with maximizing the length of fluidised soil will
Pipeline uplift is known to have occurred in result in the most efficient lowering process.
pipelines with a specific gravity of between 1.2 and 1.6. To minimize the risk of floatation in sands (or to
As suggested by Cathie et al. (1998), there are usually maximize the lowering efficiency), Powell et al.
several factors that combine to cause uplift during (2002) suggest that for rapid draining soil a specific
backfilling. A backfill plough creates considerable tur- gravity of 1.5 is sufficient but that if silt or clay are
bulence and the effect of the plough (particularly the also present (lowering the permeability) then the risk
mould boards) and the soil mass advancing rapidly of floatation is significant if the product specific
can cause uplift of a lightweight pipeline in a trench. gravity is less than 1.7.
Flow of soil down the slopes of the trench is an The subject of pipeline stability during trenching
important cause of uplift since high transient uplift and backfilling is still wide open for further insights
forces are developed when the soil impacts the and research. We caution against focusing too much
pipeline (Powell et al. 2002). Both transient hydraulic on the 2 D aspects of the problem (cross section) and
pressures and possibly wedging effects if the soil has only on soil mechanics. Uplift initiation and propaga-
a shearing resistance would act to destabilize the pipe. tion during trenching/backfilling is a 3 D problem and
The kinetic energy of the soil flow is likely to be involves pipeline structural response to bending and
greater if the trench is deep and the backfilling is hydrodynamic loads, soil transport and deposition,
rapid. High backfilling speeds, even if short lived, and consolidation. These aspects should be investi-
may create conditions in with the uplift initiates. After gated together.
initiation of an uplift feature, the gap between the
trench base and the pipe is ahead of the plough. This
in turn permits soil to flow under the pipe and propa- 7 AXIAL AND TRANSVERSE RESPONSE
gate the feature in a progressive manner. OF BURIED PIPELINES
According to Powell et al. (2002), and based on
both model testing and experience, uplift does not Axial and transverse response of a buried pipeline is
occur if the pipe weight is sufficient. They suggest important for assessing the behaviour of the pipeline
that the minimum specific gravity for a mechanic- to hydrotest or operational load conditions. Axial
cally backfilled pipeline should be 1.8. Pipelines of loads induced by internal pressures or temperatures

104

Copyright 2005 Taylor & Francis Group plc, London, UK


render a pipeline susceptible to transverse loads as a mon practice to assume that operational loads
result of its out-of-straightness i.e. imperfections. develop relatively slowly (e.g. over a period of hours
Generally the most serious imperfections are in the for temperature increases) and that both sands and
vertical plane and make the pipeline susceptible to clays can be treated as drained for axial loading
upheaval buckling. However, horizontal or inclined (Finch et al. 2000). However, this may not necessarily
imperfections can also occur due to laying around a be the case and it is prudent to consider both drained
bend or to the pipeline lying on one side of a sloping and undrained response in the design analyses unless
trench. it can be demonstrated otherwise.
Therefore, the pipeline response to both axial and For undrained conditions, the undrained shear
lateral loads is important for assessing buckling strength is the basic parameter, with the resistance
potential. An early case history presented by Nielsen given by
et al. (1990) focused industry attention on the prob-
lem and Pedersen & Jensen (1988) and Nielsen et al. (10)
(1988) suggested solution methods. Since then a lot
of experimental work has been performed to charac- Finch (1999) suggests that for clays with low shear
terize uplift resistance in various soils. We provide a strength, values of  should be 1.0 for peak resistance
summary in Section 7.2. and about 1/St where St is the sensitivity of the soil for
Transverse forces in a pipeline are a result of the a residual strength.
internal compressive forces which are unable to be Using axial resistance based on shear strength
released due to the axial restraint of the soil sur- implies a uniform strength along the length in areas of
rounding the pipe. A model for the axial response is uniform backfill. This is not the reality. Seabed fea-
therefore necessary and this is discussed first. tures and out-of-straightness have the effect of increas-
A buried pipeline may be partially in contact with ing the apparent frictional resistance due to high
relatively undisturbed soil as well as being sur- pressures on supporting areas. Finch recommends to
rounded by backfill. Axial response is reasonably consider  as 1.0 but use of equation 10 can result in
assessed assuming the pipeline is surrounded by values of  greater than 1.
backfill but considering the effect of some contact
with firmer/denser soil as part of a parametric analy- 7.1.2 Axial response
sis. Lateral response of a buried pipeline should con- The response of buried pipelines to axial loads is gen-
sider in which direction the lateral movement is erally assumed to be similar to partially buried
taking places and select soil properties accordingly. pipelines as discussed in Section 4. It is difficult to
For both axial and lateral resistance assessment, escape the analogy with axial loading of piles and
consideration should be given to whether the backfill similar approaches could be considered (e.g. Kraft et al.
will be in a drained or undrained condition. 1981, Randolph & Wroth 1978). Differences arise
because of the infinite length of the pipeline and
7.1 Axial behaviour the non-axisymmetric nature of the problem. Never-
theless, it can be reliably speculated that the soil shear
7.1.1 Ultimate axial resistance modulus would fundamentally govern the initial axial
The ultimate axial resistance of a buried pipeline in response. Since the modulus is likely to be related to
freely draining soil may be determined by consider- the stress level, the axial stiffness of the material
ing the mean normal (lateral) pressure,  on the around the pipe would vary, particularly for shallow
pipeline and the axial friction coefficient, fr (see burial. In practice, the traditional elastic-perfectly
Section 4). plastic model is generally used in finite element
Considering the normal stresses on the top, bottom analyses of buried pipes using a mobilization distance
and sides of an equivalent square leads to (Schaminee of 23 mm to mobilize ultimate resistance, values that
et al. 1990, Finch et al. 2000): are justified by some experimental work.

(8) 7.2 Transverse behaviour


Since the most critical form of transverse response is
uplift, this is addressed first. A brief treatment of
so that the axial resistance of a buried pipe is: transverse resistance in other directions is given later.

(9) 7.2.1 Uplift resistance in uniform soil


Uplift resistance models were developed initially for
For cohesive soils, a decision has to be made the pull-out capacity of anchors. Rowe & Davis (1982)
regarding drained or undrained behaviour. It is com- made theoretical and experimental investigations and

105

Copyright 2005 Taylor & Francis Group plc, London, UK


z
W
H

D
P

Figure 10. Circulation mechanism in very loose sand


(Vanden Berghe 2005).
Figure 9. Uplift wedge failure mechanism.

Table 4. Uplift factors (Schamine et al).

provide a good summary of earlier work covering Soil Uplift factor, f []


sands and clays. Subsequently, Murray & Geddes
(1987), Dickin (1988) and others developed the under- Very loose sand 0.15
Loose sand 0.4
standing mostly concentrating on uplift of anchor Gravel/rockfill 0.6
plates in sand. Recently, Merifield et al. (2001) have
provided a rigorous theoretical solution for the
undrained anchor problem.
Work on buried pipes includes Matyas & Davis test results (Schamine et al.) and particularly for the
(1983), Trautmann et al. (1985) and others. As the loose soils.
offshore industry recognized the upheaval buckling Wedge failure models can broadly be classified into
risk of buried pipelines, Schamine et al. (1990) pub- a) simple vertical slip model (e.g. Schamine et al. see
lished a fundamental set of experimental data applic- Fig. 9 with   0) and b) inclined wedge models (e.g.
able to subsea pipelines. Further experimental data White et al).
has been collected (Ng & Springman 1994, White et al. Consider a vertical block mechanism (Fig. 9 with
2001) but much is proprietary and only partially pub-   0) and assume only the soil above the crown of
lished (Finch 1999, Finch et al. 2000, Fisher et al. the pipe is active (height, H). The uplift resistance, P,
2002). per unit length for frictional (drained) behaviour is
Uplift resistance models for sand have been devel- given by:
oped from the work on anchors in which clearly
defined slip planes are apparent up to relatively large
depths and the focus was on peak uplift resistance. As (11)
shown by White et al. (op cit), a peak resistance may
be accompanied by an upward sliding block mech-
where f  K tan  and is known as the uplift factor.
anism (Fig. 9) but as a gap is formed below the pipe the
Uplift factors indicated by the Schamine tests are
boundary conditions of the problem change and the
shown in Table 4 for the limited range of H/D tested
mechanism changes to a circulation or flow around
( 4 for sands).
mechanism. Pedersen & Michelsen (1988) sketched
A variation of the frictional model (Eqn (11)) is
the same concept but without elaboration and
attributed to Pedersen in which the whole volume of
Schamine et al. (1990) identified it from their testing
soil above the pipe is involved:
programme.
In very loose sands which contract when sheared,
Vanden Berghe et al. (2005) have demonstrated this (12)
circulation mechanism numerically (Fig. 10).
Compression of the very loose sand and upward
movement makes room for the flow mechanism to where fp  K tan  (using the suffix p to differentiate
develop. The dilatancy of the soil determines at which with the f of Schamine) and 0.1 (D/H) represents the
H/D ratio the flow mechanism governs the peak weight of the soil wedges between the pipe centreline
resistance as well as the residual. and the crown.
The different mechanisms of failure go some way The equivalent vertical slip model for cohesive
to explain why the wedge failure model cannot capture (undrained) behaviour leads to:
the whole range of soil density and H/D ratios very
effectively. It may also explain some of the scatter in (13)

106

Copyright 2005 Taylor & Francis Group plc, London, UK


Palmer et al. (1990) recommend more conserva- uplift loading is often cyclic since it is associated
tive factors (0.5 for dense material and 0.1 for loose) with production cycles; upward movement may
and put forward a variation of this equation to address enable a gap to develop and particles to flow under
the problem of deep or flow failure, expressed as: the pipe, leading to upward ratcheting or creep;
very soft backfills are heterogeneous and may
behave like a frictional material.
(14) These topics are discussed below.
Discontinuous contact of the pipeline with the
For the cohesive model, the more rigorous solu- trench is normal due to irregularities in the trench
tions of Merifield et al. (2001) are preferred. Uplift level. This results in gaps (spans) under the pipe
bearing capacity factors (P/suD) have been computed which may not be fully filled with backfill soil.
using both upper and lower bound solutions account- Fortunately, this is typically associated with sag bends
ing for shallow and deep failure mechanisms. Thorne which, when loaded axially, will tend to move down
et al. (2004) has investigated in detail the issue of suc- into the soil. However, gaps could exist adjacent to a
tion behind the pipe. supported area. A gap below the pipe increases the
Bolton & Barefoot (1997) and White et al. (2001) tendency for a circulation or flow failure and there-
have justified the wedge mechanism shown in Figure fore the uplift resistance could be lower than antici-
9 by demonstrating that the angle  corresponds to pated for a uniformly bearing pipeline. Design
the dilation angle of the soil and accounting for an conservatism must be introduced for incomplete con-
increase in the vertical stress (and shear resistance) in tact with the seabed.
the vicinity of the pipe. The shear stress  along the Thermal loading is associated with production and
slip surface is expressed as: when production is halted the pipeline cools down.
Thus pipelines experience many cycles of uplift load.
If the backfill resistance is low and the pipe is able to
move sufficiently to create a gap under the pipeline,
(15)
soil particles can flow into this gap during a hot
and thus: phase. On cooling the pipe cannot return to its ori-
ginal position. Cyclic loading can thus lead to the pipe
creeping or ratcheting upwards. With loss of cover and
resistance a strain-softening resistance is experienced
and full uplift failure can occur. Nielsen et al. (1990)
postulate that this was the mechanism that occurred in
their project. Finch et al. (2000) suggest, based on 1 g
laboratory testing that ratcheting only occurs in clean
sands where the grains are not held by adhesion. In
order to limit progressive uplift the magnitude of
uplift displacements must be controlled.
Very soft clay backfills are heterogeneous and are
(16) believed to drain more rapidly than the in situ soil.
Bolton & Barefoot (1997) have shown that for the
and K0 can be taken as 1 sin crit Atlantic mud tested, both drainage is relatively rapid
The dilation angle can be assessed from the rela- and the clay dilates when sheared at very low stress
tive density of the sand, the stress level and the par- levels. This results in higher than anticipated uplift
ticle characteristics (Bolton 1986). For a range of sands resistance. They argue that adopting a frictional
between loose and dense White et al. show that the model with an uplift factor f of 0.4 can be justified for
correlation is good. In the experience of the authors, the mud tested. The authors have been involved in
this approach may overestimate uplift resistance for projects involving jet trenching of very soft clays. An
very loose sands where a negative dilation angle uplift factor of 0.3 was used based on centrifuge tests
would be applicable and where a circulation flow on the soil and no uplift problems have been reported.
mechanism occurs.
Recommended uplift factors are also given by
7.2.2 Uplift response
Finch et al. (2000) for different soil conditions.
Numerical methods used to design against upheaval
In practice, some other issues must be addressed
buckling require not only the ultimate capacity but
when considering uplift resistance:
also the displacement response to mobilize this cap-
pipelines are not always in continuous contact with acity. Results can be sensitive to this stiffness. Much
the seabed; less interest has been shown on uplift displacements

107

Copyright 2005 Taylor & Francis Group plc, London, UK


to reach failure but Finch et al. (2000) propose guide- 1.2
lines based on their experimental program. Trautmann
et al. (1985) suggest that the displacement at peak 1
resistance is between 0.51.5% of the pipeline depth.
For a 16 (0.4 m OD) pipeline buried at a depth of 1 m 0.8
(H/D  2.5) the displacement at peak would be

Py / Pyu
between 26 mm which is in agreement with Finch 0.6
et al. at that depth.
0.4 Audibert & Nyman (1977)
7.2.3 Lateral resistance Trautman & ORourke
0.2
The lateral resistance of buried pipelines is not gener- (1985)
ally important for buckling but becomes important Simplified bi-linear
0
if ground movements occur, such as by faults or 0 0.2 0.4 0.6 0.8 1 1.2
mudslides. y / yu
Pipelines buried in sand have been studied by
Audibert & Nyman (1977), Nyman (1984) and Figure 11. Lateral force-displacement curves for pipelines
Trautman & ORourke (1985). The ultimate lateral embedded in sand.
resistance can be written:
The limiting value of 10.47 reflects the transition
(17) from shallow to deep behaviour.

where the dimensionless lateral bearing capacity fac- 7.2.4 Lateral response
tor Ny depends on the relative density of the sand and Moving on to the lateral force-displacement models
on the embedment of the pipeline. Trautman & in sands, as described by Trautman & ORourke, a
ORourke (1985) showed that for loose and medium hyperbolic relationship is proposed given by
dense sands, Ny increases approximately linearly with
the embedment for H/D 8, whereupon Ny becomes (21)
constant, indicative of the transition from shallow to
deep soil failure mechanism. For dense to very dense
sands, the transition was not reached at H/D of 11. where P*y  Py/(Ny  H D) is the normalised force and
Trautmann & ORourke also demonstrated that the y*  y/yu is the normalised displacement; a and b are
values of Ny defined for the holding capacity of the model parameters. Proposed values for a and b by
anchor plates (Rowe & Davis 1982) were in good Trautman & ORourke are 0.17 and 0.83, respectively.
agreement with their own data for pipes. Rowe and Slightly different values are proposed by Audibert &
Davis showed that Ny depends primarily on the fric- Nyman (1977). The displacement at the peak resist-
tion angle and embedment ratio, and on the roughness ance yu depends on the embedment ratio (H/D) and
of the embedded structure. decreases with increasing relative density of the sand.
For homogeneous cohesive soils, the ultimate lat- The hyperbolic force displacement curve can be
eral resistance of buried pipelines can be based on the simplified into a bilinear representation. Trautman &
work of Merifield et al. 2001 for plate anchors : ORourke (1985) suggest an initial stiffness equal to
the secant stiffness at 70% of the ultimate resistance.
(18) In that case, the maximum force is reached at a dis-
placement of 0.4yu. Normalised force-displacement
where the dimensionless factor Nyu depends on the curves are plotted on Figure 11. Note that the soils
embedment of the pipeline and to a lesser extent on its modelled in this study have effective friction angles
surface roughness. between 2030 and thus would be in the relative
Considering conservatively the results of the lower density range 020%.
bound plasticity analysis quoted by Merifield et al.
(2001), the dimensionless factor Nyu can be written: 7.2.5 Resistance to inclined transverse loads
In a study related to pipelines buried in very loose
sand, Vanden Berghe et al. (2005) have shown that there
(19) is very little difference in uplift resistance when the load
direction is within about 30 of the vertical. Figure 12
shows the displacement patterns and Figure 13 depicts
the uplift factor Nz as a function of direction.
(20) The implication of this finding for upheaval
buckling is that modes of deformation in an inclined

108

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 5. Typical thermal conductivities.

Material Thermal conductivity (W/mK)

Quartz 4.09.1
Water 0.600.67
Clay (typical) 1.52.9

a) vertical b) 22.5 to the vertical will depend on the thermal conductivity of the pipeline
and its surrounds. If transient solutions are required
(for example for the heating up or shutting down of the
system) the specific heat capacity will also be required.
Hence there is a need to know the thermal properties
of the soil and make use of the low thermal conduct-
ivity where possible to provide thermal insulation.

8.1 Factors affecting thermal properties


c) 45 to the vertical d) horizontal Heat transfer can take place by conduction, convec-
tion and radiation. In saturated soils, however, heat
Figure 12. Failure mechanisms for different displacement transfer is mainly due to conduction through the solid
directions (Vanden Berghe et al. 2005). framework and the pore water (Farouki 1986). Con-
vection may be more important in coarse grained soils
8 and rockfill.
H/D=2 Heat conduction in soil can be described, for the
Phi=20- Psi= - 10
one dimensional case, by the Fourier equation:
6 Phi=25 - Psi= - 5

Phi=30 - Psi = 0
Nv [-]

4 (22)

2 where T is the temperature at time t and depth z; k is


the thermal conductivity (amount of heat that flows
through unit cross-sectional area under a unit temper-
0 ature gradient (T/z) in unit time); C is the specific
0 22.5 45 67.5 90 heat of soil; and is the density. Since the specific
Displacement Direction [] heat of particles of sand and clay is only about 20%
that of water (Geological Society of America 1942), it
Figure 13. Comparison between vertical, oblique and lat- is largely on account of the water content that soil is
eral resistance in contractive soils (Vanden Berghe et al. capable of storing heat.
2005).
The main factors that affect the thermal conduct-
ivity of saturated soil are: mineral composition, particle
direction are very likely to occur since imperfections size distribution, density/water content, and tempera-
are rarely 2-dimensional. This agrees with experience ture. In general, solids conduct heat better than liquids,
of surveyed buckles. and liquids better than gases (Brandon & Mitchell
1989).
8 THERMAL PROPERTIES OF SOILS Table 5 indicates some typical values for soil com-
ponents:
Efficient transportation of crude oil in a pipeline Mineral composition. Heat conduction through the
requires a sufficiently high temperature to maintain particles is an important mechanism of heat transfer.
low product viscosity and avoid unwanted deposition All other factors being equal, sands containing a high
of wax. Since the temperature of the soil and water percentage of quartz will have a higher thermal con-
around the pipeline is lower than the temperature of ductivity than those containing a high percentage of
the oil, heat flows from the oil to the environment. mica (Brandon & Mitchell 1989). Clays with a
Heat lost along the length of the line results in a tem- relatively high content of kaolinite have relatively low
perature drop from inlet to outlet. Pipeline design conductivities. Moreover, sands containing high per-
must ensure that this temperature drop is within centages of silica may exhibit increased thermal con-
acceptable limits. The temperature loss along the line ductivity with time, possibly due to the formation of

109

Copyright 2005 Taylor & Francis Group plc, London, UK


silica precipitants at the contact between grains Table 6. Empirical equations for determining soil conduct-
(Brandon & Mitchell 1989). ivity.
Particle size distribution. Well-graded soils con-
duct heat better than poorly graded soils because the Reference Equations (k in W/mK)
smaller grains can fit in the interstitial voids between
Kersten k  0.144  (0.9 log w0.2)  100.06364d
the larger grains, thus increasing the density and the (1949) d  dry density (in kg/m3)
mineral-to-mineral contact area (Brandon & Mitchell for silts and clayey soils
1989). Thermal conductivity in general varies with
k  0.144  (0.7 log w  0.4)  100.06364d
the grain size of the soil. At a given density and mois- for sands
ture content, the conductivity is relatively high in
Johansen k  ksat  ks(1n)kwn
coarse grained soils such as gravel or sand, somewhat
(1975) ks  kqqk01q
lower in sandy loam soils, and lowest in fine grained kw  0.6 W/mK; kq  7.7 W/mK;
soils such as silty loam or clay (Kersten 1949). k0  2.0 W/mK
Density and water content. Due to the relatively q  quartz content
high conductivity of minerals compared to water, the Simplified equation for saturated soils.
conductivity increases with density (Kersten 1949, Makowski k  (a log(w)  b)  10c
Brandon & Mitchell 1989). & Mochlinski a  0.1424 0.000465 p
Temperature. Thermal conductivity may also be (1956) b  0.0419 0.000313 p
influenced by temperature because each of the con- c  0.00062 d
stituents has temperature dependent thermal conductiv- p  weight percentage of soil finer
ities (Brandon & Mitchell 1989). All crystalline than 2 m
minerals in soils show a decrease in thermal conductiv-
ity with increasing temperature, except feldspar (clay).
The thermal conductivity of water increases with tem- used to show variation that may be anticipated
perature because of the higher level of molecular move- (Fig. 14).
ment at higher temperatures (heat transfer by collision). A detailed review is given by Farouki (1986) and
Rawat et al. 1979, Young et al. (2001) have also con-
8.2 Measuring thermal conductivity tributed. Rawat et al. suggested that maximum error
with the Kersten method was 25%. Farouki recom-
Since thermal conductivity is dependent on the spe-
mends the Johansen method for coarse to fine sand
cific mineral constituents of the soil as well as its dens-
(5% passing 2 microns). This method was found to
ity/water content it is preferable to measure the
provide the best correlation because it takes into
conductivity on soil samples in the laboratory or
account the mineralogy of the sand (which should be
in-situ. A common method is the thermal probe.
determined by x-ray diffraction). Based on laboratory
The thermal probe is a long needle that is inserted in
data, Rawat considered that the Makowski and
the soil containing both heating and temperature meas-
Mochlinskis method overestimates the conductivity
uring elements. A known amount of current is passed
unless the combined silt and clay fraction was used in
through the heater element and the resulting variation
the equations.
of temperature is measured as a function of time. The
Laboratory tests from high water content deepwa-
thermal conductivity of the soil can be deduced from
ter Gulf of Mexico clays, Indonesia and Nigeria
these measurements. The applicable procedure is
(remoulded and undisturbed) were reported by Young
described by ASTM D5334 (2000). The thermal needle
et al. (2001). Remoulded data is shown in Figure 15.
probe has been presented by various authors (Hooper &
This is of interest because it covers water contents
Lepper 1950, DeVries 1952, Woodside 1958, Falvey
that would be typical for trench backfills.
1968, Mitchell & Kao 1978, among others).
The undisturbed soil samples had thermal conduct-
ivity values ranging from 0.65 to 1.25 W/mK while
8.3 Empirical methods for determining
remoulded values were in the range 0.8 1.05 W/mK.
conductivity
In the absence of specific laboratory data, various
8.4 Selection of thermal conductivity
empirical equations are available relating the thermal
conductivity of the soil to its water content, dry density As with all design parameter selection, the use of the
and type of soil. Some empirical approaches are given parameter should be considered. Finch et al. (2000)
in Table 6. suggest the following guidelines: upper bound values
For saturated soils, the relationships presented in should be adopted for thermal insulation design, as
Table 6 can be conveniently expressed in terms of the high thermal conductivity represents high heat loss.
water content rather than the dry density or porosity. Conversely, lower bound values are applicable to
The three approaches presented in Table 6 have been upheaval buckling assessments where heat retained in

110

Copyright 2005 Taylor & Francis Group plc, London, UK


5 the pipeline will tend to increase the uplift forces
Kersten experienced by a buried pipeline.
Thermal conductivity [W/mK]

4 The thermal properties of jetted backfills may be an


Sand important design component for a deep water system.
As discussed in connection with mechanical proper-
3
ties, the changes in water content during jet trenching
are not well known. This introduces significant uncer-
2 tainty when considering the thermal properties of the
Clay
backfill. Deep water sites are commonly associated
1 with soft clays and are ideally suited to the use of jet
trenching. Even though the properties of the backfill
0 soil are significantly different from the virgin soil, the
0 20 40 60 80 backfill soil exhibits low values of thermal conductiv-
ities and sufficient strength and density to inhibit ther-
5
Johansen
mal convection currents (Young et al. 2001). In fact,
100
Thermal conductivity k [W/mK]

Quartz content, % the very soft highly plastic clays encountered at most
4 deepwater locations have three characteristics that
make them a favourable medium for flowline insula-
3 tion. First, the clays exhibit cohesion and low permea-
50
bility making them strongly resistant to thermal
2 convection (water travelling freely through the soil to
0 and from the heat source). Second, saturated clays
1 with high water contents exhibit low values of thermal
conductivity. Third, the soils can be easily jetted to
0 produce a trench with steep but stable trench walls.
0 20 40 60 80

5 9 CONCLUSIONS
Thermal conductivity [W/mK]

Clay content, % Makowski


4
0 Pipeline geotechnics deals with soil-pipeline inter-
3
action. This covers installation issues (pipeline pene-
50
tration and short-term lateral stability), axial and
2 100 lateral response to loads. It then encompasses pipeline
trenching, backfill engineering and pipeline stability
1 when buried. Particularly current topics such as riser-
soil interaction, upheaval buckling and lateral buck-
0 ling/ snaking, all of which include many load cycles,
0 20 40 60 80
Water content [%]
are all subject to ongoing investigations and joint
industry projects.
Figure 14. Thermal conductivities by various methods for
While performing this review, the authors have
saturated soils. been made aware again of the very wide range of
issues that must be faced in connection with pipeline
design. We have found no similar review papers cov-
ering the subject. Therefore, this paper claims unique-
ness and we trust it will provide others with a starting
point in many of the specific subject areas. It was
written in the midst of a busy consulting schedule and
therefore does not treat all the subject matter as fully
as we would have liked. There are likely to be some
errors that have crept in. Nevertheless, we trust that
future reviews will find this a useful starting point.

REFERENCES

ASTM D5334, 2000. Standard test method for determina-


tion of thermal conductivity of soil and soft rock by ther-
Figure 15. Remoulded conductivities (Young et al 2001). mal needle probe procedure.

111

Copyright 2005 Taylor & Francis Group plc, London, UK


Audibert, J.M.E. & Nyman, K.J. 1977. Soil restraint against Farouki, O.T. 1986. Thermal Properties of Soils. Series
horizontal motions of pipes. ASCE Journal of Geotech- on Rock and Soil Mechanic, Vol. 11, Trans tech
nical Engineering 103(GT10): 11191142. Publications: 136.
Bezuijen, A. & Mastbergen, D.R. 1988. On the construction Finch, M 1999. Upheaval buckling and floatation of rigid
of sand fill dams Part 2: soil mechanics aspects. Model- pipelines: The influence of recent geotechnical research
ling Soil-Water-Structure Interactions. Kolkman et al. on the current state of the art. Offshore Technology
(Eds), Balkema. Conference, OTC10713.
Bolton, M.D. 1986. The strength and dilatancy of sands. Finch, M., Fisher, R., Palmer, A. & Baumgard, A. 2000. An
Geotechnique 36(1): 6578. integrated approach to pipeline burial in the 21st century.
Bolton, M.D. & Barefoot, A.J. 1997. The variation of critical Proc. Deep Offshore Technology 2000.
pipeline trench back-fill properties. Proceedings of IBC Fisher, R., Powell, T., & Palmer, A.C. 2002. Full scale mod-
Conference on Risk-Based and Limit State Design and eling of subsea pipeline uplift. Physical modeling in
Operation of Pipelines. Aberdeen, May. geotechnics, Newfoundland.
Brandon, T.L. & Mitchell, J.K. 1989. Factors influencing Fontaine, E., Nauroy, J.F., Foray, P., Roux, A. & Gueveneux, H.
thermal resistivity of sands. Journal of Geotechnical 2004. Pipe-soil interaction in soft kaolinite: vertical stiff-
Engineering, Vol.115, No.12: 16831698. ness and damping. Int. Offshore and Polar Eng. Conf.,
Brennodden, H., Lieng, J.T., Sotberg, T. & Verley, R.L.P. Toulon, France :517524.
1989. An energy-based pipe-soil interaction model. Proc. Fredsoe, J 1978. Sedimentation of river navigation channels.
Offshore Technology Conf., Houston, 14 May 1989 OTC J. Hydr. Div. ASCE, Vol.104: 223236.
6057: 147158. Geological Society of America 1942. Handbook of physical
Bridge, C., Laver, K., Clukey, E. & Evans, T. 2004. Steel constants.
catenary riser touchdown point vertical interaction models. Hartlen, J and Ingers, C. 1981. Land Reclamation using Fine
Proc.Offshore Technology Conf. Houston, OTC16628. Grained Dredged Material. Proc. of the Tenth International
Bruton, D.A.S., Bolton, M.D. & Nicolson, C.T. 1998. Poseidon Conference on Soil Mechanics and Foundation Engineer-
Project Pipeline design for weak clays. Offshore Pipeline ing, Stockholm, Vol. 1: 145148.
Technology: 21. Hooper, F. C., & Lepper, F. R. 1950. Transient heat flow
Burland, J.B. 1990. On the compressibility and shear strength apparatus for the determination of heating piping and air
of natural clays. Geotechnique 40, No. 3: 329378. conditioning. Trans., American Society of Heating and
Carr, M. D., Bruton, D.A.S. & Leslie, D. 2003. Lateral buck- Ventilation Engineers 56: 309324.
ling and pipeline walking, a challenge for hot pipelines. Hight, D.W., Georgiannou, V.N., Martin, P.L. &
Offshore Pipeline Technology Conf., Amsterdam. Mundegar, A.K. 1999. Flow slides in micaceous sand.
Cathie, D., Machin J. & Overy R.F. 1996. Engineering Proc. Int. Symp. on Problematic Soils, IS-Tohoku 98,
appraisal of pipeline floatation during backfilling. Offshore Sendai, Japan, Balkema, Vol. 2: 945957.
Technology Conference, OTC8136, Houston: 197205. Johansen, O. 1975. Thermal Conductivity of Soils. PhD
Cathie, D., Barras S. & Machin, J. 1998. Backfilling Pipelines: Thesis, Trondheim, Norway. (CRREL Draft Translation
State of the Art. 21st Offshore Pipeline Technology Con- 637, 1977). ADA 044002.
ference, Oslo, February. Karthikeyan, M., Dasari, G.R. & Tan, T.S. 2001. Character-
Cathie, D. 2001. Towards Excellence in Pipeline Trenching ization of a reclaimed land site in Singapore, Soft Soil
Engineering. Offshore Pipeline Technology Conference. Engineering, Eds Lee et al.: 587592.
Cathie, D. & Wintgens, J-F. 2001. Pipeline Trenching Using Karthikeyan, M., Dasari, G.R. & Tan, T.S. 2002. In situ char-
Plows: Performance and Geotechnical Hazards. Offshore acterization of a land reclaimed using big clay lumps,
Technology Conf., OTC13145. Submitted to Canadian Geotechnical Journal.
Cassidy, M.J. 2004. Use of force-resultant models of shallow Kersten 1949. Thermal properties of soils. Engineering
foundations in offshore applications. Computational Experiment Station Bulletin 28, University of Minnesota,
Mechanics, Sept, Beijing, Springer-Verlag. Minneapolis
Clukey, E.C., Jackson C.R., Vermersch J.A., Koch S.P. & Kolbuszewski, J.J. 1948. An experimental study of the max-
Lamb, W.C. 1989. Natural densification by wave action imum and minimum porosities of sands. Proc. 2nd Int.
of sand surrounding a buried offshore pipeline. 21st Conf. Soil Mech. and Fdn Engng, Vol. 1:158165
Annual Offshore Technology Conference, OTC6151: Kraft, L.M., Ray R.P. and Kagawa T. 1981. Theoretical TZ
291300. curves. J. Geotech.Eng.Div., ASCE, Vol.107, GT11:
Dickin, E.A. 1988. Stress-displacement of buried plates and 15431561.
pipes. Centrifuge 88, Cort (ed.). Rotterdam: Balkema. Kvalstad, T.J. 1999. Soil resistance against pipelines in
De Groot, M.B., Heezen, F.T., Mastbergen, H., & Stefess jetted trenches. Proceedings of the Twelfth European
1988. Slopes and densities of hydraulically placed sands. Conference on Soil Mechanics and Geotechnical
Hydraulic Fill Structures, ASCE: 3251. Engineering. Amsterdam, The Netherlands, Vol. 2:
DeVries, D.A. 1952. A nonstationary method for determining 891898
thermal conductivity of soil in situ. Soil Sci. Soc. Am. J. Lade, P.V. 1992. Static instability and liquefaction of loose
73(2): 8389. fine sandy slopes. J. Geotech. Eng., ASCE, Vol.118,
Dunlap, W.A., Bhojanala, R.P. & Morris, D.V. 1990. Burial No.1: 5171.
of vertically loaded offshore pipelines. Offshore Technol- Leroueil, S. & Hight, D.W. 2003. Behaviour and properties
ogy Conference, OTC6375: 263270. of natural soils and soft clays. Characterisation and
Falvey, D.M. 1968. Increase accuracy of soil measurements. Engineering Properties of Natural Soils, Eds. Tan et al.
Electrical World, November. Swets & Zeitlinger, Lisse: 29254.

112

Copyright 2005 Taylor & Francis Group plc, London, UK


Leroueil, S. 2003. Linking field and laboratory soil behav- sub-sea pipelines. Proc. 7th Int. Conf. on Offshore
iour. Int. Workshop on Geotechnics of Soft Soils-Theory Mechanics and Arctic Engineering, OMAE, Houston,
and Practice. Vermeer, Schweiger, Karstunen & Cudny: Vol V:243250.
7378. Niedoroda, A.W. and Palmer A.C. 1986. Subsea trench
Lieng, J.T. & Sotberg, T. & Brennodden, H. 1988. Energy- infill. Offshore Technology Conf. OTC5340: 445452.
based pipe-soil interaction models. SINTEF Report to the Nyman, K.J. 1984. Soil resistance against oblique motions
American Gas Association. of pipes. ASCE Journal of Transportation Engineering
Locat, J., Tanaka, H., Tan, T.S., Dasari, G.R. & Lee, H. 110(2): 190202.
(2003). Natural soils: geotechnical behavior and geo- Offshore Soil Investigation Forum (OSIF) 1999. Guidance
logical knowledge. Characterisation and Engineering notes on geotechnical investigation of marine pipelines
Properties of Natural Soils, Eds. Tan et al. Swets & Rev 03.
Zeitlinger, Lisse: 328. Palmer, A.C. & Richards, D.M. 1990. Design of submarine
Loch, K. 2000. Flowline burial: an economic alternative to pipelines against upheaval buckling. Proc. Offshore
pipe-in-pipe. Proc. Offshore Technology Conf., Houston, Technology Conf., Houston, 710 May 1990 OTC 6335:
14 May 2000 OTC 12034. 551560.
Lund, K.M. 2000. Effect of increase in pipe penetration Pedersen, P.T. & Michelsen, J. 1988. Large deflection
from installation, 19th International Conference On upheaval buckling of marine pipelines. Proc. Behaviour
Offshore Mechanics & Arctic Engineering, OMAE2000, of Off-Shore Structures (BOSS), Trondheim, Norway Vol.
New Orleans. III: 965980.
Makowski, M. W. & Mochlinski, K. 1956. An Evaluation of Powell, T.R., Fisher, R., Phillips, R. & Jee, T. 2002.
Two Rapid Methods of Assessing the Thermal Resistivity Reducing backfilling risks, Offshore Site Investigation
of Soil. Proc. Inst. Elect. Engineers Vol. 103, Part A, No. and Geotechnics Diversity and Sustainability, Society
11: 453469. for Underwater Technology.
Masterbergen, D.R., Winterwerp, J.C. & Bezuijen, A. 1988. Randolph, M.F. & Wroth, C.P. 1978. Analysis of deform-
On the construction of sand fill dams Part 1:Hydraulic ation of vertically loaded piles, J. Geotech. Eng. Div.
aspects, Modelling soil-water interaction, Eds Kolkman ASCE, Vol.104, GT12:14651487.
et al: 353362. Rawat, P.C., Agarwal, S.L., Malhotra, A.K., Gulhati, S.K. &
Matyas, E.L. & Davis, J.B. 1983. Prediction of vertical earth Rao, G.V. 1979. Determination of thermal conductivity
loads on rigid pipes. Journal of Geotechnical Engineer- of soils: a need for computing heat loss through buried
ing, Vol. 109, No. 2: 190201. submarine pipeline. Proc. Offshore Technology Conf.,
Mendoza, M.J. & Hartlen, J. 1985. Compressibility of Houston, 30 April3 May 1979 OTC 3670: 27472753.
clayey soil used in land reclamation, Proceedings of the Rowe, R.K. & Davis, E.H. 1982. The behaviour of anchor
Eleventh International Conference on Soil Mechanics plates in sand. Geotechnique 32(1): 2541.
and Foundation Engineering, San Francisco, Vol. 2: Schmertmann, J.H. 1991. The mechanical aging of soils,
583586. Journal of Geotechnical Engineering, ASCE, Vol.117
Merifield, R.S., Sloan, S.W. & Yu, H.S. 2001. Stability of (9): 12881330.
plate anchors in undrained clay. Geotechnique 51(2): Schaminee, P.E.L., Zorn, N.F. & Schotman, G.J.M. 1990.
141153. Soil response for pipeline upheaval buckling analyses:
Mitchell, J.K. & Kao, T.C. 1978. Measurement of soil ther- full-scale laboratory tests and modelling. Proc. Offshore
mal resistivity. ASCE Journal of Geotechnical Engineering Technology Conf., Houston, 710 May 1990 OTC 6486:
104(5): 13071320. 563572.
Morris, D.V., Webb, R.E. & Dunlap, W.A. 1988. Self-burial Schapp, D. 1982. Natural backfill of submarine pipeline
of laterally loaded offshore pipelines in weak sediments. trenches, Offshore Oil and Gas Pipeline Technology,
Proc. Offshore Technology Conf., Houston, 25 May Amsterdam.
1988 OTC 5855: 421428. Sills, G.C. 1995. Time dependent processes in soil consoli-
Murff, J.D., Wagner, MX.A. & Randolph, M.K. 1989. Pipe dation, Compression and Consolidation of Clayey Soils,
penetration in cohesive soil. Geotechnique 39(2): 213229. Eds: Yoshikuni & Kusakabe, Balkema: 875890.
Murray, E.J. & Geddes, J.D. 1987. Uplift of anchor plates in Silva, A.J. 1974. Marine Geomechanics: Overview and
sand. Journal of Geotechnical Engineering, Vol. 113, No. Projections, In Deep Sea Sediments, Physical and
3: 202215. Mechanical Properties: 4576.
Ng, C.W.W. & Springman, S.M. 1994. Uplift resistance of Skempton, A.W. & Northey, R.D. 1952. The sensitivity of
buried pipelines in granular materials. Proceedings of clays, Geotechnique 3: 3053.
Centrifuge 94, Vol. 1: 753758. Rotterdam: Balkema. Skempton, A.W. 1970. The consolidation of clays by gravi-
Newson, T.A. & Bransby, M.F. 2004. Determination of tational compaction. Quarterly Journal of Geol. Soc.
undrained shear strength parameters for buried pipeline London, Vol. 125: 373411
stability in deltaic soft clays. Int. Offshore and Polar Eng. Small, S.W., Tamburello, R.D. & Piaseckyj, P.J. 1971.
Conf., Toulon, France :3843. Submarine pipeline support by marine sediments.
Nielsen, N-J.R, Lyngberg, B. & Petersen P.T. 1990. Offshore Technology Conf., OTC1357: 309318.
Upheaval buckling failures of insulated buried pipelines: Stoutjesdijk, T.P., de Groot, M.B. & Lindenberg, J. 1998.
a case story, 22nd Annual Offshore Technology Conference, Flow side prediction method: influence of slope geom-
OTC6488: 581591. etry, Canadian Geotechnical Journal, Vol. 35: 4354.
Nielsen, N.J.R, Petersen, P.T., Grundy A.K. & Lyngberg, B. Tornes, K, Jury, J., Ose, B.A. & Thomson, P. 2000. Axial
1988. New design criteria for upheaval creep of buried creeping of high temperature flowlines caused by soil

113

Copyright 2005 Taylor & Francis Group plc, London, UK


ratcheting. Proc. 19th Int. Conf. Offshore Mechanics and White, D.J., Barefoot, A.J. & Bolton, M.D. 2001. Centrifuge
Arctic Engineering, OMAE2000, New Orleans. modelling of upheaval buckling in sand. IJPMG-
Thorne, C.P., Wang, C.X. & Carter, J.P. 2004. Uplift capacity International Journal of Physical Modelling in Geotech-
of rapidly loaded strip anchors in uniform strength clay. nics, 2: 1928.
Geotechnique, Vol.54, No.8: 507517. Whitman, R. V. 1970. Hydraulic Fills to Support Structural
Trautmann, C.H. & ORourke, T.D. 1985. Lateral force-dis- Loads. Journal of the Soil Mechanics and Foundations
placement response of buried pipe. ASCE Journal of Division, ASCE, Vol. 96, No. SM1: 2347.
Geotechnical Engineering 111(9): 10771092. Woodside, W. 1958. Probe for thermal conductivity meas-
Mendoza, M.J., and Hartlen, J. (1985). Compressibility of urement of dry and moist materials. Heat. Piping Air
clayey soil used in land reclamation, Proceedings of the Con, Sept: 163170.
Eleventh International Conference on Soil Mechanics Yang, L.A., Tan, T.S. & Leung, C.F. 2002. One-dimensional
and Foundation Engineering, San Francisco, Vol. 2: self-weight consolidation of a lumpy clay fill,
583586. Geotechnique, Vol 52, No 10: 713725.
Vanden Berghe, J-F, Cathie, D. & Ballard, J.-C. (2005) Youd, T.L. 1973. Factors controlling maximum and min-
Pipeline uplift mechanisms using finite element analysis, imum densities of sands, Am. Soc. for Testing and
Int. Conf. Soil Mech. Foundation Eng., Osaka. Materials Spec. Tech. Pub. 523: 98112.
Van Rijn, L.C. 1993. Principles of sediment transport in rivers, Young, A.G., Osborne, R.S. & Frazer, I. 2001. Utilizing
estuaries and coastal seas, Aqua Publications, Amsterdam. thermal properties of seabed soils at cost-effective insu-
Verley, R.L.P. & Lund, K.M. 1995. A soil resistance model lation for subsea flowlines. Proc. Offshore Technology
for pipelines placed on clay soils. Proc. Offshore Conf., Houston, 30 April3 May 2001 OTC 13137.
Mechanics and Arctic Engineering Conf, Copenhagen, Zhang, J., Randolph, M.F. & Steward, D.P. 1999. An elasto-
1822 June 1995, Vol V: 225232. plastic model for pipe-soil interaction of unburied
Verley, R.L.P. & Sotberg, T. 1992. A soil resistance model pipelines. Proc. Int. Offshore and Polar Engineering
for pipelines placed on sandy soils. Proc. Offshore Conf., Brest, 30 May4 June 1999: 185192.
Mechanics and Arctic Engineering Conf, Vol V-A pipeline Zhang, J., Stewart, D.P. & Randolph, M.F. 2002. Modeling
technology: 123131. of shallowly embedded offshore pipelines in calcareous
Wagner, D.A. Murff J.D. & Brennodden, H. 1987. Pipe-soil sand. J. Geotechnical and Geoenvironmental Engineering,
interaction model. Proc. Offshore Technology Conf., ASCE, Vol.128, No.5: 363371
Houston, 2730 April 1987 OTC 5504: 181190.

114

Copyright 2005 Taylor & Francis Group plc, London, UK


An operators perspective on offshore risk assessment and geotechnical
design in geohazard-prone areas

P. Jeanjean, E. Liedtke & E.C. Clukey


BP America Inc., Houston, Texas, USA

K. Hampson & T. Evans


BP Exploration Operating Co Ltd., Sunbury, UK

ABSTRACT: This paper gives a perspective on some of the challenges faced by oil and gas operators to site
and design their facilities in geohazard prone areas. A brief overview of BPs portfolio relationship to geohazards
is given and key lessons learned are shared. The proper characterization and evaluation of geohazards entails sig-
nificant time and effort. The technology needs must be properly anticipated. Methodologies to characterize site
conditions with high resolution geophysical methods are described and the experience with various tools used in
the determination of the sediment shear strength and pore pressure regime is shared. Key steps in engineering
analyses such as slope stability evaluation, debris flow run out prediction, and fault displacements calculations
are described. Annual probabilities of failure are also estimated by innovative methods. Geotechnical design in
geohazard areas is particularly challenging because of the possible large uncertainty in soil properties. The
aleatory nature of the uncertainty is increased and may not be reduced by further site investigation. The operator
is also often faced with insufficient guidelines from design codes and may have to perform a significant amount
of design method calibration via model testing. Operating in geohazard-prone areas therefore entails making wise
decisions while properly managing increased levels of uncertainty.

1 GEOHAZARDS AND BPS PORTFOLIO repetitive sequences of sediment deposition, channel


erosion and gravitational or earthquake-triggered large-
Numerous offshore oil and gas prospects in BPs world- scale submarine sliding for over 250,000 yrs (Loncke
wide portfolio are located in geohazard prone areas, 2002). These and other shallow and deeper geological
particularly in the Caspian Sea, the Nile Delta, and the processes have resulted in a complex geology and
Gulf of Mexico. numerous present-day geohazards.
BP is currently developing oil and gas prospects in In the Gulf of Mexico, the Mad Dog and Atlantis
geohazardous areas of the Central and South Caspian fields are located along the Sigsbee Escarpment, one
basins. These prospects are named the Azeri, Chirag, of the Gulfs most prominent and complex geomor-
Gunsahli (ACG) fields and the Shah Deniz field. phological settings. Both fields straddle three distinct
Within these areas of the Caspian Sea, there exists a geological provinces: the southern edge of the northern
unique and unprecedented combination of mud volca- Gulf of Mexico continental slope, the Sigsbee Escarp-
noes, shallow faulting, soil slumping, soil instability, ment, and the upper continental rise. The export line
weak and irregularly consolidated soils, and gas seeps, from the Atlantis field is part of the Mardi Gras export
located in an area which is also moderately seismically system and crosses the Sigsbee Escarpment (Fig. 3).
active. In addition, shallow gas, over pressured water The sites are located in water depths between
bearing sands, and mud volcano breccia at depth are 1,280 m and 1,400 m above the escarpment and between
recognized as potentially hazardous to drilling activ- 2,100 m and 2,200 m below the escarpment.
ities. A seabed rendering of the Shah Deniz field is The Mad Dog field will be developed with a spar
presented on Figure 1. platform located 365 m away from the top of the escarp-
BP is also considering subsea developments in ment. The closest slope to that spar is about 200 m high
geohazardous areas offshore Egypt (Fig. 2). These and has an average angle of 19. The Atlantis field will
prospects are located in water depths of 300 to 1500 m be developed by a subsea well cluster located 1.8 km
in the West Nile Delta in areas that have experienced downslope of a 235 m high slope with an average slope

115

Copyright 2005 Taylor & Francis Group plc, London, UK


Figure 1. Seabed rendering of the Shah Deniz field, Caspian Sea (from Mildenhall and Fowler, 2001).

angle of about 21. Locally the slope angles on the


2 PURPOSE OF GEOHAZARD RISK
escarpment, both at Mad Dog and Atlantis, can exceed
ASSESSMENT
35 (Jeanjean et al. 2003).
Both fields have clearly been witness to numerous
2.1 Background
past failures of steep slopes which triggered major
debris flows. The Mad Dog spar southern mooring clus- Numerous questions immediately arise when the devel-
ter is located on the escarpment itself and in the path of opment of such geohazard prone fields is considered:
past debris flows (Fig. 3). Some of these flows ran out Why did a particular slope area fail in the past? How
distances over 7 km, at speeds back calculated to have often do slope failures occur and what is the current
been up to 100 km/hr (Niedeoroda et al. 2003). level of stability of the escarpment? What are the poten-
The seabed at both Mad Dog and Atlantis also shows tial trigger mechanisms for future failures? What would
signs of complex faulting with significant fault offsets be the consequences of slope failure? How likely is a
at the modern day seafloor. Some of the field architec- slope failure to occur over the life of the field?
ture schemes considered included crossing such faults These questions have three basic objectives: infer
by flowlines and umbilicals. what happened in the recent geological past, under-
Given the presence of these hazards, consideration stand the current situation, and predict the evolution of
must be given to the geohazard risks, however unlikely the hazard in the near future. A global understanding
in occurrence and low in consequences. of the geohazards risk can only be achieved through

116

Copyright 2005 Taylor & Francis Group plc, London, UK


Figure 2. Seabed rendering, Nile Delta.

Figure 3. Seabed rendering and field architectures of the Mad Dog field, Atlantis field, and Mardi Gras Export transporta-
tion system in the Southern Green Canyon area of the Gulf of Mexico. Red, yellow, magenta, and green symbols indicate loca-
tions of cores. Orange crosses indicate locations of deep soil borings. Numbers indicate OCS blocks in the Green Canyon
area. Scale is given by the OCS blocks which are 4.8 km by 4.8 km (3 miles by 3 miles).

117

Copyright 2005 Taylor & Francis Group plc, London, UK


the work of a multidisciplinary team where all skills seismic data has been acquired. An edge map of the
contribute to the final assessment. Geologists, geo- seabed pick from the exploration 3D volume can allow
physicists, geotechnical engineers, oceanographers, geomorphological mapping to proceed immediately. In
risk and reliability experts, finite element method addition, the shallow section of the exploration volume
(FEM) experts, all have a role to play. can be used to produce a preliminary assessment of
Although risks are an inherent part of developing hazards such as shallow water flow zones, faults, fluid
offshore fields, they are compounded in the deepwater expulsion areas, and to develop a preliminary shallow
environment because of higher capital investments, geological model of the field.
intensive use of new technology, new hardware and Once the project is mature enough so that field
new materials, new metocean settings and, of particu- development concepts can be selected, planning for
lar interest to this paper, potentially highly geohazard High Resolution 3D seismic surveys and AUV sur-
prone geological settings. veys should proceed.
The risks posed by geohazards are therefore only Ideally the geotechnical data acquisition should
one of the many risks faced by a project. All of the start only once the above surveys have been inter-
project risks must be compared and ranked, so that preted, a geomorphological characterization of the
the project activities and efforts can be prioritized. seabed and shallow soils has been performed, and a
The need to compare and rank all risks leads to a con- geological model is available. Particularly in areas of
sistent way to characterize them, usually with their highly variable terrain, the location of cores and bor-
respective annual probability of occurrence, and a ings should be carefully determined to ground truth the
measure of their consequences. models. The first step typically consists of collecting
The fundamental role of a geohazard risk assess- and analyzing long free-fall piston cores. Deeper
ment is to provide the above risk characterization, in geotechnical borings should then be targeted to fur-
as much of a quantitative manner as possible. For ther enhance the geotechnical site characterization.
slope stability issues, the annual probability of occur- Because geohazard risk evaluation is an integral
rence of a failure needs to be estimated, and its conse- part of the project sanction the challenge is to achieve
quence in terms of debris flows and turbidity current a balance between having enough definition of the
characterization and damage to infrastructure. For project, field architecture, and export routes to start
fault hazard, annual displacement rates along fault and meaningfully focus the work (i.e. not starting too
planes must be understood and hence the damage early with a poorly defined scope) and allow comple-
caused by such displacements. tion of the work prior to sanction (i.e. not finishing
the risk evaluation too late).
2.2 Purpose
The purpose of the risk assessment is therefore to
3.2 Geotechnical site investigation with
give a clear assessment of the level of risk associated
exploration drilling vessels
with geohazards, for a given field architecture.
This paper will describe some of the challenges that The first drilling vessels to be on location at a given
are faced by BP and its co-venturers in making such site are those used for the discovery and appraisals
decisions and the perspective gained in the process. wells. They are large drilling vessels and are designed
Details of the key studies for the Mad Dog and to be able to drill to substantial depths in order to delin-
Atlantis fields are detailed in the following papers, eate reservoirs.
which the reader is encouraged to read in the following However, how reliable and valuable is the data
order: Jeanjean et al. (2003), Orange et al. (2003), acquired from these exploration drilling vessels? To
Al-Khafaji et al. (2003), Slowey et al. (2003), Nowacki help answer this, data obtained with exploration drilling
et al. (2003), Lanier et al. (2003), Niedoroda et al. vessels in non-geohazard prone areas (i.e. areas
(2003), Young et al. (2003), Niedoroda et al. (2003b), where the surficial geology consists of very uniform
Brand et al. (2003), Orange et al. (2003b), Angell et al. hemipelagic sedimentation processes and where
(2003), Nadim et al. (2003), Jeanjean et al. (2003b). shallow mass wasting events are absent) has been
compared with data obtained at the same site with
geotechnically dedicated vessels. The data were
3 GEOHAZARD ASSESSMENT AND FAST
obtained from the exploration drilling vessel immedi-
PACE PROJECT SCHEDULE
ately prior to spudding the well and for the purpose of
designing and sizing jetted conductors. A few years
3.1 When to start and when to complete
later, when the field was deemed commercial, geo-
geohazard risk assessment studies
technical data was acquired using dedicated geotechni-
The geohazard assessment typically starts during the cal drilling vessels or by using free-fall Jumbo Piston
early stages of the projects once the exploration 3D Cores, for the purpose of designing the foundation

118

Copyright 2005 Taylor & Francis Group plc, London, UK


system of the subsea or floating production system. underestimated by exploration drilling vessel. Indeed,
Some of the geotechnical borings were up to 500 m on Figure 6b at a depth of 50 m (150 ft) the two
away from the location of the exploration drilling data remolded shear strength profiles match and yet the
but because of the uniform geology, comparison of undisturbed shear strength from the geotechnical ves-
the two data sets is still reasonable. sel is 50% higher than the one obtained with the
Figure 4 shows the best fit undisturbed and exploration drilling vessel. The lower soil strengths
remolded minivane profiles for the two sampling sys- obtained from samples recovered from the drill rig are
tems. It can be seen that the exploration drilling undis- attributed to greater sample disturbance.
turbed profiles are never higher than those obtained Such disparity between the data sets for the two
from geotechnical vessels. Figures 6c and f show rea- types of drilling vessels can only be expected to
sonable agreement between the data sets, Figures 6a, b, increase in geohazard-prone areas. As a result, it has
d and e clearly show that shear strength can be severely been concluded from the Gulf of Mexico experience

Best Fit Shear Undrained Shear Strength through Minivane and UU triaxial data (ksf) Best Fit Shear Undrained Shear Strength through Minivane and UU triaxial data (ksf)
0.0 0.5 1.0 1.5 2.0 2.5 0.0 0.5 1.0 1.5 2.0 2.5
0 0
Exploration Drilling vessel - Undisturbed Su
Exploration Drilling vessel - Undisturbed Su
Geotechnical Drilling vessel - Undisturbed Su
Geotechnical Drilling vessel - Undisturbed Su
Exploration Drilling vessel - Remolded Su
50 Exploration Drilling vessel - Remolded Su
Geotechnical Drilling vessel - Remolded Su 50 Geotechnical Drilling vessel - Remolded Su

100 100
Depth (ft)

Depth (ft)

150 150

200 200

(A) (B)

250 250

Best Fit Shear Undrained Shear Strength through Minivane and UU triaxial data (ksf) Undisturbed Shear Strength From Minivane (ksf)
0.0 0.5 1.0 1.5 2.0 2.5 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6
0 0

20 Exploration Drill. Vessel


Geotechnical Drilling Vessel
Exploration drilling vessel - Undisturbed Su 40
50
Exploration drilling vessel - Remolded Su
Geotechnical drilling vessel - Undisturbed Su 60
Geotechnical drilling vessel - Remolded Su
100 80
Depth (ft)

Depth (ft)

100

150 120

140

200 160

(C) 180
(D)
250 200

Undisturbed Undrained Shear Strength from Minivane (ksf)


Undisturbed Shear Strength from Minivane (ksf)
0.0 0.5 1.0 1.5 2.0
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0
0
0

Exploration Drill. Vessel 2


Exploration Drill. Vessel Exploration Drill. Vessel 1
50 Jumbo Piston Core 50
Jumbo Piston Core

100
Depth (ft)
Depth (ft)

100

150

150

200
(E)
(F)
200
250

Figure 4a-f. Comparison of shear strength profiles obtained from an exploration drilling vessel and either geotechnical
drilling vessels or Jumbo Piston Cores, at six Gulf of Mexico deepwater sites.

119

Copyright 2005 Taylor & Francis Group plc, London, UK


that the exploration-drilling vessel can provide a improved means to gather bathymetry, sonar and
lower bound data set suited for conservative design of profiler data.
jetted conductors but the use of the data for advanced The soils extreme lateral variability across the field
geotechnical design and quantitative risk assessment would necessitate in-depth geotechnical character-
would be inappropriate. ization with a system capable of efficiently sam-
However, recent projects in the Caspian have util- pling stiff debris flow material over large depths,
ized remote down-hole sampling and in-situ testing without the use of drilling vessels.
equipment which can be used from exploration drilling
After completion of this campaign, the quest for the
rigs. This equipment is compatible with the exploration
development of an Autonomous Underwater Vehicle
rig but the mode of drilling in terms of bit weight,
(AUV) started (Bingham et al. 2002). An AUV
mud pressure flow rates and fluid additives must be
became operational in early 2001, just in time for the
modified and managed in order to minimize potential
Mad Dog and Atlantis surveys (George et al. 2002)
levels of soil disturbance. This equipment has been
which gathered:
successful at obtaining high quality samples and in-
situ test data to depths of up to 450 m below mudline. Full field bathymetric model of the seabed on a
For project work in remote and developing areas of 3 m  3 m grid cell.
the world there are often issues with regard to the lim- Full field seabed sonar mosaic, at 200%  cover.
ited infrastructure available to carry out high quality Sub-bottom profiler grid on a 200 m  500 m
site investigations. For projects in the Caspian Basin, spacing.
which is a land locked sea, there is very limited infra- Arbitrary sub-bottom profiler traverses along sig-
structure available. It is very important for the assess- nificant geological features, to tie different sample
ment of geohazards that soil samples and in-situ data locations, to investigate potential debris flow run-
are of the highest quality and that quality is not com- outs, and through planned anchor locations.
promised by the use of inadequate infrastructure. In The use of the AUV allowed acquisition of these
order to achieve this goal for the Caspian sea projects data sets to be performed extremely quickly and accu-
and acquire the data fully addressing the HSE risks rately. The AUVs control systems were able to handle
associated with shallow gas, it has been necessary to the challenges of the rugged terrain, delivering a qual-
bring a dynamically positioned geotechnical vessel ity of data that would have been impossible from a
into the Caspian. This poses major issues with regard towed system. The co-location of the three acquisition
to logistics and the mobilization/demobilization costs systems within the same vehicle meant that there were
associated with such an operation are very high. no ambiguities in locating one type of data relative to
The use of drilling rigs also has potential for sam- another. Also of particular importance was that all the
pling sediments suspected to contain gas hydrates data acquired were available digitally.
with pressure samplers.
The overall experience with the use of exploration
4.2 Geotechnical site characterization
drilling vessels is that they are not appropriate to col-
lect data for advanced engineering purposes, although In 1997, the ability to continuously sample stiff surfi-
recent experience in the Caspian Sea shows promise. cial sediments was realized as being critical to under-
The operator must therefore usually rely on dedicated stand recent geological processes in geohazard-prone
geotechnical drilling vessels or long piston cores to areas. BP realized the need for a core sampler that
obtain high-quality geotechnical data. would be capable of collecting large diameter (100 mm)
long cores (i.e. 2530 m, or 75100 ft) in very soft to
stiff soils. BP therefore, through a Joint Industry
4 ROLE OF TECHNOLOGY IN SITE Funded Project (JIP), sponsored the development of
CHARACTERIZATION FOR GEOHAZARD- the free-fall Deep Sea Core Sampler (DSCS) capable
PRONE AREAS of retrieving over 24 m (80 ft) long cores in normally
consolidated soil locations without the need for
4.1 High-resolution geophysical surveys expensive and time consuming rotary drilling.
At Atlantis, the DSCS also allowed the recovery of
The results of a pilot deep tow and shallow coring
continuous cores in excess of 10 m (30 ft) in over con-
campaign in 1997 along the Sigsbee Escarpment,
solidated sediments with shear strength in excess of
resulted in a number of significant conclusions:
120 kPa (2.4 ksf) (Young et al. 2003). The capability to
The area was extremely complex and any field acquire long cores in very stiff debris flow deposits was
development would require in-depth geohazard key to understanding the timing of past debris flow
assessment. events and recent deposition rates.
Deep tow surveys were extremely inefficient in In addition, the industry now has the ability to drill
rugged and deep water terrain, requiring an deep geotechnical borings in 3,000 m (10,000 ft),

120

Copyright 2005 Taylor & Francis Group plc, London, UK


although at the time of this work, only one vessel in pressure transducers and a data logger system.
the Gulf of Mexico was able to do so. Although the tool could be deployed via a remote
Without the technology advances achieved between operated vehicle (ROV), it is usually dropped over-
1998 and 2001, mainly the coming of age of the AUV board from the vessel deck and free falls to the
technology but also improvement in geotechnical site seabed. The disposable lance embeds itself and the
characterization, it would have been very difficult to data logger records the pore pressure decay with time.
properly characterize the sites. The data logger is then retrieved via an acoustic
It is therefore paramount, and quite a challenge, for release system and the lance remains in the soil. The
operators to properly and timely identify the technol- advantage of the tool is that the pore pressure decay
ogy that will be needed for appropriate geohazard char- can be occurring while other operations are on-going.
acterization and ensure that the technology is delivered Some of the tool limitations include the limited
on time for project support. length of the lance (up to 6 m) and the fact that the
size of the area to be studied must be such that the
operator has confidence it has been hit by the free-
5 OFFSHORE GEOHAZARDS falling lance.
An example of pore pressure decay curve for a
The geohazards typically faced by offshore projects Gulf of Mexico deepwater site is shown on Figure 6.
include: The total time needed for full decay of the insertion
excess pressure is about 10 hours, at which time the
steep slopes, which can lead to shallow seated or
records indicate that the site is at a hydrostatic pore
deep seated failures which in turn trigger debris
pressure regime.
flows and turbidity currents;
faults;
mud volcanoes;
fluid expulsion features;
gas hydrates.
For lack of space, only a few of the above hazards will
be addressed herein and a brief perspective on how
they can be assessed will be given.

6 SHALLOW SEATED SLOPE FAILURES

Shallow seated failures are defined as failures for which


the failure surface is seated a few meters below the
seabed and for which the infinite slope stability frame-
work can be applied.

6.1 Pore pressure regime characterization


The pore pressure regime is always a paramount input
into any slope stability analysis. The difficulty in off-
shore sediments is that the insertion of any tool to
measure the in-situ pressure will create an excess pres-
sure which takes time to dissipate due to the low
in-situ permeability of the sediments. When dealing
with very shallow seated failures, those within a few
meters or tens of meters from the seabed, downhole
tools usually cannot be used because their length and
size would cause them to be unstable at such shallow
depths. A tool that has been successfully used to assess
the pore pressure regime in the top few meters of deep-
water slopes in the Pop Up Pore Pressure Instrument
(PUPPI) (Fig. 5). Details of the tools can be found at
http://www.geotek.co.uk.
The PUPPI consists of a disposable lance with two Figure 5. Principle of Pop Up Pore Pressure Instrument
pore pressure ports. These ports are linked to pore (PUPPI) (from http://www.geotek.co.uk).

121

Copyright 2005 Taylor & Francis Group plc, London, UK


25 4.0
Feature at 3.5m depth in
Excess Pore Pressure (kPa)

Depth in Jumbo Piston Core (m)


Lower Pore Pressure Port Zone of Core gravity core is only 2.5m
3.5
Lengthening deep in piston core
20 Upper Pore Pressure Port
3.0
Line of no disturbance
15 2.5

2.0
10
1.5

5 1.0
Zone of Core
0.5 Shortening
0 Core Shortening
0 2 4 6 8 10 0.0
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0
Time (hrs)
Depth in Large Diameter Gravity Core (m)

Figure 6. Pore pressure dissipation curves for the PUPPI


Figure 7. Example of quantification of core shortening in
two channels. Gulf of Mexico deepwater site.
jumbo piston core.

seabed on the gravity core is only 2.5 m below the


6.2 Shear strength characterization
seabed in the jumbo piston core, indicating a total core
6.2.1 Direct measurement of undrained shear shortening of 1 m in the top 2.5 m of the piston core.
strength Understanding such disturbance is critical to prop-
Although rotary drilling can be used to sample very erly use the shear strength data in slope stability ana-
surficial sediments, it has not proven to be an effective lyses but also to properly calculate sedimentation rates.
method because of time and cost required, lack of con-
tinuous sampling capability, and the fact that soft sedi- 6.2.2 Estimation of soil shear strength from
ments usually require the use of 50 mm (2 in.) liner exploration 3D seismic data
sampler that decrease sample quality. Although cores can provide invaluable data for slope
Free-fall long piston cores have proven to be a reli- stability analyses, the total number of locations that
able means of obtaining continuous core data for shal- can be sampled has to be kept within a practical limit.
low seated slope stability evaluations. Numerous piston However what is really needed is a means of estimat-
coring systems exist. Recent development in this tech- ing undrained shear strength at all locations to ensure
nology includes the Deep Sea Corer, which has been that the most critical sites have indeed been identified.
capable of obtaining 26.6 m (80 ft) cores in soft Gulf of An innovative method of estimating undrained shear
Mexico clays and a modified version of the Stationary strength from exploration 3D seismic data was devel-
Piston Corer (STACOR) (Borel et al. 2002). The qual- oped for the Mad Dog and Atlantis projects. Details of
ity of piston cores can vary significantly, largely the method are included in Brand et al. (2003).
depending on the geometry of the cutting shoe and the Piston cores are logged and their density and
type of piston (floating or truly stationary). Each sam- P-wave velocity are measured. Their acoustic imped-
pler should therefore be calibrated on a project- ance can therefore be obtained. Undrained shear
specific basis to understand how reliable the data are. strength is also measured, which allows a direct corre-
Procedures for assessing the quality of samples have lation between impedance and strength. Next the 3D
been developed (Lunne 1998). exploration seismic data is used to obtain the seafloor
In the upper parts of a piston core, disturbance can seismic amplitude at the location of the core.
include both core shortening and core lengthening. To As explained by Brand et al. (2003) the variation in
quantify such effects, one can compare multi-sensor amplitude of the seafloor peak is strongly tied to the
core logs (MSCL) from gravity cores and piston cores. change in acoustic impedance of shallow soils for each
Although short pistonless gravity cores can also expe- seismic trace. It is important to understand that the
rience core shortening if the corer becomes plugged, it seafloor peak amplitude is not a measure of the seismic
is generally accepted that the upper 2 m in a 100 mm reflection at exactly the water/soil interface. Rather, it
diameter simple gravity core in soft clays is a good is a measure of the seismic reflection controlled by the
representation of the in-situ sequence (Shultheiss 1999). impedance of the water column and average imped-
By comparing the depths below mudline of charac- ance of multiple sediment interfaces over a soil thick-
teristic features of the logs of soil density, P-wave ness controlled by the tuning frequency and resultant
velocity and magnetic susceptibility, the disturbance tuning thickness of the seismic data. The seismic
of the core can be assessed. Figure 7 summarizes the datasets used had a tuning thickness of about 10 m
results obtained at a Gulf of Mexico deepwater site. It (30 ft). Thus, the seafloor peak amplitude is depend-
can be seen that a feature located at 3.5 m below the ent upon the spatial variance of acoustic impedance

122

Copyright 2005 Taylor & Francis Group plc, London, UK


across multiple interfaces in the upper 10 m (30 ft) of 3500

7m (20ft) avg. impedance


soils across the entire study area.
The two correlations, shear strength vs acoustic 3000

(x100 g/cm2s
impedance and acoustic impedance vs seismic ampli-
tude, can then be combined to obtain a relationship
2500
between seismic amplitude and soil shear strength,
averaged over the tuning depth of the seismic data. The
relationship developed for the Atlantis prospect is 2000
shown on Figures 8 and 9.
1500
0 5000 10000 15000 20000 25000 30000
6.3 Deterministic analyses of shallow seated Amplitude
failures
6.3.1 Undrained analyses Figure 8. Correlation between seafloor peak seismic
According to infinite slope theory proposed by amplitude and average core impedance in top 7 m at Atlantis
Teunissen & Spierenburg (1997), the safety factor SF (modified from Brand et al. 2003).
of a slope against failure along a plane at depth z, can
be expressed by: 1400 67
1300
(1) 1200 57

7m (20ft) avg. shear strength (kPa)


7m (20ft) avg. shear strength (psf)

1100
where: 1000 48
sDSS
u is shear strength measured in direct simple 900
shear testing, 800 38
sTC
u is shear strength measured in triaxial compres-
700
sion testing,
28
  sTC DSS
u /su (A value of 1.3 was used at Mad Dog 600
500
and 1.25 at Atlantis.),
B is slope angle, and 400 19

vc
 is vertical effective stress at depth z. 300
With the method described in Section 6.2.2, the 200 9
operator has now a tool to be able to predict the average 100
undrained shear strength over the top 510 m continu- 0 0
ously over large areas. Slope angles can be calculated 1800 2000 2200 2400 2600 2800 3000 3200
from water bottom picks in the seismic volume. 2
7m (20ft) avg. impedance (x100 g/cm s)
Safety factors can then be calculated at all loca-
tions in a field of interest and areas of least safety fac- Figure 9. Correlation between average core impedance
tor can be identified. An example of such areas are and average shear strength in top 7 m at Atlantis (modified
shown on Figure 10. from Brand et al. 2003).

Data was collected around the Atlantis Slump E by


6.4 Probabilistic analyses of shallow seated
taking 14 cores in the areas just above, on, and just
failures estimating annual probability of
below the escarpment slope. Their locations were pur-
failure
posefully selected to capture all shallow drape failure
An example of statistical analysis of slope failure is activity in the area. For the statistical analysis, the cores
now given to demonstrate how annual probability of were assumed to form a system of check points such
failure can be estimated. Details of the method can be that no two check points captured the same shallow
found in Nadim et al. (2003). The general area near drape failure event, and all check points together cov-
Atlantis Slump E was studied in order to collect a his- ered the area of interest. Thus, the set of 14 cores were
torical record of shallow drape failures. This informa- analyzed as a system, with the aim of capturing failure
tion was then analyzed using statistical methods to frequencies within that system over the past 20,000
obtain nominal frequencies of shallow drape failures years.
throughout the recent geological history. Sediment Once all cores were dated, the identified slope fail-
deposition rates over the past 20,000 years were then ure events were grouped into 3,000-year periods to
obtained and compared with the shallow drape failure obtain relative frequencies. The results are shown on
frequency rates over the same period. Figure 11.

123

Copyright 2005 Taylor & Francis Group plc, London, UK


A smoothed line representing the sediment depos- 7 DEEP SEATED SLOPE FAILURES
ition rates (from Figure 25) has been superimposed
onto the same graph to show the remarkable agree- Deep seated failures are defined as failures for which
ment with shallow drape failure rates. The statistical the infinite slope stability framework can no longer
analyses showed that the current shallow drape failure be applied.
rate is 0.7 events/1000 years. This is equivalent to an
annual probability of failure of 710
4. 7.1 Pore pressure regime characterization
Nadim et al. (2003) showed that the probability of
shallow drape slope failure strictly within the Slump A great deal of effort was extended to characterize the
E area can be estimated to be around 210
4. They in-situ pore pressure regime of Mad Dog and Atlantis
counselled that the above numbers should be used slopes, as described by Orange et al. (2003b). Extensive
with caution, as they only provide rough estimates piezoprobe pore pressure dissipation tests were per-
due to the underlying assumptions and scarcity of formed at the Atlantis site and the slope was deter-
data. The most important perspective gained is the mined to be at a pore pressure equilibrium close to
strong correlation between the sediment deposition hydrostatic conditions. The discussion below focuses
rate and shallow drape slope failure frequency. on the Mad Dog Slump 8 results where, through per-
forming 19 piezoprobe tests, it is believed some over-
pressure exists within the slope.
The piezoprobe test consists of pushing into the soil
a spear-like probe equipped with pore pressure trans-
ducers. The probe is typically pushed to 0.9 m (3 ft) into
the formation. The process of lowering the tool through
the drill pipe and pushing it into the soil can lead to
breakage of the piezoprobe tip and/or damage to the
tool electronics. When successful, the insertion process
creates a pore pressure pulse in the formation. The
probe is then kept motionless and the induced pore
pressure is left to decay typically for less than 7 hours.
Because of time and cost consideration, the test is often
terminated before the pressure has fully dissipated.
Techniques have been developed to extrapolate the pore
pressure dissipation curve to assess the in-situ pore
pressure. The measured pore pressure regime at the two
Mad Dog borehole locations were modeled using 2D
finite difference techniques and the resulting pore
pressure field throughout the slope is depicted on
Figure 12. Most of the excess pressure is concentrated
below Horizon 25.
Figure 10. Critical areas for shallow seated failures (colors Such a pore pressure regime is then used as input
indicate various combination of slope angles and associated to slope stability analyses.
soil shear strength that give lowest estimated safety factors). A needle-shaped piezoprobe developed by Fugro
Atlantis field. has been used with relative success to infer ambient
pore fluid pressures on BP projects in the Caspian Sea
and Nile Delta.
300 16
Event Frequency (events/1000 years)
Deposition Rates (cm/1000 years)

15 Seafloor
Deposition rates (smoothed) 14
250 13
Event Frequency 12
200 11 1400 Hrz 25
Water Depth (m)

10
9
150 8
7
1500
6
100 5
4 1600
50 3
2 12 12
1 12 12 16 24
0 0 1700
0 5000 10000 15000 20000 25000 30
Time (years) 1000 1100 1200 1300 1400 1500 1600 1700 1800 1
Distance from origin (m)
Figure 11. Deposition rates and frequency of shallow
slope instability event. Figure 12. Excess head (in m) on the face of Slump 8.

124

Copyright 2005 Taylor & Francis Group plc, London, UK


Measurements of in-situ pore pressure have been been used in an attempt to interpolate between discrete
obtained in the Caspian using various versions of a sampling and in-situ data points. In-situ measurements
newly developed piezoprobe, with the current Mk II of dissolved gas and salt have also been obtained by
version successfully deployed in the East Azeri and means of the NGI BAT probe, in addition to the suite of
Shah Deniz fields. This tool has performed well, with normal down-hole tools. These integrated data sets
at least 90% dissipation attained within 4 to 6 hours of have then been used to improve profiling of soil param-
deployment. The soils in which the tool was deployed eters and also enable comparisons with the HR data
comprised clays with undrained shear strengths up to logged from exploration pilot holes.
400 kPa, with tests performed at depths of up to 450 m In addition to more traditional piston corers a down-
below mudline. hole piston corer can be used to efficiently retrieve near
Tests results have shown the presence of in-situ pore continuous samples of sediments to a depth of at least
pressures in excess of hydrostatic, especially close to 92 m in normally consolidated clays. In a recent deep-
over pressured water bearing sand units. The data water site investigation in the Gulf of Mexico 10 Fugro
acquired has been invaluable in well planning and hydraulic piston corer (FHPC) cores using a 9.1 m
casing design in the shallow section. (30 ft) barrel were obtained in one boring and 8 Shelby
tube push samples were retrieved from another boring
16.7 m (50 ft) away. The diameters of the FHPC and
7.2 Shear strength characterization
push samples were 63.5 mm and 71.9 mm (2.5 in. and
The geotechnical characterization of the Mad Dog and 2.83 in.), respectively. Each of these cores was logged
Atlantis fields was extensive and has been described using the GEOTEK Multi Sensor Core Logger
in detail by Al-Khafaji et al. (2003), Orange et al. (MSCL). Gamma density measurements for the FHPC
(2003b) and Young et al. (2003). Four deep borings and push samples are shown in Figure 13. The push
were taken at the Mad Dog field, to depths ranging sample density matched the data obtained from the
from 146.3 m (480 ft) to 426.8 m (1,400 ft) and four FHPC cores.
150 m (490 ft) borings were taken at the Atlantis field. For this work the recovery efficiency was just under
All borings included push sampling, remote vane 80% of the targeted length (10 m, or 30 ft) and the
testing, and piezoprobe testing. deck to deck deployment times for the FHPC towards
In addition to the deep borings, piston cores, up to the end of the work were about 25% longer than the
18 m (60 ft) deep were taken at 46 locations at Mad deployment time required to obtain push samples.
Dog. At Atlantis, samples were recovered at 45 loca- Recovery efficiency and deployment times are
tions along the Sigsbee Escarpment, including 5 box expected to increase as the operator becomes more
cores and 40 piston cores (up to 18.3 m deep, or 60 ft). familiar with the tool.
The samples were tested in controlled-rate-of-strain
(CRS) one-dimensional consolidation tests, static Ko Gamma Density (g/cc)
consolidated-undrained, strain-controlled, direct simple 1.20 1.40 1.60 1.80 2.00 2.20
shear tests (CKoU-DSS); and multi stage static drained 0
triaxial compression tests (CID-TC) to determine FHPC
stress history, measure drained peak and undrained push sample
shear strength properties of the clays and sands, and
obtain normalized undrained strength properties in the
SHANSEP framework. Additional advanced laboratory 25
tests such as static Ko consolidated undrained, direct
simple shear creep tests (CKoU-DSS-CR), static
Depth (ft)

TruePath Ko consolidated-undrained, triaxial com-


pression or extension tests (TPKoU-TC or TPKoU- 50
TE), and static anisotropically consolidated undrained
triaxial compression tests (CAU-TC) were also per-
formed to assess other key soil engineering parameters
(Al-Khafaji et al. 2003).
More recently, selective ring shear tests have been 75
performed to measure residual drained shear strength
parameters to help assess the potential for progressive
(delayed) slope failures along the Gulf of Mexico Mardi
Gras export pipeline and reactivated gravitational sub- 100
marine sliding in the West Nile Delta.
For some of the deep boreholes in the Caspian sea, Figure 13. Comparison of density from FHPC and push
high resolution (HR) down hole logging tools have samples.

125

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 1. Best estimate SHANSEP parameters for deter- 1600
Boring 1
ministic analyses.

Depth below sea level (m)


SEABED
SF = 1.52
1700
Site Alpha m
1800
Boring 2
Mad Dog 0.180.23 (1) 0.8 1900
Atlantis 0.25 0.67 Boring 2A
2000
(1)
Range. Layers were assigned different values.
2100
500 750 1000 1250
7.3 Deterministic analyses of deep seated Distance down line (m)
failures
The perspective gained is that proper characterization Figure 14. Safety factor and critical surface for Atlantis
Slump E (modified after NGI 2002).
of a slopes pore pressure regime and shear strength
takes a considerable amount of planning and intense
field work. Such a detailed characterization is nonethe-
less essential if the results of slope stability analysis 7.3.2 Effect of soil properties along fault planes
are to be used to make critical decisions. The majority of the supra salt sections at the Mad Dog
and Atlantis fields have been faulted as a result of the
7.3.1 Static limit equilibrium underlying salt movement. When performing a limit
Deterministic slope stability evaluations provided a equilibrium analysis, the question therefore arises:
best estimate safety factor. Details of the analyses What are the soil properties along fault planes? One
for the Mad Dog and Atlantis slopes are given in can try to perform continuous downhole cone testing
Nowacki et al. (2003). A SHANSEP model (Ladd & through a suspected fault plane or try to continuously
Foot 1974) was used to establish undrained direct sample the interval of interest. Such options are not
simple shear (DSS) strength profiles. The best esti- always practical (downhole cone testing requires the
mated parameters in this SHANSEP model are sum- ability to use a seabed reaction frame which implies
marized in Table 1. that the seabed has only modest slope angles).
Limiting equilibrium slope stability analyses were Much is unknown about the effect of slow shearing
performed for circular and non-circular surfaces, (i.e. creep rates of 0.1 to 1 mm/year; see Section 10.1)
using the Morgenstern & Price (1965) method. A few or intermittent offset along fault planes. Is the soil at its
checks on selected cases with the Spencer (1967) residual or peak shear strength? At its peak or residual
method confirmed that different and well-established friction angle? What is the effect of relaxation and con-
procedures obtain very similar results. An actual trig- solidation under the fault plane shear conditions?
ger mechanism was not a defined input in these To try to understand the importance of such issues,
analyses. Therefore both undrained and drained a sensitivity analysis was performed at the Mad Dog
analyses were carried out in order to cover different site where the shear strength along almost vertical
causes of slope failure. faults and almost horizontal potential weak planes was
A total of 7 slope profiles were evaluated at Mad reduced by an arbitrary factor (Figs 15, 16, and 17).
Dog. The slope with the lowest safety factor (Slope Details of the analysis are included in Nowacki et al.
8_2) was about 210 m high with an average slope (2003). For each combination of vertical and horizon-
angle of about 23. The calculated safety factors for tal reduction factor, a deterministic safety factor was
this slope were 1.1 for the undrained total stress calculated.
analysis and 1.3 for the drained analysis. Results are summarized in Figure 17. They show
The calculations were concentrated on one slope at that the impact of a potential reduction of shear
Atlantis. The slope was about 125 m high with an strength along a fault plane on the overall safety factor
average slope angle of about 26. The calculated of the slope depends very much on the inclination of
safety factors for this slope were 1.5 for the undrained the fault plane. For the case studied, reducing the shear
analysis 1.3 for the drained analysis (Fig. 14). strength by 28% along the mostly horizontal weak
The undrained shear strengths used to calculate plane brings the slope to failure (ie. safety factor of 1.0)
these safety factors were not reduced due to the effect but the shear strength has to be reduced by 80% along
of load duration, which was discussed above. The the vertical plane to reduce the safety factor to 1.0.
slope stability calculations were on the other hand The perspective gained is that properly character-
performed using a 2D plane strain model, which izing thrust faults with low angles from the horizontal
results in a lower safety factor compared to the actual is therefore much more important than characterising
3D situation. It was estimated that one of these effects normal faults with high degree of inclination from the
compensates for the other one. horizontal.

126

Copyright 2005 Taylor & Francis Group plc, London, UK


0.7

0.68 MAX STRESS


IN DRAPE WITH ASB-2, OCR=7.3
0.66 THICKNESS < 12 FT
0.64

0.62
ASB-1, OCR>24
0.6 ASB-1, OCR>8
ASB-1, OCR=1.5
0.58

0.56

sin '
EE-4
0.54

0.52 EE-4
0.5 Atlantis Ip<50%
Shallow Mad Dog Ip<50%
0.48 Shallow Mad Dog Ip>50%
0.46 Marlin Ip<50%
Marlin Ip<50%, Slow tests
0.44 King Ip<50%
King Ip>50%
0.42 ASB-1, OCR=1.0
ASB-1, OCR=1.0
0.4
2 3 4 5 6 7 8 9 20 30 40 50 60 70 80 90 200 300 400 500 600 700 800 900

Figure 15. Failure surface for Slump 8. No reduction in 1 10


p' [kPa]
100 1000

shear strength along fault planes and hypothetical weak f

plane (modified from Nowacki et al. 2003).


Figure 18. Drained peak friction angles for clays at various
Gulf of Mexico sites (from Nowacki et al. 2003).

7.3.3 Drained analysis


Drained analyses are also typically performed and
require the soil friction angle as an input parameter.
Figure 18 summarizes measurements from several
sites in the Gulf of Mexico. The peak friction angle is a
function of effective stress level and plasticity indexes.

8 MASS GRAVITY FLOW ANALYSIS

Details of the mass gravity flow analyses performed


for the Mad Dog and Atlantis sites have been described
Figure 16. Failure mechanisms for Slump 8. No reduction in detail by Niedoroda et al. (2003). The mass gravity
in shear strength along vertical fault. Shear strength along flows that are of interest can generally be divided into
potential horizontal weak plane (Horizon 25) reduced by two categories, debris flows and turbidity currents.
70% (modified from Nowacki et al. 2003). These are governed by different flow regimes, and
therefore require two different flow models. Debris
1.0 flows, which we consider to include mudflows, are
mass movements in which the source sediment travels
Strength reduction in vertical weak layer 1, RV

downslope, coming to rest after the initially stored


1.00

1.30

1.40

1.45

0.8 potential energy is dissipated by friction. During these


flows, the source sediment is remolded and reconsti-
tuted; the degree to which this occurs determines the
0.6 rheological and flow properties. The soil mass travels as
a visco-plastic material, with distinct stressstrain rate
SF = 1.3-1.4
characteristics.
0.4 Direct measurement of debris flow and turbidity cur-
SF = 1.0-1.3 rent speeds and other flow characteristics is not prac-
Failure
tical because they are episodic events that are difficult
SF < 1.0
0.2
to forecast. The Bing Model (Jiang & LeBlond 1993)
was used to simulate a number of debris flows observed
in the project areas. Figure 19 depicts one of the debris
flows used to calibrate the input parameters of the
0.0
0.0 0.2 0.4 0.6 0.8 1.0 model.
Strength reduction in horizontal weak layer, RH Figure 20 shows an output that is a representation of
the actual debris flows shown on Figure 19. On this fig-
Figure 17. Summary of results. Effect of potential reduc- ure there is a vertical exaggeration (about 40:1). The
tion in shear strength along near vertical faults planes or figure shows the initial and final profiles of the flow.
near horizontal weak planes on slope safety factors (modi- The horizontal axis is distance down the slope. The
fied from Nowacki et al. 2003). dashed hump is a synthetic starting condition for

127

Copyright 2005 Taylor & Francis Group plc, London, UK


Figure 20. Comparison of hindcast and observed debris
flow run out geometry for debris flow J-2 (vertical exagger-
ation of 40:1) (from Niedoroda et al. 2003).

Figure 21. Predicted debris run outs (yellow lines) and


associated turbidity current velocities for the Atlantis field
Figure 19. Geometry of debris Flow J-2, of one of the (from Niedoroda et al. 2003).
debris flows used to calibrate the input parameters of the
BING model at the Atlantis field (modified from Niedoroda
et al. 2003).
These parameters were then used to predict the run-
out distances of potential future debris flows and the
the debris flow. It approximates a stage during the velocity of the associated turbidity currents (Fig. 21).
deformation due to the initial soil mass failure where
the slide toe converts from a displaced mass to an
unstable mound. It is thought that as this happens the 9 QUANTIFICATION OF TRIGGER
soil mass becomes remolded, losing any previous MECHANISMS
structure and increasing its water content.
From this point the soil mass continues to deform The reliability assessment of an existing slope cannot
as a Bingham fluid. The models were run several times be performed without understanding the mechanisms
for each of the scenarios representing observed large by which this slope could fail. If the slope today is
debris flows. The modeling parameters (Bingham vis- stable, and has been stable for several thousand years,
cosity and yield strength) were adjusted until the why could it fail in the near future?
computed thickness and run-out distance of the flows Natural trigger mechanisms for submarine slopes
agreed with the observed values. Agreement was have been investigated by many authors and were
obtained from different examples when the Bingham summarized by Locat & Lee (2000). Human activities
yield stress and strain rates were set at about 3200 Pa involved in the development and production from the
and 9/s respectively. field needs also to be reviewed to see if they may

128

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 2. Trigger mechanisms for submarine landslides. 9.1 Natural trigger mechanisms
Natural trigger mechanisms for submarine landslides Waves, tides, earthquakes, hydrates dissociation and
(from Locat & Lee 2000) sea level changes, were studied in detail in Jeanjean
et al. (2003b) and were concluded not be credible
Reducing shear strength Increasing shear stresses mechanisms.
Earthquake Earthquake 9.1.1 Sedimentation rates
Wave Loading Wave Loading Slowey et al. (2003) showed that sedimentation rates
Tides Tides
have decreased dramatically in the last 25,000 years, in
High sedimentation rates High sedimentation rates
Gas release Gas release the study area (Fig. 25). There is also significant evi-
Sea level changes Sea level changes dence that the current environment is non-depositional,
Erosion with seafloor erosion actively occurring (Brand et al.
Diapirism 2003, Angell et al. 2003). Such a trend can hardly be
conceived of as being reversed in the near future, thus
Man-induced trigger mechanisms for submarine landslides sedimentation rates are not a credible mechanism to
(from Jeanjean et al. 2003b) cause slope failures during the life of a field (Jeanjean
2003b).
Reducing shear strength Increasing shear stresses
9.1.2 Salt tectonics
Fluid losses in shallow Weight of pieces of Orange et al. (2003) and Angell et al. (2003) clearly
section during drilling equipment on seabed showed the relationship between salt tectonic
Lack of well control Lack of well control processes and slope failures at the Mad Dog and
during SWF during SWF Atlantis sites. Jeanjean (2003b) described the details
Hydrate melting due to Anchor forces during of a 2D finite element analysis that was performed at
hot hydrocarbon storm loading Atlantis to study the effect of the salt movement on the
production seafloor geometry. The results, which should be taken
Re-injection of drill Pipeline/flowline/ as estimates, suggest that, for the very local conditions
cuttings and produced umbilical laying activities modeled, a salt migration rate of 10 mm/year would
water Accumulation of drill cuttings cause a maximum seafloor rate of uplift of 1/1000
on seabed during riserless years. Such a rate will have negligible effects on the
drilling
current seafloor geometry, when engineering time
Reservoir compaction/ frames are considered.
sea-floor subsidence
9.2 Man-induced trigger mechanisms
The drilling of numerous wells in the vicinity of the
encourage slope instability. The identified mechan- slope could affect the pore pressure regime of the
isms are summarized in Table 2. slope and thereby affect its stability.
The key trigger mechanism in the Caspian is react- The volume of water flow experienced during the
ivation of pre-existing submarine slides as a result of drilling of the riserless section of the appraisal wells at
earthquake loading. Localized near seabed failures the Mad Dog spar location has been seen to be slight to
may be caused by sediment surcharge caused by mud negligible. No fluid losses have been noticed either. It is
volcano eruptions. therefore believed that the shallow sand intervals can be
The main risk to subsea developments in the West properly controlled and the drilling of the riserless sec-
Nile Delta is currently believed to be the reactivation tion of the wells will not lead to a change in the pore
of pre-existing large-scale submarine slide packages pressure regime of the slope at the Mad Dog field.
by a combination of tectonic uplift, sediment sur- Similar conditions have been experienced at Atlantis,
charge, toe erosion and earthquake loading. drilling below the escarpment. It is generally concluded
Mechanisms capable of triggering a slope failure at that the risk of shallow water flow leading to a forma-
Mad Dog and Atlantis, within an engineering time tion dissolution, triggering slope instability, is low to
frame (defined as 3,000 years or less in this work) negligible.
were thoroughly investigated, as described by Jeanjean Because of the potential for affecting the pore pres-
(2003b). The evolution of the escarpment over a sure regime on the face of the escarpment, drill cut-
longer, geological, time period was beyond the scope tings and produced water will not be re-injected above
of this study. the salt. The re-injection wells, if any, will be sub-salt,
The discussion below focuses on Gulf of Mexico and at a sufficient depth not to influence the pore pres-
experience at Mad Dog and Atlantis. sure conditions close to the face of the escarpment.

129

Copyright 2005 Taylor & Francis Group plc, London, UK


The production of hydrocarbons will reduce the
pressure in the reservoir and thus result in reservoir
compaction. Compaction may cause the overburden to
subside and, because the reservoirs straddle the escarp-
ment, potentially increase the seafloor slope angles.
Screening analyses were performed to assess reservoir
compaction and the resulting seafloor subsidence.
Results showed that the maximum predicted seafloor
subsidence was about 0.37 m (1.2 feet) (Jeanjean
2003b). Reservoir compaction is therefore expected to
cause little seafloor subsidence and is not a credible
trigger mechanism.
When laying a flowline or an umbilical across a Figure 22. Example of logic tree assessment of fault
steep slope with the potential presence of a soft drape, parameter for probabilistic fault hazard assessment (PFHA)
the interaction between the pipe and the seafloor sedi- at Atlantis Central Graben fault (from Angell et al. 2003).
ment may increase the shear stress acting on the soil,
thereby triggering instability. The currently chosen field
architectures for the Mad Dog and Atlantis fields do not due to earthquake strong ground shaking. (See Youngs
include laying flowlines or umbilicals on such slopes. (2003) for a complete review of the methodology of
conducting a PFDHA.)
9.3 Conclusion on trigger mechanisms For this study, the input parameters are the selection
of the appropriate geologic horizon, its age, its cumu-
The conclusions from the detailed analyses of potential lative offset, the average displacement per event, and
trigger mechanisms at Mad Dog and Atlantis are very the shape parameter a of the gamma distribution used
powerful. First, it has been established that, consider- to assess the conditional probability of exceedance
ing the lack of seismicity of the area, all natural trigger (Fig. 22). Weighted alternatives were developed for all
mechanisms are slow acting, not random and none is of these parameters at each fault crossing, and hazard
plausible to cause slope instability during the life of the analyses were performed using the full range of alter-
field. The only plausible trigger mechanisms are man- native parameter sets.
induced mechanisms and these can all be negated by Two approaches are used to estimate displacement
preventive actions. per event (De). The first approach assumes that dis-
placements occur as relatively large (1.0 m) discrete
events and De is calculated by dividing the total cumu-
10 QUANTITATIVE RISK ASSESSMENT lative offset by the possible number of faulting events.
This approach was developed at the Mad Dog site
As discussed in Section 2, quantifying the risks due to where there is stratigraphic evidence of episodic fault
geohazards allows the proper integration of such risks growth.
with other project risks thereby facilitating any deci- The second approach considers the potential that
sion making process. displacement occurs by fault creep or more frequent
small events. In this case the displacement per event
10.1 Probabilistic fault displacement hazard was assumed to be between 0.1 and 2.0 m and equal
analysis (PFDHA) weight (maximum uncertainty) was assigned to values
within this range.
Seafloor faults having strong geomorphic expression A major effort was undertaken to get the best pos-
and evidence for late Quaternary activity (i.e.  sible age control for stratigraphic marker horizons in
150,000 years) are common geologic features associ- the Mad Dog and Atlantis field areas utilizing pale-
ated with the Sigsbee Escarpment. A probabilistic ontological and oxygen isotope analysis, radiocarbon
fault displacement hazard analysis (PFDHA) was con- numerical age-dating, and stratigraphic modeling
ducted as part of the site investigations for the Mad (Slowey et al. 2003).
Dog and Atlantis fields (Fig. 3). Fault slip rate is a measure of the amount of slip on
The objective was, at key field locations, to relate the fault averaged over a time period that encompasses
the annual frequency of occurrence of a fault displace- multiple ruptures.
ment event with the size of the event. Details of the The probabilistic fault displacement hazard analy-
analyses can be found in Angell et al. (2003). sis (PFDHA) addresses how frequently displacement
The methodology employed to evaluate fault dis- events occur and how large the displacements are in
placement hazard is analogous to the well-developed each event. The hazard can be represented by a dis-
formulation for probabilistic evaluation of the hazard placement hazard curve analogous to ground motion

130

Copyright 2005 Taylor & Francis Group plc, London, UK


Figure 23. Hazard results for PFHDA of Atlantis Central
Graben largest fault (modified from Angell 2003).

Figure 24. Summary of PFHA at Mad Dog field (modified


from Geomatrix, 2002).
hazard curves. The example hazard curve shown on
Figure 23 represents the hazard at a point. It relates
the amount of displacement in a single event to how 10.2 Age dating of past mass flows activities
often displacements of that magnitude or larger occur
(i.e. the frequency of exceeding a specified amount of It is of critical importance to understand when and
displacement). how often failures did occur in the past. To that effect,
Hazard from potential seafloor offset at fault cross- more than 79 samples from Mad Dog and Atlantis
ings is judged to be moderate to low. Fault offsets of were radiocarbon dated. The technique used to date
the shallowest horizons (less than 15,000 years old) the sediments has been described in details by Slowey
are typically less than 10 m to several tens of meters. et al. (2003). Radiocarbon dating is generally effec-
Fault slip rates are on the order of tenths to several tens tive for samples in the age 0 to 40,000 years. Ages
of meters per thousand years (m/kyr), also millimeters of potentially older sediments were in some instances
per year (mm/yr), with most values in the range of constrained using nanofossil biostratigraphy, specifi-
210 m/kyr. Similarly, the probabilistic annual recur- cally, calcareous nannofossils called coccoliths.
rence of 1 m events is typically less than 10
3. These results are not presented here.
The results of all the studies performed at the Mad The measured dates were used for two purposes: to
Dog site are visually summarized in Figure 24. derive the sedimentation rates in the vicinity of the
The perspective gained by these studies is that the Sigsbee Escarpment and to determine the time and
presence of potentially active faults does not preclude frequency of occurrence of large-scale slump features
safe development of seafloor facilities. To evaluate risk in the area. The sedimentation rates are shown in
associated with potential seafloor faulting, integrated Figure 25.
hazard studies can and should be conducted in the early These results indicate that sedimentation rates dur-
stages of project development, with an underlying ing the latest Holocene are, on average, on the order of
intent to understand the causative processes and quan- 12 cm/kyr. In contrast, sedimentation rates during the
titatively and explicitly evaluate the locations, magni- last glacial maximum approached 700 cm/kyr, nearly
tude and recurrence potential of displacement events. 60 times the rate that occurred during the latest
Such evaluation nevertheless requires large efforts in Holocene. Slowey et al. (2003) described in detail that
high-resolution surveys, core sampling and age dating. such a decrease is believed to be due to the reduction

131

Copyright 2005 Taylor & Francis Group plc, London, UK


800 in Mississippi meltwater outflow 13,000 BP and the
Atlantis Core CSS-1
Deposition rate (cm /1000 years)

700 Mad Dog - Core RI-1


consequent reduction in sediment supply to the Gulf
600
of Mexico, together with the effect of sea level rise
onto the continental shelf.
500
The activities of the deep seated slumps for the
400 Mad Dog Slump 8 and Atlantis Slump E have been
300 summarized in Figures 26 and 27. As can be seen, the
200 portion of the slump 8 slope analyzed with determin-
100
istic limit equilibrium technique shows no sign of
activities in the last 7,000 to 11,000 years. Also
0
0 5 10 15 20 25
Slump E shows no sign of activities in the last 9,190
Time before present (kyears) years and probably in the last 16,000 years.

Figure 25. Sedimentation rates at the Atlantis and Mad


Dog sites, above the Sigsbee Escarpment. 11 PROBABILISTIC SLOPE STABILITY
ANALYSES

11.1 Deep seated failures


Finite element analyses of the stability of typical deep-
water slopes in the North Sea have shown that a
2-wedge model of slope failure will be critical if a
weak layer with relatively low undrained shear strength
exists at the base of the slope (NGI 2002). Further
investigations showed that a simple 2-wedge model
consisting of sliding blocks and a collapsing block
(Fig. 28) may be used to investigate the stability of
slopes comprised of more or less horizontal soil layers,
even if a distinct weak layer is not present. This simple
model provides a clear insight into the main parameters
controlling the slope stability and it can be easily
adopted for reliability analysis using the first- and
second-order reliability methods (FORM and SORM).
A closed-form solution for safety factor can be
Figure 26. Age date summary of Mad Dog Slump 8. The obtained by considering the equilibrium of the two
locations of the two soils borings are shown with dots.
wedges and assuming the same safety factor SF on all
slip planes (Nadim et al. 2003):
CSS-2
Last slump: 9,970 years BP

EE-1
CSS-3 (2)
Last slump 8,510 years BP

25,450+ years w/o slump


EE-2 20,000+ years where S 1max, S 2max, and S 3max are the maximum shear
w/o slump
resistance along the sliding planes shown on Figure 29,
EE-3 EE-6
EE-4
EF-4
EF-5
20,000+ years
9190+ years w/o slump
w/o slump No Evidence
EE-5
of slump in
core
3.6 km 16070+ years w/o slump

Figure 27. Age data summary of Atlantis Slump E. Yellow


labels indicate name of piston cores. The locations of the Figure 28. 2-wedge model for slope stability (modified
two soils borings are shown with green dots. after NGI 2002).

132

Copyright 2005 Taylor & Francis Group plc, London, UK


W1 and W2 are the submerged weights of wedge 1 and Mad Dog Slump 8:
wedge 2, and P is the passive lateral resistance at the
toe of the slope. The angles  and  are defined on Mean safety factor  1.22
Figure 28. Reliability index  1.28
Form Pf  0.1
11.2 Example calculations for deep
seated failure 12 UNDERSTANDING THE RESULTS FROM
Probabilistic slope stability analyses were performed FORM ANALYSES ESTIMATING
for Mad Dog Slump 8 and Atlantis Slump E. The ANNUAL PROBABILITY OF FAILURE
starting point of the analyses was the critical failure FOR DEEP-SEATED SLIDES
mechanism identified in the deterministic slope sta-
bility calculations. For Atlantis Slump E, the 2-wedge The probability of failure given by the FORM analyses
failure mechanism predicted a deterministic safety described above does not have a return period associ-
factor of 1.54, which is slightly greater than the safety ated with it and is therefore NOT the annual probabil-
factor of 1.52 obtained with a circular mechanism. ity of failure that BP uses as the basis of assessing risk.
The following parameters were considered random When used for slopes to be built or excavated
in the analyses: submerged unit weight in each layer, (dams, embankments or levees), the interpretation is
SHANSEP parameters  and m in each layer (total of straightforward. A FORM analysis, for a given distri-
5 distinct soil layers with total thickness of 105 m), bution of slope geometry and a given distribution of
 p at top and bottom of each layer,  and m in sliding key soil properties, would give the probability that the
base layer, removed overburden at toe of the slope, slope would fail immediately after construction. But
strength anisotropy factor , and modelling uncer- the slopes under consideration here have already been
tainty parameter . In addition to these random built by nature and have been stable for thousands of
variables, the following input parameters were speci- years.
fied as fixed values (see Fig. 28): slope angle Nadim et al. (2003) proposed two possible interpret-
  32, inclination of seabed at top of slope   5, ations of FORM results that may be used to obtain the
passive lateral resistance at toe of slope P  0, over- annual probability of failure, recognizing that some of
burden at toe of slide, toe  0. The assumed distribu- these interpretations can be challenged, and offering
tions for the random variables are provided in NGI these ideas as a starting point for research into an area
(2002). The distribution of the safety factors for Mad that needs further understanding and study. The pre-
Dog Slump 8 and Atlantis Slump E are shown on ferred interpretation will now be presented.
Figure 29. First the effects of increased sedimentation on the
The results were: stability of the slope is investigated with FORM ana-
Atlantis Slump E: lyses to estimate the rate of destabilization of the slope
due to an average 0.12 m/1,000 years sedimentation rate
Mean safety factor  1.54 (Fig. 25). Figure 30 shows the changes in probability of
Reliability index  3.34 failure and mean factor of safety over the next 10,000
Form Pf  4.210
4 years. It can be seen that the mean safety factor
decreases at an almost constant rate of 0.0008/1000

1
Atlantis Slump E 4.9E-04 1.544
FORM Probability of Failure Pf
Probability of non-exceedance

0.9 Form Pf
Mad Dog Slump 8 1.543
4.8E-04 Mean Safety Factor
0.8
Mean Safety Factor

1.542
0.7 4.7E-04
1.541
0.6 4.6E-04 1.54
0.5 1.539
4.5E-04
0.4 1.538
4.4E-04
0.3 1.537
0.2 4.3E-04 1.536
0.1 4.2E-04 1.535
0 0 2000 4000 6000 8000 10000
0.5 0.6 0.7 0.8 0.9 1 1.1 1.2 1.3 1.4 1.5 1.6 1.7 1.8 1.9 2 Time in future (years)
Safety Factor
Figure 30. Increase in FORM probability of failure (Pf)
Figure 29. Cumulative distribution of safety factor from and associated decrease in mean factor of safety with time
FORM analyses for Atlantis Slump E and Mad Dog Slump 8 due to sedimentation, Atlantis Slump E (modified after
(modified from NGI 2002). Nadim et al. 2003).

133

Copyright 2005 Taylor & Francis Group plc, London, UK


years. This information will later be used for estimat- 13.1 International Centre for Geohazards (ICG)
ing annual probabilities of failure.
BP maintains close ties with the International Centre
The proposed interpretation of the FORM failure
for Geohazards (ICG), which is one of the 13 Centres of
probabilities is consistent with the theory behind the
Excellence (Senter for Fremragende Forskning, SFF)
FORM analysis. However, a number of assumptions are
established by the Research Council of Norway in
required:
2003. The Norwegian Geotechnical Institute (NGI) is
Failure occurs when the safety factor is 1.0. Since the host organization for ICG. Other partners in the
the slopes are standing today, the safety factor, centre are the University of Oslo (UiO), the Norwegian
although unknown, is greater than 1. University of Science and Technology (NTNU), NOR-
Safety factor is distributed according to the distri- SAR, and the Geological Survey of Norway (NGU).
bution calculated by FORM and presented earlier The Centres objective is to be an international centre
in this paper. of expertise on basic and applied research on geo-
The mean safety factor, and also the true safety fac- related natural hazards (geohazards).
tor, are decreasing at a rate of 0.0008/1,000 years. The activities of the center focus on:
Unsaturated soils and rain-induced landslides.
Based on the assumptions above, the question of
Risk and vulnerability analysis for geohazards.
annual probability of failure becomes the question of
Earthquake hazard, vulnerability and risk.
the likelihood that the current factor of safety will fall
Rockslides and engineering geology: stability, fail-
below unity during the following year. The current
ure, sliding and consequences.
factor of safety is unknown, but we do know its distri-
Landslides in soft sediments.
bution (distribution from FORM, but truncated to
Offshore geohazards.
reflect the fact that the slope is stable today). This
GIS applications in geohazards.
interpretation is basically a Bayesian updating proced-
Applications of Synthetic Aperture Radar (SAR)
ure where the a-priori information is that FS 1.
in geohazards.
Formally, the updated (or posterior) distribution of
Slide dynamics and mechanics of disintegration.
the factor of safety is (Nadim et al. 2003):
Tsunami modeling and prediction.
Development of graduate studies in geohazards.
(3)
13.2 BP/Scripps collaboration on geohazards
On November 11 2004, Scripps Institution of
The slope will fail during the following year only if Oceanography at the University of California, San
its current value is such that, with the rate of deterio- Diego (UCSD), and BP America Inc. announced the
ration of 0.008/1,000 years, it will fall below 1 during signing of a 3-year partnership, marking the beginning
1 year. This very simple calculation can be performed of a long-term research collaboration. The initial focus
in several slightly different ways, and annual failure of the program is to develop and evaluate new tech-
probabilities for Atlantis Slump E range between nologies to image and characterize the seafloor and sub-
10
7 and 10
9 and between 10
5 and 10
6 for Mad seafloor. Using a wide variety of surveying techniques
Dog Slump 8, depending on the assumptions made. such as electromagnetics, fiber optics, acoustics, and
The perspective gained is that additional research autonomous underwater vehicles (AUVs), Scripps and
is needed to formalize the interpretation of the results BP scientists aim to demonstrate the deployment, reli-
coming from FORM or SORM in obtaining the annual ability, and resolution of specific seafloor monitoring
probability of failure. However, both approaches give systems and further understand the fundamentals of
confidence that annual probabilities of failure are submarine slope behavior, particularly in seismically
several orders of magnitude lower than the timeless active areas.
probability of failure coming directly out of FORM. The project will instrument the Gaviota slide in the
Santa Barbara Channel, offshore California where
geomorphological indicators of possible seafloor
13 AREAS OF RESEARCH FOR IMPROVED instability exist (Fig. 31).
UNDERSTANDING OF SLOPE FAILURE
MECHANISMS 13.3 Project Offshore Deep Slope (PODS) JIP
BP is also sponsoring the PODS Joint Industry
BP actively promotes and seeks an improved under- Project (JIP). The aim of the project, initiated by
standing of the fundamental issues associated with C-CORE, is to develop a risk framework for analyz-
offshore geohazards and submarine slopes. Examples ing slope areas that is relevant to offshore Eastern
of on-going projects are descried below. Canada and other regions around the world.

134

Copyright 2005 Taylor & Francis Group plc, London, UK


10-2

Annual Probability of Failure


D/B = 5 at FS=2.0

10-3

Figure 31. Gaviota slide where seafloor monitoring will be


installed in 2005/2006 as part of the BP/Scripps collaboration.

10-4
1.5 2.0 2.5
Design Factor of Safety

Figure 33. Probability of failure of suction anchor for taut


leg mooring as a function of safety factor (from Clukey et al.
2000).

14.1 Objective Design to target reliability


One of the fundamental issues faced by operators is
Figure 32. Example of slope failure study in the centrifuge
from PODS JIP (shown with permission).
to ensure that they achieve a given level of reliability
for their facilities. Design codes and recommendations,
The following information about the project can whether in a total safety factor formulation (i.e. work-
be found at http://www.mms.gov/tarprojects and is ing stress) or in a partial loads and resistance factors
therefore in the public domain. The accomplishments formulation (i.e. LRFD) are calibrated such that the rec-
so far include: ommended factors provide a given level of reliability.
To calibrate such safety factors, reliability analyses
Data collected and summarized for the project must be performed. They describe the uncertainty
from a number of different authors and research associated with the random variables in the equations
programs. of interest and calculate the probability of non-
Case histories and loading conditions analyzed and performance. In the case of offshore foundation capac-
a synopsis of the methods used by researchers of ity calculations, soil shear strength and unit weights
these failures. are the basic random variables involved.
Studies of the Scotian Shelf and Gulf of Mexico on To define such variables, the operator must per-
geo-conditions which caused failures of the exist- form a geotechnical site investigation. The scope of it
ing slopes in those areas. greatly varies from project to project and depends on
Failures analyzed for existing slope stability equi- the geological setting, type of foundation, load mag-
librium approach. nitude to be resisted, type of equipment available.
Geotechnical aspects of gas hydrates described. Clukey et al. (2000) presented the results of reliabil-
Demonstration testing of centrifuge presented to ity studies performed to estimate the probability of fail-
gain an insight to the mechanics of slope failure. ure of a typical deepwater anchor: a suction anchor. For
Framework on risk analysis presented to identify a taut leg system, they designed a suction anchor with a
dominant failure mechanisms. deterministic safety factor of 2.0 and a depth to diam-
Basis of an experimental program designed and eter D/B ratio of 5. They then varied the safety factor by
which is operational. changing the D/B ratio. The probability of failure was
An example of slope failure observed in that cen- then computed for each safety factor and for a coeffi-
trifuge is included in Figure 33. cient of variation (COV) of 0.3 for the shear strength
intercept at the mudline and a COV of 0.25 for the soil
shear strength gradient (Fig. 33). The effect of varying
14 GEOTECHNICAL DESIGN IN the soil uncertainty was then analyzed, and quanti-
GEOHAZARD PRONE AREAS fied the increased in probability of failure associated
with the increase in shear strength COV (Fig. 34).
Once the geohazard risk assessment has been com- Using the model developed by Gilbert et al. (1999),
pleted, there still remains the challenge of designing the effect of increased distance from a boring on the
and installing the said facility. calculated probability of failure was estimated and is

135

Copyright 2005 Taylor & Francis Group plc, London, UK


10-2
Annual Probability of Failure

D/B = 5, FS=2.0

10-3

10-4 Figure 36. Location of main offshore fields with BP equity


0.2 0.35 0.5 in Gulf of Mexico.
C O V of Shear Strength Gradient

Figure 34. Probability of failure of a suction anchor for a


taut leg mooring as a function of soil uncertainty (from
Clukey et al. 2000).

10-2

D/B= 5, FS = 2.0
Annual Probability of Failure

10-3
Figure 37. Na Kika semi submersible mooring pattern and
distance from existing borings (modified from Newlin 2003).

To illustrate the variety of solutions for site investi-


gations in non geohazard-prone areas, details of the
site investigations for the Na Kika, Horn Mountain,
10-4 Thunder Horse, Holstein, and Marlin fields are given in
0 200 400 600 800 1000
Figure 36.
Distance to Soil Boring, m

Figure 35. Probability of failure of a suction anchor for a 14.2 Examples of SI in non geohazard- prone
taut leg mooring as a function of distance to nearest boring areas
(from Clukey et al. 2000).
Newlin (2003) described the site investigation and the
process adopted for the selection of the soil design
shown in Figure 34. The results are shown for the case properties (shear strength and unit weight) for the
where a suction caisson foundation would be placed anchoring system of the Na Kika semi-submersible
up to 1 km away from a soil boring location. Figure 35 production platform in the Gulf of Mexico. A salient
indicates a relatively small (factor of 2) impact on the feature of this work is that the location of the platform
probability of failure because of soil borings not being (and therefore of its mooring system) changed signif-
located at the foundation site. However, it should also icantly after the site investigation was performed and
be noted that the results shown in Figure 35 do not yet the site investigation was not repeated. Newlin
account for potential changes in the geologic stratig- described how geophysical data was used to map geo-
raphy, where the soil units could change, for example, logical units, how uncertainty in design properties
from clayey to sand/silt conditions. Therefore, the were quantified, and how design values were selected.
results reflect soil profiles and their associated Experience with the mooring on drilling rigs in the
uncertainties for relatively homogenous marine clay Na Kika field was also used to increase confidence in
deposits typical of those of non geohazard-prone sites. the selected values. A key feature of the geotechnical

136

Copyright 2005 Taylor & Francis Group plc, London, UK


N anchor SW anchor Dorsey Canyon
Petit Bois
cluster cluster
Dome

Spar

SW anchor
cluster

Figure 40. Location of the Holstein spar mooring system


in Gulf of Mexico OCS Block GC644/645, with suction
anchors, cores, CPTs, and soil borings.
Figure 38. Seabed rendering and anchor pattern at the
Horn Mountain spar location in Gulf of Mexico OCS Block
MC126/127.

Figure 41. Location of the Thunder Horse floating produc-


tion system in Gulf of Mexico OCS Block MC778, with
suction anchors and associated soil borings.

Figure 39. Details of Horn Mountain spar suction anchors


The site investigation for the Thunder Horse semi-
locations and associated geotechnical soil borings.
submersible production drilling quarter platform is
characterization is that the anchor clusters are located summarized in Figure 41. It included one boring at
between 2.2 miles and 5.5 miles from the closest soil each cluster location.
boring (Fig. 38). Last, as discussed by Jeanjean et al. (1998), the site
Other investigations described here include that investigation for the Marlin TLP site included 4 bor-
performed for the Horn Mountain spar. The mooring ings because a spar was also under consideration for
system consisted of 3 clusters of 3 anchors about the field development. The seabed location of the bor-
2,800 m away from each other. The geophysical data ings is shown in Figure 42. It should be noted that
indicated little variation in sediment morphology and although the borings were up to 2,500 m apart, the
little variation in soil properties were expected from shear strength profile shows a low scatter that is typ-
one cluster to the other. One boring was drilled at the ical of non geohazard-prone areas in the Gulf of
center of each anchor cluster (see Figs. 38 and 39). Mexico (Fig. 43).
Other examples of site investigation include the The perspective gained is that in non geohazard-
Holstein and Thunder Horse field. The Holstein spar prone areas the operators have quite a large flexibility
mooring system consists of 4 clusters of 4 anchors. in designing an appropriate geotechnical site investi-
The site investigation consisted of 1 boring and 1 gation by judiciously selecting between tools such as
CPT at the center of the spar, and of 4 long piston- borings, and/or CPTs, and/or piston cores. The clos-
cores and CPTs at various locations (Fig. 40). est information can be up to 5 miles away from the

137

Copyright 2005 Taylor & Francis Group plc, London, UK


This experience, particularly at the Na Kika site, is
consistent with the results of the reliability studies of
Clukey et al. (2000) which concluded that, in uniform
geology, the increase in probability of failure due to the
boring being up to 1 km away from the anchor is small.

14.3 Site investigation and nature of soil design


parameters uncertainty in geohazard-prone
areas
Let us now address the nature of soil uncertainty. As an
example of site investigation in geohazard-prone areas,
key features of the site investigation performed for the
Mad Dog spar mooring system will be described and
compared with the examples above from non-geohazard
prone areas.
Christian (2003) explains the differences between
Figure 42. Seabed rendering and location of the 4 geotech- aleatory and epistemic uncertainties. Epistemic
nical borings collected at the Marlin site (from Jeanjean uncertainties are due to our lack of knowledge and can
et al. 1998). The average distance between boring BH-1 and be reduced by additional information. Additional infor-
borings BH-2, BH-3 and BH-4 is 1300 m. mation may not eliminate epistemic uncertainties but,
in general, more information tends to reduce epistemic
uncertainties. Aleatory uncertainties on the other hand
are due to the random nature of a phenomenon. More
information will not reduce aleatory uncertainties
although it may establish more precisely the param-
eters governing that uncertainty. Christian (2003)
states that the problem facing geotechnical engineers
is epistemic rather than aleatory; it follows from a
lack of knowledge about materials and geometries
rather than from inherent randomness in them.
This is certainly true in non geohazard-prone areas.
Consider the case of the Na Kika, Horn Mountain,
Thunder Horse and Holstein site investigations previ-
ously described. The uncertainty associated with the
design was reduced by the appropriate number of
borings, cores, and CPTs until most of the remaining
uncertainty became epistemic and was due to the low
natural variability of the sediments. Newlin (2003)
describes an example of cost benefit analysis to deter-
mine when the cost of taking additional borings is no
longer offset by reducing epistemic uncertainties.
Figure 43. Shear strength design profiles for Marlin Is Christians statement still valid in geohazard prone
borings BH-1 through BH-4 (from Jeanjean et al. 1998). areas? Consider the example of the Mad Dog mooring
system. The North East and North West cluster are
anchor location and, given a thorough understanding located above the escarpment in an area that, except for
of the surficial geology and appropriate selection of a thin debris flow layer, has been relatively free of mass
design parameters, the foundation still can be appropri- wasting events. The Southern cluster of anchors is
ately designed (e.g. Na Kika). It should also be pointed located on the Sigsbee Escarpment itself in an area
out that in the fall of 2004, the eye of Hurricane Ivan affected by past debris flows (Fig. 44).
passed right over the Na Kika FPS and the Horn A soft clay hemipelagic drape, about 5 m thick, blan-
Mountain spar. The significant wave height was hind- kets the area. Under it, lies thick debris flow deposits.
cast to be 17 m (51 ft) at the Na Kika site and 17.6 m The scope of the site investigation is summarized in
(52.8 ft) at the Horn Mountain site, with a return period Figure 45.
well in excess of the 100 year design event (Driver A typical sub-bottom profiler line above the
2005). Yet the platforms and mooring systems did not escarpment tying the two Northern clusters together
experience major damage. is shown on Figure 46. It clearly shows continuous,

138

Copyright 2005 Taylor & Francis Group plc, London, UK


4.8 km

Cluster 1

GC782

Cluster 3

CPT Cluster 2
Borings

Figure 44. Perspective view of Mad Dog spar anchors


seabed locations.

4.8 km

Cluster 1

GC782

Figure 46. Integration of CPT results at clusters 1 and 3


Cluster 3 with sub-bottom profiler line at Mad Dog spar.

CPT Cluster 2
Borings

Figure 45. Top view of Mad Dog spar mooring pattern in


Gulf of Mexico OCS Block GC782 and associated CPT and Figure 47. Sub-bottom profiler line near Mad Dog anchor
soil borings (North is up in Fig. 45). cluster 2.

parallel and horizontal beds from one cluster to the firmly believed that no extrapolation from one anchor
other. 1 boring and 1 CPT were collected at the center location to the other could be done reliably. The scope
of each Northern cluster (Fig. 46). The CPT profiles of the field work therefore included 1 boring and 1 CPT
are overlaid on the seismic lines in Figure 46 and for each of the 4 anchors in the cluster (Figs 45 and 48).
demonstrate that, except for a thin debris flow layer, the An example of the results is given in Figure 50.
sediments strength and layering are consistent from The greatly increased variability in the soil properties
one cluster to the other. An additional boring or CPT at is obvious. The profile consists of interbedded soft
a particular location would have further reduced the clay slump units, silts and sand layers, and stiff clay
uncertainty in the soil strength but probably not by a debris flows. It is interesting to note that the stratigr-
great deal because the natural variability of the sedi- aphy inferred from the CPT and the boring also differ
ments is relatively low. (Fig. 50). The elevation of the top sand/silt unit is
A typical sub-bottom profiler line in the vicinity of lower in the boring than in the CPT, although the bor-
the Southern anchor cluster 2 is shown on Figure 47. No ing and the CPT are only 10 m apart. Potential errors
beddings can be interpreted. The only features that can in depths measurements were ruled out and it is
be interpreted from the data are the interface between believed that the depth data is reliable. One push
the hemipelagic drape and the thick debris flows below sample that was split and photographed did show silt
it. Before the site investigation was performed, it was bedding inclined at 60 from the horizontal (Fig. 50).

139

Copyright 2005 Taylor & Francis Group plc, London, UK


Qnet (ksf) Stratigraphy Stratigraphy
interpreted interpreted from
0 10 20 30 40 50 60 70 80 90 100 from CPT boring 10 meters
0 away from CPT

10

20

30

40

Depth (ft)
50

60

70

80

90

100

Figure 48. Details of site investigation for Mad Dog


anchor cluster 2, Anchor 5. Note that 1 CPT and 1 boring
were taken near the center of each anchor.

Qnet (ksf)
0 20 40 60 80
0

10 Interbedded silty sand and clays with


bedding planes inclined at 60 degrees
20 from horizontal

30 Figure 50. Comparison of inferred stratigraphy from CPT


and borings located 10 m apart. Discrepancy is explained by
40
60 inclined bedding planes in debris flow units.
50

60
These silt beds are speculated to have been part of an
Depth (ft)

70 intact block entrapped in the debris flows. The block


c rotated during the flow run-out and came to rest at
80
this steep angle. This explains why large differences
90 in depths of layers can be observed over very short
distances.
100 CPT3-A
The perspective gained is that the total uncertainty in
CPT1-A
110
CPT3-B
the soil parameters for the Mad Dog Southern cluster is
CPT-2D greater than that of the two northern clusters, despite the
120
CPT-2C fact that the scope of work included 4 borings and 4
130 CPT-2B CPTs for the southern cluster and only one boring and
CPT-2A
140
one CPT for the two northern clusters. The authors also
argue that this increased uncertainty in the Southern
Figure 49. Comparison of CPT at cluster 1 (CPT1-A) and cluster, contrary to the Northern cluster where the
3 (CPT3-A and CPT3-B) with the CPTs from cluster 2 (CPT uncertainty is epistemic, is aleatory in nature. Obvi-
2A, 2B, 2C, 2D). ously, the uncertainty at the Southern anchor locations

140

Copyright 2005 Taylor & Francis Group plc, London, UK


Undrained Shear Strength Measured by
Minivane (psf)
0 200 400 600 800 1000 1200 1400
3 10.5
indicates exact location of
shear strength measurements
10.6
3.1

10.7
3.2
10.8

Depth Below Mudline (feet)


Depth Below Mudline (m)

3.3
10.9

3.4 11

11.1
3.5

11.2
3.6
11.3

3.7
Figure 52. Example of 100 mm piston core in chaotic
11.4 debris flow deposit.

3.8 11.5
0 10 20 30 40 50 60 70
Undrained Shear Strength Measured by In geotechnical engineering, it should be noted that
Minivane (kPa) a large number of suction anchors have been instal-
led worldwide to moor floating production systems,
Figure 51. Example of highly erratic shear strength profile although no to little design guidance was available in
in debris flow core. Note that at 3.45 m, some of the data
points are only 1020 mm apart and yet the shear strength
design codes. Despite recent significant improve-
varies by a factor of 2. ments of recommended practice in API codes such as
RP2SK, some of the recommendations are still quali-
tative in nature. Therefore the design of suction
anchors in simple normally consolidated, linearly
was nevertheless greatly reduced for each location by
increasing clay shear strength profiles still requires
taking the 4 borings and CPTs. Standard practice (1
input and knowledge not found in API RP2SK. API
boring and 1 CPT per cluster) would have indeed
RP2SK still only focuses on the design recommenda-
been inappropriate and the extra field work has been
tions for suction anchors to soft clay profiles. Sand,
considered worthwhile.
silts and layered profiles are still without clear guid-
Further evidence of the random nature of debris
ance and warrant increased attention.
flow deposits is illustrated by Figure 51 which shows a
Codes also tend to focus on typical design conditions
photograph of a split core, collected in the vicinity of
because these are the ones for which there is experience
the anchor for geological purposes, with the associated
and data is available. Because geohazard-prone areas
shear strength measured by minivane for some of the
tend to present unique soil conditions, the likelihood
individual clasts present in the core. Large variation
that codes will not address the situation faced is
can be observed within a few centimeters in the core.
increased compared to non geohazard-prone areas.
Other evidence of the chaotic nature of debris flow
deposits is shown on Figure 52 where individual
clasts of different colors can be seen.
Therefore, in geohazard-prone areas, increased 14.5 Lack of validation of design methods:
soil uncertainty is present and the nature of the uncer- Model testing
tainty is now much more aleatory than epistemic. Because of the lack of existing recommended practice
that addressed the unique conditions found at the Mad
Dog prospect, BP decided to calibrate the method
14.4 Creation of industry guidelines used in the suction anchor design with centrifuge tests.
Recommended practices in codes are developed Figure 53 shows an example of such a test.
based on the knowledge and experience gained in a The perspective gained is that, in geohazard areas,
few pioneering projects. Rather than having a project design methodologies will likely require more testing
follow the guidance of the codes, the codes follow the and validation because the soil profile may fall out-
learnings of the projects. side the usual range.

141

Copyright 2005 Taylor & Francis Group plc, London, UK


This paper could not have been written without the
many companies, universities, and individuals involved
in these projects. The authors are grateful to the staff of
AOA Geophysics, BP America, BP Exploration Ltd.,
C&C Technologies, Cornell University, Fugro, GEMS,
Geomatrix, the Norwegian Geotechnical Institute,
Rockfield, TDI Brooks, and Texas A&M University
involved in these projects.
The authors are also grateful to the three reviewers
who provided valuable comments.

REFERENCES

Al-Khafaji, Z., Young, A., Degroff, B., Nowacki, F.,


Figure 53. Centrifuge test on suction anchor in layered Brooks, J. & Humphrey, G. 2003. Geotechnical Properties
clay/sand profiles. View after partial excavation. of the Sigsbee Escarpment from Soil Borings and Jumbo
Piston Cores. Proceedings, Offshore Technology Con-
ference, Houston, TX, May, paper #15158
15 CONCLUSIONS Angell, M., Hanson, K., Swan, B. & Youngs, R. 2003.
Probabilistic Fault Displacement Hazard Assessment for
Given the challenges presented in geohazard-prone Flowlines and Export Pipelines, Mad Dog and Atlantis
areas, operators must: Field Developments, Deepwater Gulf of Mexico.
Proceedings, Offshore Technology Conference, Houston,
Properly anticipate technological needs and TX, May, paper #15202
develop appropriate tools in a timely manner. Bingham, D., Drake, T., Hill, A. & Lott, R. 2002. The appli-
Perform in-depth geophysical and geotechnical cation of Autonomous Underwater Vehicle (AUV)
site characterization. Technology in the Oil Industry Vision and Experiences.
Perform numerous engineering analyses, quantify- TS4.4 Hydrographic Surveying, FIG XXII International
Congress, Washington D.C. USA, April 1926 2002.
ing displacements, level of slope stability, debris Brand, J., Angell, M., Lanier, D., Hanson, K., Lee, E. &
flows run-out distances and velocities. George, T. 2003. Indirect methods of dating seafloor activ-
Estimate annual probabilities of occurrences of ity: Geology, Regional Stratigraphic Markers, and Sea-
given events. floor Current Processes. Proceedings, Offshore Technology
Conference, Houston, TX, May, paper #15200
Provided that the above activities are properly Clukey, E.C., Banon, H. & Kulhawy, F. 2000. Reliability
scoped, experience has shown that the presence of steep Assessment of Deepwater Suction Caissons. Proceed-
slopes, large offset faults, and other hazards do not nec- ings, Offshore Technology Conference, Houston, TX,
essarily preclude the safe siting of offshore oil and gas May, paper #12192
infrastructures. Christian, J.T. 2003. Geotechnical engineering reliability:
Once adequately sited, the facilities must be How well do we know what we are doing? The 39th Karl
designed with increased aleatory soil uncertainties. A Terzaghi lecture, ASCE, JGGE, October, 2004.
lack of appropriate design code recommendations may Driver, D. 2005. Personal communication.
suggest model testing to verify the design methods for Gepmatrix 2002. Probabilitic Fault Displacement Analy-
sis Atlantis Prospect. Confidential report to BP.
the site specific conditions. George, R.A., Gee, L., Hill, A., Thomson, J. & Jeanjean, P.
Operating in geohazard-prone areas therefore 2002. High-Resolution AUV Surveys of the Eastern
entails making wise decisions while properly manag- Sigsbee Escarpment. Proceedings, Offshore Technology
ing increased levels of uncertainty. Great operators will Conference, Houston, TX, May, paper #14139
recognize the importance and value of geohazard risk Gilbert, R.B., Stong, S.J., Albrecht, R.S. & Dupin, R.M.
assessment studies to safely and confidently site their 1999. Optimizing Investigation Programs for Offshore
facilities. Foundations. Proc., 2nd International Conference on
Seabed Geotechnics, IBC Ltd., London, p. 32.
Jeanjean, P., Campbell, K. & Kalsnes, B. Use of Integrated
Study to Characterize the Marlin Deepwater Field. Proc.
ACKNOWLEDGEMENTS Society of Underwater Technology Conference,
September 1998, London.
The authors would like to thank BP America Inc. and Jeanjean, P., Hill, A.W. & Taylor, S. 2003. The Challenges of
BP Exploration Ltd. for permission to publish. Some of Siting Facilities along the Sigsbee Escarpment in the
the material presented herein was published at the 2003 Southern Green Canyon Area of the Gulf of Mexico.
Offshore Technology Conference. Framework for Integrated Studies. Keynote lecture,

142

Copyright 2005 Taylor & Francis Group plc, London, UK


Proceedings, Offshore Technology Conference, Houston, Niedoroda, A., Reed, C., Hatchett, L., Young, A. & Kasch, V.
TX, May, paper #15156 2003. Analysis of Past and Future Debris Flows and
Jeanjean, P., Hill, A.W. & Thomson, J. 2003b. The Case Turbidity Currents Generated by Slope Failures Along
for Confidently Siting Facilities along the Sigsbee the Sigsbee Escarpment. Proceedings, Offshore Technol-
Escarpment in the Southern Green Canyon Area of the ogy Conference, Houston, TX, May, paper #15162
Gulf of Mexico: Summary and Conclusions from Niedoroda, A., Jeanjean, P., Driver, D., Reed, C., Hatchett, L.,
Integrated Studies. Proceedings, Offshore Technology Briaud, J.-L. & Bryant, B. 2003b. Bottom Currents,
Conference, Houston, TX, May, paper #15269 Erosion Rates, and How to Use Them to Date Slope
Jiang, L., & LeBlond, P. H. 1993. Numerical modeling of Failures and Debris Flows Along the Sigsbee Escarp-
an underwater Bingham plastic mudslide and the waves ment. Proceedings, Offshore Technology Conference,
which it generates. Journal of Geophysical Research, Houston, TX, May, paper #15199.
p.1030310317. Norwegian Geotechnical Institute NGI. 2002. Atlantis Slope
Ladd, C.C. & Foott, R. 1974. New Design Procedures Stability Analyses. Confidential report to BP.
for Stability of Soft Clays. ASCE, Journal of the Nowacki, F., Solheim, E., Nadim, F., Liedtke, E. &
Geotechnical Engineering Division, Vol. 100, No. GT7, Andersen, K. 2003. Deterministic Slope Stability
pp. 763786. Analyses of the Sigsbee Escarpment. Proceedings,
Lanier, D., Berger, B., Kash, V., Bryant, B. & Young, A. Offshore Technology Conference, Houston, TX, May,
2003. Relationship between Near Seafloor Seismic paper #15160
Amplitude, Impedance, and Soil Shear Strength Orange, D., Angell, M., Brand, J., Thompson, J., Buddin, T.,
Properties and Use in Prediction of Shallow Seated Slope Williams, M., Hart, B. & Berger, B. 2003. Shallow
Failure. Proceedings, Offshore Technology Conference, Geological and Salt Tectonic Setting of the Mad Dog and
Houston, TX, May, paper #15161 Atlantis Field: Relationship Between Salt, Faults, and
Locat, J. & Lee, H.J. 2000. Submarine Landslides: Advances Seafloor Geomorphology. Proceedings, Offshore Technol-
and Challenges. Proceedings of the 8th International ogy Conference, Houston, TX, May, paper #15157
Symposium on Landslides, Cardiff, U.K., June. Orange, D., Saffer, D., Jeanjean, P., Al-Khafaji, Z., Riley, G. &
Loncke, L. 2002. Le Delta Profond Du Nil: Structure Et Humphrey, G. 2003b. Measurements and Modeling of
Evolution Depuis Le Messinien (Miocene Terminal). the Shallow Pore Pressure Regime at the Sigsbee Escarp-
Unpublished These de Doctorat de luniversite P. et M. ment: Successful Prediction of Overpressure and Ground-
Curie (Paris 6). Truthing with Borehole Measurements. Proceedings,
Lunne, T., Robertson, P.K. & Powell, J.J.M. 1997. Cone Offshore Technology Conference, Houston, TX, May,
Penetration Testing in Geotechnical Practice, Blackie paper #15201
Academic and Professional. Slowey, N., Bryant, B. & Bean, D. 2003. Sedimentation in
Melchers, R.E. 1999. Structural Reliability Analysis and the vicinity of the Sigsbee Escarpment during the last
Prediction, 2nd Edition John Wileys and Sons, New York, 25,000 years. Proceedings, Offshore Technology
N.Y. Conference, Houston, TX, May, paper #15159
Mildenhall, J. & Fowler, S. 2001 Mud Volcanoes and Spencer, E. 1967. A method of analysis of the stability of
Structural Development of Shah Deniz Journal of embankments assuming parallel inter-slice forces.
Petroleum Science and Engineering, No. 28, Pages Gotechnique, 17, No.1, pp. 1126.
189200. Schultheis, P. 1998. Personal communication.
Morgenstern, N.R. & Price, V.E. 1965. The analysis of the Teunissen, J.A.M. & Spierenburg, S.E.J. 1995. Stability of
stability of general slip surfaces. Gotechnique Vol. 15, infinite slopes. Technical Note, Gotechnique 45, No. 2,
No. 1, pp. 7993. 321323.
Nadim, F., Krunic, D. & Jeanjean, P. 2003. Reliability Method Young, A., Slowey, N., Bryant, B. & Gardner, S. 2003.
Applied to Slope Stability Problems: Estimating Annual Age dating of past slope failure events from C14 and nano-
Probabilities of Failure. Proceedings, Offshore Technology fossil analyses. Proceedings, Offshore Technology
Conference, Houston, TX, May, paper #15203 Conference, Houston, TX, May, paper #15204
Newlin, J. 2003. Suction Anchor Piles for the Na Kika FDS Youngs, R.R. 2003. Probabilistic Fault Displacement
Mooring System. Part 1: Site characterization and design. Hazard Analysis (PFDHA): Earthquake Spectra.
Proceedings of the International Symposium Deepwater
Mooring Systems, Houston, Texas, October.

143

Copyright 2005 Taylor & Francis Group plc, London, UK


Offshore site investigations: new frontiers

H.J. Kolk & J. Wegerif


Fugro Engineers BV, Leidschendam, The Netherlands

ABSTRACT: Major recent improvements in offshore site investigations are due to three drivers: (1) the shift
of the offshore exploration industry to deeper water, (2) the move to geologically complex and geohazard prone
areas, and (3) an increased demand for improved cable and pipeline route characterization. Associated examples
include (i) design of novel tools which operate efficiently, safely and provide high-quality data in deep water
settings, (ii) development of geotechnical equipment for geohazard sampling and in-situ testing, and (iii) appli-
cation of shallow geophysical data acquisition and reduction techniques for the upper few meters below the sea
floor for cable and pipeline route characterization. All three offshore investigation drivers, particularly deepwa-
ter geohazards, require multi-disciplinary teams for design and evaluating efficient and cost-effective investi-
gation programmes. This paper presents an overview of recent developments in site investigation techniques,
data reduction and management.

1 GENERAL INTRODUCTION worldwide in more than 600 m water depth. This


process is still continuing. Geophysical and geotech-
This paper gives an overview of major recent develop- nical investigations in such water depths required new
ments in offshore geotechnical and geophysical data soil investigation techniques since methods used in
acquisition and reduction techniques and organisation shallower water are either technically no longer feas-
of data acquisition projects. Previous overviews on ible and/or economically unattractive.
new developments were given by Power & Geise New facility concepts were required such as FPSOs
(1994) and by Kolk & Campbell (1997). A general (floating production and storage offshore) and subsea
overview of most offshore and nearshore geotechnical completion systems. The foundations for these facilities
and geophysical survey techniques is given in Fugro required detailed geotechnical data at shallow depth but
(2001). Not covered by these references are recent over a large area, as opposed to previous facilities (e.g.
advancements in investigation techniques. These can fixed platforms) which require deep investigations
be attributed to growth in the following areas: over a relatively small area (Fig. 1; Power et al. 1997).
Anchoring for floating systems, such as suction anchors,
deep water oil and gas exploration
geohazards (all water depths)
offshore cable and pipeline routes Area of Soil
improved ICT worldwide Investigation
fast-track design projects Coverage Deep water Site
HSE requirements. Survey and Soil Investigation
Coverage for FPSO
This paper emphasizes new developments related Development
to data acquisition and reduction techniques. It does
not address new developments in in-situ test data inter- Depth of Soil
pretation. Recent publications by Lunne et al. (1997), Site Survey Coverage Investigation Interest
Mayne (2005) and Randolph (2004) have all provided Shallow for FPSO Development
detailed information on this subject. Water Platform Site

Depth of Soil Investigation


2 DEEP WATER DRIVEN DEVELOPMENTS Interest for Platform Development

2.1 Introduction
During the past decade, a significant amount of hydro- Figure 1. Volume of soil to be investigated for deep water
carbon discoveries and field developments has occurred versus shallow water facilities.

145

Copyright 2005 Taylor & Francis Group plc, London, UK


suction embedded plate anchors (SEPLAs), vertical
loaded drag anchors (VLAs) and supports for seabed
structures such as templates and suction caissons, only
need geotechnical design parameters to a depth of a few
meters to tens of meters below seabed. Cables and
pipelines may only require soils data in the upper 23 m
below seabed. Deeper geotechnical data are required in
more confined areas for foundation design of tension
leg piles and well conductor/casings, and for geohazard
identification and analysis.
The geophysical and geotechnical survey indus-
trys response to the challenges associated with deep
water can be divided into two groups:
improvements in investigation techniques to solve
operational problems
introduction of tools for improved soil
characterization.
Figure 2. Fugro Explorer.

2.2 Operational solutions


2.2.1 Weight
Soil investigations in increased water depth imply
that greater lengths of drill string and cable for seabed
frames need to be handled. Thus, greater weights
needed to be carried by the derricks of soil investiga-
tion vessels. One measure taken to allow drilling to
greater depth with conventional soil investigation
vessels was to use aluminum rather than the conven-
tional 5 inch API steel drill string. Further measures
included increasing capacity by strengthening the
derricks, winches and cables. For the largest tradi-
tional geotechnical soil investigation vessels Bavenit
and Bucentaur, this has resulted in an increase of
Figure 3. PROD being launched (courtesy of Benthic
maximum drill string length from approximately
Geotech).
9001100 m (steel string) to 2100 m (combined alu-
minum and steel drill pipes).
An alternative solution was to use non-geotechnical
vessels for geotechnical investigations. An example of 3000 m long 5 inch API steel drill string. This is
this alternative is the work-over semi-submersible ample for investigations at currently identified deep-
Uncle John used in Gulf of Mexico surveys. This rig water sites.
can handle maximum 3000 m of 5 inch API steel drill An alternative method of drilling and coring/
string. Jeanjean et al. (2005) report on mixed experi- sampling/in-situ testing is used by the lightweight
ences with sample quality obtained from oil explor- PROD (Portable Remotely Operated Drill) unit. It is
ation and work-over semi-submersibles. Sample an 85 kN (8.5 tonnes) submerged weight seabed unit
disturbance in investigations is largely related to the which houses a self-contained coring and testing sys-
heave compensation system which is less sensitive on tem powered by an electrical umbilical from a support
these semi-submersibles in comparison to specialist vessel (Fig. 3; Randolph 2004, Kelleher & Randolph
soil investigation vessels. In addition, total costs for 2005).
such deployments are significantly higher than those The PROD unit is designed to operate in a max-
which use dedicated soil investigation vessels. imum water depth of 2000 m and to sample/test to 100 m
In view of these disadvantages, and limited avail- penetration below seabed. Maximum sample length is
ability of such rigs, the specialist deep water soil either 2.75 m (44 mm diameter liner samples) or 0.7 m
investigation vessel Fugro Explorer was built and (57 mm diameter thin walled tube samples). Standard
put in operation in 2003 (Fig. 2). This vessel has a 10 cm2 Cone Penetration Tests (CPT) or 60 mm diam-
draw works and heave motion compensator capacity eter ball probe tests (BPT) can be performed in 2.75 m
of 1500 kN (150 tonnes) which is equivalent to a long stroke pushes.

146

Copyright 2005 Taylor & Francis Group plc, London, UK


Mud/Water
Pump
250 Bar
Drill Floor

Recovery
by Wireline
Overshot

Drill Pipe

Seal Seating

Figure 4. Deepsea SEACALF.

2.2.2 Power
Traditionally, seabed equipment used for drill string Open Drill
stabilization and in-situ testing derives its energy via Bit
Cone
an umbilical cable from the investigation vessel. Such
Penetrometer
cables are cumbersome in deep water. Hence, alterna-
tives have been developed which have the energy Figure 5. WISON XP.
source on or near the seabed unit.
One option is to use an ROV (remotely operated
vehicle) for electrical power supply which plugs into situ test completion, it is then retrieved with a steel
the seabed unit. Another option is to provide this unit wireline/overshot system. Test and penetration data,
with special underwater batteries. The disadvantage of which are stored in solid state memory, are downloaded
the first option is that having an ROV available solely to a computer when the system is retrieved on deck.
for power supply is very expensive. A disadvantage of The 90 kN capacity Fugro XP (eXploration &
the second option is that the frame needs to be hoisted Production rig suited) suite of tools consists of a 1.5 m
to the vessel for exchanging or charging batteries. stroke CPT unit (WISON), a 1.5 m stroke Vane and
The Fugro Deepsea SEACALF (DSS; Fig. 4) is 1.0 m stroke (76 mm diameter) push and piston sam-
equipped with both options to provide greater flexi- plers. It can operate in a total depth of 3000 m below
bility with respect to ROV use. For very large the drill floor (Fig. 5; Power & Geise 1994). The
programmes of seabed testing the DSS can also be WISON XP is designed to push a 10 cm2 piezocone
powered from the surface vessel. The unit also has into the soil at a constant penetration rate of 20 mm/s.
thrusters on the frame to prevent it rotating during It has also been used to insert piezoprobes for assess-
lowering and hoisting. The unit can operate in up to ing in-situ pore water pressure.
6000 m water depth and weighs 90 kN (under water). Drilling fluid pressure is also used for percussion
Thus, CPTs to 40 m penetration and vane tests to sampling with the FUGRO CORER (FC; Zuidberg
20 m depth can be made in typical deepwater soils. et al. 1998). This tool was originally developed to
The traditional downhole testing/push sampling obtain 0.9 m long 54 mm diameter samples of compe-
equipment used in shallow water investigations is tent material, such as hard clays, dense/cemented
powered by hydraulic oil provided by an umbilical cable. sands and soft rocks. However, since its introduction,
It is impracticable to fabricate/use such cables more alternative configurations have been developed which
than 800 m long. Hence, downhole systems (e.g. the allow up to 4 m long 67 mm diameter samples. These
Fugro DOLPHIN and XP suite of tools) have been longer samplers are in particular suitable for strati-
developed which derive the energy for inserting the graphic and paleoclimatic studies. A further improve-
tool from drilling fluid pressure inside the drill string. ment is incorporation of an autoclave system for
The downhole tool is inserted in freefall mode until it sampling hydrates and gassy soils (Fugro Pressure
latches into a special bottom hole assembly. After in- Corer, FPC, Section 3.3.3 following).

147

Copyright 2005 Taylor & Francis Group plc, London, UK


Figure 7. Vessel Bavenit with STACOR on stinger.

Figure 6. Deepwater SEASCOUT.

2.2.3 Round trip duration


A major concern in deepwater investigations is to
reduce the time for lowering and retrieving both
seabed and downhole systems. Therefore, a signifi-
cant amount of time and money has been spent by
investigation contractors to (a) optimize underwater
weight, power pack battery endurance, etc. and (b)
further improve operational specifications and proced-
ures for tools, cables and winches. Frequently, single
line hoist light weight systems provided with high
speed winches are used for CPTs (SEASCOUT,
Fig. 6; Power & Geise 1994) and vane testing (e.g.
Fugros HALIBUT system) in order to accelerate Figure 8. Leveling Frame.
shallow seabed investigations.
An alternative approach to reduce total round trip data from the 9 m FHPC sampler which are similar to
duration is simply to reduce the number of required those from adjacent borehole push samples.
round trips. This has been successfully achieved by
increasing both sample and penetration length per 2.2.4 Sloping seabeds
in-situ test. This has led to increased use and further Many new offshore exploration sites are located on or
development of free-fall seabed piston corers such as near steep slopes (e.g. continental slope areas) where
the Jumbo Piston Corer (Young et al. 2000) and the conventional seabed frames for drilling or in-situ test-
STACOR (Borel et al. 2002) for deepwater seabed ing systems cannot be used. Hence, Fugro recently
sampling. Another measure to improve efficiency of designed and fabricated a Leveling Frame for perform-
such operations is the introduction of stingers for hand- ing investigations on up to 30 degree slopes (Fig. 8).
ling such corers on the vessel (Fig. 7). The unit was successfully used in a number of inves-
Examples of borehole samplers with increased sam- tigations in up to 800 m water depth.
ple length are the Fugro CORER discussed above and
the Fugro Hydraulic Piston Corer (FHPC). The latter
2.3 Improved soil characterization
tool is similar to the Advanced Piston Sampler (APS)
used in the Ocean Drilling Project (ODP; Zuidberg 2.3.1 Vertical profiling
et al. 1998). It shoots either a 4.5 m or 9 m long sam- The deepwater environment gives rise to a number of
pler at high speed into the soil to obtain 66 mm diame- geotechnical data acquisition challenges as a result of
ter samples. Jeanjean et al. (2005) provide sample test differences in water depth, foundation type and soil

148

Copyright 2005 Taylor & Francis Group plc, London, UK


Undrained shear strength (kPa)
0 5 10 15 20
0

0.2
Depth (m)

0.4
Box core -
0.6 Fallcone test
Box Core Lab.
Vane test
0.8
Deep water
Seascout CPTU
1

Figure 9. Comparison 33 cm2 CPT result with laboratory


test results on box core sample. Figure 10. T Bar, Ball, 33 cm2 and 15 cm3 piezocone
Penetrometers.
conditions. Major challenges and solutions provided
by industry are summarized below. SEASCOUT 33 cm2 CPT result with laboratory tests
Peuchen et al. (2005) present estimates of depth on an adjacent box core sample. In addition to a
accuracies relative to seabottom which can be achieved favourable comparison of these data, this graph also
for borehole and seabed investigation systems. The illustrates the occurrence of a stronger crust near
accuracy for seabed systems is primarily related to the seabed which has been observed at many other off-
distance the frame penetrated into the seabottom dur- shore locations (Ehlers et al. 2005).
ing installation. This value is dominated by the weight Penetration tests using T Bar (TBT) and Ball
and dimensions of the unit, the impact velocity during Penetrometer (BPT) have been recently introduced in
touch down and the seabed strength. Peuchen et al. offshore soil investigations for assessing undrained
estimate that, in general, the accuracy at seabed may shear strength of soft clays typical of the deepwater
vary from 0.2 m to 0.8 m. This value is adversely environment (Randolph 2004, Peuchen et al. 2005,
affected in deepwater due to very soft seabed soils. Lunne et al. 2005). The TBT is used in seabed based
However, this value can be favourably affected by the investigations only, whereas the BPT is performed in
geotechnical contractor by careful selection of frame both seabed based and downhole investigations. Both
weight and dimensions and by taking great care to these tests were developed as an alternative to
minimize impact velocity at touchdown. The accuracy CPT since theoretical considerations suggest that
of the depth of borehole samples and tests below undrained shear strength can more confidently be
seabed at deepwater sites is dominated by the water assessed from these tests than from CPTs (Randolph
depth. Peuchen et al. estimate that the depth accuracy 2004). Experience to date indicates that, contrary to
of a borehole sample or test close to seabed at a original belief, no unique, theoretical, correlations
1500 m water depth site is at best 4.5 m. This figure can be used to convert the results of TBTs and BPTs
can be significantly improved through correlation of into undrained shear strength. Research on this sub-
test results of downhole systems with those from ject is currently being conducted in a Joint Industry
seabed based system test results at the same location. Project carried out by NGI/COFS (Lunne et al. 2005).
Peuchen (2000) notes that seabed based systems also The vane shear test (VST) is an alternative in-situ
provide a greater accuracy of CPT resistance data test method for directly determining undrained shear
compared to downhole systems. This is due to more strength (peak and residual). It can be performed in a
reliable data reduction. seabed as well as in a downhole mode as noted earlier.
Most deepwater sites consist of clays and calcareous The in-situ tests VST, CPT, TBT and BPT each provide
oozes with undrained shear strengths in the order of different (generally higher) theoretical estimates of
those for normally consolidated clays. Sample hand- undrained strength than measured in laboratory tests.
ling and laboratory strength/deformation testing of Randolph (2004) attributes these differences largely to
these materials is very difficult. Therefore, great the significantly larger strain rates in these in-situ tests.
reliance is placed on assessing reliable soil parameters Another possibility may be that soil samples tested in
from in-situ tests. Enlarged CPT cones are occasion- the laboratory may be more disturbed than soils tested
ally used in combination with seabed testing systems to in-situ. Hence, current practice for such soft clays is to
enhance the accuracy of the measurements. Figure 9 use site specific empirical correlations for all tests to
from Borel et al. (2005) shows a comparison of a assess undrained shear strength (Fig. 11).

149

Copyright 2005 Taylor & Francis Group plc, London, UK


triggering system

weight-stand

pull-out cable

Figure 11. Interpreted undrained shear strength profiles


from various penetrometers.
corer tubes with liner

Long high quality seabed samples are obtained base plate


using freefall seabed piston corers. The two largest
commercial corers are the JPC (Jumbo Piston Corer, counter weight
Young et al. 2000), used in the Gulf of Mexico, and
the STACOR (Borel et al. 2005) used offshore West
Africa and other deepwater areas (Fig. 12).
Figure 12. STACOR layout.
Both samplers are designed to take up to 30 m
long, 100 mm diameter samples. Unlike the JPC, the
STACOR has a base plate reference system which Water depth (m)
allows the piston to remain effectively stationary at 0 200 400 600 800 1000 1200 1400 1600
the sediment surface during corer penetration (Borel 10
et al. 2002). The STACOR has been used for geo-
technical campaigns in deep water to 4000 m and also
Penetration (m)

in ultra-deepwater to a depth of approximately 6000 m 15


during scientific research cruises. Proven penetration
with the present system in deepwater is in the range
18 m to 22 m (Fig. 13). 20
Due to the truly stationary state of the piston, core Dalia Rosa Other
recoveries are generally in the order of 95% to 100%. Other Other Other

Disturbance index data, as defined by Lunne et al. 25


(1998), show that the quality of samples acquired
with a STACOR are good to very good and similar Figure 13. STACOR recovery data.
to borehole piston samples taken in the same envir-
onmental conditions. However, long piston corers 2.3.2 2-D and 3-D profiling
equipped with so-called pilot weight tripper give sig- An ultra high resolution (UHR) shallow seismic survey
nificantly lower quality samples. is required to give adequate quality bathymetric and sub-
In order to reduce disturbance due to sample hand- bottom profiling data of target offshore development
ling and transport, and to accelerate project sched- areas. Such a survey typically requires the acquisition of:
ules, all major deep water soil investigation vessels
suited have large geotechnical laboratories. These ultra high-resolution swathe bathymetry data
allow advanced strength and deformation testing dur- high and low frequency side scan sonar data
ing the field work. 3.5 kHz7.5 kHz sub-bottom profiler data.

150

Copyright 2005 Taylor & Francis Group plc, London, UK


Figure 14. Schematic diagram of AUV survey system.

Figure 15. Side scan sonar mosaic illustrating both circu-


Instead of surface towed equipment, it is now lar and shot-line orthogonal survey patterns surveyed with
becoming standard practice to use Autonomous Under- an AUV.
water Vehicles (AUVs) to obtain UHR data (Fig. 14;
Campbell & Burrell 2003).
Survey tools incorporated in AUVs typically
include 200 kHz swathe bathymetry and Full
Spectrum Chirp seafloor mapping systems compris-
ing integrated side scan sonar and sub-bottom pro-
filer. The resolution (frequency) from AUV systems
is generally similar to that of good quality sub-bottom
profiler data available from non-AUV systems.
However, a significant improvement over non-AUV
systems using an AUV is the ability to run virtually
any survey pattern (e.g. not just straight, orthogonal
lines as is the case with deep tow tools). This allows
surveys to be designed to meet the specific needs of
each investigation project, consistent with the type of Figure 16. Diagram showing sub-bottom profiler data dens-
foundation design under consideration and the local ity achieved with a conventional deep tow survey grid com-
geology. For example, a circular pattern of survey pared to the AUV micro 3-D seismic survey area.
lines can now be made to connect spread mooring
anchors in one continuous line to facilitate comparing
conditions from anchor-to-anchor, and with soil bor- design and construction of facilities. Common off-
ing locations (Fig. 15). Such AUV surveys facilitate shore geohazards include:
rotation of the mooring line directions around the gas hydrates
annulus of the survey footprint, and supply no unneces- shallow gas
sary data at/near the centre of the circle. slope instability
Campbell et al. (2005) describe another example active faulting and seismicity
consisting of micro 3-D surveys to detail local condi- mud volcanoes
tions at deepwater anchor pile locations (Fig. 16). shallow water flows
Although such geohazards are frequently encoun-
tered in deepwater areas, they have also been found in
3 GEOHAZARD DRIVEN DEVELOPMENTS shallower water. These geohazards have a number of
consequences with respect to geotechnical and geo-
3.1 Introduction physical surveys:
Geohazards are natural processes or phenomena that the complexity of these problems requires a different
may have an adverse effect on operations, siting, approach to programme definition and execution: A

151

Copyright 2005 Taylor & Francis Group plc, London, UK


multidiscipline team, including client, various
experts from academia and industry, propose such
programmes in consultation to optimize the effi-
ciency of the data acquisition phase.
geological and geotechnical data need to be defined
for a large volume of the earth. Hence, both data
collection, reduction and visualisation need to be
significantly enhanced.
at the same time, there may be a need for detailed
acquisition since critical conditions for geohazards
may occur very locally (e.g. geotechnical data at
faults or previous slip planes).
data typical for geohazard conditions may need to
be collected, such as:
hydrate samples
gas content and pressure
excess pore water pressure.

3.2 Organizational aspects


The experience and technology in relation to geohaz-
ard assessment has advanced significantly over the past
five years. It is now well accepted that the assessment
of each major potential geohazard requires a multi-
phase and multidisciplinary approach requiring the
input from experts from different backgrounds and
organisations (Jeanjean et al. 2005). Power et al. (2005)
propose a systematic approach to such assessments,
embracing the sum of this collective experience and Figure 17. AGES flow diagram.
expertise. This results in a cost-effective decision
making process regarding the risk, implications and data can all efficiently be combined in a data acquisi-
response to any given set of (potential) geohazards. tion programme.
This approach, called Advanced Geohazard Evaluation A significant trend that has been accelerating over
System (AGES), is summarised as a flow diagram in the past several years is the routine interpretation and
Fig. 17. This approach has a number of consequences integration of all survey-related site investigation data
with respect to geophysical and geotechnical data and the geotechnical logs on a single platform the
acquisition. A desk study, incorporating input from seismic workstation. This greatly facilitates quick
seismic data, regional geologic data and geotechnical access to disparate data sets, data integration, and a
information, is commonly performed to synthesize wider range of and more robust (including quantita-
available information prior to embarking on a rela- tive and semi-quantitative) data analyses, and improves
tively expensive site investigation. This minimises time visualization and understanding by the end user. We
in the field and optimises the scope of work. now can have at our fingertips in one (digital) project
Nowadays 3-D seismic exploration data, potentially in real-world coordinates all bathymetric, side-scan
enhanced by techniques such as short-offset reprocess- (including mosaics), all high-resolution 2-D and
ing and seismic inversion, has become a powerful tool exploration 3-D seismic data, core and boring logs,
for the creation of a full site 3-D geological/geotech- CPT logs, deep well logs, velocity data, and virtually
nical site model. This approach reduces uncertainties any other available and relevant data. As part of this
and the associated risks by identifying constraints for all-digital approach, working meetings of various
field development at an early stage, and allows defining experts in visualization rooms (to allow large-scale,
a cost-effective and realistic site investigation scope. interactive, 3-D visualization of data) are now becom-
This approach requires the use of a multi-discipline ing commonplace to address siting and design issues
team of experts which may include exploration geo- at geohazard sites. Thus, site investigation specialists
scientists, facilities engineers, geohazard specialists, are now able to characterize complex sites quickly and
geotechnical engineers and site investigation special- more robustly, with the end result being that one can
ists. This mix also ensures that geotechnical, geo- make better siting and foundation design decisions in
logical, geochemical, biological and environmental a shorter time frame than was possible five years ago.

152

Copyright 2005 Taylor & Francis Group plc, London, UK


Figure 19. Resolution of seismic inversion result (left)
compared to conventional processing result (right).

discontinuity is increased by using a longer sampler or


Figure 18. 4 m section of a 200 m Ormen Lange log. corer, such as the FC and FHPC, discussed in Section
2.2.2 and 2.2.3 respectively.
3.3 Improved characterization
3.3.2 2-D and 3-D profiling
3.3.1 Vertical profiling Geohazard analyses often require detailed character-
Conventional borehole sampling and testing obtains ization both in terms of soil stratigraphy as well as
data at limited locations of the vertical profile and soil parameter definition for large soil volumes. This
generally leaves gaps where soil conditions have to be is achieved from advanced analyses of combined high
estimated. Borehole wireline geophysical logging resolution seismic reflection data and seismic bore-
offers a means of providing a continuous vertical pro- hole logging data. This so-called seismic inversion
file through correlation of geophysical measurements technique is routinely used as the basis for estimation
with geological and geotechnical sample and in-situ of hydrocarbon reservoir properties from deep seis-
test data (Digby 2002). Routine geophysical logging mic data. However, its application for shallow geo-
consists of natural gamma, neutron and gamma logical and geotechnical characterization has been
density measurements in the drillpipe, and caliper rare. Hamilton et al. (2004) present examples of
resistivity, fluid temperature and full sonic wave enhanced stratigraphic characterization from seismic
determination in an open borehole. Wireline logging inversion in comparison to conventional seismic data
is particularly attractive to identify fault planes, pre- reduction (Fig. 19).
vious slip planes or other discontinuities in the verti- An early application of this inversion technique to
cal profile in case of potential slope stability failure obtain profiles of CPT cone resistance in highly vari-
along such planes because such discontinuities can able sands and silts was explored in the French research
not easily be detected in borehole sampling or pene- project GEOSIS (Nauroy et al. 1998). A more recent
tration testing. An example is shown in the logs application to estimate undrained shear strength from
obtained at Ormen Lange (Fig. 18). Such logs are seismic data is described by Jeanjean et al. (2005).
used to identify formation/unit boundaries and correl-
ate stratigraphy between drilling locations. They also 3.3.3 Gas hydrates
provide a range of typical log characteristics for each Large quantities of methane gas are present in seabed
horizon that can then be used as indicative of sediments as frozen deposits of gas hydrates. In
expected soil parameters but, perhaps most effect- seabed sediments, gas hydrates may occur in water
ively, they provide evidence of soil variability within depths exceeding 300 m, either in water temperatures
each unit. The continuous nature of each measure- approaching 0C or under high pressure (Fig. 20). The
ment thus highlights any anomalous horizons such as lower limit of gas hydrate occurrence depends on the
fault and slip planes that would be particularly worthy geothermal gradient, but is generally less than 2000 m
of sampling. below the solid surface (Kvenvolden 1993). Due to
Unfortunately, obtaining soil samples from such temperature and pressure restrictions on the forma-
discontinuities is very difficult in view of the limited tion of gas hydrates, their presence is restricted to arc-
sample length of conventional samples (i.e. generally tic regions and/or deepwater. Hydrate disassociation
about 1 m) and the poor depth control during drilling will occur with heating and/or pressure reduction.
and wireline sampling (Peuchen et al. 2005). However, Production wells through deepwater hydrate deposits
the probability of obtaining a sample containing such a can provide the required warming for disassociation.

153

Copyright 2005 Taylor & Francis Group plc, London, UK


Figure 20. Phase diagram methane hydrates.

The effects of disassociation on soil stability are com-


plex: they depend not only on hydrate concentration
and geotechnical properties of the surrounding soil,
but also on the in-situ structure of the hydrate.
Sampling therefore requires recovery of pressurized
cores since disassociation will occur with convention-
ally recovered samples.
The scientific world has been at the forefront of
studying the behaviour and impact of gas hydrates.
For this purpose two major government funded scien-
tific research programmes have been undertaken: the
internationally funded Ocean Drilling Project (ODP)
and the European HYACINTH project. Information
on these projects can be found on the websites
http://www.odp.tamu.edu/ and http://www.geotek.co.
uk/hyacinth/
The HYACINTH project uses two types of wireline
pressure corers developed within a previous European
Commission funded research project HYACE (an
acronym for HYdrate Autoclave Coring Equipment).
These corers, the Fugro Pressure Corer (FPC, a percus- Figure 21. Fugro Pressure Corer (FPC).
sion sampler based on the Fugro CORER discussed in
Section 2.2.2 above) and the HYACE Rotary Corer
(HRC) were developed in order to cope with a range of is sealed by a specially designed flapper valve. The
lithologies. Both corers were designed to be compatible HRC is designed to retain a pressure of up to 250 bar
with the drill string used on the vessel JOIDES and is suitable for sampling lithified sediments or rock.
Resolution by the Ocean Drilling Program. The pressure core which is preserved in either the
The FPC (Fig. 21) uses a mud driven percussion FPC or HRC autoclave is transferred under pressure
system to drive a core barrel into sediments up to 1 m into the laboratory pressure chamber in a so-called
ahead of the drill bit. The core diameter is 58 mm. On Catch Assembly. This consists of a technical part
completion of coring, recovery of the tool with a wire- and a geological part. Only the geological part
line pulls the core barrel into the autoclave, in which needs to be retained under pressure for study. The
the pressure is sealed by a specially designed flapper technical part can be extracted from the high pressure
valve. The FPC is currently designed to retain a pres- environment as soon as feasible in order for the equip-
sure of up to 250 bar. The percussion corer is suitable ment to be re-used for further corer deployments.
for use with unlithified sediments ranging from stiff To accurately model the potential effects of disas-
clays to sandy or gravelly material. In weaker sedi- sociation of the hydrated soils, examination of the
ments it acts like a push corer. samples must be made before decomposition. In this
The HRC (Fig. 22) uses an Inverse Moineau Motor respect, the use of MSCL (multi sensor core logging)
driven by mud circulation to rotate the cutting shoe up allows cores to be geophysically logged while still
to 1 m ahead of the roller cone bit. The core diameter is sealed and pressurized.
50 mm. On completion of coring, the recovery of the Logging of conventional one atmosphere (unpres-
corer with a wireline pulls the core barrel into an auto- surized) cores is usually done with the core oriented
clave, in a similar manner to the FPC, and the pressure horizontally. However, for logging pressurized cores it

154

Copyright 2005 Taylor & Francis Group plc, London, UK


of these reports along with a gallery of cruise photos
can be found on the website http://www.netl.doe.gov/
scngo/NaturalGas/hydrates/index.html
A description of various geotechnical techniques
for coring and testing is given in Zuidberg et al.
(1998). This includes a description of various in-situ
temperature measurement probes and presents results
from such probes obtained during a commercial
deepwater coring and testing programme.
Digby (2005) describes and compares the results
of borehole wireline and laboratory MSCL tech-
niques for identifying and characterizing hydrates.

3.3.4 Shallow gas


Many marine soil sediments contain gases which are
the result of either biological activities in the upper 5
to 10 m (biogenic gas) or leakage from sources at high
pressure and temperature at thousands of meters
depth (petrogenic gas). In both cases, methane is the
dominant gas type. Gas in offshore soils can be dis-
solved in the pore water or occur as free gas. The lat-
ter implies that soils contain pore fluid with dissolved
gas as well as gas bubbles (vacuoles). The occurrence
of dissolved gas and free gas depends on factors such
as gas concentration and solubility in pore water and
Figure 22. Hyace Rotary Corer (HRC). ambient pressure and temperature. Gas in dissolved
form has no effect on the mechanical properties of a
soil unless ambient pressures are reduced such that
is preferable to orient the core vertically. Since the large gas is exsoluted to form free gas (Sills & Wheeler
quantity of steel in the logging chambers of the HYACE 1992). Dissolved gas generally remains undetected
corers inhibits measurement of electrical resistivity and by geophysical surveys. Free gas reduces seismic
magnetic susceptibility, physical properties of the core compression wave velocities and consequently results
which can be logged are P-wave velocity and Gamma in anomalies on seismic profiles. This principle also
Density. underlies IFREMERs experimental Celerimeter
A United States Department of Energy (DOE) and (Vernant et al. 2004). This is a seabed push-in probe
industry sponsored joint industry project (JIP) methane for P-wave profiling.
hydrate research cruise in the Gulf of Mexico was suc- The traditional offshore method for assessing dis-
cessfully performed in the second quarter of 2005. The solved gas content of pore fluid is to use the BAT
primary objective of this JIP, which was initiated and probe for push-in pressurized sampling of the pore
managed by ChevronTexaco, is to develop technology fluid (Rad & Lunne 1991). In-situ pore water entering
and data to assist in the characterization of naturally the BAT container experiences a pressure reduction
occurring gas hydrates in the deepwater Gulf of upon sampling. This results in formation of free gas by
Mexico. Other objectives of this project are to better exsolution of any dissolved gas. The BAT system uses
understand how natural gas hydrates can affect seafloor direct sampling of headspace gas, which is decom-
stability, to gather data that can be used to study climate pressed on its way to a gas chromatograph on board
change, and to determine how the results of this proj- the soil investigation vessel. The degree of gas satur-
ect can be used to assess if and how gas hydrates act ation () is computed. An  value of 100% indicates
as a trapping mechanism for shallow oil or gas reser- pore water with full gas-saturation and a possibility
voirs. The expedition employed the semi-submersible for free gas. The method has some inherent empiri-
vessel Uncle John which carried a team of scientists cism. The accuracy of the BAT method reduces with
and engineers on a 35-day voyage. The above increasing pore water pressure because more pre-
HYACINTH/HYACE coring and laboratory equip- charge gas is required in the BAT container to achieve
ment was used. In addition the following laboratory sufficient gas sample and to prevent overpressure of
tests were done under pressurized conditions: X-ray the container. Current use is limited to pore pressures
logging, subsampling and penetration testing through equivalent to about 800 m water depth.
the sample retainer wall. During the expedition, Recently NGI developed the DGP as an alternative
reports and pictures were sent almost every day. A log to their BAT probe. This pore water sampler can

155

Copyright 2005 Taylor & Francis Group plc, London, UK


Figure 24. Fugro MkII piezoprobe.

Figure 23. Fugro Pore Water Sampler (PWS).

operate in up to 2000 m water depth and does not


require an umbilical. Hence, it can be installed using
the DOLPHIN or WISON XP. Fig. 23 shows a pore
water sampler (PWS) developed recently by Fugro
which allows sampling to the equivalent of about
3000 m water depth. The PWS measures external pore
water pressure and temperature as opposed to meas-
urements behind the container with the BAT and DGP
probes. Sample handling and analysis are according to
the Dissolved Limit Method (Christian and Cranston
1997). In particular, the PWS uses a decompression
chamber before start of gas analysis.
None of the BAT, DGP and the PWS provides reli- Figure 25. Piezoprobe test result.
able information on:
free gas content, since gas bubbles do not travel flow associated with high salt concentrations). Excess
freely through the porous filter into the sampling pore pressures are generally associated with a strength
container; thus gas concentration can be underesti- reduction and/or water flow which in turn may lead to
mated geohazards (e.g. slope instability) or conductor
free gas pressure in fine grained materials. well/casing drilling problems (Kay 2005).
In-situ pore pressures can be assessed during site
Hence, in order to obtain detailed information in case investigations using piezocones or piezoprobes,
of free gas one has to revert to pressurized soil sampling. depending on the consolidation characteristics of the
Free gas pressure is always greater than the pore soil to be tested. Excess pore pressures are generated
water pressure. Theoretical models for fine-grained during cone/probe penetration. These excess pressures
soils are available to estimate minimum gas pressures. should dissipate following tool insertion in order to
However, accurate estimates cannot yet be made from assess the at-rest in-situ pressure. Such dissipation
either theory or in-situ measurements. IFREMER, times are proportional to the square of the tools diam-
CERMES and Fugro are currently proposing a JIP for eter and the inverse of the coefficient of consolidation.
research in the area of improved understanding and In order to reduce the time required for full pore pres-
characterization of free gas. The latter includes sure dissipation in clays, tapered piezoprobes were
exploring new in-situ testing and laboratory testing designed in the past. These consisted of a piezometer
techniques. at the tip of an approximately 5 mm6 mm diameter,
about 150 mm200 mm long stem mounted below a
3.3.5 Excess pore water pressure 36 mm diameter cone tip. Unfortunately, measure-
Excess pore water pressure can occur in-situ as a ments at the tip of the probe are influenced by excess
result of sedimentation rates exceeding consolidation pore water pressures generated by the penetration of
rates (i.e. underconsolidation), previous geological the 36 mm cone (Whittle et al. 2001). Therefore, both
processes (e.g. faulting, tectonic movements, slope the tip section of the probe and the subsequent thicker
stability failures or mud volcanoes) or geochemical section of the tool should have a minimum diameter
or geophysical processes (e.g. due to osmosis or brine and a maximum length.

156

Copyright 2005 Taylor & Francis Group plc, London, UK


This has led to the new Fugro MkII piezoprobe
shown in Fig. 24, which was recently used in combin-
ation with the WISON XP system in various major
shallow and deepwater geohazard and foundation
investigations (e.g. Jeanjean et al. 2005). This probe
is provided with an additional piezometer element.
A typical test result is presented in Fig. 25. All Fugro
downhole systems and most seabed systems can be
mounted with the MKII piezoprobe. The probe has
been used on site investigations from specialist geo-
technical drilling vessels as well as from exploration
rigs in the Caspian Sea, Mediterranean and offshore
West Africa. To date, the tool has been successfully
used in water depths up to 750 m, test depths up to Figure 26. Survey layout with GAMBAS system.
410 m below the seafloor, and in clays with undrained
shear strengths of 700 kPa.
a stop-and-go motion device which enables the
sledge to remain stationary during shooting
sequences whilst the vessel continues sailing at up
4 SHALLOW SEABED INVESTIGATIONS
to 4.5 knots
a surface support where the towing winch, stop-and-
4.1 Introduction
go motion and data acquisition systems are installed
Significant changes have occurred in the past decade in a bottom to surface link by an umbilical.
methods of data acquisition and integration for shallow
The CRISP is an alternative system that does not
seabed investigations. These developments stem from
have a sledge and a stop-and-go system but uses a light
the large expansion in the offshore telecommunications
weight canister at or above the sea bottom instead.
cable installation business at the turn of the century. This
Compared to GAMBAS, this system is more suitable
required new geotechnical and geophysical investigation
for rough seabeds and high contrast soil strata. Seismic
tools for enhanced characterization of the upper 23 m
refraction tests in water depths in excess of 350 m are
of the seabed in order to improve cable burial assess-
not yet feasible due to limitations of the seismic source.
ment with respect to both reliability and efficiency. The
A disadvantage of any seismic refraction method is
new tools have subsequently been successfully used for
that it can not investigate layers which have a lower seis-
pipeline routing and installation design, and coastal civil
mic velocity than the overlying layer. Seabed towed
engineering works (e.g. pre-dredging surveys).
electrical resistivity systems can be used in such condi-
tions and also in deepwater beyond the reach of seismic
4.2 Geophysical investigations refraction systems (RHOBAS and REDAS; Puech &
Tuenter 2002). The variation of electrical resistivity is
Traditionally, a continuous vertical profile along a
indicative of variations in soil porosity and hence soil
selected cable route was determined using seismic
strength. However, its resolution is significantly less
reflection techniques. A disadvantage of this technique
than that of refraction systems and is only capable of
for cable burial assessment and pipeline design is that
detecting large changes in soil conditions.
it has limited resolution in the top few meters and gen-
Puech et al. (2004) report on a system for measur-
erally provides no information on the top 0.5 m to
ing surface waves in a marine environment. Based on
1.0 m. Therefore, the seismic refraction method was
the Multichannel Analysis of Surface Waves (MASW)
introduced into the offshore cable and pipeline market.
method, it is implemented from a sledge pulled over
This refraction method has two main advantages:
the seabed. Analysis of the surface wave data allows
Firstly, it can provide significantly improved resolution
determination of shear wave velocity profiles. Used in
in the top few meters (typically 0.2 m). Secondly, it
conjunction with high resolution seismic refraction,
allows determination of the compression wave velocity
the MASW method will greatly improve the effi-
per layer. This velocity profile offers a way to assess
ciency of continuous profiling surveys currently per-
variations of soil density and strength along an entire
formed for cable and pipeline route surveys.
survey route.
The original and most widely used high resolution
seismic refraction system is GAMBAS (Fig. 26; 4.3 Geotechnical investigations
Puech & Kolk 2000), which includes:
Cable surveys are performed over thousands of kilo-
a sledge towed on the seafloor carrying the seismic meters length and in up to a few kilometers water
source and pulling the hydrophone streamer depth. The CPT, complemented by a limited amount

157

Copyright 2005 Taylor & Francis Group plc, London, UK


thermal probes for determining thermal conductiv-
ity/heat loss
electrical conductivity probes for assessing corro-
sion potential.
In addition, various types of seabed soil (tube) sam-
plers and box corers are available. These are particularly
important for pipeline investigations if remoulded soil
properties need to be determined for backfill soils
(Cathie et al. 2005).

4.4 Data integration


A typical route survey project requires data reduction
and analyses of large amounts of data coming from
different sources. To reduce time, costs and errors,
data reduction and analyses for individual compon-
ents has become highly automated (Puech & Tuenter
2002). However, cross-checking of data coming from
multiple sources and ensuring compatibility between
various data sets requires a project team of geologists,
geophysicists and engineers. These staff are often
located at various vessels and offices all over the
world. Possibilities offered by current ICT (informa-
tion communication technology) allows project teams
be scattered worldwide yet connected by a virtual
Figure 27. SEAROBIN CPT and sampling unit. office environment.
The main end product of a route survey consists of
a set of charts summarizing seabed conditions, soil
stratigraphy, key soil parameters, and predicted burial
of soil sampling to establish ground truth, is the main
depth (or trench depth in case of pipelines) and plough
geotechnical method for characterizing soil at selected
tow forces along the selected route. Alternatively, pre-
locations along such routes. In view of the large number
diction and reporting of cable burial assessment and
of CPTs and the significant water depths, it is of great
pipeline trenchability is reported separately (Cathie &
importance that time required for CPT unit handling
Wintgens 2001).
operations is minimized. Therefore, purpose-built CPT
The next development is to link pipeline engineering
units have been developed which are lowered and
analysis software to (3-D bathymetric) survey software
hoisted on a single cable using a high speed winch. The
to allow an immediate preliminary assessment of free
SEAROBIN unit shown in Fig. 27 combines a 2 m
spans and bending moments of surface laid pipelines.
stroke CPT with a push sampler and a grab sampler
This integration allows pipeline engineers to optimize
(Hawkins & Markus 1998).
the route during the survey and, if necessary, extend or
Some cable and pipeline investigations are per-
modify the survey corridor while still on site.
formed using small penetration units provided with 1 to
2 cm2 size cone. Tests with such cones are called Mini
Cone Tests (MCT) as opposed to Cone Penetration Tests
(CPT) which are performed with the larger size refer- 5 DEVELOPMENTS FOR OTHER REASONS
ence penetrometers. Peuchen et al. (2005) discuss both
application and limitations of MCTs. In particular, The largest number of new investigation techniques is
potentially significant differences between surface due to the three drivers discussed above (i.e. deepwater,
MCT and CPT cone resistance values in sand are of geohazards and shallow seabed). Other improvements
great concern. The interpretation method for shallow to investigation techniques, or equipment modifica-
penetration CPTs in sand proposed by Puech & Foray tions, have arisen from a number of other reasons.
(2002) may be considered for both cone types. These include:
Seabed investigations for pipelines often include
general technological advances worldwide, allow-
additional and/or different probes. Examples include:
ing improvement of existing tools
large size (33 cm2) cones and T-bars for (very) soft increased pressure to accelerate the time required
seabeds for soil investigations and foundation design

158

Copyright 2005 Taylor & Francis Group plc, London, UK


increased HSE (Health, Safety and Environmental) environment during recent years. This paper has given
awareness in the offshore industry during the past a state-of-the-art review of new site investigation
5 years. techniques, data reduction and management. It has
shown that numerous favourable developments have
Significant technology advances have been made
occurred in response to these frontier challenges.
in industry in general with respect to electronics,
microprocessors and information technology. Such
developments have found numerous applications in
ACKNOWLEDGEMENTS
the geotechnical and geophysical survey industry by
improving existing investigation tools in order to
The authors gratefully acknowledge the assistance of
increase reliability, improve operational procedures
numerous colleagues within the industry who have
(e.g. through miniaturization) and reduce operational
provided input and assistance. Also the patience of the
costs.
local UWA ISFOG committee should not be forgotten.
Possibilities opened by ICT advances have also
allowed survey vessels being equipped with computer
and communication facilities similar to the land
based offices of clients, consultants and contractors. REFERENCES
These advances have allowed enhanced on-board data
processing and information exchange between all Borel, D., Puech, A., Dendani, H. & Ruijter, de. M. 2002.
team members, including continuous monitoring of High Quality Sampling for Deep Water Geotechnical
achievements and eventual adjustments of survey Engineering: the STACOR Experience. Ultra Deep
Engineering and Technology, Brest, France, June 1820,
programmes and/or (foundation) design concepts. 2002.
Laboratories on modern soil investigation vessels Borel, D., Puech, A., Dendani, H. & Colliat, J.L. 2005.
have expanded and currently incorporate advanced Deepwater Geotechnical Site Investigation Practice in
testing facilities. These ICT and laboratory facilities the Gulf of Guinea. ISFOG 2005; International
currently result in geotechnical and survey reports Symposium on Frontiers in Offshore Geotechnics, 1921
(which contain a significant amount of semi-final September 2005, Perth, Western Australia.
data) being released upon completion of field work. Campbell, K.J. & Burrell, R. 2003. Deepwater Development
Finally, one driver for new developments in geotech- Fast-Tracking: The Critical Role AUV Surveys Play in
nical equipment design and deployment is the increased Integrated Site Investigation and Geohazards Assessment.
Proceedings, 7th Annual Offshore West Africa Conference
focus on improving HSE in the offshore working envir- and Exhibition, sponsored by PennWell Corporation,
onment. This has led to two main avenues of improve- Windhoek, Namibia, 1113 March 2003.
ment within the offshore geotechnical industry: Campbell, K.J., Burrell, R., Kucera, M.S. & Audibert, J.
Implementation of Safety Management Systems 2005. Defining Fault Exclusion Zones at Proposed
Suction-Anchor Sites Using an AUV Micro 3D Seismic
(SMS) throughout organizations, from the office Survey. Offshore Technology Conference 25 May 2005,
to the vessels Houston, Texas, USA OTC Paper 17669.
Development and improvements in handling sys- Cathie, D.N. & Wintgens, J.-F. 2001. Pipeline Trenching
tems and operational procedures leading to safer Using Plows: Performance and Geotechnical Hazards.
deployment of drillpipes, geotechnical samplers Offshore Technology Conference, 30 April-3 May 2001,
and in-situ testing equipment. Houston, Texas, USA OTC Paper 13145.
Cathie, D.N., Jaeck, C., Ballard, J.-C. & Wintgens, J.-F. 2005.
These improvements have been associated with Pipeline Geotechnics State of the Art. ISFOG 2005:
considerable investment and increased operational International Symposium on Frontiers in Offshore
costs during the past few years. These costs should be Geotechnics, 1921 September 2005, Perth, Western
seen in a proper perspective: to the authors know- Australia.
ledge there has not been a single fatal work-related Christian, H.A. & Cranston, R.E. 1997. A Methodology for
offshore accident during the past 20 years with major Detecting Free Gas in Marine Sediments. Canadian
western offshore soil investigation companies. Geotechnical Journal 34(2): 293304.
Digby, A.J. 2002. Wireline Logging for Deepwater Geohazard
Assessment. Offshore Site Investigation and Geotechnics:
6 CONCLUSIONS Diversity and Sustainability: Proceedings of an
International Conference Held in London, UK, 2628
New frontier areas, consisting of oil and gas develop- November 2002, Society for Underwater Technology,
London: 391404.
ments in deepwater and geohazard prone areas, and Digby, A.J. 2005. Assessment and Quantification of the
enhanced number of cable and pipeline routes, have Hydrate Geohazard. Offshore Technology Conference
given great challenges to the survey industry during the 25 May 2005, Houston, Texas, USA OTC Paper 17223.
past decade. This industry has also seen a significantly Ehlers, C.J., Chen, J., Roberts, H.H. & Lee, Y.C. 2005 et al.,
increased focus on HSE in the offshore working 2005. The Origin of Near-Seafloor Crust Zones in

159

Copyright 2005 Taylor & Francis Group plc, London, UK


Deepwater. ISFOG 2005: International Symposium on Nauroy, J.F., Dubois, J.C., Colliat, J.L., Puech, A.,
Frontiers in Offshore Geotechnics, 1921 September Kervadec, J.P. & Meunier, J. 1998. GEOSIS: Integrated
2005, Perth, Western Australia. Approach of Geotechnical and Seismic Data for Offshore
Fugro, N.V. 2001. Geophysical & Geotechnical Techniques Site Investigations. In Robertson, P.K. and Mayne, P.W.
for the Investigation of Near-Seabed Soils & Rocks: a (Ed.), Geotechnical Site Characterization: Proceedings
Handbook for Non-specialists, Rev. 02. Fugro, N.V., of the First International Conference on Site Characteri-
Leidschendam, the Netherlands. zation, ISC98, Atlanta, Georgia, USA, 1922 April 1998
Gardner, T.N. 1988. The Acoustic Properties of Gassy Soil. 1: 497502. A.A. Balkema, Rotterdam.
PhD Thesis, University of Oxford. Peuchen, J. 2000. Deepwater Cone Penetration Tests. Offshore
Hamilton, I.W., Hartley, B., Angheluta, C. and Digby, A. Technology Conference, 14 May 2000, Houston, Texas,
2004. Turning High Resolution Geophysics Upside-down: USA OTC Paper 12094.
Application of Seismic Inversion to Site Investigation and Peuchen, J., Adrichem, J. & Hefer, P.A. 2005. Practice Notes
Geohazard Problems. Offshore Technology Conference on Push-in Penetrometers for Offshore Geotechnical
36 May 2004, Houston, Texas, USA OTC Paper 16096. Investigation. ISFOG 2005: International Symposium on
Hawkins, R.A. & Markus, A. 1998. New Developments in Frontiers in Offshore Geotechnics, 1921 September
Offshore Geotechnical Investigation. In Ardus, D.A. et al. 2005, Perth, Western Australia.
(Ed.), Offshore Site Investigation and Foundation Power, P.T. & Geise, J.M. 1994. Offshore Soil Investigation
Behaviour: New Frontiers: Proceedings of an Inter- Techniques and Equipment for the Next Century. In
national Conference Held in London, UK, 22/23/24 Chryssostomidis, C. et al. (Ed.), BOSS 94 Behaviour of
September 199., Society for Underwater technology Offshore Structures, Volume I, Elsevier, Oxford: 97109.
(SUT), London. Power, P.T., Orren, R.J. & Stephens, R.V. 1997. Integrated
Jeanjean, P., Hampson, K., Evans, T., Liedtke, E. & Clukey, E. Site & Route Assessments in Deepwater. Worldwide
2005. An Operators Perspective on Offshore Risk Assess- Deepwater Technologies Conference, London, 24th &
ment and Geotechnical Design in Geohazard-prone Areas. 25th February 1997.
ISFOG 2005: International Symposium on Frontiers in Power, P.T., Galavazi, M. & Wood, G. 2005. Geohazards
Offshore Geotechnics, 1921 September 2005, Perth, Need Not Be: Redefining Project Risk. Offshore
Western Australia. Technology Conference 25 May 2005, Houston, Texas,
Kay, S. 2005. Borehole Squeezing in Soft Clays. ISFOG USA OTC Paper 17634.
2005: International Symposium on Frontiers in Offshore Puech, A. & Kolk, H.J. 2000. Recent Developments in Cable
Geotechnics, 1921 September 2005, Perth, Western BAS Data Acquisition and Analysis. ICPC Plenary
Australia. Meeting, Copenhagen, May 3rd, 2000.
Kelleher, P.J. & Randolph, M.F. 2005. Seabed Geotechnical Puech, A. & Foray, P. 2002. Refined Model for Interpreting
Characterisation with a Ball Penetrometer Deployed Shallow Penetration CPTs in Sands. Offshore Technology
from the Portable Remotely Operated Drill. ISFOG Conference, 69 May 2002, Houston, Texas, USA OTC
2005: International Symposium on Frontiers in Offshore Paper 14275.
Geotechnics, 1921 September 2005, Perth, Western Puech, A. & Tuenter, H.-J. 2002. Continuous Burial
Australia. Assessment of Pipelines and Cables: a State-of-practice.
Kolk, H.J. & Campbell, K.J. 1997. Significant Developments Offshore Site Investigation and Geotechnics: Diversity
in Offshore Geosciences. In Vugts, J.H. (Ed.), BOSS 97: and Sustainability: Proceedings of an International
Behaviour of Offshore Structures 1: 340. Conference Held in London, UK, 2628 November 2002:
Kvenvolden, K.A. 1993. Gas Hydrates Geological 153162. Society for Underwater Technology, London.
Perspective and Global Change. American Geophysical Puech, A., Rivoallan, X. & Cherel, L. 2004. The Use of
Union, Reviews of Geophysics 31(2): 173187. Surface Waves in the Characterisation of Seabed
Lunne, T., Robertson, P.K. & Powell, J.J.M. 1997. Cone Sediments: Development of a MASW System for
Penetration Testing in Geotechnical Practice. Blackie Offshore Applications. Seatech Week, Brest, France,
Academic & Professional, London. Colloque Caractrisation In Situ des Fonds Marins, 21
Lunne, T., Berre, T. & Strandvik, S. 1998. Sample et 22 October 2004.
Disturbance Effects in Deep Water Soil Investigations. In Rad, N.S. & Lunne, T. 1991. Use of the BAT Probe for Shallow
Ardus, D.A. et al. (Ed.), Offshore Site Investigation and Gas Detection. Underwater Technology. 17(2): 1620.
Foundation Behaviour: New Frontiers: Proceedings of Randolph, M.F. 2004. Characterisation of Soft Sediments
an International Conference Held in London, UK, for Offshore Applications. In Viana da Fonseca, V. &
22/23/24 September 1998, Society for Underwater Mayne, P.W. (Ed.), Geotechnical and Geophysical Site
Technology (SUT), London: 199220. Characterization: Proceedings of the Second Inter-
Lunne, T., Randolph, M.F., Chung, S.F., Andersen, K.H. & national Conference on Site Characterization ISC-2, Porto,
Sjursen, M. 2005. Comparison of Cone and T-bar Factors Portugal, 1922 September 2004 1: 209232. Millpress,
in Two Onshore and One Offshore Clay Sediments. Rotterdam.
ISFOG 2005: International Symposium on Frontiers in Sills, G.C. & Wheeler, S.J. 1992. The Significance of Gas for
Offshore Geotechnics, 1921 September 2005, Perth, Offshore Operations, Continental Shelf Research, Vol. 12,
Western Australia. No. 10, pp. 12391250.
Mayne, P.W. 2005. Integrated Ground Behavior: In-situ and Vernant, A-M., Sultan, N. & Colliat, J-L. 2004. Etude des
Lab Tests. In Di Benedetto, H. et al. (Ed.), Deformation Proprits Acoustiques dun Sdiment Marin en Prsence
Characteristics of Geomaterials 2: 155177. Taylor & de Gaz. Journes AUM/AFM, Mcanique dans les
Francis, London. Sciences de la Mer, Brest, France, 2 & 3 September 2004.

160

Copyright 2005 Taylor & Francis Group plc, London, UK


Whittle, A.J., Sutabutr, T., Germaine, J.T. & Varney, A. 2001. Mexico. Offshore Technology Conference, 14 May 2000,
Prediction and Measurement of Pore Pressure Dissipation Houston, Texas, USA OTC Paper 12089.
for a Tapered Piezoprobe. Offshore Technology Conference, Zuidberg, H.M., Baardman, B., Kobayashi, I. & Tsuzuki, M.
30 April-3 May 2001, Houston, Texas, USA OTC Paper 1998. Geotechnical Techniques for Deep Water Coring
13155. and Testing Gas Hydrates. Methane Hydrates: Resources
Young, A.G., Honganen, C.D., Silva, A.J. & Bryant, W.R. in the Near Future?: Proceedings of the International
2000. Comparison of Geotechnical Properties from Large Symposium, Chiba City, Japan, October 2022.
Diameter Long Cores and Borings in Deep Water Gulf of

161

Copyright 2005 Taylor & Francis Group plc, London, UK


Deepwater developments: drag and plate anchors

Copyright 2005 Taylor & Francis Group plc, London, UK


Influence of anchor geometry and soil properties on numerical modeling
of drag anchor behavior in soft clay

R.M. Ruinen
Vryhof Anchors BV, Krimpen ad IJssel, The Netherlands

ABSTRACT: In recent years an increase in application of numerical models for offshore anchoring analysis
is seen. While this is common for suction and driven piles, for drag embedment anchors this is a relatively new
development. Various different analysis methods have been developed for drag embedment anchors, with the
anchor and / or soil being modeled either simple or complex. In this paper a new method for drag embedment
anchor analysis is presented and used to analyze the effect of using anchor models of different complexities. In
addition the effect of different values of the clay sensitivity on the anchor behavior is evaluated. It is found that
the difference between simple and complex models can be significant and that increasing the complexity of the
anchor model beyond a certain point will not provide a better result. In addition the results show that in soft
clays the soil sensitivity is an important parameter in analyzing the anchor behavior, as the shearing forces on
the anchor decrease with increasing soil sensitivity.

1 INTRODUCTION embedment travel approximately parallel to the direc-


tion of travel of the anchor. These areas are subjected to
While the numerical modeling of geotechnical behav- shearing forces from the soil. As the leading edges of
ior of (suction) piles is common practice, the modeling the anchor will disturb the soil during penetration, the
of drag embedment anchor behavior is still relatively shearing forces will not be generated by the undrained
new. In recent years advances have been made in the shear strength but by the remolded shear strength. A
way that drag embedment anchor behavior can be pre- numerical analysis will show the influence of using
dicted (Neubecker 1996, Eklund 1998). Methods that various values of soil sensitivity on the embedment
are available for the prediction of drag embedment behavior of drag anchors.
anchors are generally based on a simplified anchor
model (due to the complex geometry of modern drag
embedment anchors) and soil model. The purpose of 2 MODEL
this paper is to highlight two areas in the numerical
modeling of drag embedment anchor behavior that in 2.1 Soil conditions
the authors opinion have received little attention but For the analysis of the anchor penetration, a normally
are important for the understanding of anchor behavior. consolidated clay has been used, with an undrained
Firstly the influence of the anchor geometry on the shear strength (Su), submerged unit weight () and
behavior will be investigated. Modern drag embedment sensitivity (st) as shown in the table (Table 1).
anchors have complex geometries. However in numeri- To investigate the effect of the soil sensitivity (st) on
cal modeling this is in some case simplified so that the the anchor behavior, four different values have been used.
anchor consists of two planes (the shank and the fluke). These are st  1, st  2, st  4 and st  8, denoted s1,
With such simple models, the application points of the s2, s4 and s8 respectively in the calculations.
bearing and shearing forces will not correspond with
those encountered on an actual anchor. The results of a
number of numerical analyses will be shown of one
Table 1. Soil properties.
anchor modeled with various degrees of complexity, to
show the influence of the anchor geometry. Su 4  1.5  z* kPa
The second area that will be investigated will be the  4.0 kN/m3
influence of the soil sensitivity on the embedment st 1 to 8
behavior of anchors in soft clay. Drag embedment
anchors have large fluke and shank areas, which during * z is the depth below seabed.

165

Copyright 2005 Taylor & Francis Group plc, London, UK


Figure 1. Commercial drag embedment anchor.

Figure 2. Side view of anchor model with the fluke and


Table 2. Anchor models used in the analysis. shank consisting of 5 sections each.

Model Number of Number of


fluke sections shank section
into the seabed parallel to (the top of ) the fluke and
a1 1 1
the mooring line is chosen to arrive at the anchor
a2 2 2 shackle at a fixed angle (Neubecker 1996, Ruinen
a5 5 5 2004). For this analysis a more complex approach has
a10 10 10 been used whereby the anchor is seen as a free body in
the soil with 2 degrees of freedom in translation and 1
in rotation (Grote 1993, van Mierlo 2005). Using this
method the penetration direction is determined by the
2.2 Anchor and mooring line resistance of the soil acting on the anchor.
Calculation of the embedment trajectory of the
Commercially available drag embedment anchors anchor requires the equation of motion to be solved
(Fig. 1) can be characterized by their complex geome- for each time step (Equation 1).
try. The anchor selected for the analysis has a fluke
length of 2 m, fluke width of 3.5 m, a shank height of
1.6 m and a weight of 2000 kg in water. The fluke/
shank angle is set for use in very soft clay (50 angle). (1)
For the purpose of the analysis, the anchor has been
modeled with various degrees of complexity. Four dif-
ferent anchor models have been developed. In each
model the shearing and bearing areas have been where m  the mass of the anchor; I  moment of
selected such that they correspond as closely as possi- inertia; ax  acceleration in horizontal direction;
ble with the actual anchor. In the analysis the fluke and az  acceleration in vertical direction; a  angular
the shank have been modeled as a number of rectangu- acceleration; F  the total force acting in the hori-
lar sections. The number of sections chosen varies zontal or vertical direction and M  the moment act-
between 1 and 10. Table 2 shows the different anchor ing at the centre of mass. The bearing and shearing
model properties. A side view of the anchor model forces on the individual sections of the anchor are cal-
with 5 fluke and shank sections is shown in Figure 2. culated using Equations 2 and 3.
For the mooring line on the anchor, a studlink
chain has been selected for the analysis. The relevant (2)
characteristics of the chain are a diameter of 76 mm
and a submerged weight of 110 kg/m.
(3)

3 CALCULATION METHOD
where Fbearing  bearing force; Abearing  area of the
A number of methods developed up to now for the cal- section loaded in bearing; Nc  bearing capacity
culation of anchor penetration behavior have been factor; Su  undrained shear strength of the soil;
based on the assumption that the anchor penetrates Fshearing  the shearing force; Ashearing  area of the

166

Copyright 2005 Taylor & Francis Group plc, London, UK


Ti+dTi 2. calculate the force and angle in the mooring line at
the anchor shackle,
ds 3. from the acceleration vector, calculate the velocity
F ( vi1) and displacement (xi1) vectors by multi-
plying the acceleration vector with the time step
(t) as follows, v  vi  ai1  t and xi1 
Q x  v  t. i1
i i1
+d 4. with the displacement of the anchor know, the new
Ti
positionof the anchor is determined. The force
vector ( Fi1) is calculated for the new position of
Figure 3. Forces acting on mooring line segment ds.
the anchor using Equations 2 and 3 for all of the
anchor segments,
5. with the new force vector (Fi1) recalculate the

acceleration vector ( ai1;new) using Equation 1,
6. compare the calculated acceleration vector of
step 5 ( ai1;new) with the acceleration vector used
in step 3 ( ai1),
7. if the difference between the acceleration vectors
is small enough, go to the next time step (step 2).
Otherwise repeat steps 3 to 6 with a new predic-
tion for the acceleration vector. The new predic-
tion for the acceleration vector is calculated from
a and a
i1 i1;new using the Secant iteration method.

The results of the calculation procedure are the anchor


resistance in x and z direction, the line tension, the
anchor coordinates, velocity and acceleration for each
time step.

Figure 4. Forces acting on the anchor. 4 CALCULATION RESULTS

The calculations have been performed for 4 different


section loaded in shearing; and st  soil sensitivity. The soil profiles (s1 to s4) and 4 anchor models (a1, a2, a5
inverse catenary of the mooring line was calculated and a10), to investigate the effects of different anchor
by solving the differential equations (Equations 4 and model complexities and that of the soil sensitivity.
5) for the embedded mooring line length (Degenkamp Figures 5 and 6 show the results for the 4 different
1989). anchor models (a1, a2, a5 and a10) embedded in the
soil with a sensitivity of 1 (s1). The results of the cal-
culation show an increase in anchor resistance and
(4) penetration depth for increasing anchor model com-
plexity. The difference between a 5 part model and 1
is significant, while there is only a small difference
(5) between a 5 part model and a 10 part model.
When the anchor resistance generated at a certain
depth is compared, the 1 part model (a1) generates
where T  tension in the mooring line; s  length of approximately 90% to 80% of the resistance of the
embedded mooring line; F  resistance tangential to 5 part model at the same depth. With deeper pene-
mooring line; w  submerged weight of mooring tration, the difference increases. The 2 part model
line;   angle of mooring line; and Q  resistance generates approximately 95% of the resistance of the
normal to mooring line. The forces acting on a section 5 part model, while for the 10 part model this is
of mooring line ds are shown in Figure 3. Figure 4 approximately 101%. The results are summarized
shows the forces acting on the anchor for 1 fluke seg- in Table 3.
ment and 1 shank segment. Figures 7 and 8 show the results for the 5 part
The procedure for calculating the embedment trajec- anchor model embedded in the 4 different soil profiles
tory of the anchor using the above method is as follows: (s1, s2, s3, s4). The results of the calculation show a
1. start with an acceleration vector (a ) set to 0,
i1 decrease in anchor resistance and penetration depth

167

Copyright 2005 Taylor & Francis Group plc, London, UK


1100 Table 3. Anchor resistance at penetration depth for the
4 anchor models relative to model a5 in soil profile s1.
1000
Depth (m) s1 a1 (%) s1 a2 (%) s1 a5 (%) s1 a10 (%)
900
2.5 88.5 95.4 100.0 101.1
5 85.2 95.1 100.0 100.8
800 7.5 83.7 94.9 100.0 100.6
10 82.9 94.8 100.0 100.6
anchor resistance [kN]

700 12.5 82.4 94.7 100.0 100.5


15 82.1 94.6 100.0 100.5
600 17.5 81.8 94.6 100.0 100.5
20 81.6 94.5 100.0 100.5
500 22.5 81.5 94.5 100.0 100.5
25 81.4 94.5 100.0 100.5
400 27.5 81.3 94.5 100.0 100.5

300

200
1100
100
1000
0
0 10 20 30 40 50 60 900
horizontal displacement [m]
800
s1 a1 s1 a2 s1 a5 s1 a10
anchor resistance [kN]

700
Figure 5. Anchor resistance for the 4 anchor models in soil
profile 1. 600

500
horizontal displacement [m]
0 10 20 30 40 50 60 400
0
300
5
200

100
10
penetration depth [m]

0
0 10 20 30 40 50 60
15
horizontal displacement [m]
s1 a5 s2 a5 s3 a5 s4 a5
20
Figure 7. Anchor resistance for the 5 part model in the 4 soil
profiles.
25

30 for increasing soil sensitivity. When the anchor resist-


ance generated at a certain depth is compared, the soil
profile s2 (sensitivity  2) generates approximately
35 70% of the resistance of profile s1 (sensitivity  1).
s1 a1 s1 a2 s1 a5 s1 a10 For profiles s3 and s4, the generated anchor resist-
ance drops to 55% and 45% of the resistance of pro-
Figure 6. Anchor penetration depth for the 4 soil models in file 1 respectively. The results are summarized in
soil profile 1. Table 4.

168

Copyright 2005 Taylor & Francis Group plc, London, UK


horizontal displacement [m] on the embedment trajectory and anchor resistance.
From the results of the analysis the following is
0 10 20 30 40 50 60
0 concluded:
Increasing the complexity of the anchor model will
result in deeper anchor penetration and a higher
5 anchor resistance at a certain penetration depth.
The difference between a simple (1 part) model and
a more complex model (5 parts) can be up to 20%.
10 With the analyzed anchor models, the difference
between the 5 part model and the 10 part model
penetration depth [m]

was approximately 1%. With the computation time


15 increasing with increasing anchor complexity, the
difference between the 5 part and 10 part models is
significantly small to justify the use of the 5 part
model in the analyzed soil conditions.
20
The difference between the simple and complex
models increases with penetration depth.
An increased sensitivity of the soil results in a
25 shallower penetration and lower anchor resistance
together with a longer drag length. The effect is
similar for anchor models analyzed.
30 With increasing soil sensitivity, the reduction in
anchor resistance reduces relative to the soil with
sensitivity 1. This is due to the bearing resistance
35 starting to dominate the total anchor resistance
s1 a5 s2 a5 s3 a5 s4 a5 with increased soil sensitivity.

Figure 8. Anchor penetration depth model a5 in the 4 soil


profiles. REFERENCES

Table 4. Anchor resistance at penetration depth relative to Degenkamp, G. & Dutta, A. 1989. Soil Resistances to
profile s1. Embedded Anchor Chain in Soft Clay. Journal of Geo-
technical Engineering Vol 115 (10): 14201438.
Depth (m) s1 a5 (%) s2 a5 (%) s3 a5 (%) s4 a5 (%) Eklund, T. & Strm, P.J. 1998. DIGIN Users Manual ver. 5.3.
DNV Report no. 96-3637, rev. 03.
2.5 100.0 67.9 50.3 41.0 Grote, B.J.H. 1993. Simulation of kinematic behavior of
5 100.0 69.0 52.6 44.2 workanchors (continuation). Thesis. Technical University
7.5 100.0 69.7 53.7 45.5 of Delft.
10 100.0 70.1 54.4 46.3 van Mierlo, R. 2005. Anchor Trajectory Modeling. Thesis.
12.5 100.0 70.4 54.7 46.7 Technical University of Delft.
15 100.0 70.5 55.0 47.1 Neubecker, S.R. & Randolph, M.F. 1996. The performance
17.5 100.0 70.7 55.2 47.3 of drag anchor and chain systems in cohesive soil. Marine
20 100.0 70.7 55.4 47.5 Georesources and Geotechnology, Vol. 14: 7796.
22.5 100.0 70.8 55.5 47.6 Ruinen, R.M. 2004. Penetration Analysis of Drag Embedment
25 100.0 70.9 55.6 47.7 Anchors in Soft Clays. Proceedings 14th (2004) Interna-
27.5 100.0 70.9 55.6 47.8 tional Offshore and Polar Engineering Conference,
ISOPE, Toulon, Vol. II.

5 CONCLUSIONS

An analysis has been performed on the effect of various


complexities of anchor modeling and soil sensitivities

169

Copyright 2005 Taylor & Francis Group plc, London, UK


The effect of interface friction on the performance of drag-in plate anchors

S. Elkhatib & M.F. Randolph


Centre for Offshore Foundation Systems, The University of Western Australia, Perth, Australia

ABSTRACT: Finite element analysis has been used to investigate the effect of soil/anchor interface friction
on the performance of a drag-in plate anchor. Two different plate aspect ratios were considered. The plate was
subjected to combined vertical, horizontal and moment loading until plastic failure was observed. The loads at
failure were combined to form plastic yield envelopes. Using the associated flow rule and the condition of normal-
ity, anchor displacements were calculated and incorporated into an incremental procedure that enabled deter-
mination of the anchors drag trajectory. The embedment paths and installation loads were compared for friction
ratios of 0.4 and 1. The anchors with lower friction ratios achieved deeper embedment and hence higher holding
capacities.

1 INTRODUCTION installation line mooring line

The use of floating, production, offloading and stor-


fluke
age (FPSO) facilities has increased in recent times padeye
anchor
due to offshore developments moving into deeper loaded
shank normally
waters. Accordingly, the use of anchoring systems has
increased. Several anchoring options are available to (a) Installation mode (b) Normal mode
designers (Ehlers et al. 2004). Plate anchors, however,
provide a cheaper alternative due to their simpler Figure 1. Installation and loading modes of a Vryhof
installation process. There are several plate anchoring Stevmanta (Vryhof 1999).
systems available, such as drag-in plates, suction
embedded plates and dynamically embedded plates.
Essentially, they behave in the same manner during vertical, horizontal and moment loading during their
operating conditions, but vary in their installation installation and triggering process and it is necessary
process. This paper focuses on the behaviour of drag- to understand their behaviour when subjected to com-
in plate anchors during their installation process and bined loads.
their normal loading. The use of plasticity concepts and yield envelopes
Drag-in plate anchors are installed in the same for the design of foundations subjected to combined
manner as conventional anchors, by dragging into pos- loads has increased significantly over the last few years.
ition, as illustrated in Figure 1. Once the target instal- Bransby & ONeill (1999) proposed a method using
lation load is reached, the anchor is triggered into its plastic yield envelopes obtained from finite element
mooring position, where the loading is normal to the (FE) analyses of an anchor fluke. The method is based
plate (also shown in Figure 1). This mobilises the great- on the assumption that, as an anchor is dragged through
est soil resistance and hence increases the anchors soft undrained soils, failure of the soil around the
holding capacity significantly. The holding capacity anchor consists of localised plastic flow. This paper
of a drag-in plate anchor is directly related to the presents results from a FE study investigating the
depth it reaches during installation. Hence, accurate behaviour of soil surrounding an anchor plate. Since
predictions of the ultimate depth are crucial for design- shanks of drag-in plate anchors are relatively small in
ing an anchoring system. comparison to the flukes, only the fluke was modelled
Numerical studies of plate anchors have been in the FE analyses. The fluke was subjected to verti-
carried out by Rowe & Davis (1982), Merrifield et al cal, V, horizontal, H, and moment, M, loading to char-
(2001, 2003) and Salgado et al (2004). These studies, acterise a plastic yield envelope. The anchors trajectory
however, only considered the normal capacities of was then determined using an approach proposed by
plates. Drag-in plate anchors are subjected to combined Bransby & ONeill (1999), which is described in detail

171

Copyright 2005 Taylor & Francis Group plc, London, UK


h Ta
a
Vmax v
V1
v

V1
Fn
Fs
h
V(v)
H1 Displacement
M() vectors
H1 Hmax
H(h)
Figure 2. Displacement vector at loads V1 and H1.
Figure 3. Force system adopted by Bransby & ONeill
(1999).
in later sections. The effect of interface roughness
between the plate and the soil, and how it influences
the embedment characteristics of a drag anchor, was
examined. Two different plate aspect ratios were con-
sidered in the study.

2 PLASTICITY CONCEPTS

Bransby & ONeill (1999) used plasticity concepts to


analyse the behaviour of a drag anchor. FE analyses
of an anchor fluke were carried out where the anchor
was subjected to V, H and M loading until plastic failure
of the soil surrounding the fluke was observed. The
limit values were used to characterize a yield envelope 3L
in VHM space.
The plastic yield locus expresses the combination
of V, H and M loads that result in failure of the soil.
The locus can be expressed as a function of V, H and
M as follows.
6L
(1) Figure 4. The FE mesh used in the analyses.

The yield locus also enables determination of the rela-


tive values of vertical, v, horizontal, h, and rotational, 4. Calculate V, H and M using static equilibrium and
 , displacements at failure. Chen (1975) demon- check if they lie on the yield locus. If not, then
strated that if the soil surrounding the plate satisfied adjust Ta until they do.
the condition of normality (i.e. plastic flow with no 5. Use normality to determine the direction of travel
change in volume), then normality would apply for and relative rotation of the anchor.
the overall yield locus. Figure 2 illustrates the concept 6. Assume an incremental displacement, h, parallel
of normality. Consider a yield envelope in VH space. to the fluke face, and calculate v and  .
If a fluke is subjected to a combination of loads, V1 and 7. Move the anchor to the next position and repeat
H1, the displacement of the fluke, , will be normal to the procedure.
the yield envelope at these values.
Bransby & ONeill (1999) suggested the following
steps in determining an anchors trajectory. The force 3 FINITE ELEMENT MODELLING
system adopted in their analysis is shown in Figure 3.
3.1 Geometry and material properties
1. Assume an initial orientation and anchor depth.
2. Calculate the normal and sliding shank resist- Finite element analyses of a deeply buried plate were
ances, Fn and Fs. carried out using the program ABAQUS. The FE mesh
3. Assume a padeye chain tension, Ta, and calculate used in the plane strain analyses is shown in Figure 4.
a using Neubecker & Randolph (1995). It consists of first order reduced integration quadrilateral

172

Copyright 2005 Taylor & Francis Group plc, London, UK


v failure load Table 1. Maximum capacities with varying  for L/t  7.

Friction coefficient, 

Load
t h 0 0.2 0.4 0.6 0.8 1

L Vmax/Lsu 11.65 11.72 11.78 11.84 11.89 11.93


Displacement Hmax/Lsu 2.47 2.93 3.38 3.82 4.26 4.65
Mmax/L2su 1.44 1.51 1.55 1.59 1.61 1.63
Figure 5. Displacement probes used in the FE analyses.

elements. Due to the very small plate thickness, it was Table 2. Maximum capacities with varying  for L/t  20.
necessary to have a very fine mesh near the plate ends.
Friction coefficient, 
This was of particular importance when contact model-
ling was carried out and nodes at the plate/soil interface 0 0.2 0.4 0.6 0.8 1
were allowed to slide with respect to each other.
The soil was modelled as homogeneous and lin- Vmax/Lsu 11.54 11.56 11.58 11.60 11.61 11.62
early elastic, perfectly plastic, with failure determined Hmax/Lsu 1.13 1.55 1.97 2.38 2.78 3.19
using a Tresca yield criterion. A stiffness value of E/su Mmax/L2su 1.41 1.48 1.53 1.57 1.59 1.59
of 500 was used, where E is Youngs modulus and su is
the undrained shear strength. Poissons ratio was taken
as 0.49. A non-zero stress boundary condition was Tangential contact was modelled by specifying a max-
imposed on the top of the model. imum allowable shear stress, , between the soil and
The plate was modelled essentially as a rigid body, plate surfaces equivalent to su, where  is the friction
with Youngs modulus 107 times that of the soil and coefficient and su is the undrained shear strength of the
Poissons ratio of 0.15. Two different plate aspect ratios soil. When the shear stress between the surfaces reaches
(ratio of plate length to plate thickness) were considered: the maximum allowable value, the soil nodes slide along
L/t  7 and L/t  20. the plate. The friction coefficient was varied between
0 and 1 to model fully smooth to fully rough conditions
3.2 Displacement probes respectively.

Displacement controlled analyses were carried out


whereby all the plate nodes were displaced in a par- 4 RESULTS
ticular direction. The dimensions and displacement
directions of the plate are illustrated in Figure 5. In all 4.1 Maximum load capacities
cases, the displacement of the plate was continued until The maximum capacities under uniaxial loading
a plastic failure load was attained. At this stage, any obtained from the FE analyses for the two plate aspect
further increase in displacement does not result in a ratios and for  varying from 0 to 1 are presented in
load increase. Tables 1 and 2.
Different loading conditions were imposed to obtain The FE values can be compared with theoretical
a complete envelope in three dimensions. These are as predictions based on either rigorous upper bounds
follows. (for V and M loading) or limit analysis (for H load-
uniaxial loading, where the plate is subjected to only ing). ONeill et al. (2003) proposed the following
one component of loading to determine the max- relationships:
imum capacities Vmax, Hmax and Mmax;
two-dimensional loading, where the plate is sub- (2)
jected to combined VH, VM and HM loading and
the third component is kept at zero;
multiaxial loading, where the plate is subjected to
combined HM loading at fixed ratios of V/Vmax. (3)
All loads are presented as normalised values, V/Lsu,
H/Lsu and M/L2su.
(4)
3.3 Interface friction
Element faces were used to define the interacting soil Equations 2 & 4 are based on the simple kinemat-
and plate surfaces an existing feature in ABAQUS. ically admissible upper bound mechanisms illustrated in

173

Copyright 2005 Taylor & Francis Group plc, London, UK


12.2 L/t=7, Eq. 2 L/t=7, FE
Mmax L/t=20, FE
Vmax
11.7

V/Lsu
11.2
L/t=20, Eq. 2
Figure 6. UB mechanisms for V and M loading.
Rowe (1978) Merifield et al (2001)
10.7 LB and UB LB and UB
Figure 6. By equating the external rate of work done
by the anchor to the power dissipated by sliding along
the discontinuities between adjacent blocks, it is pos- 10.2
sible to obtain upper bound (UB) values for the nor- 0 0.25 0.5 0.75 1
malised loads. Equation 3 for the horizontal capacity Friction coefficient,
of the plate is based on a friction of su on the top and
bottom surfaces of the plate and a bearing capacity of 4.8 L/t=7, Eq. 3 L/t=7, FE
Ncsu at the plate ends. Nc is taken as 7.5 for plane strain L/t=7, Eq. 3 (Nc = 7.5)
conditions. (Nc = 9.1)
Rowe (1978) reported lower and upper bound solu- 3.6
tions based on FE analyses of a deeply buried strip
anchor of 10.28 and 11.42 respectively. These are for
an infinitely thin plate and are therefore independent
H/Lsu

of anchor roughness since no sliding is accounted for 2.4


at the plate ends during normal loading. Merifield et
al. (2001) also carried out a two-dimensional FE L/t=20, Eq. 3
study of an infinitely thin plate and examined sev- 1.2 (Nc = 11.7)
eral factors that influence capacity, such as embed- L/t=20, Eq. 3
ment ratio and overburden stress. They report limit (Nc = 7.5)
L/t=20, FE
values of 11.16 and 11.86 for a deeply buried plate 0
with a rough interface. The values obtained by 0 0.25 0.5 0.75 1
Rowe (1978) and Merifield et al (2001) are shown in
Friction coefficient,
Figure 7.
For horizontal loading of the plates, the discrepan- 1.7 L/t=20, FE
cies between the FE results and the limit analysis cal-
culations are considerably larger, especially for the L/t=7, Eq. 4 L/t=20, Eq. 4
thinner plate. This indicates that a value of 7.5 for Nc
1.6
is inappropriate for deeply buried plates. For L/t  7,
a better fit is obtained with Nc  9.1, while for the
M/L2su

thinner plate, Nc  11.7 provides a good fit.


1.5
As shown in Figure 7, there is a significant differ-
ence between the FE values and the UB calculations L/t=20, Eq. 5
for moment loading of the plate. The FE results
show that as  increases, the moment capacity also 1.4
increases, whereas the UB calculations (based on L/t=7, FE L/t=7, Eq. 5
the mechanism shown in Figure 6) do not depend on
. Equation 4 is based on a rotational scoop mecha- 1.3
nism where the failure surface is a circle centred at 0 0.25 0.5 0.75 1
Friction coefficient,
the plate centre with a radius equal to half the plate
diagonal.
Figure 7. Effect of interface friction on V, H and M loading.
The plastic failure mechanisms observed in ABAQUS
vary for different values of friction. For   1 (i.e.
rough conditions), the mechanism matches the one case) and   0.4 are shown in Figure 8. They consist
illustrated in Figure 6. However, for lower values of , of rotational scoops at the top and bottom surfaces,
sliding occurs on the plate surfaces and alters the centre with radius R and two scoops at the plate ends with
of the scoop. The mechanisms for   0 (frictionless radius r.

174

Copyright 2005 Taylor & Francis Group plc, London, UK


1.8
V=0 V=0

M/L2su
1.2
r r
R 0.6
R
V = 0.9Vmax V = 0.9Vmax
0
0 0.7 1.4 2.1 0 1.1 2.2 3.3
H/Lsu H/Lsu
(a) = 0 (b) = 0.4 (a) = 0.4 (b) = 1
Figure 8. Plastic failure mechanisms from ABAQUS for
Figure 10. Combined loading diagrams for L/t  20.
moment loading with   0 and 0.4.

Table 3. Constant values for different plate aspect ratios.


1.8 L/t  7 L/t  20
V=0 V=0

1.2
M/L2su

  0.4 1   0.4 1


0.6 Vmax/Lsu 11.78 11.93 11.58 11.62
V = 0.9Vmax V = 0.9Vmax Hmax/Lsu 3.38 4.65 1.97 3.19
0 Mmax/L2su 1.55 1.63 1.53 1.59
0 1.2 2.4 3.6 0 1.6 3.2 4.8 m 2.58 1.27 1.52 1.14
H/Lsu H/Lsu n 3.74 3.46 5.31 4.92
(a) = 0.4 (b) = 1 p 1.09 1.03 1.01 1.00
q 1.74 3.23 2.75 3.39
Figure 9. Combined loading diagrams for L/t  7.

5 ANCHOR INSTALLATION ANALYSIS


As  increases, the centre of the scoops approaches
the plate centre, resulting in the mechanism for   1. 5.1 Equation fitting
The maximum normalised moment capacity based on
An equation of the following form can be fit to the
the mechanisms observed in the FE analyses can be
surfaces, as suggested by Bransby & ONeill (2000).
calculated using the following equation.

(5) (6)

where R, r,  and  are as illustrated in Figure 8. The The ultimate loads Vmax, Hmax and Mmax were obtained
capacities obtained using Equation 5 for both plate from the FE analyses carried out, while the exponents
aspect ratios are shown in Figure 7. Optimal values of m, n, p and q were determined using a least squares
R, r,  and  were adopted to minimise the moment regression fit to the envelopes. Once these values had
capacity, and these agreed well with mechanisms from been determined, the incremental procedure described
the ABAQUS analyses. earlier can be used to calculate an anchors drag path.
The constant values for both plate aspect ratios are
shown in Table 3 for   0.4 and   1.
4.2 Combined loading
Interaction curves for combined loading of the thick 5.2 Performance comparison
plate (i.e. L/t  7) are shown in Figure 9 for   0.4
and   1. The curves are presented in HM space for In order to assess the performance of anchors with dif-
different ratios of V/Vmax (0, 0.2, 0.4, 0.6, 0.8 and 0.9). ferent soil friction ratios, an example is presented here.
As expected, the size of the envelopes reduces as V Consider a 10.5 m2 plate anchor (L  3.5 m) being
increases, but the shapes remain similar. installed in soft clay with a shear strength gradient of
A corresponding set of curves is shown in Figure 10 1.2 kPa/m. Different plate thicknesses are considered,
for the thinner plate (i.e. L/t  20). The same trends both with a fluke-shank angle of 55. The shank dimen-
are observed as for L/t  7. sions were a length of 4.2 m and width of 0.5 m.

175

Copyright 2005 Taylor & Francis Group plc, London, UK


Drag Dist. (x L) Drag Dist. (x L) 8 Ta/Asu, = 1 Ta/Asu, = 1 12

Norm. Line Load, Ta/Asu


0 10 20 30 0 10 20 30

Line Load, Ta (MN)


0
Fluke Depth (x L)

6 Ta, = 0.4 9

10 4 Ta, = 1
6
= 0.4 Ta/Asu, = 0.4
20 = 0.4 Ta/Asu, = 0.4
2 3
=1 =1
30 Ta, = 1 Ta, = 0.4

(a) L/t = 7 (b) L/t = 20 0 0


0 10 20 30 0 10 20 30
Drag Dist. (x L) Drag Dist. (x L)
Figure 11. Comparison of drag paths.
(a) L/t = 7 (b) L/t = 20

Figure 11 shows the drag paths calculated for all 4 Figure 12. Comparison of line loads.
cases. For L/t  7, the anchor embedment increases
by 75% when  is reduced to 0.4. For L/t  20, the
increase in embedment depth is not as significant at 1 V/Vmax, = 1 H/Hmax, = 0.4
approximately 19%.

Norm. Fluke Loads


The holding capacity, F, of an anchor when loaded 0.75
normally can be calculated using the following
H/Hmax, = 1
equation. 0.5 V/Vmax, = 0.4
V/Vmax, = 1
(7) H/Hmax, = 1
0.25
H/Hmax, = 0.4 V/Vmax, = 0.4
where A is the surface area of the anchor; Nv is a 0
bearing capacity factor (which is equal to Vmax/Lsu 0 10 20 30 0 10 20 30
given in Table 3); and su is the undrained shear strength Drag Dist. (x L) Drag Dist. (x L)
of the soil at the mid-fluke depth. Since the holding (a) L/t = 7 (b) L/t = 20
capacity of an anchor is directly proportional to su, and
hence the embedment depth, the increased embedment Figure 13. Comparison of normalised fluke loads.
increases the holding capacity of the anchor.
The installation loads are plotted against drag dis- the chain angle, a. The V/Vmax ratios are higher than
tance for all 4 cases in Figure 12. As expected, the would be expected for drag anchors since the general
load increases with increasing drag distance. The nor- understanding is that the anchors travel parallel to
malised load (Ta/Asu), also shown in Figure 12, their fluke, and are primarily subjected to horizontal
reaches a plateau following a drag distance of 10 loads. However, the high fluke-shank angle contributes
fluke lengths. This indicates that the anchor has to the high vertical loads obtained.
reached a steady state whereby the chain angle rela-
tive to the fluke (i.e. a  in Fig. 3) remains con-
stant as the anchor embeds. In all cases this value is 6 CONCLUSIONS
approximately 67.
The fluke loads V and H are presented in Figure 13. Finite element analyses have been carried out on plates
The loads are normalised with respect to their max- with different aspect ratios subjected to combined load-
imum values. For L/t  7, little difference is observed ing. The primary purpose of the study was to investi-
between the two friction ratios. The final normalised gate the effects of soil/plate interface friction on the
values, V/Vmax, were 0.82 and 0.89 for   0.4 and performance of drag-in plate anchors. In particular,
  1 respectively, while the horizontal values were the drag embedment paths were examined. The load
0.65 and 0.69. In both cases, the fluke is subjected to capacities obtained from the FE analyses were com-
higher vertical than horizontal loads. For L/t  20, bined to form plastic yield envelopes which were then
the differences are more significant between the cases. incorporated into an incremental procedure that enabled
The vertical loads are 0.71 and 0.87 for   0.4 and determination of an anchors drag trajectory. This
  1 respectively, while H/Hmax values were 0.91 and procedure made use of the associated flow rule and
0.80. M/Mmax is not presented as it is very close to zero, normality.
and hence resulting in the small rotations observed in The results obtained showed that lower friction
the embedment profiles. values achieved increases in embedment depth for
For all cases, V/Vmax increases and H/Hmax decreases both plate aspect ratios considered. Consequently, this
over the course of the drag, indicating an increase in resulted in higher performance ratios for the anchors

176

Copyright 2005 Taylor & Francis Group plc, London, UK


with lower friction, suggesting that anchors with a Merifield, R.S., Lyamin, A., Sloan, S.W. & Yu, H.S. 2003.
smooth finish such as paint would be more efficient Three-dimensional lower bound solutions for stability of
than those with untreated surfaces. plate anchors in clay. Journal of Geotechnical and
Geoenviromental Engineering, ASCE, 129(3): 243253.
Neubecker, S.R. & Randolph, M.F. 1995. Profile and fric-
tional capacity of embedded anchor chain, J. Geo-
REFERENCES technical Eng. Div., ASCE, 121(11): 787803.
ONeill, M.P., Bransby, M.F. & Randolph, M.F. 2003. Drag
Bransby, M.F. & ONeill, M.P. 1999. Drag anchor fluke-soil anchor fluke-soil interaction in clays. Canadian Geo-
interaction in clays. NUMOG VII International Symposium technical Journal 40(1): 7894.
on Numerical Models in Geomechanics. Graz, Austria. Rowe, R.K. 1978. Soil structure interaction analysis and its
Balkema: 489494. application to the prediction of anchor behaviour. PhD
Chen, W.F. 1975. Limit analysis and soil plasticity. Elsevier, Thesis. University of Sydney, Australia.
N.Y. Rowe, R.K. & Davis, E.H. 1982. The behaviour of anchor
Ehlers, C.J., Young, A.G. & Chen, J.H. 2004. Technology plates in clay. Gotechnique 32(1): 923.
assessment of deepwater anchors. Proc. Annual Offshore Salgado, R., Lyamin, A.V., Sloan, S.W. & Yu, H.S. 2004.
Technology Conf., Houston, Paper OTC 16840. Two- and three-dimensional bearing capacity of founda-
Merifield, R.S., Sloan, S.W. & Yu, H.S. 2001. Stability of tions in clay. Gotechnique 54(5): 297306.
plate anchors in undrained clay. Gotechnique 51(2): Vryhof Anchors 1999. Anchor Manual 2000. Krimpen ad
141153. Yssel, the Netherlands.

177

Copyright 2005 Taylor & Francis Group plc, London, UK


Proposed upper bound analysis for drag embedment anchors in soft clay

C.P. Aubeny, B.M. Kim & J.D. Murff


Texas A&M University, Texas, USA

ABSTRACT: This paper presents an upper bound plastic limit analysis for predicting drag anchor trajectory
and load capacity in soft clay. The shank and fluke of the anchor are idealized as simple plates. The failure
mechanism involves the motion of the anchor about a center of rotation, the coordinates of which are systemat-
ically optimized to determine the minimum load at the shackle. For a given anchor orientation, the direction of
the shackle force is varied to establish a relationship between the magnitude and direction of the shackle load.
Coupling this relationship to the Neubecker-Randolph anchor line solution produces a unique solution for the
magnitude and orientation of the shackle force. The anchor is then advanced a small increment about the optimum
center of rotation and the process is repeated. The proposed method provides a practical means to determine the
trajectory of the anchor and the anchor load capacity at any point in the trajectory.

1 INTRODUCTION involved; hence, it is likely to provide a more promising


framework for prediction of anchor trajectories than
Drag anchors are an attractive option for anchoring limit equilibrium approaches.
moorings in deep water due to their relatively low This paper presents a proposed analysis for predict-
installation cost, the extensive experience base with ing anchor trajectory using an upper bound approach
temporary moorings, and their relatively high holding from plasticity theory. A virtual work formulation is
capacities even in soft clay conditions. Further, they can employed to compute the anchor collapse load and
be easily removed and reused on other projects. How- anchor velocity at each increment of the trajectory cal-
ever, prediction of anchor capacity presents a signifi- culation. The coordinates of the instantaneous center of
cant challenge, largely because of the uncertainty in rotation of the anchor are used as optimization vari-
predicting the anchor trajectory during drag embed- ables in seeking a least upper bound collapse load. The
ment, a process that is strongly influenced by soil con- method is intended for use as a routine design tool, and
ditions, anchor geometry, and characteristics of the calculations can be performed in personal computer
anchor line. Due to this complexity, the offshore indus- spreadsheet format. Although a number of simplifying
try has largely relied on empirical methods to predict assumptions are made in some instances, the frame-
anchor penetration depth and load capacity (e.g., Naval work presented herein was developed with a view
Civil Engineering Laboratory 1987). However, empir- toward accommodating a fairly wide range of soil con-
ical approaches require a large amount of data and ditions and anchor geometries while avoiding exces-
involve considerable uncertainty in extrapolating data sive computational effort.
to different soil conditions and anchor geometries. The development of the model presented herein
Various limit equilibrium analyses developed by a num- was supported by finite element studies of soil-plate
ber of investigators for analysis of anchor penetration, interactions as well as comparisons to field and
including Stewart (1992), Neubecker & Randolph model tests of anchor drag embedment. For the pur-
(1996), and Dahlberg (1998), combined with analytical pose of brevity, these issues are not addressed here,
formulations for anchor line response (Vivatrat et al. with the major focus of this paper being a presenta-
1982, Neubecker & Randolph 1995) provide a rational tion of the general upper bound framework for pre-
basis for trajectory prediction that avoids many of the dicting anchor trajectory and load capacity.
limitations of purely empirical approaches. More recent
contributions by ONeill et al. (2003) incorporate finite
element analyses of soil-plate interactions into a plasti- 2 VIRTUAL WORK FORMULATION
city framework for predicting anchor trajectory. The lat-
ter, plasticity-based approach requires fewer intuitive The analysis considers the anchor depicted in Figure 1
assumptions regarding the various failure mechanisms comprised of a prismatic shank attached to a fluke.

179

Copyright 2005 Taylor & Francis Group plc, London, UK


Center of Rotation, C0
at (x0, y0)

vpf
Anchor
Line Force, F
Attachment
Point
Center of Rotation, Cp
Shank fts vtf
vA
ftf fpf fps Figure 2. Kinematics of anchor motions.
Fluke
npf
fbf
Rotation about center Rotation about edge
Figure 1. Schematic of embedment anchor. of fluke: npf = 6 of fluke: npf = 12

The fluke is assumed to be a plate section of arbitrary


shape subject to the restriction that it is symmetric
about the axis of the shank. The anchor motions under
consideration are considered to be restricted to planar,
rigid body motions. Thus, any instantaneous motion
of the anchor can be represented as a rotation about Distance between
some center of rotation in the plane (Den Hartog center of rotation and
1948). For any anchor motion, the anchor line attach- center of fluke:
ment point will experience an instantaneous velocity
vA, the dot product of vA with the anchor line force F Figure 3. Bearing resistance as function of center of rotation.
defining the rate of work performed on the anchor by
the anchor line.
Figure 1 shows the components of soil resistance Stresses normal to the fluke are related to soil
to anchor motions that are considered in the analysis. undrained shear strength Su by the bearing factor npf:
The bearing resistance terms fpf, fbf and fps denote the
average net ultimate stresses acting normal to the sur- (1)
faces on which they act. Soil resistance acting parallel
to the long axes of the fluke and shaft are designated
by ftf. and fts, respectively. In general, the magnitudes At a given instant of time, the bearing factor npf is
of these resistance terms vary with location on the sur- taken as constant at all points on the fluke. However,
faces on which they act; however, they are assumed to the soil strength Su can vary with depth Z; hence, fpf
be uniform in the direction normal to the plane of sym- is in general variable. Further, npf is dependent on the
metry of the anchor, i.e., in a direction into the page. location of the center of rotation tr of the fluke (Fig. 3).
The kinematics of the fluke motions (Fig. 2) can For centers of rotation beyond the edges of the fluke,
conveniently be expressed in terms of an instantaneous npf is taken equal to 12, while for a center of rotation
velocity parallel to the long axis of the fluke vt and an coinciding with the center of the fluke, npf is taken
angular velocity about a center of rotation located on equal to 6. For intermediate centers of rotation, a sec-
the long axis of the plate, Cp. Note that the center of ond order relationship is assumed:
rotation of the fluke, Cp, is distinct from the center
of rotation C0 used to describe the anchor motions in (2)
Figure 1. A local coordinate system may thus be
established, with the p and t coordinates correspond- where Lf is the fluke length.
ing to directions normal and tangential to the fluke, The unit resistances fps and fbf are formulated in a
respectively. The special cases of pure translation in manner analogous to that of Equation 1; however, the
directions normal or parallel to the plate are described associated bearing factors nps and nbf are taken as
by a center of rotation located an infinite distance being independent of the conditions of rotation. The
from the plate. A description of motions of the anchor tangential unit resistance ftf is formulated assuming
shank is treated in a similar manner. that the maximum shearing resistance parallel to the

180

Copyright 2005 Taylor & Francis Group plc, London, UK


fluke surface is simply the adhesion between the soil 8
and plate:

Moment Resistance, Nm f
6
(3)
4
where  is an adhesion factor and the resistance fac-
tor ntf equals to 2 to account for the skin resistance 2
acting on both surfaces of the fluke. The shank tan-
gential resistance fts is characterized similarly, except 0
0 2 4 6 8 10 12 14
that the resistance factor nts is defined as the ratio
of shank surface area to projected area; nts   for a Translational Resistance, Npf
cylindrical shank. The tangential components of resist-
ance are assumed to be unaffected by the conditions of Figure 4. Soil-anchor interaction relationship.
anchor rotation.
With the system of forces and stresses defined
as presented above, the anchor line force F may be (6)
related to soil resistance by the following virtual work
expression:
(7)

The resistance factors Ntf, Npf, and Nmf may be


regarded as global factors describing total soil resist-
(4) ance, in contrast to the local factors (e.g. npf) describing
soil resistance at a point. Establishing a connection
between the local and global resistance factors is pos-
where Af is the area of the wide surface of the fluke, sible by considering the interaction relationship between
Aef is the end area of the fluke, As is the area of the shank Npf and Nmf implied from Equation 2. Evaluating npf
projected normal to its long axis, vpf is the velocity of in Equation 2 for various centers of rotation, substitut-
a point on the fluke normal to its long axis, vtf is the ing in Equation 1, and integrating over the area of the
velocity of the fluke parallel to its long axis, vps is the fluke permits the evaluation of the interaction rela-
velocity of a point on the shank normal to its long tionship shown in Figure 4. This interaction relation-
axis, and vts is the velocity of the shank parallel to its ship may be considered reasonable in the sense that it
long axis. is convex everywhere and satisfies the limiting condi-
tions; i.e., for translation normal to the fluke Npf 
12, and for pure rotation about the center of the fluke
3 SOIL RESISTANCE INTERACTIONS Npf  6.
The exercise summarized by Figure 4 suggests that
Equations 2, 3, and 4 characterize the localized soil Equation 2 is reasonable. Further evaluation is needed
resistance to anchor motions. Soil resistance can alter- to verify that Equation 2 can provide reasonable Npf -
natively be characterized in terms of a total system of Nmf interactions for general conditions of anchor fluke
forces and moments acting on some component of the geometry and soil strength gradients. Such studies are
anchor (ONeill et al. 2003). The disadvantage to the currently in progress by the authors.
approach is that appropriate resistance functions and
their respective interactions require new derivations
for each anchor geometry and soil condition of interest. 4 ANCHOR LINE FORCE
However, the available total system results provide a
good benchmark for characterizing local resistance The four velocity components in Equation 4 (vpf, vtf,
functions. The validity of the localized functions for vps, vts) are unique for a given center of anchor rota-
more general conditions is a subject for further study. tion (x0, y0 in Fig. 1). Hence, for a given anchor line
As a simple example, consider the case of a rectangu- attachment angle  (Fig. 1) the coordinates of the cen-
lar fluke of length Lf with unit thickness into the page ter of rotation (x0, y0) can be optimized to obtain a
in a soil of uniform strength Su. The force acting par- minimum anchor line tension at the attachment point
allel to the long axis of the fluke (Ftf), perpendicular Ta corresponding to a F of minimum magnitude.
to the long axis (Fpf), and a resisting moment (Mf) Repeating this exercise for various  values produces
may be characterized as follows: a locus of points relating anchor line attachment angle
to the magnitude of the anchor line force as illustrated
(5) by Figure 5.

181

Copyright 2005 Taylor & Francis Group plc, London, UK


1000 0
Anchor Capacity and Line Tension (kN)

Line Load
800
-5

Embedment Depth (m)


Anchor Capacity
600
-10
= 0.5
400

Critical Angle
200 -15

0 -20
0 5 10 15 20 25 30 35 40
0 50 100 150
Load Angle (degrees from horizontal)
Drag Distance (meters)
Figure 5. Anchor line force versus direction.
Figure 6. Anchor trajectory.

Obtaining a unique solution for Ta and  is possible


using solutions relating anchor line tension to load curve intersects the anchor capacity curve. If the
attachment angle. A simple but reasonably accurate intersection point occurs at a load angle less than the
approximate solution for chain angles close to zero at critical angle, anchor motions are assumed to be purely
the seabed is given by Neubecker & Randolph (1995): translational in a direction parallel to the fluke. If the
intersection point occurs at a load greater than the
(8) critical angle, anchor rotation is assumed to occur. When
the anchor is in a translational mode, the analysis is
where za is the depth from the seabed of the anchor carried forward in 1 meter vertical increments. When
the anchor is in a rotational mode, the analysis is car-
line attachment point, and Q is the average bearing
resistance acting on the anchor line. It is emphasized ried forward for incremental rotations of 1 degree.
that Equation 8 is approximate and is not strictly Figure 6 shows an example prediction of anchor
applicable to layered soils or to deep embedment con- for anchor trajectory in a soil of uniform strength for
ditions for which the chain angle at the seabed signifi- a simple anchor comprised of a 3-m wide by 1.5-m
cantly departs from zero. The intersection point of long rectangular fluke, a 3.9-m long shank having a
Equation 8 with the upper bound Ta- relationship in 0.2-m square shank, with a 50 degree angle between
Figure 5 establishes a unique anchor line tension cor- fluke and shank. Trajectory predictions were made for
responding to a given depth za. soil sensitivity of 2; i.e., an adhesion factor   0.5.
The predictions in Figure 6 show the trajectory to be
strongly affected by soil-plate adhesion. For the simple
5 TRAJECTORY case of a uniform soil strength profile, the anchor line
tension Ta reaches its maximum level, approximately
With the anchor orientation thus determined at any 600 kN, at a relatively shallow depth and remains con-
time, the trajectory of the anchor is directly deter- stant thereafter. In more realistic situations in which
mined. This process is carried out in steps where the soil strength increases with depth, line tension of course
anchor is assumed to rotate and translate in small incre- increases much more dramatically with depth.
ments. After each increment the calculation is repeated
taking the anchors new position into account.
Figure 5 shows the load angle versus anchor capacity 6 SUMMARY AND CONCLUSIONS
curve to have two distinct portions. At shallow angles
(less than the critical angle marked on Fig. 5), load This paper presents a proposed upper bound plastic
capacity increases with increasing load angle. Above limit approach to predicting the trajectory of drag
the critical angle the load capacity curve reaches a embedment anchors. A virtual work analysis is formu-
plateau and eventually begins to decline. Different lated in terms of unit soil bearing resistance factors
algorithms for incremental penetration of the anchor (Eqs. 1 through 3). While this approach involves some
are used according to where the anchor line force uncertainties, it permits straightforward modeling of

182

Copyright 2005 Taylor & Francis Group plc, London, UK


anchor penetration for complex anchor geometries in Service (Cooperative Agreement No. 1435-01-99-
non-uniform soil strength conditions. For each incre- CA-31003), the Offshore Technology Research Center
ment of penetration analyzed in the drag embedment and their colleagues at Texas A&M University and the
process, the coordinates of the center of rotation are University of Texas.
optimized to determine the minimum collapse load
corresponding to a given anchor orientation. Consid-
eration of the anchor line orientation (Equation 8)
permits the determination of a unique anchor collapse REFERENCES
load and orientation. An attractive aspect of the
proposed procedure is that the collapse mechanism is Dahlberg, R. 1998. Design procedures for deepwater anchors
selected through an optimization procedure rather in clay. Proc. 30th Offshore Technology Conference. OTC
8837: Houston.
than relying on intuitive assumptions. Den Hartog, J.P. 1948. Mechanics, New York: Dover
The unit bearing resistance factors presented in Publications.
this paper were applied to a simple rectangular plate Naval Civil Engineering Laboratory (NCEL). 1987. Drag
to deduce the implied soil resistance interactions for Embedment Anchors for Navy Moorings, Techdata Sheet
the entire plate. The resulting interaction relationship 83-08R. Port Hueneme, California: NCEL.
(Fig. 4) appears reasonable. Studies are in progress to Neubecker, S.R. & Randolph, M.F. 1995. Profile and fric-
verify that Eqs. 1 through 3 can yield similarly reason- tional capacity of embedded anchor chain. ASCE J.
able results when applied to complex anchor geometries Geotech. Engr. 121(11): 797803.
(triangular, trapezoidal, composite shapes) and non- Neubecker, S.R. & Randolph, M.F. 1996. The performance
of embedded anchor chains systems and consequences for
uniform soil conditions. The numerical algorithm for anchor design. Proc. 28th Offshore Technology Conference.
calculating anchor trajectory during drag embedment OTC 7712: Houston.
also seems to produce plausible predictions (Fig. 6). ONeill, M.P., Bransby, M.F. & Randolph, M.F. 2003. Drag
Work is in progress to evaluate the proposed predict- anchor fluke-soil interaction in clays. Can. Geotech. J.
ive model in light of other predictive methods, labora- 40: 7894.
tory models, and field observations. Stewart , W.P. 1992. Drag embedment anchor performance
prediction in soft soils. Proc. 24th Offshore Technology
Conference. Houston: OTC 6970.
ACKNOWLEDGEMENTS Vivatrat, V., Valent, P.J., & Ponterio, A.A. 1982. The influ-
ence of chain friction on anchor pile design. Proc. 14th
The authors would like to acknowledge the support Offshore Technology Conference, Houston: 153163.
of Department of the Interior Minerals Management

183

Copyright 2005 Taylor & Francis Group plc, London, UK


Ultimate pullout capacity of SBMs VErtically Loaded Plate Anchor
(VELPA) in deep sea sediments

P.Y. Foray*, S. Alhayari** & E. Pons**


*INPG, Laboratoire, Grenoble, France and **Single Buoy Moorings Inc., Monaco

L. Thorel & N. Thetiot


Laboratoire Central des Ponts et Chausses, Nantes, France

S. Bale & E. Flavigny


INPG, Laboratoire, Grenoble, France

ABSTRACT: Plate anchors can be considered as an innovative technology and an alternative to Suction
Caissons for anchoring offshore structures in deep sea sediments. The VELPA, developed by SBM, is a rect-
angular plate type anchor driven vertically and rotated to its optimal position by applying a vertical tension on the
mooring line. An extensive study of the keying (rotation) capability and pullout behaviour of such anchors has
been carried out combining Laboratory tests, half scale field tests and centrifuge tests. The model plates were
instrumented in order to control their inclination and to measure the suction developed on the rear face of the
anchor. The results indicate that suction can take a significant part in the pullout capacity and that the VELPA
can be considered as a valuable solution for deep anchoring.

1 INTRODUCTION develop around the anchor during the application of a


pullout load and if a significant contribution of the
An increasing attention has been given in the recent suction behind the plate, inducing a reverse bearing
years to find alternatives to suction caissons for deep capacity mode, can be taken into account in the
mooring. For this purpose SBM has developed a design. Previous works were conducted by the Naval
VErtically Loaded Plate Anchor (VELPA) which is a Civil Engineering Laboratory in California by Beard
plate type anchor driven vertically in the soft sedi- (1979) and Forest et al (1995) indicating that a stand-
ments by means of the Pyrodriver. The validation of ard holding factor of 9 resulting from the deep fail-
the system implies (i) to explore the possibility of ure mechanism can be assumed if the effect of suction
installation of such plates in deep water by an eco- is neglected. But this factor can increase up to 15 with
nomical vertical driving system and (ii) to evaluate the suction. A reduction factor of 0.7 to 0.8 should be
the ultimate pullout capacity of the anchor. applied for the effect of the soil disturbance due to the
First the anchor and its fixation to the anchoring anchor installation. This factor may increase to 1 after
line has to be guided during the driving process to soil set-up. The authors propose a maximum design
allow a penetration up to 30 m in the soft clay sedi- holding factor of 9 for long term conditions and 15
ments with a minimum of applied load or impact. It is for short term conditions.
achieved by the use of the Pyrodriver which is a com- The research between SBM, Laboratoire 3S and
bustion hammer as described by Alhayari et al (2003). LCPC consisted in an extensive experimental pro-
Then a pretension and a rotation have to be applied to gram with various physical modelling of the anchor,
give the anchor its final position which should be per- combined with a numerical modelling.
pendicular to the inclination of the mooring line. The The experimental program reproduced the different
VELPA solution can be interesting if high values of phases of installation and pullout of anchors with a
the inclination of the anchoring line, up to 45 or 60, simulation of the deep sea soil conditions: soft normally
can be applied in order to reduce the global mooring consolidated clays with a low gradient in undrained
area of the floating platform. shear strength with depth, from a few kPa at the seabed
To evaluate the ultimate pullout capacity, it is to values of 30 to 40 kPa at a depth of 30 m. The dimen-
essential to analyze if deep failure mechanisms can sions of the prototype plate were supposed to be around

185

Copyright 2005 Taylor & Francis Group plc, London, UK


3 m in height (vertical plane) and 4 m in width. In a first
step, laboratory and field tests were performed by
Laboratoire 3S with models at a relatively low scale,
1/15 for the laboratory tests and 1/6 for the field tests. In
a second step a centrifuge modelling of the VELPA was
performed in the large centrifuge of LCPC, Nantes, at a
reduced 1/100 scale, but with a correct simulation of the
gradient of soil properties over 40 m deep.
A numerical modelling of the pullout behaviour of
the anchor was performed using the 2D finite element
code Plaxis. It had two objectives: (i) a calibration of the
model by simulating the different experiments, study-
ing the failure mechanisms around the anchor and the
influence of the experimental parameters, (ii) an extra-
polation to real field conditions allowing an estimation
of the ultimate pullout capacity of the anchor and the
Figure 1. Front face of the plate anchor scale model before
contribution of suction effects. The numerical model the first vertical drive in the 20 kPa clay.
was also considered as a future design tool.
More recently, the VELPA has proven its efficiency
in terms of holding capacity in very soft to firm clay of the anchor, and a pore pressure sensor at the mid-
during the full-scale tests campaign performed by dle of the back face of the plate. A large displacement
SBM offshore Angola in about 500 m water depth. sensor was placed on the anchoring line and the
This paper is focused on the results concerning the resulting load applied was measured with a 20 kN
pullout capacity of the anchors obtained from the load cell.
experimental program. The contribution of the suc-
tion is emphasised.
2.2 Testing procedure
The plate was driven vertically through a guide allow-
ing the vertical orientation to be maintained throughout
2 LABORATORY MODEL TESTS
the penetration to the required depth. A pretension load
was then applied by the action of a mechanical jack and
2.1 Experimental set-up
a pulley placed on the upper frame of the tank. During
A large experimental tank (1 m width, 2 m length and this step the anchor is rotated to the appropriate pos-
1 m depth) was built in the laboratory and filled with ition (perpendicular to the pullout angle). Once the
homogeneous clay with several undrained shear anchor had rotated to the required final position, the
strength: 1 g laboratory tests cannot reproduce the pullout system (pulley, screw jack and geared motor)
gradient in soil mechanical properties observed in the was moved to give to the anchoring line the correct
field conditions over the 20 to 30 m of potential pene- inclination.
tration of the anchor. Therefore it was decided to fill The pullout test was performed applying a dis-
the tank with an homogeneous clay with a constant placement to the anchoring line at a constant speed of
undrained shear resistance. Three tanks were pre- 4 mm/min and measuring the resulting load. This rate
pared: tank n1 with Su  1 kPa to reproduce the con- ensures undrained conditions in the soil around the
ditions close to the seabed, tank n2 with Su  4 kPa anchor. The load was sustained until a displacement
to reproduce the conditions at a few meters under the of 150 to 300 mm was achieved.
seabed and tank n3 with Su  20 kPa to reproduce In each tank, 6 to 9 tests were performed with the
the conditions at a depth of around 20 meters. mid-point of the anchor at an initial depth between 40
The model consists in a steel plate anchor with and 70 cm. The anchoring angles  of the mooring
four fixation points and equipped with two keying line with respect to the horizontal plane varied from
in flaps which helped the anchor rotate during the 25 to 90.
pretension phase. The dimensions of the model were
20 cm height and 30 cm width with a thickness of
2.3 Experimental results
1.4 cm, corresponding to an efficient anchorage sur-
face of S  0.06 m2. Cable links for the four connect- Typical pullout curves are presented in Figure 2,
ing points on the anchor were used because of the together with the mobilization of the suction at the
small dimensions of the model. centre point of the rear face of the anchor.
As shown in Figure 1, the model was instrumented The values of the Ultimate Pullout Capacities
with an inclinometer that gave the transverse inclination (UPC) obtained for all the tests in the three containers

186

Copyright 2005 Taylor & Francis Group plc, London, UK


1100,0 11,0 Table 2. Maximum contribution of suction.
1000,0 10,0
900,0 9,0
Suction Contribution
Tank n kPa Factor %
800,0 8,0

Suction (kPa)
700,0 7,0 1 4.5 (for   90) 73
Force (N)

600,0
Force (N)
6,0 2 10 (for   41) 61
500,0 Suction 5,0 3 38 (for   90) 20
400,0 4,0
300,0 3,0
a suction factor C(Pi), as:
200,0 2,0
100,0 1,0 (2)
0,0 0,0
0,0 50,0 100,0 150,0 200,0 250,0 300,0 where F(Pi)  Suction measured  effective area of
Displacement (mm) the anchor.
The contribution factors C(Pi) reported in Table 2
Figure 2. Pullout test 7 in tank n2. Load and Suction are very high, confirming the over-estimation of the
curves. Inclination of the anchor   45. real suction load resulting from the integration of the
maximum suction over the whole anchors area.
Table 1. Summary of the Ultimate Pullout Capacities (UPC) It can be noted that the relative contribution of suc-
for the laboratory tests. tion decreases when the shear strength Su increases,
but its absolute value increases.
Su UPC Holding Factor Higher suction effects should be obtained at larger
Tank n kPa N Nc depths without drainage paths in field conditions.
During these laboratory tests, encouraging results
1 0.81.1 300460 5.47.8
concerning the technical feasibility of the anchorage
2 3.54.5 9171150 44.8
3 20 740011600 6.29.5 were displayed and the VELPA has proved its ability
to be installed with good precision.

are summarized in Table 1. Global holding factors Nc


were calculated according to the formula: 3 FIELD TESTS
(1) 3.1 Testing procedure
where Su  average undrained shear strength of the The half-scale field tests were performed in a site
soil at the mid-depth of the anchor, and Seff  effective with an homogeneous layer of clay over 6 meters deep
area of the plate. The average values of Nc are reported in the post-glacial sediments of Bourgets Lake, close
in Table 1. to Grenoble. The average value of the undrained shear
Nc factors obtained were lower than expected for strength of the clay is 33 kPa, which is exactly in the
tanks n1 and n2. One reason is that the ratio between range of the expected values for the real field condi-
the height of the anchor to the depth of the tank (1 m) tions at a depth of 30 m.
did not allow to develop a full deep failure mech- The model plate represented Figure 3 had a width of
anism, with circular slip lines surrounding the anchor. 0.675 m, a height of 0.5 m and a thickness of 33 mm. It
The second reason is that drainage paths for the suc- was instrumented with two inclinometers in order to fol-
tion can more easily develop in the laboratory experi- low the rotation of the plate during all the positioning
ments through failure plane reaching the surface. With phases and the pullout tests, and with two pore pressure
the development of deep failure mechanism, higher sensors in order to measure the suction on its rear face.
values of the holding factor Nc should be induced. Four anchoring points ensured the stability of the
The different tests have shown that for an anchor- plate and the perpendicular to the anchoring line in its
age inclination  40 the anchorage angle has no final position. A 200 kN load cell and a displacement
influence on the UPC, and for  40, a global sensor were fixed along the anchoring cable. A steel
increase in the value of the UPC of up to 21% was follower was designed to drive and guide the anchor
observed for tanks n1 and n2. 4.50 meters deep by the vertical action of the arm of an
hydraulic power shovel. The pullout load was applied
either through the arm of a 300 kN power shovel or
2.4 Contribution of suction
with a 200 kN mechanical jack.
Assuming a uniform distribution of the suction over Five complete tests were performed, including the
the plate area, the measured value was used to define different phases: initial vertical driving, initial rotation

187

Copyright 2005 Taylor & Francis Group plc, London, UK


Pore pressure -
-100 100
Kyowa
Pore pressure -
-80 Entran 80
Load

-60 60

Pore pressure (kPa)

Load (kN)
-40 40

-20 20
0 20 40 60 80 100 120
0 0
Displacement (cm)
20 -20

40 -40

Figure 3. Field instrumented plate. Figure 5. Pullout test n1 on the half scale field model.
Load and suction as functions of displacement
-120 120
Pore pressure - Kyowa Initial anchoring angle:   52.8;
Pore pressure - Entran depth before the pullout test: pini  2.55 m.
Load
-100 100

-80 80
were applied. Figure 5 shows the same curves as a
function of the displacement of the mooring line for
-60 60 Test 1 and an inclination of 53, where a more regular
Pore pressure (kPa)

loading rate was applied.


-40 40 It can be observed that the two pore pressure sensors
Load (kN)

react exactly in the same way and at the same rate than
-20 20
the load cell: a strong increase in the load corresponds
to a strong increase in the suction behind the plate.
0 100 200 300 400 500 600 700 800 900 The initial values of the pressure were close to
0 0
30 kPa, corresponding to the hydrostatic pressure. For
time (s) the whole series of the field tests, the suction reached
20 -20 values around 40 kPa, corresponding to a contribution
between 15% and 20% of the total load. This contribu-
40 -40 tion remained nearly constant either during a continu-
ous loading or during fast loading impulsions, showing
that the pullout capacity of the anchor is a combination
60 -60
of suction mobilization and mechanisms of passive
Figure 4. Pullout test n2 on the half scale field model. Initial pressure in front of the plate and reverse bearing cap-
anchoring angle:   37.5; depth before the pullout test: acity in its rear part.
pini  2.93 m. The ultimate pullout capacities obtained for the field
tests were between 80 and 100 kN, corresponding to a
and pretension, final positioning at a given loading holding capacity factor Nc  7.5 to 9.3. These relatively
inclination and pullout testing. The initial depth of the low UPC values can be attributed to the fact that the
middle point of the anchor was either 3.75 m or anchor did not generate a complete deep failure mech-
4.25 m. The pullout tests were performed with values anism. The anchor depth after pretension was between 5
of the inclination of the anchoring line of 35, 38, and 6 times its height, which may be sufficient to mobil-
40, 45 and 53. Unloading steps and strong changes ize a complete reverse bearing capacity model, as sug-
in the pullout rate were also applied in some tests to gested by the numerical approach and the analytical
simulate storm conditions. results. But possible cracks or drainage paths around the
plate may have occur during these tests, reducing the
suction and inducing a global lower value of the holding
3.2 Pullout test results
factor Nc.
The total pullout load and suction curves as a function These tests were technically a success. All the phases:
of time are presented in Figure 4 for Tests 2 and an driving, pretensioning/rotation and pullout actions
inclination of 37.5. In this test, fast loading cycles showed a very good feasibility.

188

Copyright 2005 Taylor & Francis Group plc, London, UK


120 50
P113
100 P103 0

Pressure variation (kPa)


Pullout Force (daN)
80 -50
Suction Pi2

60 -100

Force
40 -150

Figure 6. View of the centrifuge model anchors.


20 -200

0 -250
0 50 100 150
Pretensioning
position Pullout position Displacement (mm)
Hydraulic jack
Follower Force
Chain sensor
Figure 8. Pullout test of a pre-embedded anchor with an
Fork inclination of 45.
10
Cable Displacement
520 mm
45 Electric jack 4.2 Pullout tests Pre-embedded plates
The height of the soil mass was 430 mm and the
900 mm anchors were buried at a depth of 210 mm (220 mm
from the base and 260 mm from the sides of the con-
tainer). Also two pore pressure sensors were placed
Figure 7. Container configuration for the complete tests. within the soil around anchor 2 and three around the
non-instrumented anchor 1.
The graphs shown Figure 8 display the force-
displacement curve and the curves of relative pore
4 CENTRIFUGE TESTS
water pressure for each sensor for the pullout test of
anchor 2. The maximum ultimate pullout capacities
4.1 Testing setup
were 6.1 MN for anchor 1 and 9.5 MN for anchor 2 at
The material used for these experiments was a high the prototype scale, with corresponding values of Nc
kaolin clay; Speswhite clay. The cylindrical container respectively of 28 and 42.
had a diameter of 900 mm and a height of 520 mm and These values are much higher than those found for
was filled in layers of about 100 mm, each being con- the laboratory and field tests. The forcedisplacement
solidated for about one week in order to obtain the curve shows a peak followed by a drop in force and
shear strength profile required. An additional in flight finally a tendency towards a residual value, whereas
consolidation was performed at 100 g, where the clay for the previous laboratory tests the force increases
consolidates under its self-weight. A control of the with displacement up to a plateau.
gradient in undrained shear strength was achieved per- The maximum relative depression measured by the
forming in flight model CPT. The average profile sensor Pi2 on the rear part of anchor 2 is
200 kPa.
obtained by this method was in the range of: Su  Assuming a uniform distribution of this depression
0.8  z (Su in kPa, z  depth in m, prototype). (overestimation of the suction), a contribution due to
The anchors, with dimensions of 4  3  0.4 m suction behind the anchor of 27% of the UPC is
are modelled at a scale 1:100 by duralumin plaques, obtained. The difference in the form of the curves and
as shown on Figure 6. The anchor lines are modelled the significant peak depression behind the anchor sug-
by 1.2 mm diameter stainless steel cables, attached to gests that in the centrifuge, and with buried anchors
the anchors at 4 points by cables of 0.75 mm. (no remoulding of the soil), a more complete mobil-
Of the two anchors, anchor 1 was non-instrumented isation of the reverse bearing capacity failure is
and anchor 2 was equipped with a pore pressure sensor; achieved. The drop in force and suction after the peak
Kyowa PS-2KA. is probably caused by the development of a fissure up
A first series of tests consisted of two pullout tests to the surface of the soil. The residual values of Nc,
of pre-embedded anchors, positioned at an inclination respectively 15 and 20.6, obtained for large displace-
of 45. The second series consisted of two complete ments, are much closer to the values obtained at 1 g,
tests of the anchor system, i.e., driving, pretensioning where the reverse bearing capacity failure mode was
at 80 and pullout at 45 (Fig. 7). not fully mobilised.

189

Copyright 2005 Taylor & Francis Group plc, London, UK


1,2 -240 even after failure. Because of this, and contrary to the
tests on the buried anchors, the constant suction allows
1 -200 the force to maintain a constant value.

Suction (kPa)
0,8 -160
Force (kN)

0,6 -120 5 CONCLUSIONS


Load
0,4 -80 The pretension method using a quasi-vertical inclin-
Suction
ation of the anchoring line gave satisfactory start of
0,2 -40
the rotation of the anchor. It was verified that the final
0 0 inclination of the anchor is controlled by the inclin-
0 20 40 60 80 100 120 140 ation of the anchoring line.
Displacement (mm) The laboratory and field tests allowed to validate
the installation procedure. The depth of the anchors
Figure 9. Pullout test of the driven and pretensioned did not allow a complete deep failure mechanism.
anchor. Final inclination of the mooring line: 45. Nevertheless, suctions in the order of 40 kPa could be
measured behind the plate, giving a suction contribu-
tion of 15%20% to the ultimate pullout capacity.
4.3 Pullout of pretensioned plates
Centrifuge tests on pre-installed anchors give val-
The driving of the anchor was achieved in the same ues of Nc of 42 and 28, while pullout tests after the
way as for the in-situ experiments. A hydraulic jack, simulation of the complete installation of the anchor
attached to a follower and a fork which makes con- give Nc values of 31 and 24 and suctions in the same
tact with the anchor (see Fig. 7), drives the anchor order of the field tests.
into the soil at a speed of about 35 mm/s. After the The promising results of the laboratory and field
driving operation, the follower is removed. models have encouraged SBM to undertake full-scale
The pretensioning is undertaken using the traction tests offshore Angola. The tests were performed suc-
cable at an angle of 80 from the horizontal in order to cessfully thus validating the VELPA concept for the
permit the use of the minimum force possible. After mooring of offshore structures.
the pretension, the angle of the traction cable was Further research in the following areas would be of
changed to 45 and a pullout test was realised. interest: long-term loading, displacement under
The graphs on Figure 9 show the forcedisplacement working load and the effects of the dissipation of suc-
curve for the pullout test, as well as the curve of relative tion, cyclic and shock loading, 3D numerical model-
depression for the Kyowa sensor in anchor 2. ling, soil set-up effects.
The ultimate pullout capacities were 6.6 MN for
anchor 1 and 4.9 MN for anchor 2 at the prototype scale,
corresponding to Nc values of 31 and 24. REFERENCES
The maximum relative depression measured
behind anchor 2 was
59 kPa, corresponding to a suc- Alhayari, S., Van Foeken, R. 2003. VErtically Loaded Plate
tion contribution to the UPC of 14%. Anchor (VELPA) for Deepwater Taut Moorings. 15th DOT
Globally, these values are lower than for the pull- International Conference and Exhibition, Marseille,
out tests on the buried anchors. This difference can be France. November 2003.
explained by the remoulding of the soil around the Beard, R.M. 1979. Long-Term Holding Capacity of Statically
anchor during the phases of driving and pretension. Loaded Anchors in Cohesive Soils. Civil Engineering
Laboratory, Naval Construction Battalion Center, Port
Another important difference is that for the com- Hueneme, California, Report No. TN-1545. January 1979.
plete tests, the form of the forcedisplacement curves Forest, J., Taylor, R., Bowman, L. 1995. Design Guide for
found in the 1 g laboratory tests is reproduced, with a Pile-Driven Plate Anchors. Naval Facilities Engineering
more or less constant plateau of force. In this case Service Center, Port Hueneme, California, Report No.
there is no drop in suction after the failure point, the TR-2039-OCN. March 1995.
relative depression is maintained at about
50 kPa,

190

Copyright 2005 Taylor & Francis Group plc, London, UK


Effects of long term loading on storm capacity of
vertically loaded anchors

H.A. Taiebat
University of Technology Sydney, NSW, Australia

C.P. Thorne & J.P. Carter


The University of Sydney, NSW, Australia

ABSTRACT: In this paper attempts have been made to find the effects of sequential loadings on the ultimate
uplift capacity of a typical horizontal plate anchor loaded vertically. A series of finite element analyses was per-
formed where the anchor was subjected to different initial loading under undrained conditions followed by con-
solidation under sustained loading. The anchor was then loaded under undrained conditions up to failure. The
anchor is idealised as a circular plate embedded in a homogeneous soil. The soil is represented by the Modified
Cam Clay material model with an undrained shear strength varying linearly with depth. No allowance has been
made to model break away between the soil and the anchor. The results of the finite element analyses show that
if a low value of the initial loading is applied to the anchor under undrained conditions followed by consolida-
tion, the ultimate undrained uplift resistance of the anchor increases. However, for relatively high values of the
initial loading the soil fails during the consolidation period.

1 INTRODUCTION reconfigured so that the anchor pulling line forms a


right angle to the fluke, thus increasing the available
Both the exploration and exploitation of hydrocarbons uplift capacity by a factor of about 2. These anchors
are heading more towards deeper waters. Tension leg, have different innovative designs to satisfy the needs
taut and semi-taut leg and semi-submersible platforms for load resistance as well as convenient installation
are among the options that are used increasingly in procedure. Unlike the conventional drag anchors, the
deep waters. With the greater depths there are greater capacity of the new anchors is not dependent on the
needs for more robust anchoring systems to transfer direction of pull so that even direct vertical loading
predominantly tensile forces to the ocean floor. can be sustained.
Vertically loaded anchors (or VLAs) are among the Anchors are usually subjected to a permanent ten-
few technologies currently used for this purpose in sile loading, which is a fraction of the maximum
deep waters. undrained capacity of the anchor, and is required for
The last 20 years has seen a great increase in the the stability of the platform. They are also subjected
capacity of drag embedment anchors, mostly due to to temporary fluctuating loadings applied due to
improvements in design that have allowed a greater storms and surface or deep currents. The permanent
penetration in softer seabeds. This higher capacity has load is significantly less than storm loading and is
in turn allowed the extension of development into usually applied in a relatively short period of time and
areas with much deeper water and more hostile envir- then remains approximately constant with time. This
onmental conditions. will result in dissipation of the initial excess pore
Over the last 10 years or so an entirely new breed pressures and consolidation of the soil around the
of drag embedment anchor has been developed. These anchor so that when storm loading is applied it might
anchors are configured at installation as drag embed- be expected that the response to loading differs from
ment anchors. They are placed and installed in much that for a single first time load.
the same way as the conventional drag anchors; the The initial capacity of such anchors has been stud-
anchors can be penetrated to the required depth by drag ied both theoretically (e.g. Rowe & Davis 1982,
forces or by direct penetration. However after pene- Pyrah et al. 1985, Booker & Small 1987, Small et al.
trating to the required depth the new anchors can be 1988, Merifield et al. 2003, Thorne et al. 2004) and

191

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 1. Material properties for the Modified Cam-Clay
model. 0

Property Value -18m

Anchor
M 1.0
Elastic stiffness,  0.05 -20m
-20m
Plastic stiffness modulus,  0.30
Poissons ratio,  0.25
Void ratio at unit p in e-ln p space, ecs 2.20
Submerged unit weight,  6 kN/m3 -22m
Coefficient of permeability, k 1  10
9 m/sec 0 3m
-40m
0 25m
experimentally, by model scale or full size tests (e.g.
Dahlberg & Strom 1999). Little work has been pub-
lished on the effects of long term loading on anchors. Figure 1. Finite element mesh used in the analyses.
In conventional anchors, it is accepted that a minor
increase in ultimate undrained capacity can be expected defines the overall geometry of the finite element
if the anchor is left under a load less than its ultimate model.
undrained capacity for a long enough time to allow for The deep anchor plate was assumed to be rigid and
dissipation of pore pressures to occur before the anchor impermeable. This latter condition is important because
is finally taken to failure. This paper explores theoreti- if drainage paths exist through the anchor, dissipation
cally whether this effect is present for VLAs and what of pore pressures will occur very much faster as water
the implications are for movement of the anchor and its flows from the high pressure areas above the anchor
ultimate capacity. to the low pressure areas below.

2 ANALYTICAL MODEL 3 FINITE ELEMENT RESULTS

The typical anchor examined in this study is idealised A series of finite element analyses was performed to
as a 2.5 m diameter circular plate (although real investigate the behaviour of the anchor under various
VLAs may have a rather complex shape and often an types of loading. The results of the analyses are pre-
innovative design) embedded at a depth of H  20 m sented in the form of load deflection curves for each
below the seabed, i.e. relatively deep. The seabed soil type of loading.
is assumed to obey the Modified Cam Clay material
model. Elastic-perfectly plastic material models, such 3.1 Undrained and fully drained loading
as Tresca and Mohr-Coulomb, do not show any stress
The first analyses were to determine the capacity of
softening/hardening, and therefore are not suitable to
the anchor loaded under undrained conditions, i.e.
model the softening/hardening that may occur in the
very rapid loading, and fully drained conditions, i.e.
soil below/above the anchor as a result of pore pressure
long term loading. The load deflection curves for
changes. The material properties used to define the
both cases are shown in Figure 2. In this figure Pu
Modified Cam Clay model were chosen as those of a
represents the undrained ultimate pullout capacity of
typical soft seabed soil and are presented in Table 1.
the anchor, P is the load level,  is the vertical dis-
For the purposes of the calculation a normally con-
placement of the anchor normalized by the anchor
solidated soil was assumed giving an undrained shear
diameter, D. The analysis performed under undrained
strength gradient of 1.75 kPa/m. For calculation of the
conditions results in a clearly defined ultimate pull-
initial effective stress state, the coefficient of lateral
out capacity of Pu  2540 kN at a vertical displace-
earth pressure, Ko, was assumed to be 1.
ment of about 0.5 m. However, the drained analysis
The geometry of the problem under investigation is
results in a more abrupt failure at a vertical load of
axi-symmetric. The finite element program AFENA
1440 kN, and a vertical displacement of 0.19 m, after
(Carter & Balaam 1995) has been used for the analy-
which the displacements become unreasonably large.
ses of the anchor. The finite element mesh used in the
Overall the behaviour under drained loading is softer
analyses consists of 425 isoparametric (8 noded) ele-
than that of undrained loading.
ments. A thin layer of elements has been used around
the anchor in order to capture the effects of local
3.2 Sequential loading
shearing close to the foundation. A schematic repre-
sentation of the axi-symmetric finite element mesh A series of finite element analyses was performed to
used in the analysis is shown in Figure 1, which also show the behaviour when the anchor was initially

192

Copyright 2005 Taylor & Francis Group plc, London, UK


Figure 2. Load-deflection curve of anchor under undrained Figure 4. Load deflection curves predicted under low
and fully drained conditions. sequential loadings.

under the steadily maintained load of 40% of the ultim-


ate undrained capacity. The anchor displacement has
been plotted against a non-dimensional time factor, T,
defined here as T  kt/D, where t denotes real time.
For a typical permeability of k  1 10
9 m/sec, the
time for 90% consolidation is about 116 days, although
50% of the consolidation would occur in about 23
days. The rate will depend on the existence or other-
wise of soil structure. For example, sand or silt or shell
seams in the clay can increase the permeability of the
soil mass by several orders. The rate will also be influ-
enced by the amount of disturbance in the soil sur-
rounding the anchor during installation because large
Figure 3. Displacement versus time factor under 40% of scale shearing could disrupt such natural drainage fea-
the undrained pullout capacity. tures and decrease the permeability of the soil mass.
Figure 4 shows the load deflection curves for
analyses under initial loadings of 20%, 40%, and 50%
loaded quickly to a percentage of the undrained pull- of the ultimate undrained pullout capacity. The initial
out capacity, and then this load was maintained at a loading follows the undrained loading curve, and
steady value until the pore pressures were fully dissi- then, as consolidation under constant loading takes
pated. Following this stage, the anchor was subjected place, the deflection increases to that of the drained
to further loading under undrained conditions, i.e. rap- loading curve. Subsequent undrained loading results
idly, until failure occurred. The finite element analyses in a relatively stiff response and finally a maximum
show, in general, two different types of response; one load is reached that slightly exceeds the first time
when the initial undrained loading is relatively low, ultimate undrained capacity. The structural loading
and the other when the initial undrained loading is rel- on the anchor is determined by the total pressure dis-
atively high. tributions above and below the anchor. Figures 5 and
6 show the pressure distribution across a radius of the
3.2.1 Sequential loading, low initial loading anchor at the end of the initial loading of 40%
The behaviour of the anchor under initial loads that undrained capacity, and at the end of the consolida-
are below 55% of the undrained pullout capacity is tion under that load, respectively. The pressures have
presented in this section. As a specific example, the been normalized by the initial vertical effective stress
behaviour of the anchor under 40% initial loading is in the soil at anchor level before application of any
presented in detail. loading, i. It can be seen that while there is only
Under the initial undrained loading, there is an minor variation of effective stresses at the end of
immediate displacement during the loading which is undrained loading, the difference between the effect-
followed by a gradual increase in deflection with time ive stresses above and below the anchor becomes
during the consolidation period under the constant large and the effective vertical stress below the anchor
load. The rate of displacement decreases and eventu- approaches zero at the end of the consolidation
ally ceases after some time. Figure 3 shows the behav- period, which is an indication of possible breakaway
iour of the anchor during the consolidation period and formation of a gap below the edge of the anchor.

193

Copyright 2005 Taylor & Francis Group plc, London, UK


3.00 Effective vertical stress above 1.50
Total vertical stress above
2.50 Effective vertical stress below
Total vertical stress below
2.00
1.00
Pressure /'i

Pressure /'i
1.50

1.00

0.50 0.50 Initial 40% loading

0.00 First time undrained

-0.50
0.0 0.5 1.0 1.5 2.0 0.00
Radial distance (m) 0.0 0.5 1.0 1.5 2.0
Radial distance (m)

Figure 5. Variation of stresses across a radius of the anchor


at 40% initial loading before consolidation. Figure 7. Variation of the effective stresses at failure above
the anchor under 40% staged loading and under the first
2.50 time undrained loading.
Effective vertical stress above
2.00
Effective vertical stress below
1.50
Pressure /'i

1.00

0.50

0.00

-0.50
0.0 0.5 1.0 1.5 2.0
Radial distance (m)

Figure 6. Variation of stresses across a radius of the anchor


at 40% initial loading after consolidation.
Figure 8. Displacement versus time factor under 60% of
As indicated previously, the ultimate pullout the undrained pullout capacity.
capacity of the anchor under staged loading, after
consolidation at 40% of the undrained capacity, is down, as occurred at lower initial loads, instead it
slightly larger than the undrained pullout capacity of accelerated before failure (Figure 8). This behaviour
the anchor. The results of the analyses do not show is typical for the anchor under initial loading above
any significant difference between the effective 55% of the undrained capacity. To evaluate this
stresses predicted at failure below the plate for the behaviour in detailed, the performance of soil around
two cases of staged loading and first time undrained the anchor under 60% initial loading will be exam-
loading. However, the effective stress predicted at ined here.
failure above the anchor under staged loading is Figures 9 to 12 show distributions of the total verti-
larger than that predicted under the first time cal stress, the minimum principal stress, the shearing
undrained loading. This has the effect of increasing resistance and the size of the yield surface, pc, in the
the shearing strength of the soil above the anchor. soil around the anchor at a stage before the deflection
Figure 7 shows the distribution of the effective during consolidation accelerated with time. The
stresses at failure above the anchor, across a radius of stresses and the size of the yield surface have been
the anchor for both cases. normalized by the initial vertical stress that existed at
anchor level. The shearing resistance of the soil
3.2.2 Sequential loading, high initial loading around the anchor has also been normalized by the ini-
When the anchor was subjected to an initial sustained tial shearing resistance of the soil around the anchor. It
loading above 55% of the ultimate undrained cap- may be seen that beneath the anchor the vertical effect-
acity, say 60%, the behaviour of the anchor was ini- ive stress and the minimum principal stress approach
tially the same as for the loads below 55%. When zero indicating that the anchor is probably in the process
consolidation had proceeded to a time defined by of breaking away from the soil beneath.
T  0.036, i.e. about the time required for 90% of Figures 11 and 12 show that a conical zone has
consolidation, the rate of displacement did not slow formed beneath the anchor where the size of the yield

194

Copyright 2005 Taylor & Francis Group plc, London, UK


Figure 9. Distribution of the vertical effective stress, nor-
malized by ,
i in the soil around anchor during consolida- Figure 12. Distribution of the size of the yield surface, p,
c
tion under 60% initial loading at T  0.004. normalized by ,
i in the soil around anchor during consoli-
dation under 60% initial loading at T  0.004.

Figure 13. Load deflection curves predicted under high


sequential loadings.
Figure 10. Distribution of the minimum principal stress,
normalized by ,
i in the soil around anchor during consoli- surface, pc, is negligible and the soil has almost no
dation under 60% initial loading at T  0.004. shearing resistance. The size of the yield surface for
the soil just under the anchor is also negligible.
Figure 13 shows the load deflection curves for
staged loading for 2 values of the initial loading. The
curves for 60% and 80% initial loading are termin-
ated before the end of the consolidation period. It is
believed that this corresponds to the anchor separat-
ing from the soil beneath.
The change in anchor capacity when the anchor
separates from the soil beneath has been investigated
both theoretically and experimentally. Rowe & Davis
(1982) and Thorne et al. (2004) have shown that, for
undrained conditions, separation occurs when the
effective overburden pressure on the anchor is less
than 6 to 7 times the undrained shear strength of the
soil. This criterion appears to be valid for the case
considered in this paper. Experimental comparisons
Figure 11. Distribution of the normalized shearing have also been made between anchors where dissipa-
strength/sui of the soil around anchor during consolidation tion of pore pressures is prevented and where deliber-
under 60% initial loading at T  0.004. ate venting of the soil beneath the anchor has been

195

Copyright 2005 Taylor & Francis Group plc, London, UK


provided (Baba et al. 1989, Das & Singh 1994, and it has received to date. In particular, the long term
Shin et al. 1994). These comparisons show that the load capacity needs to be evaluated with more
ultimate capacity decreases from around 12 to 13 precision both experimentally and theoretically.
times the undrained shear strength to around 8 to 9 (7) Because real VLAs have tynes on the leading edge
times the shear strength if venting of the underside of the fluke it is likely that the dissipation of pore
occurs. Thus dissipation of the negative pore pressures pressures around a real anchor will be rather more
beneath the anchor might be expected to result in a rapid than for the circular anchor considered here.
capacity in the order of 60% to 70% of that with no If any drainage path exists through the anchor the
dissipation. Based on this observation and the results times for dissipation will decrease significantly.
of the analyses described here it may be concluded
that the true long term capacity of an anchor preloaded
in the manner assumed is in the range of 50% to 60% ACKNOWLEDGEMENT
of the undrained pullout capacity.
The research described in this paper was conducted as
part of the work of the Special Research Centre for
4 CONCLUSIONS Offshore Foundation Systems, established and sup-
ported under the Australian Research Councils
There are a number of conclusions and recommenda- Research Centres Program.
tions that can be made based on the results of the
study presented here.
(1) Long term loading of up to 55% of the undrained REFERENCES
pullout capacity increases the ultimate undrained
staged loading capacity by up to 20%. Baba, H. U., Gulhati, S. K. & Datta, M. 1989. Suction effect in
(2) Long term loading at loads greater than 55% of the plate anchors in soft clays. Proceedings 12th International
undrained pullout capacity is likely to result in sep- Conference on Soil Mechanics and Foundation
Engineering, Rio de Janeiro, Vol. 1, pp. 409412.
aration of the anchor from the soil beneath with a Booker, J. R. & Small, J. S. 1987. The time deflection behav-
large increase in deflection and even failure. iour of a rigid under-reamed anchor in a deep clay layer.
(3) The time for the dissipation of pore pressures and International Journal for Numerical and Analytical
failure to occur for the anchor at a relatively large Methods in Geomechanics, Vol. 11, pp. 269281.
initial loading depends on the loading level. In a Carter, J. P. & Balaam, N. P. 1995. AFENA Users Manual.
stage loading under an initial loading of 60% of Centre for Geotechnical Research, Department of Civil
the ultimate undrained capacity, failure occurred Engineering, University of Sydney, Australia.
at T  0.036 while under an initial loading of Dahlberg, R. & Strom, P. J. 1999. Unique onshore tests of
80%, it occurred at T  0.004. deepwater drag-in plate anchors. Ocean Technology
Conference, pp. 713723.
(4) The time for a real anchor to reach these time fac- Das, B. M. & Singh, G. 1994. Uplift capacity of plate
tors depends on the soil mass permeability. For a anchors in clay. Proceedings 4th International Offshore
2.5 m diameter anchor and a fairly typical value and Polar Engineering Conference, Osaka, Japan, Vol. 1,
of permeability of k  10
9 m/sec, the time for pp. 436442.
50% consolidation is about 23 days while for Merifield, R. S., Lyamin, A. V., Sloan, S. W. & Yu, H. S. 2003.
90% consolidation it will be 116 days. Three dimensional solutions for stability for plate anchors
(5) Soil permeability and uniformity are likely to vary in clay. Journal of Geotechnical and Geoenvironmental
widely from site to site. The rate of consolidation Engineering, ASCE, Vol. 29, No. 3, pp. 243253.
will depend primarily on the existence or other- Pyrah, I. C., Hird, C. C. & Tanaka, Y. 1985. Consolidation
behaviour of a single under-ream anchor. Proceedings
wise of soil structure. For example, sand or silt or 5th International Conference on Numerical Methods in
shell seams in the clay can increase the permeabil- Geomechanics, Nagoya, pp. 629636.
ity of the soil mass by several orders. The rate will Rowe, R. K. & Davis, E. H. 1982. The behaviour of anchor
also be influenced by the amount of disturbance in plates in clay. Gotechnique, Vol. 32, No. 1, pp. 923.
the soil surrounding the anchor during installation Shin, E. C., Das, R. N., Omar, M. T., Das, B. M. & Cook, E. E.
because large scale shearing could disrupt such 1994. Mud suction force in the uplift of plate anchors in
natural drainage features and decrease the perme- clay. Proceedings 5th International Offshore and Polar
ability of the soil mass. It is disadvantageous to site Engineering Conference, Vol. 1, pp. 462466.
VLAs in areas of thin sand or silt seams. Small, J. S., Thorne, C. P. & Ta, L. 1988. Effect of pore pres-
sure dissipation on the behaviour of plate anchors in clay.
(6) Based on the data presented in this paper, the Proceedings 8th International Polar and Offshore
fully drained long term capacity of the anchor is Engineering Conference, Montreal, pp. 497504.
estimated to lie between 50% and 60% of the Thorne, C. P., Wang, C. X. & Carter, J. P. 2004. Uplift capacity
undrained pullout capacity. This aspect of the of rapidly loaded strip anchors in uniform strength clay
behaviour of anchors deserves more attention than Gotechnique, Vol. 54, No. 8, pp. 507517.

196

Copyright 2005 Taylor & Francis Group plc, London, UK


An experimental and numerical study of rate effects for
plate anchors in clay

M.J. Rattley1, B.M. Lehane2, D.J. Richards1 & C. Gaudin2


1
University of Southampton, England
2
University of Western Australia, Perth, Australia

ABSTRACT: Existing design charts for evaluation of the uplift capacity of plate anchors are difficult to apply
in practice as it is often unclear if the applied loading will lead to an undrained response and, if it does, how the
average/operational undrained strength of the clay should be assessed. This paper examines explicitly the effects
of partial drainage during load application in a series of centrifuge model tests in overconsolidated kaolin and
reports predictions from complementary coupled-consolidation finite element analyses. Good agreement between
the experimental results and numerical predictions is demonstrated, and inferred normalised velocities marking
the uplift rates at which drained and undrained behaviour may be assumed are shown to be in agreement with those
inferred from variable rate T-bar penetration tests.

1 INTRODUCTION i.e. where it is assumed that the interface at the base


of the anchor cannot sustain tension due to suction.
There are many applications, both offshore and onshore, Back analyses and physical modelling studies for
where foundation systems are required to resist uplift National Grid Transco, UK (NGT-UK), of the perform-
loads. In such instances, plate anchors are often used as ance of high voltage transmission tower foundations
a solution to transmit the vertical forces to surrounding subjected to uplift loads from broken conductor events
soil at various depths. A considerable number of studies have indicated that the assumption of a breakaway
concerning the uplift performance of plate anchors have condition in clay may be overly conservative in many
been published, although until relatively recently, most instances (Woods 1999, Richards 2002). From this
experimental research focused on anchor capacity in work, the existence of a tensile capacity at the anchor
sand. Earlier studies for anchors in clay (Meyerhof & base, the assumption of totally undrained conditions
Adams 1968, Vesic 1971, Davie & Sutherland 1977, during loading (i.e. no drainage during load applica-
Das 1978) have presented analytical solutions, often tion) and the presence of viscous effects were identi-
supplemented by experimental tests. These studies fied as being three areas in need of research prior to
express the ultimate undrained uplift capacity of a strip the development of new design methods for transmis-
anchor (Fult per unit length) in a uniform clay as: sion tower foundations. This research, which also has
obvious applications for the Offshore Industry, is being
(1) conducted jointly by the University of Southampton
and the University of Western Australia.
where B is the width of the anchor, Nc is the uplift This paper presents a component of the NGT-UK
capacity factor, su is the operational undrained shear research dealing with cases where clay is used as back-
strength,  is the unit weight of the soil and H is the fill material. In particular, the results of centrifuge
depth of the anchor below the surface. tests on plate anchors in overconsolidated clay sub-
More recent studies into the capacities of anchors in jected to various pull-out rates are compared with pre-
clay include those described by Rowe & Davis (1982), dictions from numerical finite element (FE) analyses.
Merifield et al. (1999, 2001) and Thorne et al. (2004). Unlike most reported total stress FE analyses, these
These primarily consider the capacity of strip (rather employ an effective stress approach and a coupled
than circular or square) anchors and adopt a total stress consolidation model. The paper also uses the results
approach and an isotropic undrained soil strength, from high speed T-bar penetration tests to assess the
which remains constant with depth. They also focus likely performance of plate anchors under very fast
on what has been termed the breakaway condition (undrained) loading rates.

197

Copyright 2005 Taylor & Francis Group plc, London, UK


2 CENTRIFUGE TESTS

The programme of centrifuge tests described in this


paper was undertaken primarily to assess the effect of
the rate of loading on the uplift capacity of model plate
anchors in overconsolidated clay. Tests were performed
in the geotechnical drum centrifuge at the University
of Western Australia (UWA). A complete description
of this facility is given by Stewart et al. (1998).
The model anchors used in the tests were fabricated
from steel with 5 mm thick square bases and side widths
(B) of 30 mm and 45 mm; the embedment depth of
the anchors (H) was 45 mm and the anchor stem was
6 mm in diameter. Case history data assembled by
Thorne et al. (2004) indicate that the H/B range of the
anchors (1.0 to 1.5) is such that a shallow mode of
failure may be expected.

2.1 Sample preparation


Figure 1. T-bar and idealised shear strength profiles.
The kaolin clay sample was prepared in two stages.
The clay was mixed initially under a vacuum at a water
content of approximately 120% (twice the liquid limit)
to form a homogeneous slurry. It was then placed at 5 for the foundation soil). For this centrifuge sample,
20 g on top of a pre-placed 15 mm thick sand drainage equation (2) also predicts a very similar strength to
layer via a delivery nozzle in a self-levelling and non- that measured in triaxial compression tests on kaolin
erosive manner. The sample was then allowed to con- (Lehane & Gaudin 2005) i.e. sutc  suT-bar. The ide-
solidate under its self-weight at 250 g. Subsequent to alised shear strength profile used for numerical
an initial series of tests (performed as part of a separate analysis is also shown on Figure 1.
NGT-UK investigation in 2 quadrants of the sample),
the upper 50 mm of the clay layer was removed and 2.2 Uplift tests
the model anchors were founded on the new surface
of the clay at 1 g. At two test sites (one for each of the A total of seven tests were performed at uplift rates of
two plate widths investigated), the entire thickness of 0.03, 3, 30 and 100 mm/s for anchors with B  30 mm
clay was removed and replaced with a fine silica sand; and at 3 mm/s for anchors with B  45 mm. Each
these two footings were subsequently placed on the anchor was caught in flight through the use of a spe-
surface of this sand layer. cially designed hook mechanism and then released
Following placement of all the anchors, a second following completion of the test. This method elim-
layer of clay slurry backfill was poured at 20 g. Care inated the need to stop the centrifuge between tests,
was taken to allow the clay to self level over each thus avoiding cycles of swelling and re-consolidation
anchor footing and minimise possible disturbance. The prior to each test.
sample was then consolidated at 150 g before reducing
the centrifuge acceleration to the test acceleration value 2.3 Results
of 50 g. At this stage, the height of the backfill clay
layer was 45 mm above the base of the footings. A summary of the measured capacities for the 7 uplift
The use of different maximum acceleration levels tests completed is shown in Figure 2. These capacities
(150 g and 250 g) during the two phases of sample are plotted against the pull-out rate and expressed as
preparation resulted in a non-uniform undrained the ultimate capacity developed (Fu) less the footing
strength profile, as indicated by the T-bar undrained weight (W) divided by the plate area (B2). It is evident
strength (suT-bar) profiles (with NT-bar  10.5) shown on that:
Figure 1. These strengths are well represented by the 1 The anchor capacity increases with the uplift rate,
following equation: displaying approximately 60% increase as the uplift
rate increases from 0.03 mm/s to 100 mm/s.
(2) 2 Fast loaded anchors founded on clay have an appre-
ciably higher capacity than those founded on sand,
where v is the vertical effective stress and OCR is confirming the presence of tensile resistance at the
the overconsolidation ratio (3 for the backfill and anchor baseclay interface.

198

Copyright 2005 Taylor & Francis Group plc, London, UK


150 16 100
B=30mm, clay base
90
14

T-bar resistance at plate level (kPa)


B=45mm, sand base
80
B=30mm, sand base

Base resistance of plate (kPa)


12
B=45mm, clay base 70
100
(Fu -W)/B2 (kPa)

10 60

8 50

40
50 6
Partial drainage 30
in clay 4
20
Fully undrained T-bar Plate anchor
2 10
0
0.01 0.1 1 10 100 0 0
0.01 0.1 1 10 100
Uplift rate (mm/s)
Velocity (Equivalent diameters per
Figure 2. Summary of measured capacities. second)

Figure 3. Variation of T-bar and plate anchor resistance


with velocity.
3 Subtraction of the resistance offered by footings
founded on sand (for which no suction is assumed
possible) from those founded on clay indicates that
Figure 2 (for which the maximum uplift rate was
the tensile stress developed at ultimate conditions for
100 mm/s  3 D/s) at rates less than about 3 D/s. The
both the footings with B  30 mm and B  45 mm
base resistance offered by the anchors is also seen to
is about 70 kPa (at a loading rate of 3 mm/s). This
be between 6 and 7 times the T-bar strength i.e. Nc for
stress is broadly compatible with that expected from
the base component of resistance is 6 to 7.
an inverted undrained bearing capacity failure.
It is noteworthy, however, that the rate dependence
4 The rate of increase in capacity of the anchors at
of the T-bar resistance does not remain at 510% per
uplift velocities greater than or equal to 3 mm/s
log cycle but actually increases with rate reaching a
equates to approximately 6.5% per log cycle; this
value of about 40% per log cycle at penetration rates
increase can be explained entirely by viscous effects
of between 12 D/s and 20 D/s. Such strong rate effects
and consequently it may be inferred that the clay
have significant implications and are under further
response was fully undrained at a rate of 3 mm/s.
investigation at UWA.
5 The relatively low capacity of the anchor loaded at
a rate of 0.03 mm/s may be attributed to partial
drainage reducing the available base tensile capacity.
3 NUMERICAL ANALYSIS
A series of T-bar tests (with diameter, D  5 mm)
were also conducted in the centrifuge sample at pene- A series of axisymmetric effective stress analyses were
tration rates varying from 0.3 mm/s to 100 mm/s. The undertaken using the OASYS finite element program
T-bar resistance at 45 mm depth (i.e. coincident with SAFE (Oasys 2002). The model employed the same
the base of the anchors) are normalised by NT-bar  anchor dimensions as those used in the centrifuge tests
10.5 and plotted against rate on Figure 3; the rate was for 30 mm square plate anchors (but using an equiva-
normalised by the T-bar diameter to allow comparison lent plate diameter of 33.8 mm) to ensure as close a
with undrained anchor tests. comparison as possible to the test conditions i.e. the
The T-bar strengths indicate a minimum resistance analyses modelled the centrifuge tests rather than
at between 1 mm/s and 3 mm/s (0.2 D/s and 0.6 D/s), their equivalent prototypes and the soil unit weight
suggesting that penetration rates above this level are was the only parameter adjusted (factored by 50 to
fully undrained and that rate effects are controlled correspond with the centrifugal acceleration applied
solely by viscous effects. The T-bar resistances indi- in the tests). The analysis employed an isotropic elastic-
cate a comparable rate dependence to those seen in perfectly plastic constitutive model for the soil, which

199

Copyright 2005 Taylor & Francis Group plc, London, UK


was assumed to have a Mohr-Coulomb failure criter- 3.1 Analyses for partial drainage
ion (with c  0). A fully rough interface between the
To assess the effect of uplift rate on the capacity of the
anchor and soil was assumed and no limit on the suc-
model anchors due to effects of partial drainage (e.g.
tion that could develop in the clay was imposed. The
at a rate of 0.03 mm/s; see Figure 2), consolidating
finite element mesh comprised 1120, 8-node elem-
analyses were performed using the Oasys SAFE pro-
ents. Preliminary (plane strain) analyses were per-
gram. All input parameters for these analyses remained
formed using this mesh to confirm compatibility of
as specified for the undrained analysis. However, for
the derived solutions with those published for strip
these cases, a value for the soil permeability was
anchors by Rowe & Davis (1982).
required. As a first estimate, this was assumed con-
The T-bar strength (measured at 1 mm/s) was nom-
stant and equal to 1  10
8 m/s.
inally assumed to represent the undrained strength in
Each increment in a SAFE analysis represents an
triaxial compression. This was inputted to the pro-
increment of time. A total time step was inputted cor-
gram indirectly by specifying separate friction angles
responding to the total level of imposed displacement
() for the backfill and the underlying clay so that
at the anchor head, thus allowing analyses to be per-
the strength in triaxial compression was equivalent to
formed for constant uplift rates. SAFE employs an
that measured in the T-bar test. The actual response of
additional iterative process to that used in drained and
the centrifuge clay (with OCRs between 3 and 5) under
undrained analysis which runs simultaneously for the
undrained loading is expected to be similar to that
consolidating analysis. The time increment adopted for
indicated on Figure 4. Preliminary effective stress
each step of the analysis is selected by the program
undrained FE analyses confirmed the suitability of
during the iterative process, thus making it possible to
this approach.
allow large time increments at stages of the computa-
The initial FE analyses modelled (i) fully undrained
tion where the pore pressures are changing only slowly.
conditions above and below the anchor base and (ii)
For a given state, as the time increment is increased,
undrained conditions in the backfill with the anchor
so volume changes will increase and consequently
base in free draining sand (with no capacity to gener-
changes in effective stress caused by volume change
ate suction). These analyses effectively replicated the
increase. The program iterates to find time incre-
two plate tests performed at 3 mm/s; see Figure 2. The
ments which, for all elements give maximum changes
analyses incrementally increased the displacements at
in effective stress as a specified percentage of current
the top of the anchor stem up to a maximum displace-
mean normal effective stress (p). Consolidating
ment of 5 mm.
analyses were performed for uplift rates varying
A comparison of the measured and predicted per-
from 3  10
5 to 100 mm/s.
formance for these two limiting cases is shown in
Figure 5. It is evident that the capacity for the break-
away condition (i.e. no suction) and the fully bonded 3.2 Pore pressure predictions
condition (i.e. full suction) are modelled well both in
Predicted pore water pressure fields at three loading
relative terms and in absolute magnitudes. The soil
displacement rates are presented in Figure 6. It can be
stiffness values were adjusted in the analyses to obtain
observed from Figures 6(a) and 6(c) that the increase
the apparent good agreement between the predicted and
measured load-displacement response. Best agreement
was found for a (relatively low) shear modulus to T-bar
strength ratio (G/suT-bar) of 25.

q q

2su m = 6sin
3 - sin

p
h= v= p

m = 6sin
3 + sin

Figure 4. Criteria for the assignment of  in replicating Figure 5. Comparison of Fu vs. anchor displacement for
the idealised strength profile. drained and undrained failure (G/suT-bar  25).

200

Copyright 2005 Taylor & Francis Group plc, London, UK


in anchor capacity may be attributed to suctions gen- 100 kPa less than those for the drained case. Pore
erated at the base of the anchor during uplift. Average pressures at the intermediate loading rate (0.03 mm/s)
pore water pressures at failure at the base of the foot- plotted in Figure 6(b) fall between those at the
ing displaced at an undrained rate are approximately drained and undrained rates.

Figure 6. Pore pressure fields for (a) 3  10


5 mm/s (b) 0.03 mm/s and (c) 30 mm/s uplift velocities.

201

Copyright 2005 Taylor & Francis Group plc, London, UK


140 numerical predictive approach but also indicates that
the initial estimate for the permeability (k) of 1 
10
8 m/s for overconsolidated kaolin was appropriate.
120 If the back figured G value at the anchor base of
350 kPa is assumed representative of the average mod-
ulus operating during pull-out tests (see Figure 5), the
100 selected k value equates to an average coefficient of
consolidation (co) of 1  10
6 m2/s (30 m2/year).
This cv value is about five times higher than the re-
80 loading/overconsolidated value inferred from cen-
Fu (N)

trifuge piezocone dissipation tests in kaolin at


OCR  1 (House et al. 2001).
60 Randolph (2004), and others, have proposed the
use of the normalised velocity (V  vD/cv) when
Theoretical assessing the degree of partial consolidation that may
40 Experimental be expected during steady state penetration of in-situ
test tools. Conditions were essentially undrained in
the centrifuge tests when v 0.3 mm/s and hence
20 when V 10, and essentially drained for v 
0.0003 mm/s and V  0.01. These V limits are the
same as those observed by House et al. (2001) during
0 T-bar penetration tests in normally consolidated kaolin.
1E-051E-04 0.001 0.01 0.1 1 10 100
v (mm/s)
4 CONCLUSIONS
Figure 7. Predicted and measured variation of uplift
capacity with uplift rate.
Centrifuge experiments and parallel finite element
studies of relatively shallow plate anchors in overcon-
3.3 Uplift velocity analysis solidated clay have indicated that:
The ultimate uplift capacities predicted by the con- 1 Full suction can develop if the uplift normalised
solidating analyses display a clear trend of increasing velocity (V) is in excess of 10.
with an increase in uplift velocity. Predictions are pre- 2 Rate effects at very high velocities (e.g. V  100)
sented in Figure 7 and reveal the limiting cases of (low are due to viscous effects and may be far more sig-
capacity) fully drained response and (high capacity) nificant than the rate capacity increase of 5 to 20%
fully undrained-full suction response. At intermedi- per log cycle often adopted in practice.
ate, partially drained rates (e.g. 0.03 mm/s), some 3 Effective stress coupled FE analyses using a sim-
suctions, as shown on Figure 6(b), develop and the ple elastic-perfectly plastic Mohr Coulomb model
available capacity falls between that of the drained for clay with OCR of 4  1 are an expedient means
and undrained limits. It should be appreciated that the of assessing the effects of partial drainage on
FE analyses did not model viscous effects and there- anchor capacity.
fore no increase in capacity is predicted at rates in
excess of about 0.3 mm/s.
As seen in Figure 7, the predicted capacities are in ACKNOWLEDGEMENTS
good agreement with those recorded for the centrifuge
model tests at the same uplift velocities. A trend line The support provided by National Grid Transco, UK
with the following form is fitted through the predic- towards the research presented in this paper is grate-
tions on Figure 7 to indicate the general trend: fully acknowledged.

(3)
REFERENCES
where a is the upper limit for Fu (i.e. undrained limit) Das, B.M. 1978. Model tests for uplift capacity of founda-
and b, c, d and e are curve-fitting constants with the tions in clay. Soils and foundations, 18(2), 1724.
drained (lower) limit being equivalent to c/e. Davie, J.R. & Sutherland, H.B. 1977. Uplift Resistance of
The agreement between measurements and predic- cohesive soils. Journal of Soil Mechanics, Foundations
tions seen on Figure 7 confirms the suitability of the Division, ASCE 103( 9), 935952.

202

Copyright 2005 Taylor & Francis Group plc, London, UK


House, A.R., Oliveira J.R. & Randolph M.F. 2001, Evaluating Richards, D.J. 2002. The determination of dynamic resist-
the coefficient of consolidation using penetration tests. ance of transmission tower footings to uplift using cen-
Int. Journal of Physical Modelling in Geotechnics, 1(3), trifuge modelling techniques. Report No. NGDR/C/
1726. RD003/Report/F1. Dept of Civ. Eng., University of
Lehane B.M. & Gaudin C. 2005. Effects of drained pre- Southampton, UK.
loading on the performance of shallow foundations on Rowe, R.K. & Davis, E.H. 1982. The behaviour of anchor
overconsolidated clay. Proc. 24th Int. Conf. Offshore plates in clay. Geotechnique, 32(1), 923.
Mechanics and Artic Engineering, Halkidiki, Greece, Stewart, D.P., Boyle, R.S. & Randolph, M.F. 1998. Experience
June 2005 (in press). with a new drum centrifuge. Proc. Int. Conf. Centrifuge
Merifield, R.S., Sloane, S.W. & Yu, H.S. 2001. Stability 1998, Tokyo, 1, 3540.
of plate anchors in undrained clay. Geotechnique 51(2), Thorne, C.P. Wang, C.X. & Carter, J.P. 2004. Uplift capacity
141153. of rapidly loaded strip anchors in uniform strength clay.
Meyerhof, G.G. & Adams, J.I. 1968. The ultimate uplift Geotechnique, 54.(8), 507517.
capacity of foundations. Canadian Geotechnical Journal, Vesic, A.S. 1971. Breakout resistance of objects embedded in
5(4), 225244. the ocean bottom. Journal of Soil Mechanics, Foundations
Oasys 2002. SAFE Users manual, Ove Arup & Partners, Division, ASCE, 97( 9), 11851205.
London W1P, UK. Woods, R.I. 1999. The dynamic resistance of transmission
Randolph, M.F. 2004. Characterisation of soft sediments for tower footings A review of current understanding and
offshore applications. Proc. ISC-2 on Geotechnical and recommendations for further investigation. Report NO.
Geophysical Site Characterisation, 1, 209249. CBGE/9900/0. Dept. of Civil Eng., Univ. of Surrey, UK.

203

Copyright 2005 Taylor & Francis Group plc, London, UK


Vertical pullout behavior of plate anchors in uniform clay

Zhenhe Song & Yuxia Hu


Department of Civil Engineering, Curtin University of Technology, WA, Australia

ABSTRACT: With the development of offshore engineering, plate anchors such as SEPLA (Suction Embedded
Plate Anchor) and VDPA (Vertically Driven Plate Anchor), are relatively new innovations that make use of different
installation methods to embed a plate-like anchor into the soil. In this paper, the pullout behaviours of strip
anchor and circular anchor have been studied numerically in uniform clay. There are small strain analysis and
large deformation analysis performed. In small strain analysis, the plate anchor is wished to place for its pullout
capacity investigation. In large deformation analysis, a continuous pullout of anchor is simulated. Plate anchors with
fully attached base and vented base are considered. It is found that when soil unit weight is considered in the
analysis, the overburden pressure adjacent to the plate plays an important role on the embedment depth where plate
base loses its suction and soil separates from it.

1 INTRODUCTION circular anchors in soft clay to determine the breakout


factors and the variation of suction force with embed-
In recent years oil and gas mining has moved into ment ratio. But the tests were 1 g floor model test, and
increasingly deeper water in search of undeveloped the gravity effect of soil was limited in the tests.
fields. As water depths approach and exceed 500 m, Rowe & Davis (1982) carried out a numerical
conventional offshore foundation systems become study of anchors embedded in clay in both immediate
inefficient and ineffective in stabilizing platforms and (vented) and no breakaway (fully attached) of the soil
floating production storage units. Consequently, many behind the anchor. In their study, an elasto-plastic finite
types of anchoring systems are being developed and element analysis was used to determine the breakout
used in order to withstand large mooring forces. The factors for horizontal and vertical strip anchors and
SEPLA (Suction Embedded Plate Anchor) and VDPA horizontal circular anchors. However, when their results
(Vertically Driven Plate Anchor) are two such systems. were compared with the recent lower bound solutions,
These kinds of plate anchors can be used in a wide the finite element results showed a great underestima-
range of soils, in shallow to ultra deep water, and can tion due to imposing arbitrary displacement criterion
operate with extremely high capacity under varying for ultimate breakout factor (Merifield et al. 2001).
loading conditions and directions. Papers by Merifield et al. (1999a, 1999b, 2001,
Over the last four decades a number of researchers 2003) used finite-element formulation of limit analysis
have proposed approximate, inexactly correlating tech- to reach results of upper and lower bound limit theorem
niques to estimate the uplift capacity of plate anchors solutions. The analyses in these papers had broad
in various types of soil. These results can be found in scopes that covered different anchor shapes at various
the works of Vesic (1971), Meyerhof & Adams (1968), embedment ratios in both homogeneous and non-
Meyerhof (1973), Das (1978, 1980) and Das et al. homogeneous soils. Based on the lower bound finite
(1994). Most of the results from these studies of element results, a straightforward numeric way to evalu-
anchors in clay consist of either simple analytical solu- ate the net ultimate uplift capacity of plate anchors
tions or are derived empirically from laboratory scale has been proposed. However the suction force of the
model tests. anchors in these studies has not been considered, and
Das (1978, 1980) have suggested procedures, based the soil flow mechanism and soil weight effect were
on model laboratory test, to estimate the ultimate studied separately. This separation might not repre-
uplift capacity of square, rectangular, and strip anchors sent field situation where soil flow mechanism and
embedded vertically or horizontally in clay. These tests soil weight act together. This is discussed more in this
were mostly performed in soft clays with a limited paper.
number of tests performed in stiff clays. Das et al. At the same time, Martin & Randolph (2001) have
(1994) conducted a number of laboratory tests on reported upper bound and lower bound solutions for

205

Copyright 2005 Taylor & Francis Group plc, London, UK


an ultra thin plate with no-breakaway (fully attached).
It was assumed that there was no restriction on the Z
development of tensile stress, whether in the soil itself H
or on the base of the plate. When the upper bound
solutions reached the lower bound solutions, exact
solutions were resulted as Nc  12.42 for smooth cir-
cular plate anchor and Nc  13.31 for rough circular B(D)
plate anchor. su
This paper aims to investigate continuous pullout Attached or Vented
behaviours of strip and circular plate anchors in uni- Homogenous soil
form clay. In numerical analysis, the breakout factor of
pre-embedded anchor and continuous pullout anchor E=500su
are studied. Anchor at different embedment with fully
attached base and vented base are set up to study the
soil flow mechanism and soil weight effect together.
Figure 1. Problem definition.

2 RESEARCH METHODS material with Tresca yield criterion. Poisson ratio


  0.49 and friction and dilation angles     0
2.1 H-adaptive RITSS method are set to simulate undrained condition. Both strip
anchor and circular anchor are analysed with different
Although small strain analysis can provide results on embedment ratio H/B or H/D up to 8, where H is plate
clearly defined soil conditions and reliable bearing embedment depth, B is the width of strip plate and
capacity, the continuous pullout analysis using FE D is the diameter of circular plate. Soil is homogeneous
method can simulate the continuous pullout behaviour with E/su  500, where E is Youngs modulus and su
of the plate anchor. is the undrained shear strength of soil. Soil weight is
Remeshing and Interpolation Technique with Small not included in these pre-embedment analyses, thus the
Strain model (RITSS), which has been proved work- FE results can be reasonably compared with any exist-
ing well for simulation of large deformation problems ing small scale laboratory test data. This is because the
of soil (Hu & Randolph 1998a,b), is used to simulate overburden pressure of soil is insignificant in small
the continuous pullout process of anchor. As the name scale (or 1 g) laboratory test, when the anchor size is
implies, this method is to use small strain model to reduced about 100 times to its prototype.
simulate large deformation of soil with frequent
remeshing. The detailed development of the method
can be found in Hu & Randolph (1998a,b). 3.1 Strip plate anchor
The strip plate is considered as 4 m wide (B) and
2.2 Soil-anchor attachment 0.2 m thick. Both fully attached (A) base and vented (V)
base are considered. The fully attached base can sus-
In this paper, strip and circular plate anchors are inves- tain a large suction force without separation, and the
tigated on its pullout capacity with two anchor base vented base cant sustain any suction force at all, thus
conditions: fully attached (A) and vented base (V) in seperation will occur when tension exists. A rough
clay. In the fully attached case, it is assumed that the soil interface between soil and plate is assumed.
anchor interface can sustain enough tension to ensure The FE results are shown in Figure 2 as a term
that the anchor remains full contact with soil all the time of breakout factor. The breakout factor can be calcu-
during pullout. In the vented case, which is equivalent lated as:
to immediate breakaway, the opposite is assumed.
Thus upon loading, there is no adhesion or suction
between soil and anchor allowed. Therefore, the suction (1)
impact on anchor pullout capacity can be examined.

where qu is the ultimate pull-out pressure, Qu is the


3 SMALL STRAIN ANALYSIS ultimate pull-out force and A is the plate area (per
metre run for strip plate). The current FE results have
In pre-embedment analysis, anchor is wished to place been compared with existing laboratory test data (Das
without the consideration of soil disturbance during 1980, Rowe et al. 1982), analytical solutions (Merifield
pullout. Soil is simulated as an elastic-perfectly plastic et al. 2001) and FE results (Rowe & Davis 1982).

206

Copyright 2005 Taylor & Francis Group plc, London, UK


14 16

12 14

10 12
Rowe(A) Merriffiled UB Current(A)
10
8
Nc0 Nc0 Current(A)
8
6 Current(V)
6 Das test(A)
4 Current(V) Das test(V)
Das (1980) L/B=5 4 Martin
Merrifiled LB
2 Rowe (1982) L/B=5 Rowe(V) Meriffield LB
Rowe (1982) L/B=8 2 Rowe(V) k4
0 0
0 2 4 6 8 10 0 2 4 6 8 10
H/B
H/D

Figure 2. Bearing capacities for rough strip anchors (small


Figure 3. Bearing capacities for rough circular anchors
strain).
(small strain).

For plate anchor with fully attached base (A), the Figure 3 shows the breakout factor results of cur-
current FE results have the same trend of the FE rent FE analysis with other existing data. For circular
results from Rowe & Davis (1982). However, Rowe anchor with full attachment (A), the current FE results
and Daviss results are 5% below the current FE are compared with the laboratory test data (Das, et al.
result. In Rowe & Davis FE analysis, failure was con- 1994) and the exact analytical solution 13.31 (Martin &
sidered to have been reached at a plate displacement Randolph 2001) for deeply embedded rough plate
that was a selected multiple of that which would have anchor. The current FE result catches the same trend
been reached had conditions remained entirely elastic. of the laboratory test data. The limit breakout factor
This definition is arbitrary in terms of the choice of Nc  14.33 for deeply embedded anchor at H/D 2.
the multiple to be used. The ultimate capacity might This value is 7% higher than the exact solution 13.31.
be predicted when the soil was still in the range of con- This is due to the thickness effect with rough plate
tained plastic deformation prior to ultimate collapse. analysed.
For plate anchor with vented base (V), a large For circular anchor with vented base (V), the current
underestimation can be observed from Rowe & Davis FE results agree well with the laboratory test data
FE resutls (1982) when embedment ratio (H/B) is (Das, et al. 1994) When H/D 5. It shows a similar
greater than 3. The reason of under estimation is again trend with the lower bound solution from Merifield
due to the truncation of ultimate load as explained et al., (2003). The limit breakout factor for deeply
above. However, all the other results stay well together embedded anchor is reached at H/D  6. The FE results
until H/B  6. When H/B is larger than 6, a limit from Rowe & Davis (1982) again gave an underesti-
breakout factor is achieved by current FE results. The mation about 60% due to the truncation in their pullout
limit breakout factor was reached by Dass (1980) capacity estimation.
laboratory test data at H/B  8. This may be due to
the soil weight effect. Although the soil weight effect
4 CONTINUOUS PULLOUT ANALYSIS
was insignificant in small scale test, it could still affect
the results slightly. In the laboratory test by Rowe &
In continuous pullout analysis, strip and circular
Davis (1982), there was little difference observed in
anchors are studied. Anchors with fully attached base
pullout capacity for anchors with aspect ratios between
and vented base are simulated to investigate the soil-
5 and 8. Thus a conclusion was reached that an rect-
plate separation behaviour. The soil is homogenous
angular anchor with aspect ration larger than 5 could
material and modelled with friction and dilation angles
be taken as a strip anchor. This appears reasonable
   0, undrained shear strength varying from
since the test result with L/B  5 stays close to other
3 to 25 kPa, and the ratio of Youngs modulus to shear
strip anchors results.
strength E/su  500. The anchor width for strip
anchor and anchor diameter for circular anchor are set
3.2 Circular plate anchor up as 4 m, 6 m and 8 m. The saturated soil self-weight
is taken as 17 kN/m3 or effective soil unit weight
The diameter for the circular plate anchor considered
7 kN/m3 for soil under water.
is 4 m and thickness 0.2 m. Both fully attached (A)
base and vented (V) base are considered. The current
4.1 Flow mechanism
FE results have been compared with existing labora-
tory test data (Das et al. 1994) and analytical solutions Figure 4 displays the development of soil flow mech-
(Martin & Randolph 2001, Merifield et al. 2003). anism during continuous pullout of strip anchors with

207

Copyright 2005 Taylor & Francis Group plc, London, UK


Figure 4. Flow mechanisms of strip plate anchors with (H/B)initial  5.

fully attached base (A) and vented base (V), when the For all the other cases of strip anchor and circular
plates are embedded at H/B  5 initially. The anchor anchor, the development of the flow mechanism shows
width is B  5 m and the soil strength is su  4.5 kPa. the same process from deep failure mechanism to shal-
Figure 4(a) depicts the soil flow around a strip anchor low failure mechanism. When shallow failure mech-
with fully attached base when the plate is pullout to anism is formed, the soil-plate separation occurs for
H/B  4. The soil flow mechanism is fully localised. anchors with vented base. The anchor embedment
The flow mechanism in Figure 4(b) for vented depth where soil-plate separation starts (separation
case shows an identical flow mechanism to Figure depth, Hs) is discussed in the following section.
4(a). This means there is no soil-plate separation.
Once the anchor has reached the embedment depth of
4.2 Separation depth (Hs )
H/B  0.25 (Figs 4(c), 4(d)), a soil-plate separation is
observed with vented base. However, the soil still fol- From the preliminary study of Mehryar & Hu (2002),
lows the anchor with fully attached base. At the same it has been reported that v/su  57 was critical on
time, the anchor is very close to the soil surface, thus whether separation will occur during pullout, where
a surface heave has formed. Therefore, it is apparent only one plate diameter (D  4.5 for circular plate) was
that the pullout capacity factor in small strain analysis considered. In order to study the plate width/diameter
cant take the soil heave into account. Moreover, the effect together with the overburden-strength ratio effect,
embedment depth where soil-plate separation occurs a stability number D/su is introduced in searching a
is affected by soil strength and soil self weight unique presentation of separation depth.
together. This cant be examined in the small strain Figure 5 displays the breakout factor Nc of circular
analysis with weightless soil. plate with D  4.5 m in uniform soil. In Figure 5(a),

208

Copyright 2005 Taylor & Francis Group plc, London, UK


Figure 5 (a). Breakout factor for su  3 kPa.

Figure 6. Soil failure mechanism when Hinitial Hs.

separation depth is found at Hs/D  0.55. For anchors


with H/D  2 and 1, same response is observed as
that in Figure 5(a). However, when anchor was initially
embedded more shallowly than the separation depth
((H/D)initial  0.5), the breakout factor of attached
anchor increases at the beginning of the pullout and
joins the response of the deeply embedded anchor with
further pullout. However, for vented anchor, the sep-
aration developed right after the pullout action with
Figure 5 (b). Breakout factor for su  5 kPa. lower breakout factor obtained. The breakout factor at
the immediate separation for an initially shallowly
soil strength was set up as su  3 kPa. The anchor with embedded anchor (Ncsh) comes from the shear resist-
attached base and anchor with vented base are analysed ance of the shear plan formed above the anchor
with two initial embedment ratios (H/D)initial  1 and (Qushear) and the soil weight above the anchor (Qusoil)
0.5. From the results of anchor with attached base of (Fig. 6). This can be expressed as the following:
(H/D)initial  1, it can be seen that the breakout factor
is constant (Nc  14) until it starts to decrease at the
embedment of H/D  0.6. This marks the transition
point from a deeply embedded anchor to a shallowly
embedded anchor. When anchor is initially embedded at
(2)
a shallower depth with (H/D)initial  0.5 with attached
base, the breakout factor increases at the initial pullout or for strip plate anchor
until it reaches the value of the deeply embedded
anchor, and coincides with it in the rest of pullout.
When anchor base was vented, the breakout factor (3)
shows exactly the same response at the initial pullout
until the embedment reaches H/D  0.3. This indicates
the depth of soil separation from anchor base, hence The higher Nc value obtained for deeply embedded
Hs/D  0.3 for this case. With further pullout, the anchors after separation point is due to the soil heave
breakout factors decrease linearly and more rapidly formed (Fig. 4). From Figure 5, it is clear that the
than that of attached anchor. Although the anchors depth of soil-anchor base separation is critical to the
were initially embedded in different depth, they both pullout behaviour of anchor. For an anchor initially
were initially embedded deeper than the separation embedded deeper than the separation depth, the anchor
point, thus the separation happened at the same depth. base suction can be lost when separation occurs. For
Figure 5(b) shows the pullout anchor analysis with an anchor initially embedded shallower than the sep-
three initial embedment (H/D)initial  2, 1 and 0.5 in aration depth, the separation happens straight after
the soil of su  5 kPa. The transition depth from deep the pullout action with the maximum breakout factor
anchor to shallow anchor is found at H/D  0.9. The can be calculated by Equation 2 for circular anchor or

209

Copyright 2005 Taylor & Francis Group plc, London, UK


5 the exclusion of soil weight. The strip anchor reaches its
SW17 B=4.0m(strip)
SW17 B=6.0m(strip)
deep failure mechanism at H/B  3 for attached base
4
SW7 B=4.0m(strip) and H/B  6 for vented base. The circular anchor
3
Fitted strip
SW17 D=4.5m (circular)
becomes deeply embedded at H/D  2 for attached
Hs/B SW7 D=4.5m (circular) base and H/D  6 for vented base.
(Hs/D) SW17 D=6.0m (circular)
2 SW17 D=8.0m (circular) In large deformation analysis, soil flow mech-
SW17 D=4.0m (circular) anism shows two stages for anchor with fully attached
1 Fitted circular
base: deep embedment, where the pullout capacity
of anchor is constant (Nc  14.33); shallow embed-
0
0 10 20 30 40 50 60 ment, where the pullout capacity of anchor is decreas-
B/su(D/su) ing during pullout (Nc 14.33). However, for anchor
with vented base, there is a third stage after reaching
Figure 7. Separation depth of strip anchor and circular the shallow embedment: soil-plate separation. Once
anchor with vented base. the separation occurs, the pullout capacity decreases
dramatically. Thus it is a critical point separation
depth (Hs).
Equation 3 for strip anchor. For all the analyses with The separation depth of strip and circular anchors
different anchor size and soil strength, the separation can be predicted by Eqs 4 and 5. The separation depth
depth against stability number D/su is plotted in ratio (Hs/B or Hs/D) is decreasing with increasing sta-
Figure 7. It is apparent that the separation depth is bility number (B/su or D/su). For any anchor ini-
decreasing with decreasing su, thus increasing D/su. tially embedded deeper than Hs, the pullout capacity
The fitted curve for circular plate anchor gives a for- can be predicted by anchor with fully attached base,
mula shown below: and the pullout capacity starts to drop dramatically
after the Hs is reached. For any anchor initially
embedded shallower than Hs, the maximum pullout
(4) capacity can be calculated by Eqs. 2 and 3 due to the
formation of vertical shear plane above the plate.

with correlation coefficient R2  0.99. This formula ACKNOWLEDGEMENTS


can be easily used to predict the separation depth when
soil conditions are known. However, care should be The research presented here is supported by the
taken when choosing  as an input. Since for satu- Australian Research Council through the ARC dis-
rated clay,  is 17 kN/m3 due to the air infill between covery scheme (DP0344019). This support is grate-
the plate base and soil after plate-soil separation, and fully acknowledged.
 is 7 kN/m3 due to the water infill after plate-soil
separation for under water condition. These two  val-
ues give a large difference in estimated Hs. Thus, the
REFERENCES
soil unit weight needs to be chosen carefully accord-
ing to field situations. Aubeny, C. P., Murff, D. J. and Roesset, J. M. 2001.
Similar observations have been obtained for strip Geotechnical Issues in Deep and Ultra Deep waters, 10th
anchor analysis. The separation depth versus B/su is Int. Conf. On Computer Methods and Advances in Geo-
shown in Figure 7, with correlation coefficient R2  mechanics, Tucson, USA.
0.99. It can be seen that it has the same trend as that Das, B. M. 1978. Model tests for uplift capacity of founda-
of the circular plate anchors. tions in clay, Soils and Foundations, Japan, 18(2), 1724.
Das, B. M. 1980. A procedure for estimation of ultimate
capacity of foundations in clay, Soils and Foundations,
(5) Japan, 20(1), 7782.
Das, B. M., Shin, E. C., Dass, R. N. and Omar, M. T. 1994.
Suction force below plate anchors in soft clay, Marine
Georesources and Geotechnology, 12, 7181.
5 CONCLUSIONS Hu, Y. and Randolph, M. F. 1998a. A practical numerical
approach for large deformation problems of soil, Int.
J. Numer. And Anal. Methods in Geomech., 22, 327350.
The pullout behaviours of strip and circular plate Hu, Y. and Randolph, M. F. 1998b. H-adaptive FE analysis
anchors in uniform clay are investigated in both small of elasto-plastic non-homogeneous soil with large deform-
strain analysis and large deformation analysis. ation, Int. J. Compu. Geomech., 23(1/2), 6184.
In small strain analysis, the pullout capacity results Martin, C. M. and Randolph, M. F. 2001. Applications of the
agree well with the small scale laboratory tests due to lower and upper bound theorems of plasticity to collapse

210

Copyright 2005 Taylor & Francis Group plc, London, UK


of circular foundations, Proc. 10th IACMAG Conf., Merifield, R. S., Lyamin, A., Sloan, S. W. and Yu, H. S.
Tucson, USA, 2, 14171428. 2003. Three-dimensional lower bound solutions for
Mehryar, Z., Hu, Y. and Randolph, M. F. 2002. Pullout stability of plate anchors in clay, Journal of geotechnical
capacity of circular plate anchor in clay FE analysis, Proc. and geoenviromental engineering, 129(3), 243253.
8th Int. Symp. On Numerical Models in Geomechanics Meyerhof, G. G. and Adams, J. I. 1968. The ultimate uplift
(NUMOG), Rome, Italy. capacity of foundations, Canadian Geotechnical Journal,
Merifield, R. S., Lyamin, A., Sloan, S. W. and Yu, H. S. 5(4), 225244.
1999a. Three dimensional lower bound solutions for the Meyerhof, G. G. 1973. Uplift resistance of inclined anchors
stability of plate anchors in clay, Research report and piles, Proc. 8th Int. Conf. On Soil Mechanics and
(178.02.1999). Foundation Engn., Moscow, 2(1), 167172.
Merifield, R. S., Lyamin, A., Sloan, S. W. and Yu, H. S. Rowe, R. K. and Davis, E. H. 1982. The behaviour of anchor
1999b. Three dimensional lower bound solutions for the plates in clay, Geotechnique, 32(1), 923.
stability of plate anchors in clay, Proceedings of the 7th Vesic, A. S. 1971. Breakout resistance of objects embedded
International Symposium on Numerical Models in in ocean bottom, J. Soil Mechanics and Foundation Div.,
Geomechanics (NUMOG), Graz, 481487. ASCE, 97(9), 11831205.
Merifield, R. S., Sloan, S. W. and Yu, H. S. 2001. Stability of
plate anchors in undtrained clay, Geotechnique, 51(2),
141153.

211

Copyright 2005 Taylor & Francis Group plc, London, UK


Pullout capacity of circular plate anchors in double-layered clays

Dr. Jun Liu


Lecturer, School of Civil and Hydraulic Engineering, Dalian University of Technology

Dr. Yuxia Hu
Senior Lecturer, Department of Civil Engineering, Curtin University of Technology

Liling Wu
Research Student, School of Civil and Hydraulic Engineering, Dalian University of Technology

ABSTRACT: Plate anchors are commonly used as foundations for structures that require uplift resistance.
Although the pullout capacity of plate anchors in clay has been studied extensively, most of the studies were
concerned with single layer soils. Natural soils are often formed in discrete layers while the understanding of
plate anchor performance in multiple layer soils is relatively poor. In this study, the pullout capacity of a circu-
lar plate anchor with pre-embedments and the continuous pullout performance of the anchor are investigated
numerically using an elasto-plastic finite element method. In the continuous pullout analysis the Remeshing
and Interpolating Technique with Small Strain model (RITSS) has been used. The plate anchor is embedded in
a stiff clay layer overlain by a soft clay layer. It is found that the pullout capacity of plate anchor depends on the
distance between the plate and the soil layer boundary and the soil layer strength ratios. The critical distance is
where the pullout capacity of plate starts to decrease when it moves closer to the interface of the soil layers. This
critical distance decreases with increasing soil strength ratio (top soil strength/bottom soil strength). There are
two stiff soil cones formed at both upper and lower sides of the plate as it is pulled into the top soft soil layer
from the underlaying stiff soil layer. Comparing with the findings for plate in single soil layer with top soil
strength, due to the stiff soil cones formed, an increased pullout capacity and delayed soil-plate separation are
observed at soil strength ratios less than 2/3.

1 INTRODUCTION properties. In contrast to the experimental studies, very


few numerical analyses have been performed to investi-
The uplift capacity of anchors in soils has been stud- gate the pullout capacity of anchors in clay. Merifield
ied extensively in the past. However, the majority of (2001) applied upper bound and lower bound finite
past research has been experimental and aimed at element method to investigate the pullout capacity of
anchors embedded in a single layer of homogeneous anchors in clay. Others used displacement finite element
clay. Common anchor shapes studied include square, methods to examine the behavior of anchors in clay
circle or rectangle shapes. Most of available anchor (Ashbee 1969, Davie & Sutherland 1977, Dewaikar
design guidelines suggest using rectangle anchors as 1988). All these studies, however, were limited to small
a base anchor shape and then obtaining the results of strain analysis of single layered, isotropic and homoge-
other anchor shapes through empirical shape factors. neous soils. However, natural soils are often formed in
The existing results on the pullout capacity of discrete layers having significantly different properties.
anchors in clay were either derived from simple analyt- For anchors in double-layered clays, it is expected that
ical solutions or from laboratory model tests. These the bearing capacity depends on the soil strength ratio,
results can be found in the works of Adams & Hayes and the distance between the plate anchor and the soil
(1967), Meyerhof & Adams (1968), Vesic (1971, 1972), layer interface. In design practice, methods for calculat-
Meyerhof (1973), Das (1978, 1980), Ranjan & Arora ing the bearing capacity of foundations on multiple
(1980), and Das et al (1985a, 1985b, 1989). The pullout layer soils range from averaging the strength parameters
capacity of anchors is typically expressed in terms (Bowles 1988) to semi empirical approaches based on
of a breakout factor, which is a function of the anchor experimental studies (Meyerhof & Hanna 1978). Finite
shape, embedment depth, overburden pressure and soil element method has also been applied to this problem

213

Copyright 2005 Taylor & Francis Group plc, London, UK


Soil Surface using the AFENA finite element package, Carter &
Balaam 1995) and * is the strain recovered by SPR
(Superconvergent Patch Recovery, Zienkiewciez & Zhu
Layer 1 (su1) 1993). This measurement of error is easy to use due to
d its non-dimensional character.
H Once the discretisation error is obtained using
Layer 2 (su2)
Equation 1, the existing mesh can be refined. However,
t it is not straightforward to set up a unique criterion for
mesh refinement in plastic analyses (Zienkiewiez et al.
D 1995). Instead, the minimum element size (hmin) is
specified. The following procedure has been found to
work well for the refinement:
z 1 Generate a coarse mesh using a specified element
density function, such as the exponential function
Figure 1. Plate in layered soil: soft over stiff. (Hu & Randolph 1998a);
2 Compute an initial strain field h using AFENA
(Carter & Balaam 1995);
(Burd & Frydman 1997, Merfield 1999, Wang & Carter 3 Calculate the recovered strain field * using SPR;
2002). However, these studies were limited to footings 4 Estimate the strain error ei* in each element over
with rigid shafts. the mesh using Equation 1;
This paper studies the pullout capacity of pre- 5 Half the size of the elements in regions where
embedded and continuous pullout plate anchor in ei* 0.5emax;
double-layered clays using elasto-plastic finite element 6 Use this updated mesh density to create a new mesh;
method. The soil profile has been set up as a soft soil 7 Check if the minimum element size hmin has reached
layer (Layer 1) overlaying a strong soil layer (Layer 2), the specified value hmin;
with soil strength ratio (su1/su2) varying from 0.1 to 0.8 8 If yes, the mesh refinement stops; otherwise go back
(in which a uniform soil has a strength ratio of 1). The to step 2 until the specified minimum element size is
case set-up is illustrated in Figure 1. In the analyses reached.
presented in this study, the plate is initially embedded
in the underlaying strong layer 6D (D is plate diameter) 2.2 H-adaptive RITSS method
from the soil surface.
In continuous pullout analysis, the h-adaptive RITSS FE
method is used. As the name implies, incremental small
2 NUMERICAL METHOD strain analyses are used to solve large deformation prob-
lems by frequent remeshing and interpolation. The
2.1 H-adaptive finite element method stress field and the soil properties of the current mesh
are interpolated into a new mesh after each remeshing.
In finite element analyses, accuracy is often a major
In h-adaptive RITSS analyses, the above 8-step refine-
concern. It is well accepted that adaptivity is an effective
ment approach is used in the initial mesh generation
way to reduce the error created by the discretisation of
before remeshing starts. During each remeshing, the
the computational domain. For an optimal mesh using
mesh density field is only checked once to generate
h-adaptivity (which reduces the element sizes in a high
a new mesh according to the updated domain boundary,
error region), the element sizes are adjusted according
since the mesh is already close to optimal after the ini-
to the error distribution. Thus the discretisation error has
tial refinements. The following procedure is used for
to be estimated.
large penetration analyses:
Hu & Randolph (1998b) developed an SPR-strain
error estimator that has been demonstrated working 1 Generate an optimal mesh for the initial computa-
very well in elasto-plastic analysis, with the error in tion domain using the 8-step refinement;
element i being defined as 2 Calculate the FE results with N steps of small strain
analysis using AFENA;
3 Update the domain boundary, check the mesh dens-
(1) ity field using the error estimated and generate a
new mesh;
4 Check if the displacement of the plate has reached
where i is the area of element i in two-dimensional the required value; if not, go to step 2; otherwise stop.
analyses (or volume of element in three-dimensional Steps (2) to (4) are repeated until the desired dis-
analyses), h is the strain from the FE solution (here placement is obtained. The number of steps of small

214

Copyright 2005 Taylor & Francis Group plc, London, UK


strain analysis between each remeshing is designed in 15
such a way that the total penetration of N steps remains
in the small strain range and is less than a half of the 12 single layer 13.31
minimum element size (0.5hmin). su1/su2
9
0.1

Nc
0.2
3 ANALYSIS APPROACH AND FINITE 6 1/3
ELEMENT MODELS 0.5
2/3
3 0.8
The clay is modelled as an undrained elasto-plastic Martin(13.31)
material with Tresca yield criterion. The anchor is 0
assumed to be rigid. For all the analyses, the soil-plate 0 1 2
interface is simulated using elasto-plastic nodal-joint H/D
elements. The diameter of the plate anchor D is 5 m, and
the thickness of the plate anchor t is 0.2 m. Six-node tri- Figure 2. Pullout capacity factor in layered soil.
angular elements are used and the smallest element size
of the mesh adjacent to the anchor (hmin) is less than
0.1 m (hmin/D 2%). The anchor is considered rough in 1.4
all cases. The tension strength between the soil and the 1.2
plate is zero, thus the soil is allowed to separate from the 1
anchor when the nodal joint elements are under tension.
Hcr/D
0.8
Since undrained conditions are assumed, friction and 0.6
dilation angles   0, and a uniform stiffness ratio 0.4
E/su 500 (where E is Youngs modulus and su is the 0.2
undrained shear strength) are specified. The saturated 0
soil self-weight has been taken as  17 kN/m3. The 0 0.2 0.4 0.6 0.8 1
value of Poissons ratio is   0.49. su1/su2

Figure 3. Critical embedment in lower stiff layer.


4 RESULTS AND DISCUSSION

The pullout capacity of a plate anchor can be expressed and reaches to an ultimate value at a critical embedment
by a pullout factor Nc: (Hcr). This means that if a plate is embedded at a depth
deeper than Hcr, the pullout capacity is only dependent
(2) on the lower stiff soil layer with an ultimate Nc  13.86,
which is very close to the exact solution by Martin &
Randolph (2001) (Nc  13.31) for an ultra thin plate in
where F is uplift force, A is plate anchor area and su is homogeneous soil. If the plate is embedded at a depth
the undrained soil shear strength at the anchor pos- closer to the layer boundary (H Hcr), a reduction in
ition. For all cases, the plate anchor is initially embed- pullout capacity is observed and the effect of top soft
ded in the lower stiff soil layer with total embedment soil layer can not be ignored.
ratio d/D  6. By changing the position of the layer Figure 3 shows the critical embedment ratio (Hcr/D)
boundary, the pullout capacity verses the anchor dis- under different soil strength ratios (su1/su2). It is appar-
tance to the layer boundary (H) is studied (Fig. 1). ent that the critical embedment ratio decreases with
increasing soil strength ratio. The influence region of
the top soft soil layer is deeper when soil strength ratio
4.1 Small strain analysis
is small. The influence region reaches to 1.2D below the
The undrained shear strength of the lower stiff soil layer soil layer boundary when su1/su2  0.1. With lower soil
(Fig. 1) is chosen as su2  25 kPa. Soil strength ratio, strength ratio, the pullout capacity reduction is much
su1/su2 is selected as 0.1, 0.2, 1/3, 0.5, 2/3 and 0.8 by greater (Fig. 2).
varying su1. Figure 2 shows the results of the pullout
capacity factor (Nc) vs embedment ratio (H/D) at differ-
4.2 Continuous pullout analysis
ent strength ratios su1/su2. Soil weight is not considered
in the analysis, since it has no effect on the small strain Continuous pullout analysis is carried out with soft clay
analysis. overlaying stiff clay. The RITSS is utilised in this analy-
It is clear from Figure 2 that the pullout capacity fac- sis. The method has been demonstrated to work well
tor Nc increases with increasing embedment ratio (H/D), in soil deformation analysis with single layer soils

215

Copyright 2005 Taylor & Francis Group plc, London, UK


1 0
0.1x13.31

0.2x13.31

1/3x13.31
1/3x13.31

0.5x13.31

2/3x13.31

0.8x13.31

Nc=13.31
0
Stage 4 -1

-1

su1/su2=0.5
-2 Hinitial/D = 2
-2 Stage 3
Separation

0.2

1/3
d/D

d/D
-3
-3
su1/su2=1
Layer Boundary
-4
1/3

2/3
0.1

0.2

0.5

0.8

-4

-5
Stage 2 Hinitial/D = 1
-5
Layer Boundary Layer Boundary
Stage 1
-6
0 2 4 6 8 10 12 14 -6
Nc value 0 2 4 6 8 10 12 14
Nc
Figure 4. Continuous pullout responses in layered soils at
Hinitial/D  1. Figure 5. Continuous pullout responses with Hinitial/D 1
and 2.

(Hu & Randolph 1998b). Here, the method has been four stages, which are indicated for su1/su2  0.5 case in
extended to double layer soils. Figure 4: Stage 1 Stable in the lower layer (H  Hcr):
Figure 4 displays the FE results with Hinitial/D  1.0 the Nc values are stabilised in the lower stiff soil layer as
and d/D  6. The soil strength ratio varies from 0.1 to 1 the capacity in a single layer soil; Stage 2 Passing
with su2  10 kN being kept constant. Soil unit weight the layer boundary (H Hcr): when the plate anchor is
is taken as  17 kN/m3. A rough interface between moving closer to the layer boundary, the Nc values begin
the plate and the soil is assumed. There is no tension to decrease until the plate anchor is pulled into the top
strength considered between the plate and soil. Thus layer completely; Stage 3 Stable in the top soft layer:
separation between soil and plate is allowed when the the Nc value remains constant within top soil layer.
nodal joint elements are under tension. However, they are higher than the values of single soil
From Figure 4, it can be seen that when the soil layer ((su1/su2)  13.31), especially at lower strength
strength ratio su1/su2  0.1, the initial embedment is less ratios (su1/su2 2/3). This is due to the stiff soil cones
than the critical embedment (Hinitial/D Hcr/D). Thus a attached to the plate, which is explained next; Stage 4
reduced pullout capacity factor is predicted from the Separation: the plate anchor is approaching the soil sur-
beginning of the pullout in comparison with the pull- face. Thus a separation between plate base and soil
out capacity factor of a single layer of soil (su1/su2  1). underneath it occurs due to the lose of suction force.
However, for all the other strength ratios, Nc starts with The higher pullout capacity in stage 3 can be demon-
13.31 obtained from single soil layer. This Nc value is strated in Figure 6. At the initial pullout stage (Fig.
the same as the exact solution for an ultra thin plate, 6(a)), the layer boundary starts to deform. Once the
however, it is lower than the one from small strain analy- plate passes the layer boundary (Fig. 6(b)), the top soft
sis (Nc  13.86). This may be due to the soil weight soil flows to the plate base with a soil cone of the bot-
considered in this continuous pullout analysis, thus a tom stiff soil being left at the top side of the plate. With
buoyancy effect is expected due to the plate thickness. further pullout another soil cone of stiff soil is formed at
Figure 5 shows the plate pullout response with initial the bottom side of the anchor (Fig. 6(c), (d)). Thus two
embedment ratios Hinitial/D  1 and 2 respectively. It is soil cones of stiff soil are attached to both sides of the
clear that after the plate reaches its critical embedment plate and moving with the plate. Therefore, they con-
ratio (Hcr/D) the pullout behaviour of plate is related to tribute to the increase of pullout capacity of plate when
the distance between the plate and the layer boundary. it is completely embedded in the top soft soil layer.
The total embedment depth (d/D) has very little influ- Figure 7 shows the stiff soil cones under different
ence on the plate pullout capacity. strength ratios and different initial embedments. It can
Summarizing the results in Figure 4 and 5, it can be be seen that the initial embedment depth (Hinitial) has
seen that the uplift resistance in two layered soils has very little influence on the fully developed cone size.

216

Copyright 2005 Taylor & Francis Group plc, London, UK


(a) uplift = 0.16D (b) uplift = 1.2D
su1/su2=0.5, H/D=1.0 su1/su2=0.5, H/D=2.0

(a) (b)

(c) uplift = 2.5D (d) uplift = 5.4D

Figure 6. Soil flow mechanism during plate continuous pull- su1/su2=1/3, H/D=2.0
out at Hinitial/D  1 and su1/su2  0.5 (blue/dark: layer 1, green/ su1/su2=1/3, H/D=1.0
light: layer 2).
(c) (d)

Thus, in Figure 5 there is no great difference in pull-


out capacity with different initial embedment (Hinitial).
However, with decreasing soil strength ratio, the upper
soil cone size increases and lower cone size decreases.
The cone size increase above the plate is larger than the
size decrease beneath the plate. Therefore, in Figure 4,
the increase in the pullout capacity of plate pulled into
the top soil is significant at small soil strength ratios.
Soil-plate separation depth (d s ) in Stage 4 (Fig. 4) is
related to soil unit weight (), soil strength (su1), soil
strength ratio (su1/su2) and plate diameter (D). The sep-
aration depth of plate in a uniform soil has been inves-
tigated by Song and Hu (2005). Equation 3 shows su1/su2=0.2, H/D=2.0
the findings of the relationship between the separation su1/su2=0.2, H/D=1.0
depth (ds /D) and soil stability number (D/su) of circu-
lar plate in uniform soils. (e) (f)

Figure 7. The cone size with different strength ratios and


(3) initial embedment depth at uplift distance  2.8 D.

Figure 8 shows the separation depth against the top (su1/su2 2/3), the separation depth in double layer soils
soil strength (su1) and soil strength ratio (su1/su2) in this are less the ones in single layer soils. This is because the
study. In single soil layer analysis (Song and Hu, 2005), stiff soil cone attached to the lower side of the plate has
the strength of the uniform soil (su) is used. It can be delayed the separation. Nevertheless, the overall trend
seen that in double soil layer, the separation depth is the is the same that the separation depth increases with
same as the one in single soil layer for higher strength increasing soil strength of the soil layer where the plate
ratios (su1/su2 2/3). However, for lower strength ratios is embedded. The initial embedment ratio (Hinitial/D  1

217

Copyright 2005 Taylor & Francis Group plc, London, UK


su1/su2 plate. This increase is only significant at strength
0 0.2 0.4 0.6 0.8 1.0 1.2 ratios of su1/su2 2/3.
1.2 4 The initial embedment depth (Hinitial) has little influ-
Single soil layer (Song & Hu) ence on the cone size, thus it has little influence on
1 Double soil layer (This study) the pullout capacity response and soil-plate separ-
0.8
ation depth.
5 Larger sizes of stiff soil cones are observed with
ds/D

0.6 lower soil strength ratios. This also affects the soil-
plate separation depth. Delayed soil-plate separa-
0.4 tions are evident with lower soil strength ratios of
su1/su2 2/3. Therefore, for any double layer soils
0.2
with su1/su2 2/3, an increased plate pullout cap-
0 acity in third stage and delayed separation in fourth
0 2 4 6 8 10 12 stage during plate pullout should be expected.
su1 or su2 (kPa)

Figure 8. Soil-plate separation depth in double layered soils.


ACKNOWLEDGEMENTS

The research presented here is supported by Australian


and 2) has little effect on the soil plate separation depth,
Research Council through the Large ARC discov-
since they are influential on the size of stiff soil cones
ery scheme (DP0344019). This support is gratefully
attached to the plate.
acknowledged.

5 CONCLUSIONS
REFERENCES
In this paper, the elasto-plastic finite element method
has been used to study the pullout behaviour of a circu- Adams, J. I. & Hayes, D. C. 1967. The uplift capacity of shal-
low foundations. Ontario Hydro-Research Quarterly: 113.
lar plate anchor embedded in double layer soils. The
Ashbee, R. A. 1969. A uniaxial analysis for use in uplift
anchor was assumed rough. Soil profiles were set up as foundation calculations. Report RD/L/R 1608, central
a soft clay layer overlaying a stiff clay layer. A circular Electricity Research Laboratory.
plate anchor was initially embedded in the stiff layer. Bowles, J. E. 1988. Foundation Analysis and Design. New
The pullout capacity of anchors embedded in different York: McGraw-Hill.
depth and during continuous pullout was studied. The Burd, H. J. & Frydman, S. 1997. Bearing capacity of plane-
critical distance of the embedded plate to the soil layer strain footings on layered soils. Canadian Geotechnical
boundary and the separation depth where soil separated Journal (34): 241253.
from the plate base were also investigated. Some find- Carter, J. P. & Balaam, N. 1995. AFENA Users Manual,
Geotechnical Research Centre, The University of Sydney.
ings are listed below:
Das, B. M. 1978. Model tests for uplift capacity of founda-
1 In plate pre-embedment analysis, the pullout capac- tions in clay. Soils and Found., Japan, 18(2): 1724.
ity of the plate anchor depends on the distance of Das, B. M. 1980. A procedure for estimation of ultimate cap-
the plate to the soil layer boundary (H). The critical acity of foundation in clay. Soil and Found. 20(1): 7782.
Das, B. M., Tarquin A. J., and Moreno R. 1985a. Model tests for
distance (Hcr) is where the plate pullout capacity
pullout resistance of vertical anchors in clay. Civil Eng. for
starts to decrease due to the closeness of the plate Practicing and Design Engineers., Pergamon Press, New
towards the soil layer boundary. The relative critical York, 4(2): 191209.
distance (Hcr/D) decreases with increasing strength Das, B. M., Moreno R., and Dallo K. F. 1985b. Ultimate
ratio (su1/su2), it varies from 1.2 for su1/su2  0.1 to pullout capacity of shallow vertical anchors in clay. Soils
0.4 for su1/su2  0.8. and Foundations, 25(2): 148152.
2 In plate continuous pullout analysis, the pullout Das, B. M. & Puri, V. K. 1989. Holding capacity of inclined
capacity of the plate anchor shows four stages: square plate anchors in clay. Soil and Found., 29(3):
stable in lower layer (HHcr); passing the boundary 138144.
Davie, J. R. & Sutherland, H. B. 1977. Uplift resistance of
(H Hcr); stable in top layer and soil-plate separation.
cohesive soils. Journal of the Soil Mechanics and
3 In first and third stages, the pullout capacity of Foundations Division 103(9): 935952. ASCE.
anchor can be estimated by using single soil layer Dewaikar, D. M. 1988. Finite element analysis of certain
parameters. However, an increased pullout capacity aspects of deep anchors in cohesive soil medium.
comparing with the one for single soil layer is due to Numerical methods in Geomechanics (Innsbruck 1988):
the two stiff soil cones attached to both sides of the 14371441. Rotterdam: Swobada, Balkema.

218

Copyright 2005 Taylor & Francis Group plc, London, UK


Hu, Y. and Randolph, M. F. 1998a. A practical numerical Ranjan, G. & Arora, V. B. 1980. Model studies on anchors
approach for large deformation problems of soil, Int. J. under horizontal pull in clay. Proc. 3rd Aust, N.Z Conf.
Numer. And Anal. Methods in Geomech., 22, 327350. Geomech. 1: 6570. Wellington, N.Z.
Hu, Y. and Randolph, M. F. 1998b. H-adaptive FE analysis Song, Z. & Hu, Y. 2005. Vertical pullout behavior of plate
of elasto-plastic non-homogeneous soil with large deform- anchors in uniform clay. Proc. Int. Symp. on Frontiers in
ation, Int. J. Compu. Geomech., 23(1/2), 6184. Offshrore Geot.
Martin, C.M. & Randolph, M.F. 2001. Applications of the Vesic, A. S. 1971. Break-out resistance of objects embed-
lower and upper bound theorems of plasticity to collapse of ded in ocean bottom. Journal of the Soil Mechanics and
circular foundations, Proc. 10th IACMAG Conf., Tucson. Foundations Division 97(9): 11831205. ASCE.
Merifield, R. S., Sloan, S. W. & Yu, H. S. 1999. Rigorous Vesic, A. S. 1972. Expansion of cavities in finite soil mass.
plasticity solutions for the bearing of two-layered clays. Journal of the Soil Mechanics and Foundations Division
Geotechnique 49(4): 471490. 98(3): 265290. ASCE.
Merifield, R. S., Sloan S W & Yu, H S. 2001. Stability of plate Wang, C. X. & Carter, J. P. 2002. Deep penetration of strip
anchor in undrained clay. Geotechnique 51(2): 141153. and circular footings into layered clays. The International
Meyerhof, G. G. and Adams J. I. 1968. The ultimate uplift Journal of Geomechanics 2(2): 205232.
capacity of foundations. Canadian Geotechnical Journal Zienkiewiecz, O. Z. and Zhu, J. Z. 1993. The superconver-
5(4): 225244. gent patch recovery and a posteriori error estimates. Part
Meyerhof, G. G. 1973. Uplift resistance of inclined anchor 1: The recovery technique, Int. J. Numer. Meth. Engng.,
and piles. Proceedings, 8th International Conference on 33, 13311364.
Soil Mechanics and Foundation Engineering 2: 167172. Zienkiewiecz, O. Z., Huang, M. and Pastor, M. 1995.
Moscow. Localization problems in plasticity using finite element
Meryerhof, G. G. & Hanna, A. M. 1978. Ultimate bearing solutions with adaptive remeshing, Int. J. Numer. And Anal.
capacity of foundations on layered soils under inclined Methods in Geomech., 19(2), 127148.
load. Can. Geotech. J. 15: 565572.

219

Copyright 2005 Taylor & Francis Group plc, London, UK


The use of inflatable anchors in offshore sandy soils

T.A. Newson
Geotechnical Research Centre, The University of Western Ontario, London, Ontario, Canada

F.W. Smith
Division of Civil Engineering, University of Dundee, Nethergate, Dundee, Scotland, U.K.

P. Brunning
Stolt Offshore MS. Ltd., Bucksburn House, Bucksburn, Aberdeen, Scotland, U.K.

ABSTRACT: Whilst many remotely operated vehicles (ROVs) have the necessary hydraulic control systems
to employ anchors, previous attempts to develop seabed fixity have had variable success. This project had the
aim of determining whether a flexible, inflatable anchor system would provide sufficient uplift capacity to fix
ROVs during offshore activities. A series of scaled physical model and full scale prototype tests have been used
to assess the performance of the proposed system. A limited range of anchor designs and operating conditions
were investigated for a generic sandy offshore soil. The experimental data showed that the pullout capacity
increased for higher membrane pressures. The normalized pullout capacity and mobilization distance for peak
load was found to be comparable with the small scale tests conducted. Based on the data shown in this study, the
inflatable anchor system shows considerable promise for offshore use in sandy soils.

1 INTRODUCTION paper describes the full-scale prototype experiments


(using a generic sandy soil), the inflatable anchor sys-
Since many remotely operated vehicles (ROVs) are neu- tem and the testing apparatus.
trally or positively buoyant, any activities that require
any significant reaction load, e.g. in situ soil testing, are
not possible without additional anchoring or clump 2 EXPERIMENTAL METHODOLOGY
weights. Whilst the majority of ROVs used by the off-
shore oil and gas industries have the necessary hydraulic Full-scale anchor testing was carried out on the labora-
and pneumatic control systems to employ anchors, pre- tory floor in a large cylindrical steel container of
vious attempts to develop seabed fixity have had vari- 2400 mm internal diameter and 3000 mm height, see
able success. These include standard anchor systems, Figure 1a. A series of tests involving constant velocity
such as helical screw, suction, duckbill and plate pullout of an inflatable anchor, with varying inflation
anchors. Any viable alternative must provide a cheap pressures was conducted. Two large stroke (500 mm)
and reusable system that will provide sufficient pull- screw jacks were used to extract the anchor and these
out capacity and be able to operate in the demanding were fixed to the top of the container using cross beams,
deep offshore environment. as shown in Figures 1a and 1b. The anchor was pulled
This project had the aim of determining whether a out of the soil using a rigid supporting beam system,
flexible, inflatable anchor system may provide suffi- shown with the anchor in Figure 1b.
cient uplift capacity to fix ROVs during offshore activ- The inflatable anchor system and geometry is
ities. A series of scaled physical model and full-scale shown in Figures 1b and 2. The anchor consists of two
prototype tests have been used to assess the perform- nested cylindrical steel tubes (of 1010 mm and 340 mm
ance of the proposed anchor system in terms of pullout length), which contain an internal inflatable section of
capacity and mobilization distance. A limited range of length 760 mm with two rubber membranes (with ini-
anchor designs and operating conditions were investi- tial diameter of 250 mm). Fluid or air can be pumped
gated to provide data for this assessment. The scaled into the annular space between the tube and membranes
physical model tests have been described elsewhere to inflate them. A second mode of operation is possible
and details can be found in Newson et al. (2003). This using this system, where the membrane section can be

221

Copyright 2005 Taylor & Francis Group plc, London, UK


shortened and an internal piston causes expansion of the anchor as it is pulled out of the soil. Two 5 tonne
this section (the system was not tested in this mode). load cells are used to measure the load applied to the
The fluid (water) in the rubber membrane is pressurised anchor as it is pulled from the soil. A PowerDaq data
by means of a hand pump (with an in-line pressure acquisition board is used to monitor all of the trans-
transducer) and volume measured using a graduated ducers. Analogue inputs are used to monitor the pres-
reservoir. sure, travel and load.
A large stroke (500 mm) linear strain displacement The anchor tests were conducted on Congleton silica
transducer was used to monitor the displacement of sand. This has a uniform grading with D50  0.2 mm
(see Fig. 3), specific gravity Gs  2.65 and sub-
rounded particles. The angle of repose of the loose
soil was measured to be 3234 and this is believed to
be close to the critical state angle of friction, crit. The
range of densities for this soil are found to vary between
max  1.85 t/m3 and min  1.30 t/m3. Dry sand beds
were constructed by air pluviation using a controlled
rate, to ensure uniform and repeatable soil densities.
Small pots were used in trials and measured after
tests to verify the densities. Although the edges of the
container were unlined, since the container was large
enough to provide at least 10 diameters of soil on either
side of the anchor, any influence due to the rigid bound-
ary was thought to be insignificant. The side walls of
the container were also strain gauged and monitored as
a safety precaution.
Figure 1a. Cylindrical soil testing chamber. Sand was pluviated to a depth of 200 mm in the base
of the cylindrical container and the anchor was pos-
itioned on the surface of the soil. Further sand was then
added until the final sample height gave the required
pipe embedment depth, hence this represented a
wished-in-place condition. Whilst the sand sample
was being placed around the anchor, it was important
that the anchor was free to settle vertically so that a
net vertical load was not applied to the anchor before
the pullout test commenced. This was to prevent the
mobilization distance to peak load being underesti-
mated. Pullout for all of the tests was conducted at a
fixed rate of 10 mm per minute. Since dry sand beds
were used for the prototype testing, the results are for
a drained case and the effective body weights of the
material will be greater than in a field situation.
Figure 1b. Inflatable anchor system and supporting beam.

100
To pressure control/volume
measurement system 80
F
% Passing

F 60
Soil
surface
40
H
20
Inflatable
L membranes
0
0.01 0.1 1 10
D Particle size (mm)

Figure 2. Geometry of anchor and loading system. Figure 3. Particle size distribution of Congleton sand.

222

Copyright 2005 Taylor & Francis Group plc, London, UK


3 EXPERIMENTAL RESULTS previous model tests, the pullout load is seen to increase
with inflation pressure and relatively large mobilization
3.1 Calibration of the anchor distances are observed for peak loading. The compari-
son between the uninflated anchor and the two inflated
Correct interpretation of the pressure-volume curves
tests show quite a marked increase in pullout capacity.
requires an estimate of pressure losses in the system due
Comparisons between some of the earlier model
to hoop stresses in the membrane. Calibration of the
anchor tests (Newson et al. 2003) and the prototype
membrane resistance was conducted by expanding the
tests are shown in Figures 6a to 6c. The normalized
fluid filled (water) membrane in free air. The calibra-
tion curve is shown in Figure 4, along with data show- 100
ing the increase in membrane diameter with pressure. Model

Pullout capacity, Nu
The volume-pressure data in subsequent sections has 80 Prototype
been corrected for this effect.
60
3.2 Inflation pressure (P) 40
The effect of varying the initial inflation pressures on
20
the measured pullout force against anchor displacement
is shown in Figure 5. The corrected inflation pressures 0
applied were 0 kPa, 74 kPa and 101 kPa. For each of 0 2 4 6 8
these tests the embedment depth (H) was 450 mm (giv- Embedment depth, H/D
ing H/L  0.6) and the soil density ranged from 1.35 to
1.45 t/m3. These tests show peak pullout loads from 2 to Figure 6a. Dimensionless pullout capacity against normal-
7.9 kN and mobilization distances from a few millime- ized embedment depth.
tres to approximately 125 mm. In common, with the

100 400 100


Pressure Model
Pullout capacity, Nu

90 350 80 Prototype
80 Diameter
Diameter, D (mm)
Pressure, P (kPa)

300
70 60
60 250
50 200 40
40 150
30 20
100
20
50 0
10 0 0.05 0.1 0.15
0 0
Displacement, df/(H+L)
0 5 10 15
Vol (litres) Figure 6b. Dimensionless pullout capacity against normal-
ized displacement.
Figure 4. Membrane calibration curve in free air.

45
10
101 kPa 40 Model
Pullout capacity, Nu

35 Prototype
8 74 kPa
0 kPa 30
Load (kN)

6 25
20
4
15
2 10
5
0 0
0 50 100 150 200 250 0 10 20 30 40 50
Displacement (mm) Inflation pressure, P/v'

Figure 5. Variation in anchor pullout capacity with inflation Figure 6c. Dimensionless pullout capacity against normal-
pressure (P). ized inflation pressure.

223

Copyright 2005 Taylor & Francis Group plc, London, UK


anchor pullout capacity (see Equation 1 below) that was 70
suggested by Dickin & Leung (1990), is plotted against 60

Pressure, P (kPa)
normalized embedment depth H/D in Figure 6a and this
shows increasing load capacity with embedment. 50
40
30
(1)
20 PT1
PT2
10 PT3
where F is the pullout force, D is the membrane diam-
eter, L is the membrane length, H is the embedment to 0
the top of the membrane and  is the effective unit 0 1 2 3 4 5
weight of the soil. Volume (litres)
Whilst the pullout capacity of the full-scale anchor
tests was lower than that for the model anchor tests, Figure 7. Sand pressure-volume relationships during sleeve
the normalized full-scale embedments are quite small inflation proof tests PT 1 to 3.
and the general trend seems to agree with the previ-
ous work conducted. normalized embedments (H/D) ranging from 1 to 7.
Figure 6b shows the normalized pullout capacity Conventional cylindrical plate anchors were found to
plotted against dimensionless displacement (for the have Nu of approximately 5, for H/D values of 4
peak loading). The mobilization distances for the model (Ovesen 1981). Lee (1991) investigated cylindrical low
tests are again larger than the full-scale tests, but follow pressure grouted anchors and found values of Nu in the
the general trend found with the earlier model tests. In range of 2 to 6, for normalized embedments less than
addition, the grade, thickness and roughness of rubber H/D  6. Multihelix helical anchors provided more
was found to affect the model tests, hence a propor- pullout resistance, but still gave lower capacity with
tion of the variation between the two test groups may Nu 10 for values of H/D  6 (Ghaly & Hanna 1994).
also lie with these factors. Comparison with the observed mobilization dis-
The results of increasing the inflation pressure of tances (df) for peak load were found to be quite high
the anchor system are shown in Figure 6c. for the majority of tests for the inflatable anchor with
Normalized pullout capacity is plotted against blad- df /(H  L) being approximately 810%, which com-
der pressure P, normalized by the initial in situ vertical pares to 13% for the range of aforementioned anchor
effective stress. types.
The highest values of dimensionless pressure are
seen to be found with the model tests, which would
have very low initial stresses due to the small abso- 3.4 Volume-pressure relationships
lute embedment depths. The normalization has been Typical volume-pressure curves (determined from the
achieved using vertical effective stress and it may anchor system during inflation proof tests) for the sand
have been more logical to have used horizontal effect- used in this study are shown in Figure 7. Analysis and
ive stresses; assuming a normally consolidated mater- interpretation of this data can provide a range of param-
ial, these numbers should be therefore 2 to 2.5 times eters, e.g. elastic (G, ), angles of friction and dilation,
greater. The general trend shown in Figure 6(c) is for lateral limit and in situ lateral pressures, etc. For these
increasing pullout capacity with increasing normal- curves the shear modulus (G) is estimated to range from
ized inflation pressure. This appears to be consistent 1600 to 4150 kPa, which is comparable with the meas-
between the two sets of data. Assuming this relation- urements made in the small scale study and is quite
ship holds for over-consolidated soils, then higher rela- reasonable for sandy soil in this state; although it should
tive initial horizontal effective stresses would reduce be recognized that the process of measurement modi-
the pullout loads (assuming installation effects are fies the structure of the soil to some extent.
negligible).

4 PREDICTION OF PERFORMANCE
3.3 Comparison with other anchor systems
The pullout loads for the inflatable anchor system We may interpret the pullout behaviour in a simplistic
compare very favourably with values for different types manner using Equation 2, and estimate the pullout
of anchor found in the literature using similar sand capacity (F) of the anchor due to the inflatable mem-
grading and relative densities. Shallow enlarged base brane using:
piles modeled by Dickin & Leung (1990) were found
to have normalized pullout capacity, Nu 5 for (2)

224

Copyright 2005 Taylor & Francis Group plc, London, UK


where n is the effective normal stress, A is the inflatable anchor system increases for higher inflation
membrane surface area ( .D.L), D is the membrane pressures. For a full-scale system, four such anchors
diameter, L is the membrane length and  is the soil- attached to an ROV would provide over three tonnes of
membrane interface friction angle (. ). reaction force. The performance of the inflatable sys-
If the membrane remains uninflated, then the nor- tem is far better than would have been observed in hel-
mal stress (n) acting will be purely from horizontal ical screw or plate anchors of equivalent geometry.
stresses (h  K v) due to the soil self-weight The pullout capacity of the anchors in sand has been
stresses (v  [H  L/2]), where H is the expressed in a dimensionless form and gives a range
embedment to the top of the membrane,  is the effec- of Nu for the above test programme of approximately
tive unit weight of the soil, K is an earth pressure 59, which compares very favourably to values for
coefficient (K  1
sin ) and  is the angle of other forms of anchor. The mobilization distances (df)
internal friction. For this case, we can assume that for peak load were found to be quite high for the major-
  35, L  0.76 m, H  0.45 m,   14 kN/m3, ity of tests with df /(H  L) being approximately
D  0.25 m and   0.3, then Equation 2 predicts a 810%, which compares to 13% for other anchor
pullout capacity of F  1.3 kN, which is slightly systems. The additional benefits of monitoring the
lower than that measured in the test programme, but pressurevolume relationship during inflation have
does not account for the friction on the steel section been demonstrated, allowing the determination of a
of the anchor. range of soil parameters to provide information for
If the membrane is inflated, then we need to estimate design and for optimization of waiting time prior to
the increase in the membrane diameter (D), which may loading the anchors.
be accomplished using cylindrical cavity expansion The mechanisms of soil failure around the anchor
theory (e.g. Carter et al. 1986) or from monitoring the system are still unknown and this aspect should be
volume change of the inflatable section of the anchor. investigated further using numerical analysis. This
The radial strain (r  r/ro) of the expanding cavity should provide more accurate pullout estimates for util-
due to an internal pressure (p) can be found from: ization of the anchor system, will aid improvements
of the operating methodology and equipment setup/
design.
Installation of the anchors in the field would most
(3) practically be accomplished using a system similar to
that of the self-boring pressuremeter (e.g. Hughes et al.
1977). A cutter would be required at the foot of the
where Kp is the passive earth pressure coefficient, po anchor rotating inside an internally tapered shoe; soil
and ro are the initial cavity pressure and radius respect- cuttings would need to be removed inside the instru-
ively, G is the shear modulus, re is the radial stress ment using flushing fluid (provided by the ROV or
at the elastic-plastic interface and A14 are constants using the surrounding seawater). This should ensure
depending on Poissons ratio (), the angle of internal minimal disturbance to the soil outside the anchor,
friction () and the angle of dilation (). although this aspect has been questioned by a number
If we assume that the inflation pressure p  105 kPa of researchers (e.g. Bruzzi et al. 1986). Alternatively
and assuming G  2500 kPa, po  12 kPa,   0.35, a water jet-based insertion system may be considered,
  0, then the diameter of the bladder will increase e.g. Benoit et al. (1995). Further work would be
to D  0.265 m with cylindrical cavity expansion. This required to assess the effect of soil disturbance due to
compares to a diameter, D  0.29 m estimated from the installation of the anchor and the performance of the
measured membrane volume changes. Accounting for system in strain-softening materials, such as cemented
the latter increase in membrane area and assuming the sands.
normal stresses are equal to the corrected inflation pres-
sure, then Equation 2 predicts that the pullout capacity
of the full-scale anchor will be F  12.6 kN, which is
an over-prediction, but is comparable with the experi- REFERENCES
mental data.
Benoit, J., Attwood, M., Findlay, C. & Hilliard, B. 1995.
Evaluation of jetting insertion for the self-boring pres-
suremeter Canadian Geotechnical Journal 32(1): 2239.
5 CONCLUSIONS Bruzzi, D., Ghionna, V., Jamiolkowski, M., Lancellotta, R. &
Manfredini, G. 1986. Self-boring pressuremeter in Po River
As was reported in the initial study, the data shows sand. In Proceedings, Pressuremeter and its marine appli-
that the inflatable anchor system displays considerable cations. American Society for Testing and Materials, Spe-
promise for use in loose offshore sands. As expected the cial Technical Publication STP 950, 283302.
experimental data show that the pullout capacity of the

225

Copyright 2005 Taylor & Francis Group plc, London, UK


Carter, J.P., Booker, J.R. & Yeung, S.K. 1986. Cavity expansion Lee, C. 1991. Pullout capacity of expanded anchorages.
in cohesive frictional soils. Geotechnique 36(3): 349358. In Centrifuge 91 Conference, Ed. (Ko), Balkema,
Dickin, E.A. & Leung, C.F. 1990. Performance of piles with Rotterdam, 145151.
enlarged bases subject to uplift forces. Canadian Geo- Newson, T.A., Smith, F.W., Brunning, P. & Gallagher, S. 2003.
technical Journal 27: 546556. An experimental study of inflatable offshore anchors. In
Ghaly, A. & Hanna, A. 1994. Ultimate pullout resistance of ISOPE 2003 Conference, Honolulu, Hawaii, (Paper #
single vertical anchors. Canadian Geotechnical Journal 2003-JSC-127).
31: 661672. Ovesen, N.K. 1981. Centrifuge tests of the uplift capacity of
Hughes, J.M., Wroth, C.P. & Windle, D. 1977. Pressuremeter anchors. In Proc 10th International Conference on Soil
tests in sands. Geotechnique 27(4): 455477. Mechanics and Foundation Engineering, Stockholm: 1,
717722.

226

Copyright 2005 Taylor & Francis Group plc, London, UK


Deepwater developments: suction caissons

Copyright 2005 Taylor & Francis Group plc, London, UK


Suction caisson soil displacement during installation

E.C. Clukey
BP America, Houston, USA

ABSTRACT: Previous work has suggested that the external skin friction for the portion of a suction caisson
installed by pumping out the water inside the caisson (referred to as suction) is reduced compared to the exter-
nal skin friction for a driven or pushed pile. The primary reason given for this reduction is that the suction pene-
tration tends to displace most of the soil at the caisson tip towards the inside rather than outside of the caisson.
Presently, however, a consensus does not exist regarding the proposed movement of soil inside a suction cais-
son and the resulting loss in skin friction. To better understand the soil displacement, four caissons installed in
the Gulf of Mexico in mostly normally consolidated soils were examined. By comparing the observed penetra-
tion outside the caisson with the point where an inner plate stopped penetration, the amount of displaced soil
that moved inside the caisson at the caisson tip was determined. The four caissons were 6.50 m in diameter and
were installed to about 24.0 m below the mud line. The external wall of the caisson was flush. The wall thick-
ness at the caisson tip was 51 mm. However, only the inner 22 mm was flat. The remainder (29 mm) had an
external 4:1 taper. The results from the field measurements suggest that during suction penetration, as a min-
imum, all the soil displaced by the 22 mm flat portion of the tip moved inside the caisson. This observation,
therefore, tends to support the basis for reduced external skin friction.

1 INTRODUCTION effect. Andersen et al. (2004) also investigated the


potential impact of a 45 outward facing bevel on the
Suction caissons have been a primary foundation sys- amount of soil pushed to the outside and its possible
tem in deepwater Gulf of Mexico, especially in water effect on the skin friction. Although the bevel seems
depths greater than 1300 m. They have been used for to have increased the soil pushed to the outside, there
several different applications including taut to semi- appears to be only a marginal increase in the octahe-
taut legged mooring systems, manifold foundations, dral stresses. More work was considered necessary to
and anchorages for pipeline end terminals (PLETS). better understand the impact of this bevel.
For mooring applications the uplift capacity of the One method for possibly determining the relative
caisson will often control the design. This uplift capac- amount of soil displaced to the outside of the caisson
ity is comprised of three parts: (1) the submerged is through field measurements during actual installa-
weight of the caisson and appurtenances, (2) the reverse tions. This paper examines the results of four 6.5 m
end bearing at the bottom of the caisson and (3) the diameter 24.0 m long (design embedment) suction cais-
skin friction along the external sides of the caisson. sons. These caissons were unique versus other suction
The external skin friction is influenced by the dis- caissons installed by BP in that inside the caisson a
placement of soil during the installation process. flat steel plate was placed across the caisson top. The
Because the wall thicknesses and installation processes distance from the caisson tip to the bottom of the steel
for a suction caisson are different compared to a driven plate was 24.00 m. Structural members to support the
or pushed pile, the external skin friction of the caisson caisson head were placed above this plate. The caisson
can also be different from the estimated skin friction penetration was stopped when the minimum design
for a driven pile. Andersen and Jostad (2003) suggested penetration was observed on the outside of the caisson.
that, based on finite element analyses, during suction In each case mud was also observed being pumped
installation more soil is displaced to the inside rather from inside the caisson, indicating that the soil plug had
than the outside of the caisson, thereby reducing the contacted the top plate. The flat plate and observation
mean octahedral stresses and the skin friction on the of mud exiting the caisson allowed for a more accurate
outside of the caisson. They have recommended that estimate of the soil-plug heave at the final caisson pene-
in the absence of site specific soils data, the skin fric- tration. With this implied position of the soil plug, the
tion should be reduced by 35% to account for this outside measured final caisson embedment, and the

229

Copyright 2005 Taylor & Francis Group plc, London, UK


internal structural details of the inside of the caisson, (Fig. 1). The plate schedule, however, varied along the
the relative amount of soil displaced inside the cais- length of the caisson as shown on Table 1. Since the
son by the caisson wall could be determined. outside diameter was constant, the inside diameter
This paper examines the results obtained from the varied according to the changes in the wall schedule.
installation of these four piles and the implications of The wall thickness for section 2 (Table 1) was vari-
the measured soil-plug heave on the displacement of able due to the presence of an internal structural
soil into or outside the caisson. The next section of the brace. This brace was required for structural support
paper describes in more detail the caisson geometry. of an external lifting padeye and was comprised of a
This is followed with a discussion of the field meas- 25 mm thick steel vertically oriented plate that
urements obtained to monitor the installation. The spanned from the padeye to about 2/3s of the caisson
results obtained from these measurements and pos- diameter. Four wing walls at 22 angles from the ver-
sible errors in the interpretation are then discussed. tical plate attached the internal brace to the inside of
Finally, conclusions with respect to the movement of the caisson. The total volume occupied by the internal
soil inside and outside the caisson are presented. brace and the wing walls is 0.36 m3.
In addition, the wall schedules shown on Table 1
there is also an outward facing taper (Table 1) of 4
2 CAISSON GEOMETRY (vertical) to 1 (horizontal), or 76 from the horizontal,
at the caisson tip. However, the taper does not begin
As previously described, the caisson was 6.5 m in flush with the inside diameter. Rather the internal
diameter (OD) with a 24.0 m long design embedment 22 mm is flat with the outer 29 mm tapered.

3 FIELD MEASUREMENTS

The field measurements made to monitor the installa-


tion of the caissons were:
1. The caisson penetration was determined by observ-
ing paint marks on the outside of the caisson with
an ROV camera (Fig. 2). The paint marks were
305 mm apart with the thickness of each line 76 mm.
It is roughly estimated that the accuracy of observ-
ing a single point as the caisson penetrates into the
soil is about 10 mm. Note that the 50 on the side of
Figure 1. 6.5 m diameter manifold suction caisson. the caisson in picture is the depth mark in feet.
2. The suction pressures (differential) are measured
as the caisson penetrates into the soil to determine if
Table 1. Details of caisson geometry. unusually high installation pressures are required.
Wall
Distance from thickness
Section pile tip (m.) (mm) Comments

1 0 to 2.06 50.8 4 on 1 outward


bevel on outer
29 mm.
2 2.06 to 4.50 22.4 Wall thicknesses
(38%), vary due to
38.1 presence of internal
(50%), structural brace.
50.8
(12%)*
3 4.50 to 6.93 22.4
4 6.93 to 9.37 25.5
5 9.37 to 11.81 31.8
6 11.81 to 21.56 38.1
7 21.26 to 24.00 44.5

* Percentages are based on the % of the outside


circumference. Figure 2. Visual observation of caisson penetration.

230

Copyright 2005 Taylor & Francis Group plc, London, UK


The differential pressure is the difference between a less than predicted, suggesting that even more soil
point outside the caisson and the hose connecting the moved inside the caisson than predicted based on the
pump used to evacuate water from inside the caisson. 50% (self-weight)/100% (suction) split described
3. The total flow or volume of water evacuated from above.
the inside of the caisson is measured. Although the
flow meter use was considered highly accurate,
there appears to be about a 3 to 4% error in the cali- 5 POTENTIAL ERRORS IN
bration supplied by the manufacturer. This is based INTERPRETATIONS
on a number of comparisons between the penetra-
tion depth and the volume of water extracted for a Several potential sources of error in the interpretation
number of BP caissons. This error is believed to be of the analyses could impact the assessment that, as a
related to the exit pipe not being sufficiently long to minimum, all the soil displaced by the flat portion of
prevent turbulence effects on the discharge side of the caisson tip enters the caisson. These potential
the pump. However, as will be shown below, the cali- sources of error and their likelihood of occurring are
bration does appear to be constant with penetration. discussed below:
Seafloor gradient and caisson tilt: The seafloor
4 RESULTS OF INSTALLATION gradients and the tilt (pitch/roll) orientation of the
caisson will affect the observed penetration on the
The self weight penetration and the final penetration outside of the caisson. If the seafloor is sloping
for the four installed caissons are shown on Table 2. towards or the caisson is tilting away from the obser-
As discussed, when the caisson penetration was vation point, then the average caisson penetration will
stopped, mud was observed exiting from the ROV be underestimated. The corrections for seafloor gra-
pump, indicating that the flat plate at the top of the dient and caisson tilt (pitch/roll) are shown on Table 3.
caisson had contacted the mud line. The slope, pitch and roll in Table 3 are all in
The predicted values were based on the following degrees. The slope angle is the angle toward the point
assumptions: (1) The recommendations suggested by where the outside of the caisson was observed. For
Andersen and Jostad apply only to the flat portion of the DC41 caissons, the observation point was towards
the caisson tip therefore 50% of the soil displaced the south, while for the DC32 and DC33 caissons the
by the flat part of the tip (22 mm) enters inside the observation point was to the east. Positive slope indi-
caisson during self-weight penetration and 100% cates that the dip angle is towards the observation
enters during suction penetration; (2) the soil com- point.
pletely fills the inside volume of the caisson as it Positive pitch indicates the caisson is tilting
moves into the caisson; (3) additional soil does not towards the observation point while positive roll indi-
move upwards into the caisson resulting from the cates that the caisson is tilting to the right of the
potential differential movement or failure of the soil observation point (starboard side).
plug, possibly at higher suction pressures. Scour at top of soil plug: As the soil plug moves
The slight differences in the predicted penetration upwards and gets relatively close to the top plate the
result from the differences in the self weight vs. the horizontal water velocities increase. These water veloc-
suction penetration. Caisson DC41W penetrated to ities can potentially erode the soil at the top of the soil
deeper than the design penetration of 24 m. This was plug and create a soil-water suspension prior to the
possible since the caisson extended about another 2 m soil plug contacting the top plate. If this scour/ ero-
above the top plate. sion occurs the soil-plug heave and ultimately the soil
The observation can be made from Table 2 that all entering the caisson could be overestimated.
the caissons except DC41W had final penetrations To investigate the potential error from this effect
the erosion potential was determined from:
Table 2. Caisson penetration depths. i. the discharge volume from the pump near the end
of suction installation.
Self-wt. Final Predicted ii. determination of the horizontal velocities as a
Caisson penetration, observed penetration,
function of the radial distance from the inlet port
Designation m. penetration, m. m.
on the flat plate when the soil plug was 30 mm
DC32 15.70 23.93 23.97 from the top plate.
DC33 15.85 23.87 23.97 iii. the scour potential of Gulf of Mexico clay as
DC41E 14.45 23.84 23.96 defined by Nierdoroda et al. (2003).
DC41W 15.51 24.02 23.97 iv. an assessment of the volume of eroded material,
Ave. 15.38 23.92 23.97 based on areas at the top of soil plug subjected to
different rates of erosion.

231

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 3. Summary of seafloor gradient and tilt corrections.

Caisson Pitch & slope


designation Slope Pitch Roll corr. (m) Roll corr. (m) Total corr. (m)

DC41E 2.0 1.0 0.3 0.06


.01 0.05
DC41W 2.0 1.2 0.0 0.04 0.00 0.04
DC33 0.7 0.7 0.0 0.00 0.00 0.00
DC32 1.3 0.5 0.2 0.05
.01 0.04
Ave. 1.5 0.9 0.1 0.04 0.00 0.03

Toward the end of suction installation the dis- Table 4. Average horizontal velocities and scour rates.
charge velocity was about 2.40 m3/min. Given the
size of the inlet port, this resulted in an upward water Distance from Horiz. velocity Scour rate
velocity of 1.76 m/s at the inlet port. The horizontal inlet (mm) (m/s) (mm/hr)
water velocity, which can cause erosion, will vary
84 2.47 20,000
with the radial distance from the center of the inlet 101 2.05 1,000
port and the soil plug-top plate separation distance. 127 1.64 60
If the horizontal velocity at 84 mm were equal to 152 1.36 5
the inlet port velocity the distance from the soil plug 172 1.17 1
to the top plate would be 42 mm and the scour rate
would be 40 to 50 mm/hr. This scour rate would not
to sufficient to cause significant erosion of the soil
plug since the time to erode the clay is small (caisson top plate, the potential for error in the soil plug position
penetration 1.2 mm/s). due to erosion is estimated to be about 10 to 20 mm.
Niedoroda et al. (2003) present information on scour Finally, it is noted that although the scour data
velocities for Gulf of Mexico soils with velocities rang- cited by Niedoroda et al. were for Gulf of Mexico
ing from 0.3 m/s to 2.0 m/s. For velocities above 1.3 m/s, soils, the scour tests were performed on soils with sig-
their data shows a linear trend of velocity vs. the log of nificantly greater shear strengths than the soils in
scour rate in mm of scour/hr. However, it should be which the caissons were installed. Even though scour
noted that the data suggests a very high gradient with rate may not directly correlate with shear strength
very large changes in the scour rate. Extrapolation to (Briaund, 2005), the differences between the soils
velocities much beyond the actual data was considered tested in the lab and those at the installation site in the
unwarranted. Therefore, the maximum velocity consid- field suggest that additional testing may be warranted.
ered in the scour assessment was 2.5 m/s. Table 4 shows Local seafloor high point near pump inlet port:
the estimated average horizontal velocities and scour If a local seafloor high point existed at the pump inlet
rates as a function of the distance from the inlet port port, the observed embedment would again tend to
when the soil plug is 30 mm from the top plate. overestimate the amount of soil entering inside the
These results suggest that high erosion rates will caisson. However, with four caissons and with this
only occur within the inlet port radius (84 mm). For potential error being random, in contrast to the potential
the soil plug-top plate separation of 30 mm, the erosion error from the seafloor gradients, caisson tilt and ero-
of soil directly underneath the inlet port is subjected sion of the soil plug, the likelihood of this effect signifi-
to the scour rate of 20,000 mm/hr. If the time during cantly biasing the average results is considered low.
which scour can occur is estimated to be approxi- Caisson not completely filled with mud: There is
mately 20 seconds, then the volume of eroded soils the potential that because of the variable wall sched-
would be .001 m3. This volume of material is much ule that as it penetrates, the caisson would not com-
less than the volume of hose from the inlet port to the pletely fill with soil. This would occur if the soil plug
pump (0.032 m3). Therefore, it is unlikely that any sig- did not expand to accommodate wall thickness
nificant muddy water would have been observed changes as the caisson penetrates. The most likely
when the plug was within 30 mm of the top plate. point where this would occur would be in the first
Estimates of the erosion of the soil plug for soil plug- section where the wall thickness decreases, resulting
top plate separations of less than 30 mm werent con- in a greater inside diameter. Section 2 (Table 1) is
sidered since the scour velocities increased to values therefore the most likely section, where wall thickness
well beyond the range reported by Niedoroda et al. changes could impact the assessment of soil moving
Based on the analyses and the potential for significantly inside the caisson. If the entire section had trapped
enhanced erosion as the soil plug gets very close to the water instead of soil in the annulus this would result

232

Copyright 2005 Taylor & Francis Group plc, London, UK


DC 41W and soil plug erosion appear to be systematic. The
Total Flow, m^3 other potential errors either appear to be small or ran-
0 100 200 300 dom. The effects of random errors would be min-
15 imized by the averaging done for the four caissons.
16 For the systematic errors the net overall average
17 error for the seafloor gradient and tilt is 0.03 m. This
Depth below Seafloor, m

18 error should be added to the overall observed embed-


19 ment depth and would imply that less soil entered the
20 caisson than the initial reading would suggest. If this
21 correction is added to the observed embedment depth
22 0.30 m. offset the adjusted average embedment depth would be
23 0.60 m. offset 23.95 m.
24 Measuredl R2 = 0.99990 The correction for the scour potential (0.01 to
0.02 m) suggests that less soil-plug heave occurred
25
than would be implied from the initial readings. This
Figure 3. Total flow versus penetration. correction could also approximately be considered by
increasing the observed penetration by the same
amount. Therefore, the final corrected average embed-
in the caisson having only 25 mm less embedment, ment would range from 23.96 to 23.97 m. The lower
implying less soil entered the caisson. However, since bound number of this range would suggest that 105%
the shear strength of the soil is not sufficient to pre- of the flat portion of the caisson tip displaced soil into
vent lateral spreading and ultimately failure without the caisson during suction penetration. The higher
lateral support of the plug, the soil is likely to expand number (23.97) suggests that 100% of the displaced
to fill this volume. In addition, Section 2 his internal soil from the flat portion of the tip entered the cais-
stiffeners that also tend to push soil laterally to fill up son, since it matches exactly the average predicted
the annulus volume. The volume of the stiffeners is penetration (Table 2) which, as discussed, assumed
about half of the volume required to fill the extra this condition.
annulus volume. Therefore, it does not appear that If the flat portion of the caisson tip displaced only
changing wall thicknesses will introduce significant 50% of the soil into the caisson then the caisson
errors in the estimates of the soil entering the caisson. embedment would have been 24.02 m. Therefore, the
Plug pulled upwards by suction forces: Another difference between the final corrected average embed-
potential error could occur if the soil plug were pulled ment and the implied embedment for this case is 0.05
upwards by the suction force applied during suction to 0.06 m.
penetration. This upwards movement could occur as Finally, it should also be noted that even if the ero-
suction pressure increases and the soil plug begins to sion adjustment is increased to 0.04 m, 80% of the
fail. However, as shown on (Fig. 3), the amount of water soil displaced by the flat portion of the tip would be
evacuated by the pump per meter of penetration is con- displaced into the caisson. Therefore, even with this
stant throughout the entire embedment process. This conservative adjustment it appears that significantly
constant trend is based on the linear trend in the data. If more soil, displaced by the flat portion of the tip,
the suction forces were sufficient to pull the soil plug moves inside rather than outside the caisson. The
upwards, additional flow would occur and the data results also suggest that the 4:1 taper at the tip effect-
would begin to trend to the 0.30 m and 0.61 m offset ively pushes soil towards the outside of the caisson.
lines shown on the figure. These offsets represent the
amount of soil plug movement in terms of the flow vol-
ume if the soil plug moved upwards 0.30 and 0.60 m 7 SUMMARY AND CONCLUSIONS
respectively. The data, however, show that the amount of
soil moving into caisson is constant, which implies that Based on observations made during the installation of
additional soil, other than that displaced by the caisson four suction caissons installed in mostly normally con-
wall, did not move into the caisson. The high R2 value solidated clays it appears likely that more soil moves
for the measured data also indicates that the volume of inside the caisson during suction penetration compared
soil entering the caisson is constant with penetration. to the case for a driven or pushed pile where about 50%
of the displaced volume at the tip would be expected to
move into the pile or caisson. For the caissons exam-
6 INTERPRETATION OF SOIL-PLUG HEAVE ined in this study the caisson tip had a flat portion and
a very steep 4 on 1 taper. The installation observa-
For the potential errors discussed above, only the tions suggest that about 100 to 105% of the flat por-
errors that result from seafloor gradients, caisson tilt, tion of the tip moves inside the caisson. The results

233

Copyright 2005 Taylor & Francis Group plc, London, UK


also suggest that the outward taper at the caisson tip significant observations during the caisson installa-
effectively pushes soil to the outside of the caisson. tion. John Roth provided details of the final seafloor
The potential for errors in the interpretation of the gradients and the caisson tilt. Their help is gratefully
observed measurements were discussed. The most sig- acknowledged. Finally, Pierre Beynet of BP provided
nificant and systematic errors appear to be associated very useful comments and discussions with the author
with the seafloor gradient, caisson tilt and soil-plug on soil erosion. His help is greatly appreciated.
scour. Additional scour tests on soils with similar shear
strengths as those at the installation site would help to
better determine the error caused by soil-plug erosion.
The difference between the caisson penetrations REFERENCES
for the cases where 50 and 100% of the flat portion
Briaud, J-L. 2005. personal communication
of caisson tip displaced soil inside the caisson is Andersen, K. H. & Jostad, H. P. 2003. Shear Strength
0.05 to 0.06 m. This difference, with the corrections Along Outside Wall of Suction Anchor in Clay After
described in this paper, is considered sufficient to pro- Installation. Proceedings. XII ISOPE Conference, Kyushu,
vide reasonable estimates of soil-plug heave. Japan, May
Andersen, K. H., Jostad, H. P., Andresen, L. & Clukey, E. C.
2004. Effect of Skirt Tip Geometry on Set-up Outside
ACKNOWLEDGMENTS Suction Anchors in Soft Clay. Proceedings, 24th
International OMAE Conference, Vancouver, June
The author acknowledges support for this work and Niedoroda, A. W., Reed, C. W., Hatchett, L., Jeanjean, J.,
Driver, D., Briaud, J-L. & Bryant, W. 2003. Bottom
permission to publish the results from BPs Gulf of Currents, Erosion Rates, and How to Use Them to Date
Mexico Deepwater Development Business Unit. Slope Failures Along the Sigsbee Escarpment.
Mark Alexander of Technip and Dave Brewster pro- Proceedings of the Offshore Technology Conference,
vided details of the caisson configuration as well as OTC 15199, Houston, May

234

Copyright 2005 Taylor & Francis Group plc, London, UK


Lessons learned from several suction caisson installation projects in clay1

Y.C. Lee, J.M.E. Audibert & K.-M. Tjok


Fugro-McClelland Geosciences, Inc., Houston, Texas, USA

ABSTRACT: During the last five years, a large number of suction caissons have been installed worldwide to
support sub-sea facilities and to anchor floating facilities. For example, SPAR and semi-submersible hulls have
been anchored to the seafloor using suction caissons, arranged in clusters of three to four caissons each, in water
depths ranging from approximately 1,200 m to 1,800 m. During actual installation of suction caissons, unex-
pected problems may be encountered. This paper attempts to bridge the gap between the theoretical predictions
and field observations by discussing the problems experienced and lessons learned by the authors during the
installation of many suction caissons at sites in the Gulf of Mexico and offshore West Africa. The problems and
lessons include effects of caisson penetration rate on the shear transfer between the soil and caisson interface,
the importance of pump calibration and under-pressure correction, caisson tip bevel effects, rotation of the caisson
during installation, field adjustment of target penetration, and the relationship between padeye depth and design
robustness. Experience gained during these caisson installations1 is used to suggest improved suction caisson
installation predictions and procedures.

1 INTRODUCTION the suction caissons are designed using a single rep-


resentative shear strength profile for all the anchors,
During the last five years, a large number of suction unless soil conditions are known to vary extensively
caissons have been installed worldwide to support sub- between clusters (e.g., rafted blocks may exist within
sea facilities (e.g., manifolds, SDUs, UTAs) and for a buried landslide unit).
anchoring floating facilities (e.g., FPSs, FPSOs, Using a single shear strength profile to obtain a pre-
SPARs). Although all the caisson installations are con- diction of the required under-pressure for all caissons
sidered successful, various installation difficulties were is not realistic when spatial variation of soil properties
encountered during the installation of some caissons. may occur. In some cases, re-evaluation of the shear
For example, the caisson at one anchor location had to strength profile at each of the anchor cluster was per-
be re-installed three times, once due to excessive rota- formed, based on cluster-specific data (e.g., large piston
tion and once due to the inability to maintain a suction corers and piezo-cone penetration tests), in order to pro-
seal, and in other cases caissons had to be installed short vide a more realistic estimate of the required suction.
of their target penetration. Finally, the caisson wall of This paper attempts to look at several aspects of
some caissons buckled during installation. Experience1 the installation prediction and monitoring, in order to
gained from overcoming these installation difficulties provide an improved theoretical prediction model.
is used in this paper to suggest improvements in the
design and installation predictions of suction caissons.
The design of suction caissons and prediction of the 2 PENETRATION RATE EFFECTS ON
required and allowable suction pressures are based on SHEAR TRANSFER
the estimate of the shear transfer between the soil and
caisson interface, which in turn depends on the shear Based on our experience with several suction caisson
strength profile and shear transfer factor. Generally, installations in the Gulf of Mexico and offshore West
Africa, the shear transfer factors appear to be different
1 during the self-weight penetration and the suction pene-
The authors believe that sharing the experience they gained
during specific projects will be useful to the engineering tration phases of caissons installation. It is believed
profession. The data and information presented herein have that the likely cause for the two different shear transfer
been made generic to maintain confidentiality of the spe- factors is the difference in penetration rate of the suc-
cific projects. Any resemblance to a particular project is tion caisson into the seafloor, as further explained in
purely coincidental. the following sections.

235

Copyright 2005 Taylor & Francis Group plc, London, UK


during suction penetration. It is believed that, under
the higher penetration speeds experienced during self-
weight penetration, continuous pore water pressure
build-up creates a thin film of water along the soil-
caisson interface, which acts as a lubricant, thus
reducing the shear transfer factor to values below
1/St. Such observation is not new and is similar to the
experience gained with driven piles, whereby the
shear transfer factor during driving has been found to
be smaller than that during self-weight penetration.
Similar observations were also made when installing
caissons of very similar sizes at two near-by sites off-
shore West Africa, where the soil sensitivity (St) is in
the order of 4 to 6. The only difference was the instal-
lation speeds used in the two projects. At Site A, self-
weight penetration speeds varied between 3 to 5 m/min,
and all the suction caissons reached their target pene-
tration (between 10 m and 15 m BML) without the use
of suction. The shear transfer factor was back-calcu-
lated to be in the order of 0.15 to 0.18. However, at
near-by Site B, the suction caissons were installed at
much slower speeds (ranging from 0.6 to 2.4 m/min),
and the suction caissons reached the end of self-weight
penetration at shallower depths (between 8.5 m and
11 m). For Site B, the shear transfer factor was
back-calculated to be in the order of 0.20 to 0.22.
Therefore, again, it would appear that a deeper self-
weight penetration depth and a lower shear transfer
factor are associated with a higher rate of penetration.

2.3 Required under-pressure offset


As seen from the recorded suction data from two dif-
ferent GOM sites (Figs. 1a and b), a noticeable jump
(in the order of 48 kPa) in the required suction was
Figure 1. Theoretical predictions and field observation of observed at the start of the suction penetration phase,
required suction pressure for two Gulf of Mexico sites. for all the SPAR suction caissons. Although the suc-
tion data follow the predicted trend based on the shear
transfer factor of 0.33, the required under-pressures
2.1 Penetration rates
were consistently offset from the theoretical predic-
Typically, the suction caissons self-weight penetra- tion by about 48 kPa. As such, it appeared prudent to
tion rate is in the order of 1 to 3 m/min, while it is in modify the prediction mechanism to account for the
the order of 0.05 to 0.1 m/min during the suction pene- difference, i.e., offset the theoretical predictions to
tration phase. Therefore, self-weight penetration speeds match the field observations (shown as bold dash
are about 20 to 30 times faster than those achievable curves in Figures 1a and b).
during the suction penetration phase.
2.4 Revised shear transfer model
2.2 Shear transfer
In light of the installation data collected during various
Based on our experience with suction caisson instal- suction caisson installation monitoring, a new hypoth-
lation in typical Gulf of Mexico clays, the shear trans- esis is advanced. As mentioned earlier, it is believed that
fer factors for the self-weight and suction penetration the shear transfer factor is dependent on the rate of
phases were found to be approximately 0.23 and 0.33, penetration. Therefore, in principle, if the self-weight
respectively. Typically, the average soil sensitivity (St) penetration phase could be slowed down to penetration
of Gulf of Mexico clays is in the order of 3. Therefore, speeds similar to those of the suction phase, the shear
it can be seen that the shear transfer is less than 1/St transfer factor during the self-weight penetration phase
during self-weight penetration, while it is right at 1/St should be identical to that during the suction phase

236

Copyright 2005 Taylor & Francis Group plc, London, UK


(i.e.,   1/St). This would imply that the required are lowest. In some extreme cases, it was observed
underpressure is based on a single shear transfer factor that the correction was twice the actual suction (e.g.,
(typically, 1/St  0.33 for the Gulf of Mexico projects). the registered suction was 145 kPa, the correction was
However, as can be seen from the field monitoring 97 kPa, for an actual suction on the caissons interior
data from different GOM sites, the under-pressure of only 48 kPa). The corrections become relatively
needs to catch-up with the 0.33-shear-transfer pres- less significant as the suction phase progresses, since
sure curve (bold curve in Figures 1a and b), in order to the flow rates decrease and the measured differential
start the penetration process. This would explain the pressures are higher.
jump start underpressure described earlier. At deeper To obtain a calibration curve, the pumping system
self-weight penetrations depths, keeping all other is operated in open water, preferably at a depth close
variables the same, the amount of jump start under- to the seafloor, so as to account for ambient pressure
pressure would be expected to be higher. effects (if any). The differential pressures and the
flow rates are recorded as the pump is ramped up to
full power and then ramped down back to zero.
2.5 Hindcast verification
Based on installation data from various GOM pro-
jects, the predictions based on separate self-weight 4 BEVEL EFFECTS
penetration and suction penetration shear transfer fac-
tors with a jump start underpressure are essentially On three GOM projects, the suction caissons were
identical to the predictions based on the single shear fabricated with a 45-degree outward bevel, across
transfer factor of 0.33. This agreement validates our half the wall thickness at the caissons tip (Fig. 2b).
postulation of the shear transfer factors dependency The aim was to reduce the soil plug heave within the
on rate of penetration and the single shear transfer caisson, by pushing a larger percentage of soil dis-
factor hypothesis. placed by the caisson volume to the outside of the
caisson.
One project offered a unique opportunity to evalu-
3 PUMP CALIBRATION AND UNDER- ate the effects of a bevel at the caisson tip, as all but
PRESSURE CORRECTION one of the suction caissons were installed with the 45
half bevel described above, while one caisson had a
Before the start of any suction caisson installation, the full cross-section bevel, as further discussed below.
ROV pump system must be calibrated, in order to deter- Based on the experience gained during the sequen-
mine the correction that must be applied to the regis- tial installation of the first three caissons within one of
tered suction. Such calibration is needed because the anchor clusters (presented as locations A, B and C
suction values significantly in excess of the actual suc- in Fig. 1a), it became apparent that the soil resistance
tion applied to the caissons interior will be registered was getting progressively higher (due to the progres-
by the pumping system instrumentation. This over- sively shallower encounter of a buried landslide
registration (or parasitic pressure) is mostly due to deposit), and that it would not be possible to install
a strong venturi effect present at the inlet of the pres- the fourth caisson to the minimum required penetra-
sure gage (or transducer), and to a lesser degree to tion without sucking too close to wall-buckling pres-
hydraulic losses over the short distance (in the order sures. Therefore, it was decided to shorten the fourth
of 0.6 m) between the end of the suction stab and the caisson (location D) by removing a short section from
pressure measurement point. its bottom can. By reducing the length of the suction
The magnitude of the parasitic suction depends on
several factors, including, (1) the fluid velocity (directly
related to pumping rate and suction hose diameter),
(2) how far back the pressure gage/transducer is located
away from the inside of the caisson, (3) the presence
of any elbows or restrictions between the caisson and
the location of the pressure gage/transducer inlet, and
(4) the local geometry of the pressure gage/transducer
port (i.e., is inlet recessed from, or protruding above,
the inner surface of the suction hose, are there any burrs
or other imperfections at the inlet as a result of the
drilling and tapping of the inlet, etc.).
The suction pressure correction is particularly
important at the start of the suction phase, when the Figure 2. Effects of skirt tip geometry on percentage of
flow rates are highest and the differential pressures soil entering caisson.

237

Copyright 2005 Taylor & Francis Group plc, London, UK


caisson, its buckling limits would be improved sub- long caissons (i.e., L/D ratios 6). During some of
stantially, and the pad eye would end up at an appro- the GOM caisson installations, rotation of some of
priate embedment depth ratio promoting translational the caissons gradually became so excessive that the
behavior of the caisson under load. suction caissons had to be extracted from their initial
The section was removed from the bottom of the location and moved to an alternate location, or
fourth caisson by torch-cutting it on board the derrick embedment had to be terminated short of the design
barge. This created a sharper (55-degree from the hori- target penetration to prevent accumulating rotations
zontal) bevel across the full cross-section of the cais- in excess of the maximum allowable (7.5).
sons wall (see Fig. 2c). Although lighter than the other Suction caissons have large soil resistance to torque,
caisson, this fourth caisson reached a self-weight pene- as a result of their large size. This implies that a large
tration, which was 1 to 2 m deeper than those reached torque is required to force the suction caisson to rotate.
by the other three caissons installed within the same Two possible causes are described and discussed below.
cluster.
There were no reasons to suspect that there might
be any substantial differences in the shear strength
profiles at the four caissons forming that particular
cluster, except for the slightly shallower presence of
the top of the buried landslide (located some 10 m
below the self-weight penetration range). Comparing
the suction phase installation records for the four suc-
tion caissons (Fig. 1a), further confirmed that the shear
strength profile at the fourth anchor location did not
appear to be significantly different from those encoun-
tered at the first three anchor locations. Also, the
installation procedure for the last caisson was identi-
cal to that used for the other three caissons. Therefore,
the increase in self-weight penetration for the fourth
caisson cannot be attributed to any change in shear
strength, and it can only be concluded that the deeper
self-weight penetration of the fourth caisson is due to
the presence of a full bevel.
Finally, it is worth noting that the soil plug heave
for the caisson with the full and sharper bevel (loca-
tion D in Fig. 1a) was only 0.15 m, while the soil
heave for all the other caissons ranged from 0.4 to
1.0 m (with an average heave of 0.7 m). The observed Figure 3a. Torque induced by subsea currents on mooring
0.15 m heave compares favorably with the theoretical chain.
estimate of 0.25 m, assuming that 100% of the soil
displaced by the caisson wall is pushed to the outside
of the caisson. Therefore, it can be concluded that the
sharper outward bevel was very effective in forcing
the soil to the outside of the caisson.

5 CAISSON ROTATION POSSIBLE CAUSES


AND CONTROL

Random amounts of rotation were observed during


the installation of a number of suction caissons at vari-
ous sites. Others have also reported observing large
suction caisson rotations (e.g. Newlin 2003). Typically,
anchor suction caissons are designed for a maximum
rotation of 7.5 from their target heading, imposed
to prevent out-of-plane eccentric loading from over-
stressing the pad eye.
The risk of accumulating excessive rotation during
penetration into the seafloor increases for unusually Figure 3b. Caisson isolated from subsea current forces.

238

Copyright 2005 Taylor & Francis Group plc, London, UK


The first possible cause considers that the pad eye can be due to one or more reasons, singly or in com-
and/or the internal stiffeners may be slightly inclined bination, including accumulation of excessive tilt or
from the vertical, and act as a vane (or fin) inducing rotation, reaching the allowable under-pressure against
an unbalanced torque. However, in view of their small soil plug failure or against caisson buckling, or failure
size relative to the entire caisson, and unless the pad of the installation equipment.
eyes or the stiffeners were noticeably inclined, it is The decision to stop penetration of a particular
unlikely that they could create sufficient torque to caisson short of its target embedment can only be
induce the large rotations observed. Moreover, this done safely and cost-effectively if it can be supported
hypothesis would imply a unidirectional rotation and by appropriate real-time information. This points to
would not be able to explain the changes in rotation the need for having in place a competent monitoring
direction observed during a given caisson installation. program and software package to perform real time
The second cause considers that subsea currents can re-analysis of the suction caisson to confirm its
exert significant drag forces on the mooring chain and acceptability in the present configuration.
cable, as shown in Figure 3a. These lateral forces could The caissons axial capacity and lateral behavior
induce a substantial torque at the head of the caisson, under the design mooring loads specific to that par-
and the observed changes in rotation direction could ticular anchor leg need to be checked. The caissons
be easily explained by changes in the subsea currents pullout capacity and lateral deflection behavior under
direction and intensity. An opportunity to indirectly various loading for the actual tilt, twist and embed-
validate this hypothesis occurred during installation ment achieved during installation need to be evalu-
of one of the caissons at one of the GOM sites. For the ated when assessing the acceptability of a particular
installation of that particular caisson, the mooring suction caisson in its as-installed condition.
chain was laid on the seabed (Fig. 3b). This effectively To help reach sound and timely decisions, while
isolated the suction caisson from any lateral forces under the extreme pressures and time constraints asso-
that may have been exerted on the mooring chain/cable. ciated with any offshore construction project, one must
No measurable rotation was observed for that particular have prepared a concise decision tree addressing pos-
caisson. sible installation problems and their consequences.
Finally, although it is only an indirect confirmation Such a decision tree should define what would consti-
of our hypothesis, our experience has been that no tute an acceptable caisson, considering various com-
rotation was ever observed when installing caissons bination of each of the installation factors, and help
used as foundations for sub-sea facilities (i.e., with no make prompt decisions on the acceptability of a suc-
chain that could be acted upon by currents). tion caisson terminated short of its target penetration.
To build the decision tree, the design target penetra-
tion is first determined based on the maximum allow-
6 EMBEDMENT SHORT OF TARGET able tilt and rotation of the suction caisson, and the
pullout capacity desired for the maximum mooring line
Generally, the design of suction caissons for an anchor load conditions. The absolute minimum penetration is
system is based on the most loaded caisson, i.e., one then determined, based on the allowable pullout cap-
size fits all design. This means that except for the most acity and lateral deflection of the suction caisson for
loaded caisson, all the other caissons are to some the specific mooring line load conditions for that par-
degree over-designed and, thus, could stand to be ticular anchor leg. Reduced embedment dictates tighter
installed with less embedment. tolerance on the allowable tilt and rotation limits for
During an actual suction caisson installation, it may the caisson. Intermediate penetration scenarios can be
become necessary to stop embedment of a particular considered, and the corresponding allowances of the
suction caisson short of its design penetration. This tilt and rotation of the caisson determined. The allow-
able tilt and rotation criteria can be tabulated (Table 1)
Table 1. An example of a pile acceptability matrix. to allow quick on-site decisions to be made in order to
determine the acceptability of a suction caisson, or
Tilt (m) the necessity of a pump out procedure and relocation.
Penetration
allowance* 2.5 5.0 7.5
7 PADEYE DEPTH OPTIMIZATION VS.
Rotation ()
DESIGN ROBUSTNESS
2.5 3.0 1.5 0.6
5.0 1.5 0.6 0.0
7.5 0.6 0.0 0.0 The design of suction caisson anchors generally
includes optimization of the pad eye depth, so as to
* Negative values denote penetration allowances less enhance the performance of the suction caisson under
than target penetration. the design mooring loads, by involving two conflicting

239

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 2. Illustration of the effect of pad eye depth on design robustness.

Shorter caisson (Figure 4a) Longer caisson (Figure 4b)


D  4.8 m, E  24 m, E/D  5 D  4.8 m, E  35 m, E/D  7.3

Pad Pad
Pad eye Desired Pad eye Desired
Embedment eye depth backward Embedment eye depth backward
Achieved depth, depth, ratio rotation depth, depth ratio rotation
embedment E (m) PD (m) PD/L (%) or plowing E (m) h(m) PD/L (%) or plowing

Target 24.0 16.0 66.7 Yes 35.0 21.0 60.0 Yes


0.6 m short of Target 23.4 15.4 65.8 Yes 34.4 20.4 59.3 No
1.5 m short of Target 22.5 14.5 64.4 Yes 33.5 19.5 58.2 No
2.1 m short of Target 21.9 13.9 63.5 Yes 32.9 18.9 57.4 No
3.0 m short of Target 21.0 13.0 61.9 No 32.0 18.0 56.3 No

The drawback of a fully optimized pad eye depth


design (i.e., 55 to 60% of the embedment depth) is
that such a design is much less robust to accommo-
date field decisions to reduce the achieved embed-
ment depth. This is because an optimized caisson
would quickly tend to rotate forward (i.e., toward the
floating structure) for small embedment reductions as
illustrated in Table 2, and on Figures 4a and 4b.

8 CONCLUSIONS/RECOMMENDATIONS

This paper has discussed several aspects of the prob-


lems encountered, and lessons learned, during the
installation of many suction caissons, including:
1 The specific shear strength profile at each anchor
cluster should be utilized to develop a more realis-
Figure 4. Lateral displacement of (a) Long and tic estimate of the required under-pressure, as the
(b) Short suction caisson under design loading. pressure predictions are based principally on the
shear strength profile.
2 The shear transfer factors at the soil-caisson inter-
goals, (1) reduce the pad eye depth, to flatten the chain face are dependent on the rate of penetration of the
trajectory (reverse catenary) through the soil and, suction caisson. The shear transfer factor increases
minimize the axial uplift, and (2) increase the pad eye (up to 1/St) as the rate of penetration decreases.
depth, to promote pure translation (or plowing) or a Therefore, self-weight penetration and required
slight backward rotation (beneficial to reduce any underpressure predictions need to take into account
tendency for a separation to form on the back side of the rate of penetration.
the caisson), and avoid forward rotation of the caisson 3 The measured differential pressure does not accur-
(toward the floating structure). ately represent the under-pressure existing inside
For caissons with L/D ratios 5, the optimization the suction caisson. There is a need to calibrate the
process results in the pad eye being located approxi- pumping system in order to determine the correc-
mately at 2/3 the embedment depth. For caissons with tion that must be applied to the registered suction.
L/D ratios 6, the 2/3-rule of thumb would still pro- 4 A jump start pressure appeared to consistently
vide the desired backward rotation behavior. However, offset the theoretical suction predictions in order to
as suction caissons get longer and more slender, there match the field observations. This supports our
is a greater need for the optimum pad eye depth to postulation of the shear transfer factors dependency
become shallower (i.e. 55 to 60% of the embedment on rate of penetration and the single shear trans-
depth), in order to minimize the steepness of the mooring fer hypothesis.
chain reverse catenary, and, thus, to reduce the axial 5 The observed soil plug heave for a caisson with a
upward pullout. full bevel was only 0.15 m compared to the average

240

Copyright 2005 Taylor & Francis Group plc, London, UK


soil heave of about 0.7 m for all the other caissons, REFERENCES
implying that the full bevel was very effective in
forcing the soil to the outside of the caisson. Audibert, J. M. E., Clukey, E. and Huang, J. 2003. Suction
6 Excessive rotation during penetration into the Caisson Installation At Horn Mountain A Case History.
seafloor can impede the installation process, pos- Proceedings of The Thirteenth (2003) International
Offshore and Polar Engineering Conference, Honolulu,
sibly due to lateral forces induced by subsea cur- Hawaii, USA.
rents on the mooring chains. Huang, J., Cao, J. and Audibert, J. M. E. 2003. Geotechnical
7 In the event of an embedment short of target, a deci- Design Of Suction Caisson In Clay. Proceedings of The
sion tree needs to be in place to reach sound and Thirteenth (2003) International Offshore and Polar
timely decisions. Engineering Conference, Honolulu, Hawaii, USA.
8 Pad eye depth optimization improves the perform- Newlin, J.A. 2003. Suction Anchor Piles for the Na Kika
ance of the suction caisson anchors during its oper- FDS Mooring System. Proceedings, International
ational life, at the expense the expense of robustness Symposium on Deepwater Mooring Systems: Concepts,
of the design during installation. Design, Analysis and Materials, American Society of
Civil Engineers, Houston USA.

ACKNOWLEDGEMENTS

The authors are grateful to Fugros management for


supporting the preparation, publication and presenta-
tion of this paper at the ISFOG conference.

241

Copyright 2005 Taylor & Francis Group plc, London, UK


Centrifuge tests on axial capacity of suction caissons in clay

W. Chen & M.F. Randolph


Centre for Offshore Foundation Systems, The University of Western Australia, Perth, WA, Australia

ABSTRACT: The axial capacity of suction caissons during installation and pullout was investigated for cais-
sons installed either by jacking or by suction. The study was based on centrifuge model tests carried out in both
normally consolidated (NC) and lightly overconsolidated (LOC) kaolin clay. The sensitivity of the soil was
determined from cyclic T-bar tests and it was found that the fully remoulded shear strength gave good estimates
of the shaft friction during caisson installation. The measured installation resistance during suction installation
was similar to that during jacked installation in both types of clay. No essential difference was found in the long
term uplift capacity of caissons installed by jacking or by suction, according to unsealed pullout tests in the NC
clay, and sealed pullout tests in the LOC clay. Lower bound external shaft friction ratios during pullout after
consolidation were derived, together with consistent end-bearing factors.

1 INTRODUCTION 2 PHYSICAL MODELLING

Recently, suction caissons were successfully installed Jacked and suction installation tests on the caissons
in water depths of around 2000 m as anchoring piles were all carried out at 120 g using the beam centrifuge
for the Na Kika Floating Production System (FPS) in at the University of Western Australia. Two model suc-
the Mississippi Canyon Area of the Gulf of Mexico tion caissons were made of 6061 T6 aluminium, with
(Newlin 2003b). The critical design case for these an outside diameter of 30 mm, a height, L, of 120 mm,
caissons gave a loading angle at the caisson of 36 and a wall thickness of 0.5 mm, thus representing
(Newlin 2003a), for which the uplift capacity of the a prototype caisson 14.4 m long, 3.6 m in diameter
caisson would be critical (Clukey & Phillips 2002). and 0.06 m wall thickness. The surfaces of the cais-
However, currently there is no specific design code sons were anodized after sandblasting. A pad-eye was
for the axial capacity of suction caissons in soft marine located at 0.4 L (48.5 mm, model scale) from the tip
clay, other than that for open-ended driven piles (API of each caisson. Caisson 1, tested in the NC clay, had
RP2A 1993). A design approach was recently pro- two stages of internal stiffener designed to accommo-
posed suggesting that suction installation would lead date miniature pressure cells (Chen & Randolph
to lower external friction values than if the caisson 2004a). The stiffener increased the wall thickness to
was jacked into position (Andersen & Jostad 2002). 1 mm between 41 mm and 56.5 mm from the tip, and
While contrary conclusions were obtained from cen- then to 1.5 mm for a further 7 mm (opposite the pad-
trifuge tests in NC clay (Chen & Randolph 2004a), eye) before reverting to the 0.5 mm wall. Caisson 2,
the thin-walled nature of suction caissons would still tested in the LOC clay, was similar to Caisson 1, except
suggest lower shaft friction than for conventional that the total pressure transducers were situated lower
driven piles. in the caisson. The 1.5 mm thick wall thickness
In this paper, results from a series of centrifuge started at 35 mm above the tip, and the wall thickness
tests in both normally consolidated (NC) clay and then decreased in stages to 1 mm at 45 mm from the
lightly overconsolidated (LOC) clay are described. tip, and then to 0.5 mm at 56.5 mm from the tip.
The caissons were installed either entirely by jacking, Methods of preparing the NC kaolin clay sample fol-
or by jacking to approximately 50% penetration (simu- lowed those described by House (2002). The LOC sam-
lating self-weight penetration) before applying suc- ple was consolidated at 180 g and then tested at 120 g,
tion to achieve full penetration. The vertical capacity targeting an OCR of 1.5. The sensitivity of the sample
of the suction caissons was measured after allowing was investigated by cyclic T-bar tests as suggested by
full consolidation, either with the top lid vented or Watson et al. (2000), and the result of a typical test car-
with a sealed lid. ried in the LOC clay are shown below in Figure 1. As

243

Copyright 2005 Taylor & Francis Group plc, London, UK


su (kPa) depth = 8 m Penetration resistance (kPa)
1.00 depth = 9 m
-30 -10 10 30 depth = 10 m 0 50 100 150 200
0.90

Remoulded ratio
depth = 11 m 0
0 depth = 12 m
0.80 B13JCC, by jacking,
2 2
Depth of tip (m)

Remoulded to k=1.64 kPa/m, alpha=0.42


0.70 ~40% of the
4 4

Embedment (m)
1st out 1st in 0.60 original strength B13SCC, by suction,
6 6 k=1.64 kPa/m, alpha=0.43
0.50
8 Theoretical: Nc=7.5,
0.40 8
10 2nd in alpha=0.42
0.30 10
12 1 3 5 7 9 11
14 No. of cycles 12
14
Figure 1. Cyclic T-bar test in LOC clay (OCR  1.5). 16

Table 1. Key soil properties for NC and LOC kaolin clay. Figure 2. Penetration resistance for caissons installed by
jacking and by suction in LOC clay (OCR  1.5).
Property NC LOC

Undrained shear strength gradient, dsu/dz 1.2 1.7 resistances were just greater than 150 kPa. For the
Effective density,  (kN/m3) 6.8 7.2 suction-installed caisson, there is a slight fluctuation
Strength ratio, su/vo 0.18 0.24 in resistance at a depth of 7 m, as the control system
Coefficient of earth pressure at rest, K0 0.65 0.70 switched from jacked (self-weight) installation to suc-
Sensitivity factor, St 22.8 2.5 tion installation.
The caisson was stopped when there was a sudden
jump in the readings of the internal pore pressure trans-
shown in the right-hand graph, after 11 cycles of pene- ducer on top of the caisson, indicating contact between
tration, the undrained strength of the soil over the depth the soil and the caisson lid. The ultimate depth of instal-
range 8 to 12 m had reduced to a relatively steady lation (Linstalled) was found to be similar for both
value of 40% of its original value. Allowing for the methods of installation in NC clay (Chen & Randolph
fact that some softening occurs during the initial pene- 2004a). For the tests in LOC clay, the final depths of
tration of the T-bar (Einav & Randolph 2005), the installation were 13.9 m and 14.0 m respectively, for
sensitivity of the LOC clay may be estimated as just suction installation and jacked installation, with a dif-
greater than 2.5. Similar cyclic T-bar tests carried out ference of less than 1% (Fig. 2).
in the NC clay suggested a sensitivity of 2 to 2.8. The penetration resistance of the caisson in clay is
Soil properties for the NC kaolin clay sample can composed of two parts: tip resistance and shaft fric-
be found in Chen & Randolph (2004a). A summary of tion. As is customary, these components are assumed
key soil parameters for the NC and LOC clays used proportional to the local shear strength of the clay,
here is given in Table 1. using a bearing factor, Nc, and friction ratio, . It was
shown by Chen & Randolph (2004a) that an Nc value
of 7.5 appears appropriate during caisson penetration,
3 PENETRATION RESISTANCE although the deduced value is affected slightly by
assumptions regarding the internal shaft resistance
Chen & Randolph (2004a) compared the penetration above the pad-eye stiffener. In the analyses, allowance
resistance of caissons installed in NC clay by jacking was made for remoulding of the soil and incomplete
or using suction, and found close agreement in the f low of the soil back against the internal caisson wall
overall resistance profiles. above the stiffener (Erbrich & Hefer 2002), consider-
Both installation methods were also used in the LOC ing 3 possible modes: 1) full attachment; 2) no attach-
clay. The measured penetration resistance, p, defined ment; and 3) partial attachment, with some water
as the net penetration force divided by the gross cross- trapped between the soil and the wall (Figs 3ac).
sectional area of the caisson, is depicted in Figure 2 Full attachment is considered possible for very soft
for these two types of installation. The two tests were clay with high sensitivity, where the strength of the
carried out in soil with an identical undrained strength soil is greatly decreased after passing the stiffener. In
gradient of 1.64 kPa/m. that case, the shaft friction inside and outside the cais-
It can be seen in Figure 2 that the penetration resist- son would be taken as the remoulded strength. The full
ance is very similar regardless of the method of instal- detachment case is likely only for overconsolidated
lation, especially for penetration greater than 10 m. clay, with high strength ratio, su/v0 and low sensitivity.
At the end of installation, at a depth of 14 m, both For the NC or LOC clay used here, with medium

244

Copyright 2005 Taylor & Francis Group plc, London, UK


Capacity (kPa)
i-a= i-a.su i-a= 0 kPa Reduced -300 -250 -200 -150 -100 -50 0 50 100 150
i-a= i-b i-a 0
B11SOC 2
Water B2JOC

Embedment (m)
4
B9SOI Installation
6
B2JOI
8
p = P/A
10
Real capacity 12
(a) Full attachment (b) No attachment (c) Partial attachment Pseudo 14
(very soft) (very stiff) (assumed here) capacity for 16
tests OC Consolidation
Figure 3. Flow mechanism of the soil inside the caisson.
Figure 4. Uplift resistance during unsealed pullout in
NC clay.
strength ratio and sensitivity, the soil plug is expected
to relax gradually back towards the caisson wall, but
is likely to trap some water between the soil and the Table 2. Shaft friction ratio during unsealed pullout of the
internal wall. Therefore, the shaft friction is likely to caisson in NC clay (assuming ext  int).
be much lower above the stiffener than below, although
it will not be zero. zmax dsu/dz pmax
Back-analysis of the axial resistance measured in Test No. (m) (kPa/m) (kPa) pmax/s-u 
NC clay revealed a difference of 10% on the deduced
 values during installation by varying the shaft fric- B2JOI 14.18 1.17 102 12.3 0.38
tion from zero to full strength above the stiffener. As B9SOI 14.14 1.20 110 13.0 0.40
a compromise, a nominal value of 0.5 kPa has been Average (OI) 14.16 1.18 106 12.7 0.39
B2JOC 14.45 1.17 174 20.6 0.65
assumed for the average internal shaft friction above B11SOC 14.32 1.28 194 21.2 0.70
the upper edge of the stiffener for caissons in NC clay Average (OC) 14.39 1.23 184 20.9 0.68
(Fig. 3c). For the caisson tests in the LOC clay (OCR
1.5), considering the difference in both the strength Note: J for jacking, S for suction, O for open (vented),
gradient and the geometry of the internal stiffener, an I for immediate pullout, C for after consolidation.
average nominal shaft friction of 1 kPa was adopted
in that region. Varying the friction above the stiffener
between 0 and 2 kPa resulted in a difference of only The consolidation period is equivalent to a prototype
15% in the  value back-figured from the measured time of 1.7 years and the caisson was allowed to settle
installation data. freely under a constant (nominal self-weight) load
With the above assumptions, Chen & Randolph during that period.
(2004a) found the shaft friction ratio for caissons in The uplift resistance, expressed as the ratio of
NC clay to be 0.30 to 0.45. This is consistent with the uplift force to the gross cross-sectional area of the
sensitivity of St  2 to 2.8 deduced from cyclic T-bar caisson, is plotted against the embedment of the cais-
tests, following the normal industry assumption of sons in Figure 4, for unsealed pullout in NC clay, with
 1/St (API RP2A 1993). For tests in the LOC and without allowing consolidation. Similar values of
clay (Fig. 2), the penetration resistance could be sim- uplift capacity were obtained for both jacked caisson
ulated closely using shaft friction ratios of 0.42 for and suction-installed caissons.
jacked installation and 0.43 for suction installation. For the immediate tests, the pullout capacities were
These  values are also in good agreement with the both around 100 kPa. When analysing the uplift
measured sensitivity of 2.5. capacity, it is convenient to assume the same values
of  along the full length of the caisson, inside and
outside, since it is difficult to predict the shaft friction
4 UPLIFT CAPACITY DURING PULLOUT
above the internal stiffener. The  values thus back-
figured from the two tests are shown in Table 2,
4.1 Unsealed pullout in NC clay
and average 0.39. This value is similar to that dur-
After the caisson was installed to the target position, ing installation, which is as expected with no strength
unsealed pullout tests were performed either immedi- regain following remoulding.
ately after installation (denoted by OI), or after a Also plotted in Figure 4 are two unsealed pullout
consolidation period of 1 hour (denoted by OC). tests following consolidation, one for each type of

245

Copyright 2005 Taylor & Francis Group plc, London, UK


installation. It can be seen that at the beginning of the Capacity (kPa)
pullout, the axial capacity increased rapidly to a very -400 -300 -200 -100 0 100 200
high value, and then dropped abruptly before decreas- 0
ing smoothly with further movement of the caisson. 2
B11SCC
This pseudo capacity as indicated on the figure is
B12SCC 4

Embedment (m)
thought to be the result of transient suction developing
between the caisson top lid and the soil plug during con- 6
solidation; this suction is not sustained since the cais- Pullout Installation
8
sons were vented. The real capacity after the reduction
10
is considered to reflect the pullout capacity based on
sliding along the internal and external caisson surface. 12
Note that similar behaviour was not observed during 14
the immediate pullout tests, showing the influence of 16
consolidation on conditions within the caisson.
The post-consolidation tests show a significant Figure 5. Sealed pullout capacity versus embedment of
increase in the axial capacity, reaching around 184 kPa caissons in NC clay.
or 1.7 times that without consolidation. Assuming
similar shaft friction on both sides of the caisson leads
to back-figured  values of 0.65 for the jacked installa-   effective unit weight of soil
tion and 0.70 for the suction installation, with an aver- zbase  depth of caisson base
age value of 0.68 (Table 2). The increase is about 70% Abase  area of sealed base of the caisson
s  average undrained shear strength along the
compared to the value during immediate pullout. Close u

agreement was found between the two types of installa- caisson shaft
tion both in the resistance profiles and the derived  Wplug  submerged weight of the soil plug inside the
values. The reason why the uplift capacity, pmax, in the caisson
suction installation is slightly larger than that for jacked ext  external shaft friction factor.
installation (194 kPa versus 174 kPa) is the larger Sealed pullout tests were carried out for both suction-
strength gradient for the former (1.28 kPa/m versus installed caissons and jacked caissons. Unfortunately,
1.17 kPa/m). In fact, normalising the capacities by in some of the tests installed by jacking, misrouting of
dividing by the average shear strength, su, gives close external drainage lines led to cavitation, preventing a
values of 21.2 for the suction-installed caisson and 20.6 proper closed condition being achieved and redu-
for the jacked caisson. It should be noted that the ultim- cing the pullout capacities. Accordingly, the routing of
ate embedment of the caisson, zmax, was used when the drainage lines was subsequently adjusted to avoid
analysing the pullout capacity; zmax is generally larger this problem. Two of the successful sealed pullout
than Linstalled, due to settlements in consolidation. tests B11SCC and B12SCC, which were installed by
suction, will be reported here. In these tests, the cais-
4.2 Sealed pullout in NC clay son was first installed by suction, after which 1 hour
of consolidation was allowed at 120 g; then the caisson
Sealed pullout tests were carried out in the NC clay was pulled out vertically at a velocity of 0.3 mm/s.
after 1 hour consolidation. During sealed pullout, the The pullout resistance versus embedment of the cais-
drainage valve was closed to ensure the soil plug son for tests B11SCC and B12SCC is shown in the
remained inside the caisson during pullout. For sealed left side of Figure 5.
pullout, the uplift capacity is composed of three parts: It can be seen in Figure 5 that the absolute uplift
1) reverse end bearing resistance, 2) external shaft capacity of the sealed pullout test B11SCC is 319 kPa,
friction, and 3) the submerged weight of the soil plug while that for B12SCC is 288 kPa. The latter value is
and the caisson body. For the general case of a caisson around 90% of that measured in test B11SCC, which
with n stiffeners, the resistance p, defined as the may be attributed to the different soil strength values
axial force divided by the cross sectional area of the for those two tests. The strength gradient, dsu/dz, was
caisson, can be expressed as: 1.26 kPa/m for the former and 1.17 kPa/m for the
later. As shown in Table 3, however, the average nor-
malized sealed pullout capacities after consolidation
for these two tests, are 33.3 for B11SCC and 34.2 for
(1) B12SCC respectively. The average value of 33.8 is
1.6 times that during unsealed pullout under similar
where: conditions. This difference is due to significant
Nc  bearing capacity factor for caisson base reverse ending bearing capacity, which exceeds the
su base  undrained shear strength at base internal shaft friction.

246

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 3. Upper bound Nc values during sealed pullout of Table 4. Values of  during sealed pullout after consolida-
the caisson in NC clay. tion in NC clay for different assumed Nc values.

zmax dsu/dz pmax Test No. Nc  9 Nc  10 Nc  10.5 Nc  11 Nc  12


Test No. (m) (kPa/m) (kPa) pmax/s-u  Nc
B11SCC 1.26 1.10 1.02 0.94 0.78
B11SCC 13.84 1.26 325 33.3 0.70 12.5 B12SCC 1.13 0.99 0.93 0.86 0.73
B12SCC 14.41 1.17 292 34.2 0.70 12.3 Average 1.20 1.05 0.98 0.90 0.76
Average 14.11 1.22 309 33.8 0.70 12.4
(CC)

Note: CC denotes closed (sealed) pullout after is 10% greater than the  value derived from the
consolidation. unsealed caisson tests.
The  values suggested above are rather larger
than those proposed by Andersen & Jostad (2004) for
Another significant difference between sealed and suction-installed caissons, assuming that the caisson
vented failure is that the maximum pullout capacity wall is accommodated by soil moving inside the cais-
was developed after around 0.9 m of movement dur- son during suction installation. There are also other
ing the sealed pullout, while it was reached at a very aspects, such as the relatively long consolidation times
small displacement during the unsealed pullout. It can (prototype times of 1 year to achieve an estimated 90%
be inferred that the failure mode is ductile for the sealed consolidation), which suggest that significant outward
pullout but brittle for the unsealed pullout where the soil displacements occur during suction installation.
resistance is primarily in shaft friction. Field measurements reported by Newlin (2003b) also
When deriving the reverse end bearing resistance suggest little inward soil displacement occurs at typi-
factor during pullout after consolidation, the external cal levels of suction applied during installation.
 value can be assumed to be the same as that estimated
during the unsealed pullout (based on equal internal and
4.3 Sealed pullout in LOC clay
external shaft friction). It is also convenient from a
design perspective to calculate the end-bearing resist- Sealed pullout tests were carried out successfully in
ance factor relative to the shear strength at the original the LOC clay (OCR of 1.5), for both jacked and suc-
(maximum) embedment depth, ignoring the 0.9 m of tion-installed caissons. A consolidation period of one
upward movement prior to mobilising the peak cap- hour (1.7 years at prototype scale) was allowed fol-
acity. With that assumption, and adopting an external lowing installation. The velocity of pullout was also
 value of 0.70, the reverse end bearing resistance set to be 0.3 mm/sec, which is the same as that used
factor Nc was calculated from Equation 1 as 12.5 (Table for the tests in NC clay. The uplift resistance versus
3). Also shown in the table is a parallel test B12SCC, caisson embedment is shown in the left part of Figure 6,
for which an Nc value of 12.3 was obtained. for the test B13JCC installed by jacking, and test
However, the external  value of 0.7 should be B13SCC installed by suction. It can be seen in the
considered a lower bound, since excess pore pressures graph that the uplift resistance profiles are very simi-
inside the caisson will take longer to dissipate com- lar for both types of installation.
pared to those outside the caisson due to the longer The maximum uplift resistances were found to be
drainage paths. Also, the severe disturbance caused 379 kPa and 389 kPa respectively for tests B13JCC
by the internal stiffener is likely to lead to longer and B13SCC. The corresponding normalized uplift
recovery times for the internal shaft friction, com- capacities are 33.2 and 34.1, respectively, for the
pared to that on the external wall. The lower bound above two tests, with a difference of only 3%. As a
external value of  estimated assuming equal internal result, it can be inferred that there is little difference
and external friction will lead to an upper bound between the sealed pullout capacities of caissons
value of Nc for the sealed caisson. The average Nc installed by jacking or by suction. Interestingly, the
value of 12.4 obtained here above is slightly larger average normalized uplift capacity is found to be
than the value of 12 considered as an upper bound by close to that measured in the NC clay, indicating that
Chen & Randolph (2004b). the larger axial capacity of caissons in the LOC clay
Table 4 shows the values of external shaft friction is mainly attributable to the higher strength gradients
ratio, , deduced from the tests on sealed caissons, for in the OC clay.
different assumptions regarding the reverse end-bearing The external shaft friction ratio  can be deter-
resistance. The average  values are surprisingly mined from the measured pullout capacity according
high, exceeding unity for Nc values less than 10.5. Even to Equation 1, by assuming different Nc values, as
taking Nc  12, which is considered a likely upper shown in Table 5. It can be seen in the table that the
limit, an average  value of 0.76 was obtained, which deduced  values rise to greater than unity if Nc is set

247

Copyright 2005 Taylor & Francis Group plc, London, UK


Capacity (kPa) the caisson lid were found to be similar for cais-
-500 -400 -300 -200 -100 0 100 200 sons installed by jacking and by suction.
0 4 No essential difference was found in the unsealed
B13JCC, by jacking 2 uplift capacity of jacked or suction-installed cais-
B13SCC, by suction sons, either during immediate pullout or during
Embedment (m)

4
extraction after full consolidation.
p = P/A 6 5 Assuming identical shaft friction inside and out-
Installation
Pullout
8 side of the caisson,  values during unsealed pull-
10 out after consolidation in NC clay were 0.70,
12 some 70% larger than during immediate pullout.
6 The sealed pullout resistance after consolidation in
14
NC clay gave combinations of external  and Nc
16 values of (0.70, 12.4) or (0.76, 12), with  exceed-
ing unity for Nc of 10 or less.
Figure 6. Sealed pullout resistance versus embedment of
caisson in LOC clay (OCR  1.5). 7 No difference was found on the sealed uplift
capacity of jacked or suction-installed caissons
after consolidation in the LOC clay. Adopting an
Table 5.  values during sealed pullout after consolidation Nc value of 12 led to a corresponding  value of
in LOC clay for different assumed Nc values. 0.73, or slightly less than that for the NC clay.

Nc Although not shown here, similar radial total stress


zmax dsu/dz changes were measured around the caissons for both
Test No. (m) (kPa/m) pmax/s-u 12 10.5 9.5 types of installation (Chen & Randolph, 2004a,b).
This suggests that the major part of the caisson wall is
B13JCC 14.01 1.64 33.2 0.68 0.90 1.05 accommodated by soil being displaced outside the cais-
B13SCC 13.92 1.64 34.1 0.77 0.99 1.14 son, rather than being drawn into the soil plug, during
Average 13.97 1.64 33.7 0.73 0.95 1.10 both suction installation and jacked installation.

to be less than 10. Adopting a possible upper bound


Nc value of 12, as assumed for the NC clay, would ACKNOWLEDGEMENTS
result in  values of 0.68 and 0.77, respectively for
jacked installation and suction installation. The aver- The work presented in this paper forms part of
age  of 0.73 is slightly smaller than the value of 0.76 the activities of the Centre for Offshore Foundation
obtained in the NC clay. This would agree with the Systems (COFS), established and supported under
trend in most pile design guidelines for the shaft fric- the Australian Research Councils Research Centres
tion ratio to decrease with increase OCR or strength Program. This support is gratefully acknowledged.
ratio, su/v0. Special thanks go to Mr. D. Herley and Dr. A. House
for the assistance during the centrifuge tests.

5 CONCLUSIONS
REFERENCES
The penetration resistance and uplift capacity of cais-
sons installed by jacking and by suction in both nor- Andersen, K.H. & Jostad, H.P. 2002. Shear strength along
mally consolidated (NC) and lightly overconsolidated outside wall of suction anchors in clay after installation.
(LOC) kaolin clay have been investigated through a In J.S. Chung et al. (eds). Proc. 12th Int. Offshore and
series of centrifuge tests. The following conclusions Polar Engrg. Conf., ISOPE2002, 785794.
Andersen, K.H. & Jostad, H.P. 2004. Shear strength along
have been drawn from the results:
inside of suction anchor skirt wall in clay. Proc. 36th
1 The sensitivity of the clay was estimated from Annual Offshore Tech. Conf., Huston, Paper OTC 16844,
cyclic T-bar tests, giving St values of 2 to 2.8 for 113.
the NC clay and 2.5 for the LOC clay. API (1993). RP2A: Recommended practice for planning,
designing and constructing fixed offshore platforms. API
2 During caisson installation, the  value was found
Rec. Practice 2A. American Petroleum Institute, Dallas,
to average around 0.39 in NC clay and 0.42 in the Texas.
LOC clay, and therefore consistent with a value of Chen, W. & Randolph, M.F. 2004a. Radial stress changes
1/St as commonly assumed for pile installation. around caissons installed in clay by jacking and by
3 The penetration resistance and the maximum depth suction. In J.S. Chung et al. (eds). 14th Int. Offshore and
of installation before contact of the soil plug with Polar Engng. Conf., ISOPE 2004, Toulon, France, 2004.

248

Copyright 2005 Taylor & Francis Group plc, London, UK


Chen, W. & Randolph, M.F. 2004b. Radial stress changes Newlin, J.A. 2003a. Suction anchor piles for the Na Kika
and axial capacity for suction caissons in soft clay. (sub- FDS mooring system. Part 1: Site characterization and
mitted for publication in Geotechnique) design. In J. Zhang & R.S. Mercier (eds). Deepwater
Clukey, E.C. & Philips, R. 2002. Centrifuge model tests to Mooring Systems: Concepts, Design, Analysis, and
verify suction caisson capacities for taut and semi-taut Materials, ASCE. Houston, USA, Oct., 2003, 2854.
legged mooring systems. Proc. Int. Conf. on Deep Offshore Newlin, J.A. 2003b. Suction anchor piles for the Na Kika
Tech. New Orleans, USA, Nov., 116. FDS mooring system. Part 2: Installation performance.
Einav, I. & Randolph, M.F. 2005. Combining upper bound In J. Zhang & R.S. Mercier (eds). Deepwater Mooring
and strain path methods for evaluating penetration resist- Systems: Concepts, Design, Analysis, and Materials,
ance. Int. J. Num. Methods in Eng., in press. ASCE. Houston, USA, Oct., 2003, 5557.
Erbrich, C. & Hefer, P.A. 2002. Installation of Laminaria Watson, P.G., Suemasa, N. & Randolph, M.F. 2000. In
suction caissons-a case history. Proc. of the 34th Annual J.S. Chung et al. (eds). Evaluating undrained shear
Offshore Tech. Conf., Houston, Paper OTC 14240, strength using the vane shear apparatus. Proc. 10th Int.
21572170. Conf. on Offshore and Polar Engng, ISOPE2000,
House, A. 2002. Suction caisson foundations for buoyant Seattle, 2, 485493.
offshore facilities. PhD Thesis, The University of Western
Australia.

249

Copyright 2005 Taylor & Francis Group plc, London, UK


Evaluation of recovery of wall friction after penetration of
skirts with laboratory and field tests

Y. Yoshida, N. Masui & M. Ito


Obayashi Corporation, Tokyo, Japan

ABSTRACT: When skirt suction foundations are installed into cohesive soils, wall friction resistance on
skirts reduces due to strength decrease of remolded soil along the skirts and the shear strength gradually
recovers through dissipation of excess pore water pressure and thixotropy. Experiments in a laboratory and a
field have been conducted to evaluate the recovery for alluvium soft clay in Osaka Bay, Japan. Laboratory tests
under 1 g and 25 g fields demonstrate that the pore water pressure dissipation has smaller effect on the recovery
of wall friction than thixotropy for the clay. The result of the field tests using a larger scale skirt corresponds
with that of the laboratory tests and nearly the same recovery rate of the wall friction is obtained.

1 INTRODUCTION

Skirt suction foundations have been applied to the


substructures of offshore concrete platforms in the
North Sea and recently they were used for breakwaters
in Japan (Zen et al. 2003). The skirts are usually cylin- Suction head
drical units made of concrete or steel with a closed
top shaped as dome or plate. The skirts are penetrated
into the seabed by means of selfweight and ballast and
at the final stage by a differential hydrostatic pressure
Bottom slab
generated by pumping up water from inside the skirts.
The underpressure is called suction. A schematic Suction
view of the skirt suction foundations is shown in
Figure 1. Underpressure
When skirts are penetrated into cohesive soils, the Seabed Skirt
shear strength of a thin zone of the clay along the
skirts reduces by remolding and generation of excess
pore water pressure. It is assumed that the shear
Figure 1. Schematic view of skirt suction foundations.
strength reduces more the higher the sensitivity of clay
is. However the shear strength is known to increase
with time after penetration and the phenomenon is because of the shear strength increase at the interface
called as set-up effect. The causes are believed to be between skirt and soil. Therefore for skirt suction
dissipation of excess pore pressure and thixotropy, foundations, which are mainly supported by skin fric-
which is defined as recovery of shear strength to the tion related to the shear strength of soil, understand-
original value after softening by remolding. Those ing of the recovery is essential to estimate not only
effects on recovery are expected to vary with time vertical capacity but also pull-out capacity. From the
(e.g. Andersen et al. 2002). view point of geotechnical design of skirt suction
Vertical loads applied to the skirt suction founda- foundations, questions are how fast the shear strength
tions on soft clay ground are firstly supported by the recovers and what the governing factor for the strength
development of excess pore water pressure inside the regain is.
skirt and friction along the outside of the skirts. Then This paper presents experimental results acquired
some part of the loads is shifted to the tip resistance from a series of laboratory and field tests using small
of the skirt and friction along the inside of the skirt and large sizes of skirts penetrated into soft clays.

251

Copyright 2005 Taylor & Francis Group plc, London, UK


2 LABORATORY TEST RESULTS longer set-up time is, the larger the friction factor is
acquired.
Two sets of laboratory test have been conducted prior
to field tests. The clay used in the lab tests were trans-
2.2 Test in 25 g field
ported from near the location of the field tests and its
physical properties are shown in Table 1. Test speci- A plastic skirt with double cylindrical walls was pen-
mens made of plastic were penetrated repeatedly at etrated by jack after set-up periods of 0.5, 3, and 15
certain recovery intervals and each time the specimen hrs under 25 g field. Those set-up times are equivalent
was raised after the penetration in order to measure to 13, 78, and 390 days under 1 g field respectively
friction at the same depth. The first lab test was under considering the scaling law for consolidation time.
1 g field and the second test under 25 g field employed The outer and inner walls have a diameter of 370 mm
a centrifuge to accelerate the dissipation of water and and 150 mm each and the thickness of the walls is
consolidation of the clay. 4 mm. Considering the change of the shear strength
In this research the recovery of shear strength is during the test as illustrated in Figure 4, 11 kN/m2 and
evaluated in terms of friction factor calculated with 14 kN/m2 are used for set-up periods of 0.5, 3 hrs and
the Equation 1. Undisturbed shear strength is used in set-up period of 15 hrs respectively when estimating
the calculations. friction factors.
Figures 5 to 7 shows friction resistances obtained
(1) during the tests. The small vibrations in the curves are
due to discrete measurements and steady penetration
where Qf  friction resistance;   friction factor; was confirmed.
su  undrained shear strength; and Af  contact area Friction factor  at the first penetration and raising
with soil. is 0.3 and 0.2 respectively as shown in Figure 5 and a
Strictly speaking the friction factors contain the
effects of the material property related to the rough-
ness of the skirt material in addition to the change of
the shear strength of the soil. However, since the
effect of the material property is expected small and
the form of the equation is thought typical in practical
design, the equation is used in this research.
The two tests were already reported (Masui et al.
2004) and the results are briefly described below.

2.1 Test in 1 g field


The test specimen was 300 mm long, 100 mm wide
and 5 mm thick plastic plate and it was penetrated by
jack twice at certain intervals for recovery.
The observed friction resistance is shown in Figure 2
with the shear strength of the clay. From 0.3 at the ini-
tial penetration friction factor  is increased to 0.7 at the
beginning of the 2nd penetration. The rapid decrease Figure 2. Friction factors at 1 g field test.
after the peak is estimated due to gradual increase of
remolded soil area along the skirt. Instead of the suc-
ceeding decrease, it is thought to be safe to count the 1.0
set-up effect within small settlement such as design
0.8
Friction factor,

for serviceability limit state.


Friction factors corresponding to all tested set-up 0.6
periods are graphically illustrated in Figure 3. The
0.4
Table 1. Physical properties of clay for lab tests.
0.2
Liquid limit wL 64%
Plastic limit wP 29% 0.0
0 30 60 90
Plasticity index Ip 35
Time (day)
Soil particle density s 2.64 g/cm3
Sensitivity St 2.5
Figure 3. Recovery of friction with time.

252

Copyright 2005 Taylor & Francis Group plc, London, UK


reason of the slightly smaller  at raising is presum- friction factors after 0.5 and 3 hrs were almost 0.3
ably the fact that the clay is subject to the cyclic indicating relatively faster recovery of the shear
remolding by the raising operation following the pene- strength. Because of decrease in friction due to fur-
tration operation. ther remolding, friction factor gradually reduces with
Friction at re-penetration after set-up periods of increase in the penetration depth as shown in the curves
0.5, 3, and 15 hrs are shown in Figures 6 and 7. The below the depth of 70 mm in Figure 6 and below
90 mm in Figure 7.
Figure 7 shows that even after 15 hrs (equivalent to
su(kN/m2) 390 days under 1 g field) the friction factor became
0 10 20 only in between 0.3 and 0.4. The increase is much
0
smaller compared to the 1 g test result shown in

20

40
depth (mm)

before test
60
after test

80

100

120

Figure 4. Undrained shear strength of clay at 25 g field test.

Figure 6. Friction factors after set-up periods of 0.5 and


3 hrs.

Figure 5. Friction factors at the initial penetration and


raising. Figure 7. Friction after set-up period of 15 hrs.

253

Copyright 2005 Taylor & Francis Group plc, London, UK


Figure 3. Reasons for the discrepancy are estimated buoyancy compensator skirt waling
as follows.
Figure 8 shows that the dissipation of pore water
pressure was almost completed after 10 hrs equivalent sheet piles
to 260 days under 1 g field. The smaller increase com-
pared with the 1 g field test results from uncompleted
thixotropy process because the scaling law is not
applicable to thixotropy. Namely the increase due to
the pore water pressure dissipation is not significant
in comparison with that due to thixotropy. pit water level

3 FIELD TEST ground level

After the laboratory test, a much larger scale of skirt buoyancy


was penetrated at field to evaluate the recovery of compensator
shear strength of clay in actual field (Ito et al. 2004).

3.1 Test conditions excavation bottom


The test site is located in a reclaimed land along alluvial clay
skirt
Osaka bay area in Japan. A test pit was made by excav-
ating the land up to alluvial clay, which had been
dredged to reclaim about 40 years ago. The physical
properties are shown in Table 2.
Figures 9, 10 and 11 illustrate the test pit and a test
specimen. The test specimen was made of steel and Figure 9. Test setting.
contains a buoyancy compensator (the hexagonal col-
umn in Figure 11) and a skirt part of three cylinders.
The diameter of the cylinder is 1.0 m and the thick-
ness of the skirt is 12 mm.
After pouring water in the pit, a set of penetration
and raising phases was repeated three times with
intervals for friction recovery. Figure 12 illustrated
the procedure together with penetration and raising
depths and interval time.

140
Pore pressure (kPa)

120

100 Figure 10. Test pit.

80
Hydrostatic pressure = 79kPa
60
0:00 3:00 6:00 9:00 12:00 15:00
Time (hr)

Figure 8. Pore pressure dissipation measured at skirt tip


during the set-up period of 15 hrs.

Table 2. Physical properties of clay for field test.

Liquid limit wL 7284%


Plasticity index Ip 5161
Wet density t 1.64 g/cm3
Soil particle density s 2.69 g/cm3
Figure 11. Test specimen.

254

Copyright 2005 Taylor & Francis Group plc, London, UK


The undrained shear strength of the clay was illus- for test 1 and the strength measured at test 3 is adopted
trated in Figure 13. Compared with the strength before for the other tests.
excavation the strength decreased at test 1 (2 weeks
after the pit excavation). The strength decrease pre-
3.2 Test results
sumably resulting from swelling continued until test 2
and later it kept almost constant. Therefore shear Resistance through the all tests is illustrated in
strength at test 1 is used in estimating friction factor Figure 14. No settlement was observed during the
set-up periods although effective weight of the speci-
men became approx. 40 kN temporarily by change of
Penetration Raising
pit water level.
Figure 15 shows the friction observed at penetra-
(a few hours)
tion phases. The values of friction are calculated by
Test 1 subtracting tip resistance measured with soil pressure
1.3m 1.15m meters.
0.15m
3weeks
Resistance (kN)
Penetration Raising
-200 -100 0 100 200
0.0
(a few hours)
Test 1
Test 2
0.2 Test 2
1.25m 1.20m Test 3
0.05m
8
8weeks
penetrate by 10cm 0.4
Penetration depth (m)

Penetration Raising
0.6
(a few hours)
Test 3
0.8
1.3m 1.20m
0.10m
penetrate by 10cm
1.0
Figure 12. Test procedure.
1.2

Su (kN/m2)
1.4
0 20 40 60
0.00
Resistance (kN)
-200 -100 0 100 200
0.20 1.10
Test 1
Test 2
0.40 Test 3
1.15
Penetration depth (m)

Before excavation
0.60 Test 1
GL ( -m)

Test 2 1.20
Test 3
0.80
1.25
1.00
1.30
1.20
1.35
1.40
Figure 14. Resistance through all tests (above: all phases,
Figure 13. Undrained shear strength of clay. below: an enlarged view of deep penetration depth).

255

Copyright 2005 Taylor & Francis Group plc, London, UK


Friction (kN) 1.0
0 50 100 150 200
0.0 0.8

Friction factor,
Test1
0.2 Test 2 (after 3 weeks) 0.6
Test 3 (after 8 weeks)

0.4 0.4
Penetration depth (m)

0.2 laboratory test


0.6
field test
0.0
0.8 0 30 60 90 120
Time (day)
1.0
Figure 17. Recovery of friction factors with time.

1.2
Friction (kN)
1.4 -200 -150 -100 -50 0
0.8
Figure 15. Friction observed at penetration phases. Test 1
Calculated 0.9

Penetration depth (m)


= 0.4 0.35 0.3
Friction (kN) 1.0
0 50 100 150 200
0.8 1.1
Test 1
0.9 Calculated
1.2
Penetration depth (m)

1
1.3
1.1
1.4
1.2
Friction (kN)
1.3 -200 -150 -100 -50 0
= 0.3 0.35 0.4 0.8
1.4 Test 2 (after 3wks)
Test 3 (after 8wks) 0.9

Penetration depth (m)


Friction (kN)
Calculated
0 50 100 150 200 1.0
0.8
Test 2 (after 3wks)
0.9 Test 3 (after 8wks)
1.1
Penetration depth (m)

Calculated
1.0 1.2

1.1 1.3
= 0.6 0.35 0.4
1.2 1.4

1.3 Figure 18. Friction factors at raising phases.


= 0.4 0.5 0.6 0.7
1.4
friction factor at the 1st penetration (test 1) is 0.35 to
Figure 16. Friction factors at penetration phases. 0.4 and the factors after set-up are increased to 0.55
for test 2 (set-up period of 3 weeks) and 0.68 for test
Figure 16 shows the frictions at penetration phases 3 (that of 8 weeks).
together with lines of friction estimated with certain Recovery of the friction factor with time is shown
friction factors using the Equation 1. Comparison together with the laboratory test results in Figure 17.
between the estimated friction factors indicates that The results of both tests are in good agreement and

256

Copyright 2005 Taylor & Francis Group plc, London, UK


validate the recovery of the shear strength of the effect. The field test results almost correspond with
tested clay. the laboratory results and it validated the recovery
Figure 18 illustrates measured friction at raising characteristics of the tested clay.
phases with estimated friction. The friction factor at
test 1 is 0.35 to 0.4, which was almost the same as that
observed at penetration. In test 2 and 3 the friction is
decreased from 0.55 to 0.45 and from 0.68 to 0.50 REFERENCES
compared with penetration phases. The reduction is
Andersen, K.H. & Jostad, H.P. 2002. Shear strength along
assumed due to remolding at the preceding penetra- outside wall of suction anchors in clay after installation.
tion and it revealed recovered friction reduces easily Proceedings of 12th International Offshore and Polar
by re-penetration. Engineering Conference
Ito, M., Masui, N., Nagao, H. & Mitachi, T. 2004. Field per-
formance test of skirt suction foundations (part 3 recov-
4 CONCLUSION ery of wall friction. Proceedings of 59th JSCE Annual
Meeting (in Japanese)
Laboratory and field tests have been conducted to Masui, N., Ito, M., Inoue, A., Shimada, Y. & Hermstad, J.
evaluate the recovery of friction on skirts in soft clay 2004. Recovery of Wall Friction after Penetration in Skirt
Suction Foundation. Proceedings of 14th International
after remolding caused at penetration. Offshore and Polar Engineering Conference
In case of soft alluvium clay in Osaka Bay, the lab- Zen, K. Chen, G. Kasama, K. & Deguchi, S. 2003. Experi-
oratory results clearly demonstrate that thixotropy sig- mental and analytical studies on bearing capacity of suction
nificantly contributes the recovery of shear strength foundation. Proceedings of 13th International Offshore
and the dissipation of excess pore pressure has a minor and Polar Engineering Conference

257

Copyright 2005 Taylor & Francis Group plc, London, UK


Study of sand heave formation in suction caissons using Particle Image
Velocimetry (PIV)

Manh N. Tran1, Mark F. Randolph2 & David W. Airey1


Centre for Offshore Foundation Systems
1
University of Sydney, Sydney, Australia
2
University of Western Australia, Perth, Australia

ABSTRACT: One of the key aspects that has attracted interest when using suction caissons in sand is the con-
trol of sand heave during installation. Although this plays a major role in the installation success, the mecha-
nism of sand heave is not fully understood. A test apparatus, where a half-caisson model was used, was
developed aiming to study this phenomenon directly. The experiments were conducted in dense silica sand and
were carried out with different caisson wall penetration depths. Deformations of the soil skeleton were captured
in each test by a high-resolution digital camera. Using particle image velocimetry (PIV), displacement and
velocity vector fields were established, which helped identify different zones of sand movement. A theoretical
study was conducted in parallel using the finite element package PLAXIS. The results were then compared with
the PIV results, and found to be in good agreement with the test observations.

1 INTRODUCTION restrained. The tests were, however, conducted at dif-


ferent wall embedment.
Suction caissons have become increasingly popular in
the offshore industry in the last two decades, due to the
simplicity and effectiveness of the installation method. 2 PARTICLE IMAGE VELOCIMETRY (PIV)
Despite these advantages, the potential for excessive
heave of the soil plug during installation in dense Progress in camera technology and PIV technologies
sand remains a major concern. Field observations have in geotechnical engineering have helped make it pos-
shown very different degrees of plug heave, varying sible to directly study sand movements. Originally
from the negligible heave in the Draupner E platform used in fluid mechanics, PIV was later applied in the
installation (Tjelta 1995) to the infamous excessive geotechnical field to measure soil deformation to a
heave in the Gorm project (Senpere & Auvergne 1982). high level of precision and accuracy, comparable to
Sand heave was also recorded in laboratory tests, both local instrumentation (White et al. 2003). Analyses
under normal gravity conditions (Tran et al. 2004) and presented in this paper were conducted using GeoPIV
at elevated acceleration levels in a geotechnical cen- software, developed at Cambridge University Engi-
trifuge (Allersma et al. 1997). Although sand heave neering Department, UK (White & Take 2002).
appeared to be a function of several parameters in these Although offering high resolution measurements,
studies, the mechanism of formation was not fully there are a number of factors that can affect and reduce
understood. the performance in tests using PIV. These are caused
This paper therefore aims to provide better knowl- by the PIV analysis itself and with optical effects that
edge of the phenomenon by the direct study of how cause image distortion, including noncoplanarity of
sand heave occurs inside the caisson during unob- the CCD and object planes, camera lens distortion,
structed installation. In particular, a test apparatus and refraction through viewing windows (White et al.
using a half-caisson model was developed. In all tests, 2003). It is therefore important to evaluate the accu-
sand movement was captured and subsequently racy of deduced movements in each experimental set-
analysed using particle image velocimetry (PIV) to up. In theory, with the current equipment and the
investigate plug formation. The study presented here parameters used in subsequent PIV analyses, as
is limited to fixed tests, where the half-caisson was described in detail in the following section, a precision

259

Copyright 2005 Taylor & Francis Group plc, London, UK


better than 1/100000th of the field of view (FOV) D and wall length L, and had a wall thickness t of
could be easily reached. Noting that a typical FOV 1.2 mm. A continuous O-ring, 1 mm in diameter, was
was around 150 mm, this corresponds to an error of fitted along the caisson wall and top to prevent leak-
less than 1.5 m. age during pumping. A water evacuation valve, which
was left open during pushed-in penetration, was also
available. It may be noted that the model caisson has
3 TEST APPARATUS AND PROCEDURE a wall thickness-to-diameter ratio t/D of 1.2%, which
is higher than typical prototype values of 0.30.5%.
3.1 Soil characterisation However this was necessary because the 1.2 mm thick-
ness was the smallest width that allowed an O-ring to
Superfine silica sand was used in this test series. The
be fitted on. The caisson was held in place by a guide
soil is commercially available, with grain size distri-
system, consisting of 3 arms with wheels. Each arm
bution of the material being shown in Figure 1. Other
could independently move inward and outward in
soil properties are summarised in Table 1.
radial direction. Details can be seen in Figures 2a,b.
The sample was prepared by first filling the test
The caisson was connected to a hose to allow water
chamber with water. Dry sand was then carefully
to be pumped out. The flow rate was controlled by
rained down at a slow rate from a scoop. The sample
varying the degree of opening of a butterfly valve.
was densified by applying vibration to both the bot-
Flow was achieved using the water head difference
tom and the chamber sides. The density of the prepared
between the test chamber and the hose outlet, with the
sample was then measured and found to correspond
rate recorded continuously using an electronic bal-
to a relative density Dr of 90%. Saturation was also
ance. Taking the water density as 1 g/cm3, the water
checked by visual inspection of the test chamber, and
volume in cm3 was therefore equivalent to the weight
by stirring a submerged sample prepared separately
in grams. Differential pressure was measured using a
using the same technique. With no air bubbles visible,
differential pressure gauge. Since it could not be sub-
it was likely that a satisfactory saturated condition
merged, pressure was measured by connecting the gauge
was being achieved.
to 2 soft plastic tubes. With one tube embedded inside
the caisson, and the other positioned on the caisson
3.2 Half-caisson model top, the differential pressure could be measured. Tests
were conducted in a chamber of 370  220  400 mm
A half-caisson model was developed to investigate
(width  thickness  height). A transparent Perspex
the formation of sand heave during the application of
window was fitted in the front face to allow observa-
pumping. The caisson was 100 mm in both diameter
tions to be made. The sample depth was 200 mm and
water depth was 170 mm.
100% During each test, sand movement was captured
by close-range photography using a 4-megapixel
80% (2272  1704) digital still camera. It was set up at
0.30.5 m from the test chamber Perspex window.
% passing

60% Consecutive images were recorded at the rate of around

40%
Differential pres- 3 arms (can move independently
20% sure tube outlet in radial direction)
Pump outlet
0%
0.01 0.10 1.00
Particle size (mm)

Evacuation valve
Figure 1. Grain size distribution curve for superfine silica
outlet
sand.
Continuous
L O-ring t
Table 1. Properties of superfine silica sand.

Specific gravity Gs 2.67 D


D50 0.18 mm
Min. saturated density s min 1.94  103 kg/m3 (a) (b)
Max. saturated density s max 2.10  103 kg/m3
Permeability k 0.953.6  10
4 m/s Figure 2. (a) Half-caisson (b) Test chamber and guide
system.

260

Copyright 2005 Taylor & Francis Group plc, London, UK


2 frames/second. Details of the experimental set-up PIV analyses were hence conducted on the results
are illustrated in Figure 3. taken immediately after pumping, using a patch size
of 40  40 pixels. In real object-space, this was roughly
equivalent to 2.5  2.5 mm. The centre-to-centre dis-
3.3 Test procedure
tance of the patches varied from 2040 pixels depend-
All tests followed the procedure given below: ing on the detail required.
(a) The sand surface was levelled, and the sample
densified by vibration.
4 RESULTS AND DISCUSSION
(b) The half-caisson model was slowly pushed to
penetrate to the desired embedment depth, with
4.1 Sand displacement
the evacuation valve left open.
(c) The valve was then closed. Guide arms were Figures 46 present displacement fields from tests
adjusted to press the caisson against the Perspex with different penetration depths. To assist with the
window, hence locking it in place. Care was taken interpretation the displacements have been scaled up
not to apply excessive force as it could result in as indicated on the figures. The sand displacement
caisson deformation. field for shallow penetration depth L/D of 0.1 (i.e.
(d) The camera and light were then set up to capture L  10 mm) is shown in Figure 4a. It can be seen
the zone of interest. The camera was then started, from the results that there are two distinct areas of
followed by the application of pumping by open-
ing the butterfly valve.
It was observed that immediately after pumping, the
suction quickly increased to a peak, and sand move-
ment was recorded, but with no sign of piping channel
formation. However, if pumping was maintained, pip-
ing channel was seen to quickly form, (which was
also evident from a rapid drop in suction) and lead to
localised piping failure. Assuming that the continu-
ous caisson installation is a series of discrete move-
ments (self-balance between the driving force and soil
resistance), the sand heave occurs at any wall embed-
ment depth, when the continuous process is broken
into snapshots, is hence instantaneous in that snap-
shot, and best represented in this fixed test by the
results taken immediately after starting pumping, i.e.
when first sand movements were recorded. The later
results, where pumping was continued until piping,
represent the situation where the caisson movement
was obstructed by, for example, boulders, and is out
of the scope of this paper.

Differential pressure tubes

170 m
Water
200 m head
difference

Butterfly valve to apply


and control pumping
Qpump
measured by an electronic balance
Figure 4. (a) Displacement field for L/D  0.1 (b) At right
Figure 3. Experimental set-up. hand sides tip.

261

Copyright 2005 Taylor & Francis Group plc, London, UK


caisson wall, will flow into the caisson during suction
installation. Observations here suggest that this is
unlikely to be the case. Rather they suggest that the
majority of the sand heave in the wedge formed along
the inner caisson wall is the result of its own expan-
sion in volume, not an extra supply of sand due to
inflow. Although unexpected, the results are, however,
explainable. Under suction, sand adjacent to the inner
caisson wall will experience the most extreme seep-
age flow conditions due to the location of the shortest
hydraulic path along the caisson wall. This in turn sig-
nificantly loosens sand in the region, causing it to
expand in volume. For dense sand, the expansion can
be quite substantial. Furthermore, as sand moves
upward, the material immediately adjacent to the
inner caisson wall undergoes shearing. This action
Figure 5. Displacement field for L/D  0.2.
will lead to dilation of the dense sand, causing addi-
tional loosening and expansion of the sand inside the
caisson in the vicinity of the walls.
Figures 5 and 6 show test results for deeper wall
embedment depths of L/D of 0.2 and 0.3 respectively.
Unlike previously, where no movement was observed
for the central plug, considerable upward movement of
the whole sand plug occurred at these deeper penetra-
tion levels. There are large displacements within wedge
zones along the inner wall in both cases. It is noted that
the displacement zone only extends to a relatively small
distance below the caisson wall embedment. Apart
from zones next to the caisson wall, the displacement
level appears to be quite uniform for the rest of the plug
inside the caisson, but reduces quickly with further
depth below the tip embedment level. Only modest sand
displacement is recorded in the outer caisson wall areas
and below the caisson tip, showing a similar trend to the
Figure 6. Displacement field for L/D  0.3. previous observations at shallow penetration.
The results are interesting considering the high
suction pressure recorded for each case. The nor-
sand movement: a large displacement zone in the two malised pressures p/
L, where p is the measured
regions (a single semi-toroidal region) close to the peak pressure, 
is the soil submerged unit weight,
inner caisson wall, and the remaining field with taken as 10 kN/m3, and L is the wall penetration
essentially no movement. Largest displacements are depth, are 7, 20 and 16.7 for L/D  0.1, 0.2 and 0.3
observed along the wall, but these reduce quite rap- respectively. The results clearly indicate that, at least
idly towards the plug centre, where no sand move- for the soil type and test conditions here, no excessive
ment was recorded. This field of movement forms a sand inflow occurs even under suction pressures that
wedge shape adjacent to the inner wall. are much greater than needed to create piping condi-
More details of the displacements around the cais- tions within the soil plug. Rather, the sand heave
sons right hand side, shown in Fig. 4a, are shown in formed during the recording period, either in the form
Figure 4b. While there is a clear trend that the sand of the small wedges seen in the shallow penetration
tends to flow around the tip into the caisson compart- case, or the whole plug motion at greater caisson
ment, the level of inflow movement is very small embedment depths, is primarily due to volumetric
compared to that along the inner wall. The zone of expansion of the sand. Only a very small amount of
influence, where movement is still recorded, extends sand inflow contributes to the overall sand heave.
to only a small depth below the caisson tip. At the
same time, almost no movement is observed in
4.2 Sand movement velocity
regions adjacent to the outside caisson wall.
The results are interesting, as it is expected that a The sand velocity distribution is discussed in this sec-
large amount of sand, especially from along the outer tion. Figure 7a shows the sand velocity for the test

262

Copyright 2005 Taylor & Francis Group plc, London, UK


mm 15 mm 30
10
5 0.20
0
0 20 40 60 80 100 120 20 0.16
mm
(a) L/D = 0.1, Qseepage / Ahalf-caisson = 0.44 mm/s.
0.12
mm
10
0.08
40
0
0 10 20 30 40 50 60
mm
20
(a) L/D = 0.1, p/L = 7.

0 mm
0 20 40 60 80 100 120 40
mm 1.0
(b) L/D = 0.2, Qseepage / Ahalf-caisson = 1.80 mm/s. 0.8

mm

60 20 0.6

40
0.4

20 0
0 20 40 60
mm
0 (b) L/D = 0.2, p/L = 20.
0 20 40 60 80 100 120 140
mm
mm 60
(c) L/D = 0.3, Qseepage / Ahalf-caisson = 1.78 mm/s.

Figure 7. Sand velocity (mm/s).


1.0
40 0.8
with L/D  0.1. High velocities were measured in
regions immediately next to the caisson wall, with the
highest observed near the inner wall tip level. Results
0.6
for deeper penetration depths of L/D of 0.2 and 0.3, as 20
seen in Figures 7b and c respectively, show a very sim-
ilar trend, with highest velocities found near the tip of 0.4
the caisson wall. The formation angle of the high veloc- 0.2
ity zone (the wedge zone) appears to be very similar
0
despite different wall embedment level. However, this 0 20 40 60
trend is only evident for L/D ratios of up to 0.3 as mm
reported in this study, and may require further verifi- (c) L/D = 0.3,p/L = 16.7.
cation before any general conclusion can be made.
It should be noted that the sand movement seen Figure 8. Seepage velocity (mm/s) by PLAXIS.
here is created by seepage. Indeed, when seepage flows
through sand, it tends to drag the sand grains with it
(through frictional force), eventually mobilises the The seepage velocity distribution on the right hand
grains once the drag force is large enough. Therefore, half of each wall embedment case is shown in Figures
study of seepage flow may help explain the recorded 8a,b and c. The numerical simulations clearly show the
sand movement pattern. In this paper, the theoretical close relationship between the theoretical velocity of
seepage was obtained from numerical simulations the seepage flow and the measured velocity of the
using the finite element package PLAXIS, with input sand in the caisson. It can be seen from these PLAXIS
parameters such as test geometry and suction pres- results that seepage velocities around the caisson tip
sures based on those measured in the experiment. A are very high, with extreme flow conditions near tip
uniform value of permeability ko  1.0 10
4 m/s was level. Considering the direct influence of seepage
used (which is typical for the soil density obtained). velocity on the sand stability as mentioned before,

263

Copyright 2005 Taylor & Francis Group plc, London, UK


mm effective stress on the soil skeleton, hence restricting
k = 2ko its movement.
30 0.16 It is also noted that the theoretical seepage flow
0.20
distribution for the inside region, as shown in Figure 8,
0.12
20 compares well with the recorded sand velocity pat-
tern. However, the values of seepage velocity from
the PLAXIS results are smaller than the sand veloci-
10
ties in the high velocity zones (i.e. in the wedge
0.08
regions). This is unexpected as seepage flow velocity
0 is expected to be higher or equal to the sand velocities
0 10 20 30 40 50 60 because the soil is transported by seepage. The results
mm
(a) L/D = 0.1, p/L = 7. may be due to adopting a uniform value of perme-
ability in PLAXIS. In reality, sand expansion will
mm increase the soil permeability in these high flow
40 velocity zones, in turn leading to faster seepage flow.
k = 2ko Indeed, PLAXIS simulations as seen in Figures 9a,b
0.8
and c, where wedges with twice the initial permeabil-
0.6
ity was used (i.e. kwedge  2ko  2.0 10
4 m/s)
20 clearly show much higher seepage velocity in these
zones. The flow distribution in these results also
agrees better with the measured sand velocity pattern.
0.4 The philosophy of this fixed test is to explore what
0
happens in the most critical situations, and determine
0 20 40 60 if the results can be safely interpreted for normal mov-
mm ing cases. It is believed that the fixed caisson experi-
(b) L/D = 0.2, p/L = 20. ment represents the worst installation condition, where
a very high suction pressure is suddenly applied while
mm no movement, hence extended cut-off length, is per-
mitted. These in turns create a very significant prefer-
k = 2ko ential flow region (as seen clearly in the wedge zones),
40 and best condition for triggering a piping failure. The
0.8
1.0 expectation is that the sand outside is most likely to
0.6 move into the caisson. In other words, seepage flow has
20 its best chance to pull the sand inside. However the
results obtained clearly show that this is not the case,
0.4 0.2 and that the heave is mostly caused by plug expan-
0 sion. This result, to some extent, can be used to apply
0 20 40 60 to the moving cases. Indeed, considering effective
mm
(c) L/D = 0.3, p/L = 16.7. heave (i.e. minus the caisson wall volume) as the result
of sand inflow and expansion, if the heave in these
Figure 9. Seepage velocity (mm/s) by PLAXIS with extreme cases is mainly caused by plug expansion,
higher wedge permeability. then in the moving case where there is more restric-
tion to sand inflow due to the continuous extension of
the cut-off wall length, and much lower suction pres-
sures, it is most likely that sand heave will be created
this leads to the impression that under the applied in the same manner.
suction, significant sand mobilisation will occur not
only in the region immediately inside the wall, but
also in the region outside the caisson wall, possibly by 5 CONCLUSIONS
flowing into the caisson.
While these results are consistent with the meas- The experimental program presented here is the first-
ured sand velocities inside the caisson, as seen in ever direct study of sand heave formation within the
Figure 7, it is clearly not the case for the outside zone suction caisson compartment. It demonstrated that
where no significant movement was recorded. This non-intrusive measurement of sand movement could
may be because, unlike the inside zone where upward be achieved using PIV. The tests showed different sand
flow lifts the soil up and creates volume expansion movement patterns for shallow and deep caisson wall
and boiling, the outside downward flow increases the embedment. They also indicated that the majority of

264

Copyright 2005 Taylor & Francis Group plc, London, UK


the sand heave observed was caused by dilation of the Research Scholarship from the University of Sydney,
sand plug, possibly enhanced by sand dilation adja- Australia.
cent to the caisson wall, rather than sand inflow.
Theoretical modelling of seepage velocities com-
pared quite favourably with the measured sand move- REFERENCES
ment. Although only conducted using small-scale
model, the experiments have enabled observation of Allersma, H.G.B., Plenevaux, F.J.A. and Wintgens, J.F.P.C.
the mechanisms underlying plug formation in instal- 1997. Simulation of suction pile installation in sand in a
lation in sand. The results presented here are for the centrifuge. Proc. 7th International Offshore and Polar
Engineering Conference, Honolulu, USA: 761766.
restrained caisson, but further tests are being planned, Senpere, D. and Auvergne, G.A. 1982. Suction anchor piles
in which the caisson is free to move. A proven alternative to driving or drilling. Offshore
Technology Conference, Houston, USA. Paper number:
OTC 4206.
Tjelta, T.I. 1995. Geotechnical experience from the installa-
ACKNOWLEDGEMENTS tion of the Europipe jacket with bucket foundations.
Offshore Technology Conference, Houston, USA. Paper
The authors would like to express special thanks to number: OTC 7795.
Dr David White for his permission and help in using Tran, M.N., Randolph, M.F. and Airey, D.W. 2004.
the GeoPIV software. Discussions and advice from Experimental study of suction installation of caissons in
Dr Conleth OLoughlin are acknowledged. Also, thanks dense sand. Proc. 23rd International Conference on
to Mr John Bagrie, who has helped in developing the Offshore Mechanics and Artic Engineering, Vancouver,
testing equipment. The work presented in this paper is Canada. Paper number: OMAE04-51076.
White, D.J. and Take, W.A. 2002. GeoPIV: Particle Image
part of the research program conducted at the Centre Velocimetry (PIV) software for use in geotechnical test-
for Offshore Foundation Systems (COFS) within the ing. Software manual.
University of Western Australia and the University of White, D.J., Take, W.A. and Bolton, M.D. 2003. Soil deforma-
Sydney, funded by the Australian Research Council. tion measurement using particle image velocimetry (PIV)
The first author holds an International Postgraduate and photogrammetry. Gotechnique 53(7): 619631.

265

Copyright 2005 Taylor & Francis Group plc, London, UK


Electrokinetic and electrochemical stabilization of caissons
in calcareous sand

A. Rittirong & J.Q. Shang


Department of Civil and Environmental Engineering, University of Western Ontario, London, Canada

M.A. Ismail & M.F. Randolph


Centre for Offshore Foundation Systems, University of Western Australia, Crawley, WA, Australia

ABSTRACT: A laboratory experimental study was conducted on electrokinetic and electrochemical stabi-
lization of model caissons embedded in offshore calcareous sand. Four tests were carried out to study effects of
the electrode configuration, polarity reversal and chemical stabilization agent on the pullout resistance of the
model caisson. The results after the electrokinetic and electrochemical treatments indicated that the pullout
resistance of the model caisson increased by up to 120% as compared to the control model. Polarity reversal and
application of CaCl2 as a stabilization agent further enhanced the effectiveness of the electrical treatment. The
improvement of the soil properties was further assessed by X-ray diffraction analysis (XRD), X-ray fluorescence
analysis (XRF) and electron microscopic imaging.

1 INTRODUCTION has been well documented (e.g. Bjerrum et al. 1967,


Wrixon & Cooper 1998, Lo et al. 2000). Micic et al.
Calcareous sediments, typically containing more than (2003) reported a significant improvement in load car-
50% calcium carbonate, cover approximately 50% of rying capacity of a skirt foundation model treated by
the seabed in the world (Sverdrup et al. 1942). Many electrokinetics. Recently the stabilization of model
offshore petroleum platforms are erected on calcare- caissons by electrokinetics in calcareous sand has been
ous sediments with grain sizes ranging from sand to investigated (Mohamedelhassan et al. 2005). The elec-
clay. The characteristics of calcareous sand have been trodes were installed surrounding the caisson to
studied extensively after problems encountered with improve shaft friction. Calcium chloride in the elec-
pile foundations at North Rankin offshore platform trodes associated with electrokinetic reactions gener-
(King & Lodge 1988). The installation of foundations, ated cementation.
especially driven piles and caissons, crushes soil par- For stabilization of offshore foundations, the instal-
ticles and destroys cementation bonds of the calcare- lation of electrodes surrounding the caisson can be
ous soil, leading to low bearing and pullout capacity. difficult and costly for deep-water operations. To
Furthermore, since calcareous soils are often highly improve the applicability of the technique for off-
compressible due to the high void ratio (Coop & shore foundations, the current research is focusing on
Atkinson 1993), the skin friction on foundations such a single central electrode installed inside the caisson
as piles becomes very low as reported by Male et al. model. In this configuration, the caisson and electrode
(1988). Therefore, to enhance the bearing capacity of can be installed simultaneously in the field application.
foundations embedded in calcareous soils, it is of cru- A series of four tests were carried out by applying a
cial importance to enhance the interface bonding dc voltage to the embedded caisson models with and
between the foundation and the soil. without chemical enhancement, referred to as electro-
Electrokinetic stabilization is a ground improvement kinetic treatment and electrochemical treatment,
technique used mostly for fine-grained soils, such as respectively. The main objectives of this study are
silts and clays. Casagrande (1941) first applied the (1) to study effects of electrical polarity and electrode
technique of electrokinetics to stabilize soft silty clays. configuration on the efficiency of treatment; and
Since then this technique has been applied to many (2) to study the relation between the electric field and
projects including pile foundations (Milligan 1995). pullout resistance enhancement generated by cemen-
Strengthening soft marine clay by using electrokinetics tation of soil solids.

267

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 1. Summary of properties of the calcareous sand and Table 2. Experimental program.
seawater.
Polarity
Calcareous sand Value Seawater Value Test No. Polarity scheme reversal CaCl2 (g)

Sand size, % 99.9 Na, mg/l 10965 Test 1 Untreated caisson (control)
Fine grain size, % 1 K, mg/l 409 Test 2a Initially cathodic caisson Hour 48 0
D10, mm 0.17 Ca2, mg/l 469 Test 2b Initially cathodic caisson Hour 26
D30, mm 0.22 Mg2, mg/l 1433 Test 3a Anodic caisson No 0
D60, mm 0.29 Fe2-Sol, mg/l 1 Test 3b Anodic caisson No
Coefficient of 1.65 Cl
, mg/l 20015 Test 4a Initially cathodic caisson Hour 48 447
uniformity, Cu S2
, mg/l 938 Test 4b Initially cathodic caisson Hour 26
Coefficient of 0.94 HCO3
, mgCaCO3/l 125
curvature, Cc CO2

3 , mgCaO3/l 5
Specific gravity, Gs 2.76 SO42
2809
Carbonate, % 80.6 OH
, mgCaCO3/l 5
Quartz, % 16.4 pH 8.0

2 CALCAREOUS SOIL AND SEAWATER

The sand and seawater were recovered from the coast


of Western Australia. The properties of the sand and
seawater are summarized in Table 1. The sand is poorly
graded and, based on the X-ray diffraction analysis,
consists of 80.6% carbonates and 16.4% quartz. For
the seawater, the main cation and anion are sodium,
Na, and chloride, Cl
, respectively. The seawater has
a pH of 8.0 and electrical conductivity of 49.5 mS/cm.
The soil sample saturated with the seawater has elec-
trical conductivity of 15.14 mS/cm.

3 EXPERIMENTAL PROGRAM

The experimental program comprising four tests is


summarized in Table 2. Test 1 was a control test in
which the caisson model was embedded in the satu-
rated calcareous sand without any treatment, thereby
providing a reference value of the pullout capacity. The
other three tests with active treatment were conducted
in two phases: Phase A (pre-failure treatment) involved
electrical treatment of the model caisson following Figure 1. Caisson model.
installation and Phase B (post-failure treatment) inves-
tigated the effects of electrical treatment after the capacity (Butterfield & Johnston 1980). On the other
caisson had been subjected to pullout failure. In hand, it is well-known that the steel caisson would suf-
Phase B test, the caisson was pushed back to the orig- fer significant corrosion due to electrochemical reac-
inal position before undergoing electrical treatment tions during the treatment. Secondly, the configuration
again. Each phase of the test lasted 7 days. of using the caisson as a cathode was not investigated
In particular, Test 2 was the electrokinetic treatment. because previous studies clearly demonstrated that
Initially the caisson was a cathode (negative polarity) this configuration substantially decreased the pile
and the electrode was an anode (positive polarity). capacity (Butterfield & Johnston 1980). The electrode
The polarity was reversed after 48 hours and 26 hours configuration of Test 4 was identical to that in Test 2,
of treatment for Phase A and Phase B, respectively. In with the exception of using calcium chloride as a
Test 3, the model caisson was set as an anode and cementing agent. Granular calcium chloride of 0.447 kg
polarity of the current remained the same during the was filled into the 50 mm diameter central electrode
treatment. This test was designed based on two con- before installation.
siderations. Firstly, it was reported that anodic piles in The caisson model was installed in a steel drum
soft clay had significant increase in load carrying coated by a plastic sheet as shown in Figure 1. A gravel

268

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 3. Summary of test results. 40

Negative Test 2 Phase A


Positive
Dry unit Loss in Pullout 30 Test 3 Phase A
caisson caisson
Test weight Energy mass of resistance Soil Test 4 Phase A
No. kN/m3 W-h caisson % N plug 20

Measured current, A
Polarity reversal
Test 1 12.96 0.0 500 Yes 10
Test 2a 12.84 15029 957 Yes
Test 2b 13523 6.2 1022 Yes 0
Test 3a 12.88 9820 925 Yes
Test 3b 7515 12.3 629 Yes -10
Test 4a 12.95 10510 1113 Yes
Test 4b 9963 1.5 1126 Yes
-20

-30
0 20 40 60 80 100 120 140 160 180
drainage layer was placed on the bottom of the drum Treatment time, hours
for drainage control through a valve. The drum was (a) Pre-failure treatment
then placed on the rotating disk of the sand hopper
allowing dry sand to fall freely. The falling height was 40
set at 0.65 m. The unit weights of the calcareous sand Negative Positive Test 2 Phase B
samples used in the tests are summarized in Table 3. 30 caisson caisson
Test 3 Phase B
Test 4 Phase B
The unit weights were consistent and were of typical
density for loose sand. After depositing the first 20 Polarity reversal
Measured current, A

130 mm of the calcareous sand in the drum, the elec-


trode was placed at the center of the drum, as shown 10

in Figure 1, and the sand was continuously deposited


0
up to 280 mm.
The model caisson was then placed in the drum
-10
and the deposition of sand continued until the model
caisson was embedded 75 mm under the sand. The -20
loading plate and a dead-weight surcharge equivalent
to 4.5 kPa were then placed on the top of the sand. -30
Seawater was introduced from the top with the bot- 0 20 40 60 80 100 120 140 160 180
tom drainage valve open until no further air bubbles Treatment time, hour
were observed emerging from the bottom drainage (a) Post-failure treatment
channel. At this point, the valve was closed and the
sand was considered to be nearly, if not fully satu- Figure 2. Measured currents.
rated. A constant voltage of 13 V was applied across
the caisson and the electrode in all treatment tests.
The treatment for each phase lasted 7 days with cur-
of starting the treatment and the minimum current of
rent intermittence intervals of 2 min on and 2 min off,
2 A was reached after 48 hours of the treatment, at
controlled by a programmable timer. After the com-
which time the polarity was reversed. After the polarity
pletion of each phase, a pullout test was performed at
reversal, the current suddenly increased and reached
a constant displacement rate of 0.5 mm/min. During
23 A after 6 hours of the reversal and then dropped to
the pullout test, the water level of seawater was kept at
10 A before the test was terminated. The pullout test
30 mm above the surface of the calcareous sand.
results are shown in Table 3 and Figure 3. The pullout
resistances of the untreated caisson and the treated
caisson were 500 N and 957 N, respectively, repre-
4 RESULTS AND DISCUSSION
senting an increase of 90%. In Test 2 Phase B (post-
failure treatment), the polarity was reversed after 26
4.1 Electrokinetic treatment
hours of the treatment and the current was less than
The effects of the electrokinetic treatment were inves- that in Phase A. The pullout resistance after Phase B
tigated in Test 2. The caisson was initially set as a treatment was further increased to 1022 N, slightly
cathode (negative polarity). In Test 2 Phase A (pre- higher than that in Phase A. This indicates that the
failure treatment), the measured current with time pullout resistance of the caisson was fully recovered
ranged from 30 A to 2 A as shown in Figure 2. The after the electrokinetic treatment, even after the caisson
maximum current of 30 A was reached after 2 hours had been subjected to pullout failure. It was further

269

Copyright 2005 Taylor & Francis Group plc, London, UK


1400
Test 1 Control
Test 2 Phase-A
1200
Test 2 Phase-B
Test 3 Phase-A
1000 Test 3 Phase-B
Pullout load, N

800

600

400

200

0
0 5 10 15 20 25 30 35 40
Pullout distance, mm

Figure 3. Pullout resistance influenced by polarity reversal. Figure 4. Pullout resistance influenced by chemical
enhancement.
noticed from Figure 3 that there was no abrupt post-
peak drop in the pullout resistance. The corrosion of the failure of the caisson tip due to corrosion. Further, a
model caisson was measured by weighing the caisson soil pH value of 3.41 was measured at the caisson tip.
before and after treatment. The loss of mass by corro- After the specimen was removed from the drum, weak
sion was 6.2% by weight, which was most noticeable cementation was found in the soil surrounding the cais-
at the bottom of the caisson. son bottom. From comparison of the results of Tests 2
and 3, it is evident that the polarity reversal is essential
to control corrosion and acidity of soil surrounding
4.2 Electrical polarity the caisson and to achieve better effectiveness.
In order to study the effects of current polarity, the
model caisson was set as an anode for the entire 4.3 Electrochemical treatment
period of treatment in Test 3. The record of current is The effects of electrochemical treatment with calcium
shown in Figure 2. The measured current ranged from chloride (CaCl2) were investigated in Test 4. The elec-
7 A to 21 A. The maximum current of 21 A was trode configuration and polarity reversal were identical
reached after 3 hours of Phase A treatment. The cur- to those in Test 2. The perforated central electrode was
rent gradually decreased to 7 A before Phase A was filled with calcium chloride granules before installa-
terminated. The current in Phase B was initially less tion. When a voltage was applied on the electrode, the
than that in Phase A and then dropped to the same solution of calcium chloride was delivered into soil
value of 7 A as that of Phase A within 24 hours of pores by the electrical current. The measured current in
treatment. The pullout resistances versus pullout dis- Test 4 was noticeably lower than in Test 2 (no chemical)
placement are plotted in Figure 3. In Test 3 Phase A, as shown in Figure 2, apart from at the start where the
the initial slope of the pullout load versus displace- maximum current was 30 A, the same as in Test 2
ment curve was slightly flatter than that of Test 2 Phase A. The results of pullout tests are shown in
Phase A (See Figure 3). The displacement at the ulti- Figure 4. The maximum pullout resistance was 1113 N
mate pullout resistance in Test 3 Phase A was larger after Phase A treatment, which was 220% of the pull-
than that of Test 2 Phase A. The residual resistance of out resistance of Test 1 (control) and 16% higher than
both tests tended to be approximately the same, about that of Test 2 Phase A. The energy consumptions in
600 N. In Test 3 Phase A, a sudden drop of the pullout Test 4 were lower by 35% and 45% for Phase A and
load was observed, which resulted from tensile failure Phase B, respectively, compared to those in Test 2. The
of the caisson bottom due to corrosion. The loss of caisson corrosion after the same treatment time was
mass in the model caisson due to corrosion was 12.3%, 1.5%, as compared to 6.2% for Test 2. The reduced
again most noticeable at the tip of the caisson. This corrosion was consistent with the reduced power con-
might have caused the loss of tensile strength of the sumption for the treatment.
steel caisson. The ultimate pullout capacity of 925 N
was slightly less than that of Test 2 Phase A.
4.4 Electric field configuration
Test 3 Phase B (see Figure 3) showed much lower
pullout resistance, compared to that of Test 3 Phase A The effects of electric field intensity, E (V/m), was
and that of Test 2 Phase B. This was partly a result of investigated by means of finite element analysis. The

270

Copyright 2005 Taylor & Francis Group plc, London, UK


C
L Table 4. Results of qualitative XRD and XRF analyses.
Strength
E (V/m)
Strong Chemical Original Test 2 Test 4
cementation Mineral formula sand wt.% wt.% wt. %
25.0
Albite NaAlSi3O8 2.8
Ankerite Ca(Fe, Mg) 0.5
22.5
(CO3)2
Aragonite CaCO3 21.2 19.1 15.4
20.0 Caisson Drum Calcite CaCO3 59.4 59.8 56.7
FeO(OH) FeO(OH) 2.0
Gypsum CaSO42H2O 2.5 0.3
17.5 Hematite Fe2O3 2.7 2.8
Quartz SiO2 16.4 6.3 10.0
15.0 Siderite FeCO3 9.4 12.0
Weak Total 99.8 99.8 99.7
E=

cementation Subtotal of new minerals 14.6 17.1


12.5
12.5
V/m

10.0

7.5
Electrode
5.0

2.5

25.0 V/m
0.0

Figure 5. Electric field configuration based on finite ele- (a)


ment analysis.

software QuickField was used to simulate the two-


dimensional distribution of the electric field. A volt-
age of 13 V was imposed between the electrode and
the caisson. The measured electrical resistivity of the
saturated calcareous sand is 0.67 m. The axisym-
metric electric field distribution is shown in Figure 5.
The analysis of electric fields was compared with the
location of the cementation. It was found that in Test 2
and Test 4, weak cementation was formed correspon-
ding to E  12.5 V/m. In addition, highly cemented
soil solids were found where the electric field intensity
was greater than 25 V/m. (b)
In Test 3, because of acidity of the soil, lightly
cemented soil solids were found at the caisson surface. Figure 6. Electro-microscopic images; (a) Untreated cal-
careous sand (Mohamedelhassan et al. 2004); (b) Treated
sand after electrokinetic treatment (Test 2).
4.5 Change in soil mineralogy
Change in the mineralogy of the calcareous sand after in Table 4. The original sand primarily consists of
treatment was examined by X-ray diffraction analysis carbonates (aragonite and calcite) and quartz. After
(XRD), and X-ray fluorescence analysis (XRF). Two the electrokinetic treatment (Test 2) and electrochem-
soil samples were taken from the cemented soil ical treatment (Test 4), new minerals found are com-
attached to the caisson. The results are summarized posed of gypsum (CaSO4 2H2O), hematite (Fe2O3),

271

Copyright 2005 Taylor & Francis Group plc, London, UK


and siderite (FeCO3). Small amount of FeO(OH) was results that illustrate the potential of electrokinetic
found after Test 4 electrochemical treatment. The and electrochemical stabilization for caissons embed-
amount of new minerals ranged from 14.6% to 17.1%, ded in weak calcareous soils.
with siderite being the main component. This indicates
that the increase in the pullout resistance and cemen-
tation after treatment are, at least partially, a result of
the newly generated minerals. REFERENCES
An inspection of the compositions newly identi-
fied minerals showed that almost all consisted of iron Bjerrum, L., Moum, J. & Eide, O. 1967. Application of
(Fe). This indicates that corrosion of the caisson model electro-osmosis to a foundation problem in a Norwegian
contributed to the enhancement of pullout resistance Quick Clay. Geotechnique 17: 214235.
Butterfield, R. & Johnston, I.W. 1980. The influence of
and cementation. To investigate the microscopic soil electro-osmosis on metallic piles in clay. Geotechnique
structure, electro-microscopic imaging was performed, 30(1): 1738.
as shown in Figure 6. Figure 6(a) shows particles of Casagrande, L. 1941. Zur Frage der Entwsserung
the original calcareous sand. After electrokinetic treat- feinkrniger Bden (on the problem of drainage of fine
ment, the formation of amorphous compounds as soils). Deutsche Wasserwirtschaft No.11.
cementation bonds is clearly seen in Figure 6(b). The Coop, M.R. & Atkinson, J.H. 1993. The mechanics of
sand was virtually cemented by the treatment. The cemented carbonate sands. Geotechnique 43(1): 5367.
amorphous compounds are found as the cementing King, R. & Lodge, M. 1988. North-West shelf development
agents to bond sand particles. Furthermore, electroki- the foundation engineering challenge. In R.J. Jewell &
M.S. Khorshid (eds), Engineering for calcareous sedi-
netic treatment dissolved some particle surfaces and ments, Vol. 2: 333342. Rotterdam: Balkema.
then formed tight cementation bonds. Lo, K.Y., Shang, J.Q. & Micic, S. 2000. Electrokinetic
strengthening of soft marine clays. International Journal
of Offshore and Polar Engineers 10(2): 137144.
5 CONCLUSIONS Male, R., Haggerty, B.C. & Khorshid, M.S. 1988.
Engineering management of the foundation modification
A laboratory experimental study was conducted on programme. Engineering for Calcareous Sediments:
the electrokinetic and electrochemical stabilization of Proceedings of the International Conference on Calcare-
model caissons embedded in a calcareous sand sub- ous Sediments, Vol.2: 807836. Rotterdam, Balkema.
merged under seawater. Electrokinetic treatment with Micic, S., Shang, J.Q. & Lo, K.Y. 2003. Improvement of
polarity reversal increased the pullout resistance by load-carrying capacity of offshore foundations by elec-
up to 90% as compared to that of the control test. In trokinetics. Canadian Geotechnical Journal 40: 949963.
addition, it was found that the polarity reversal is impor- Milligan, V. 1995. First application of electro-osmosis to
improve friction pile capacity-three decades later.
tant in the treatment, in order to minimize the amount Proceedings of the Institution of Civil Engineers,
of corrosion at the tip of the caisson. After the elec- Geotechnical Engineering 113(2): 112116.
trochemical treatment, the pull resistance of the Mohamedelhassan, E., Shang, J.Q., Ismail, M.A. &
model caisson was further increased by up to 120% as Randolph, M.F. 2005. Electrochemical cementation of
compared to the control test with a significant reduc- calcareous sand for offshore foundations. International
tion of energy consumption and corrosion as that of Journal of Offshore and Polar Engineering, (scheduled
the electrokinetic treatment test. The analysis on the for March 2005 issue).
electrical field distribution reveals that cementation Sverdrup, H.U., Johnson, M.W. & Fleming, R.H. 1942. The
was generated where the magnitude of electric field oceans, their physics, chemistry and general biology.
Englewood Cliffs, N.J, Prentice-Hall.
was greater than 12.5 V/m, whereas strong cementa- Wrixon. R.C. & Cooper, G.A. 1998. Theoretical and practical
tion took place where the electrical field intensity was application guidelines for using electrokinetics to improve
greater than 25 V/m. Most of new minerals contain casing support in soft marine sediments. Proceedings,
iron resulting from corrosion of the steel caisson dur- IASC/SPE Asia Pacific Drilling Technology Conference:
ing the treatment. This study provides encouraging 7783. APDT.

272

Copyright 2005 Taylor & Francis Group plc, London, UK


Vertical uplift capacity of suction caisson in clay

L. Thorel, J. Garnier, G. Rault & A. Bisson


Laboratoire Central des Ponts et Chausses, Centre de Nantes, Route de Bouaye Bouguenais Cedex, France

ABSTRACT: The main theoretical methods used for designing suction caissons under vertical static loads are
presented and discussed in this paper. Depending on the suction pressure developed under the head of the cais-
son and on the drainage conditions, three different failure mechanisms are usually taken into account. The lower
limit of the bearing capacity is given by a sliding failure mechanism where no passive suction is generated.
A reverse end bearing mechanism, which could happen when the caisson is pulled out at a fast rate, with mobiliza-
tion of large passive suction, would lead to a larger uplift capacity. Series of centrifuge tests have been carried
out with special devices allowing to install the caissons in-flight, to load them and to perform several CPT tests
without stopping the centrifuge. Results of the centrifuge model tests are compared to the predictions of the
main theoretical methods and the effect of the aspect ratio (slenderness) is discussed.

1 INTRODUCTION The comparison focuses on the aspect ratio effect


(slenderness of the suction caisson).
The floating production offshore platforms require to
improve and to innovate in new deep water anchoring
systems: suction pile is a technology frequently used 2 ANALYTICAL MODELS
(e.g. Colliat 1999, Iskander et al. 2002). Those anchors
may be submitted to complex loading including inclined Several analytical models, based on force balance prin-
and cyclic loads. Depending on the platform type (TLP ciples, have been developed in the recent years. They
or SPAR), the load mooring point may be either at the assume simplification of the problem, but also that the
top of the caisson or at different depths on the side wall. failure mode is a sliding failure, tensile failure or a
The ultimate vertical pullout capacity of suction reverse bearing capacity failure.
caisson has been investigated by several authors (e.g. Figure 1 shows a schematic suction caisson embed-
Rasmussen et al. 1991, Christensen et al. 1991, Renzi ded at a depth D in the marine soil under a depth of
et al. 1991, Steensen-Bach 1992, Cluckey & Morrison water h.
1993, Dyvik et al. 1993, Morrison et al. 1994, Clukey In order to estimate the drainage condition, Deng &
et al. 1995, Deng. & Carter 2000, Rahman et al. 2001, Carter (2000) have introduced the adimensionnal
Allersma et al. 2003, Clukey et al. 2004). parameter Tk, which links soil properties, load rate and
Depending on the loads velocity and on the drainage a conventional length (i.e. diameter, as the drainage
conditions, three different failure mechanisms have paths follow the higher pressure gradients, preferably
been identified: sliding failure, tensile failure, reverse horizontally than vertically):
bearing capacity failure. The lower limit could be given
by a sliding failure mechanism where no passive suc-
(1)
tion will be mobilized. A reverse end bearing mech-
anism, which could happen when the caisson is pulled
out at a rapid rate, with generation of large passive with: cv: soil coefficient of consolidation
suction would lead to a larger uplift capacity. v: loading velocity (displacement rate).
Theoretical formulas are listed and compared to This simple parameter, varying over several orders
experimental data obtained on centrifuge small scale of magnitude may be linked to the probable failure
model. The large LCPCs centrifuge in operation in mechanism (Table 1).
Nantes (Corte & Garnier 1986) has indeed been used The Deng & Carter FEM analysis was done on
for 15 years in the area of suction caisson anchoring clay beds showing an increase of shear strength with
systems (Garnier 2002a, Rault et al. 2004). depth that corresponds to most of the field conditions.

273

Copyright 2005 Taylor & Francis Group plc, London, UK


Fu h
Mudline

Figure 1. Schematic suction caisson.

Table 1. Draining conditions (from Deng & Carter, 2000).


Figure 2. Schematic sliding failure.
Failure
Tk Drainage mechanism

0.6 Tk Drained Sliding (Fig.2)


0.002 Tk 0.6 Partially Tension (Fig.3)
drained
Tk 0.002 Undrained Reverse (Fig.4)
bearing
capacity

However, the selection of the value to be used for cv in


equation (1) remains very difficult.

2.1 Sliding failure


In this mechanism, only the caisson is pulled out of the
subsoil (Fig. 2). The soil plug remains at the same place.
Thus shear failure along the exterior and the interior Figure 3. Tensile failure mechanism.
wall of the pile occurs. Friction forces inside (Ti) and
outside (Te) the caisson have to be taken into account.
Water pressure induces forces outside and inside the 2.3 Reverse bearing capacity failure
suction caisson (Fuout et Fuint). A tip resistance Futip
may be included in the calculation, but it is often of low This mechanism includes the same kind of forces than
amplitude (small thickness of the caisson wall com- the previous one except the force due to soil failure,
pared with its diameter) regarding the other parameters. which is deduced from bearing capacity formula, but
with an reversed mechanism. The caisson is pulled
out with the soil plug. Thus the overall pullout load is
2.2 Tensile failure estimated from a bearing capacity force and from the
skin friction on the external wall.
In this case (Fig. 3), the soil plug moves upward with
Centrifuge tests performed on suction caisson
the suction caisson. The plug weight (Ws) has to be
embedded in clay show that failure occurs following
taken into account. Only friction outside the caisson
such kind of mechanism.
plays a part. In addition, a tensile force exists at the
The following formulas, written by several authors
bottom of the plug (Ft). This force is mainly linked to
during the recent years, have been based on the equi-
the soil characteristics (su).
librium of the forces. The differences come mainly
For both the sliding and tensile failure mechanisms,
from:
several formulas may be found in the literature
giving the bearing capacity F1 or F2 (Christensen the evaluation of the skin friction between the suc-
et al. 1991a & 1991b, Sage Engineering 2001 where tion caisson and the soil,
the Bureau Veritas one is used). the calculation of the bearing capacity.

274

Copyright 2005 Taylor & Francis Group plc, London, UK


F3 with

Fuout
Wp

Ws

Te

NCuAe-qtipAe

Figure 4. Schematic reverse bearing capacity mechanism. 2.3.3 Cluckey & Morisson (1993)

Those parameters have been determined empirically


in the cases of Christensen et al. (1991a & 1991b), (5)
Bureau Veritas (Sage engineering, 2001), Cluckey &
Morisson (1993) or with the help of finite element with
models (Renzi et al. 1991, Deng & Carter 2000, Nc   2 bearing capacity factor
Rahman et al. 2001). s  1.2 caisson shape factor
d  1  0.18 tan
1 (B/L) depth factor
2.3.1 Christensen et al. 1991a & 1991b i  6T/suNc B2 inclination factor (T: horizontal load)
These authors suggest that the equilibrium should be
written as: Most of the parameters are deduced from experi-
mental data (e.g. sol-caisson friction).

2.3.4 Renzi et al. (1991)


(2)
where:
Fout  Awh (6)
su(z) is the undrained shear strength profile
Te  e su Al (e external skin friction factor) where:
N  min{9;6.2(1  0.35 D/B)}bearing capacity fac- Nc  9 bearing capacity factor
tor, based on Brinch Hansen (1961, 1970)   0.3 skin friction factor
formulas
qtip  hw + Ds pressure at pile tip level. 2.3.5 Deng & Carter (2000)
After simplifications, the equilibrium equation
becomes:
(7)
(3)
where:
2.3.2 Bureau veritas (sage engineering, 2001) cs  1.2 shape factor
This formula includes the same force analysis than ce  1  0.4(D/B) embedment factor
the previous one, but each term includes several par-
ameters that are linked to the soil properties and to the
2.3.6 Rahman et al. (2001)
caisson geometry.
(8)

where:
Nc  8(D/B)
0.1833 bearing capacity factor
(4) dc  1  0.4 tan
1(D/B) embedment factor

275

Copyright 2005 Taylor & Francis Group plc, London, UK


Figure 5. LCPC centrifuge.
Figure 6. Model caisson used in the centrifuge tests carried
out by Morrison et al. (1994).
3 CENTRIFUGE TESTS

For many years, tests on suction caisson in saturated


clays have been performed in the LCPC centrifuge, in
the framework of several research programmes (e.g.
Clukey & Morrison 1993, Puech et al. 1993, Morisson
et al. 1994, Garnier et al. 1995, Raines & Garnier 2004,
Rault et al. 2004).
The large LCPC geotechnical centrifuge (Cort &
Garnier, 1986) has a radius (distance axis/basket plat-
form) of 5.50 m, a maximum payload of 2 tons and a
maximum centrifuge acceleration of 200 g (Fig. 5).
An on-board robot has been added to the facility allow-
ing to perform several tests on the model without stop-
ping the centrifuge. In 2005, an earthquake actuator
will also be put in operation allowing to simulate strong
earthquake effects on the centrifuge models.
Most of the experimental data has been obtained on
Figure 7. Pull out test of suction caisson Comparison
soil models made of Kaolin Speswhite Clay, submitted between field and centrifuge model tests carried out on
to high g-level (100 g), corresponding to a prototype Lysaker and Speswhite clay (from Morrison et al. 1994).
100 times larger. It has indeed been demonstrated that,
as long as the shear strength profiles are the same, the
results of undrained loading tests of suction caissons (calculated for load controlled tests), the characteris-
are very similar, whatever the type of clay used. Three tics of the clay and the suction caisson diameter give
series of pull out loading tests of suction caissons Tk 0.002. This means undrained conditions and so
where reported by Morrison et al. (1994): a reverse bearing mechanism at failure. The load is
applied with an hydraulic servo-jack or with electric
one 1-g field test carried out by NGI at the Lysaker
jacks (Fig. 8) during the centrifuge test.
site;
Caissons of different aspect ratios are installed
one 10-g centrifuge model test performed by LCPC
in-flight and tested under monotonic vertical loading
using the Lysaker clay;
up to the failure. The largest tested suction caisson
two 10-g centrifuge model tests performed by LCPC
corresponds to 30 m high prototype (Fig. 9).
using Speswhite clay.
Before starting the caisson pull out test, the clay
Figure 6 shows the model caisson used in these cen- shear strength is measured in-flight using the movable
trifuge tests and results of the pull out loading tests on-board CPT test device (Fig. 9). The undrained shear
are compared in Figure 7, as curves of force versus dis- strength is calculated from the tip resistance qc using
placement for three different test series (values are the following relationship (Garnier 2002b), based on
given in model units). shear vane versus CPT correlation tests:
In the present series of centrifuge tests, load con-
trolled or displacement controlled, the loading velocity su(z)  qc(z)/18.5

276

Copyright 2005 Taylor & Francis Group plc, London, UK


120
Experiment
Deng & Carter
100 Renzi et al.
Cluckey & Morisson
Christensen et al.

Failure load [MN]


80
Veritas

60

40

20
Prototype scale
0
1 2 3 4 5 6 7
Aspect ratio [-]

Figure 10. Failure load under vertical loading versus aspect


ratio of the suction caisson.

The formulas presented in Section 2.4 are then


used to calculate the theoretical bearing capacities
and results are compared with experimental data in
Figure 10, in prototype scale.
Even if some scattering may be observed, the agree-
ment between theoretical and experimental data is good.
The differences between experiments and theory does-
nt exceed 30% for short caissons, but may reach 80%
for high slenderness. The best theoretical predictions
Figure 8. View of the whole device for a suction caisson come from methods proposed by Rahman et al. (2001),
test in the swinging basket of the centrifuge. Vertical static Cluckey & Morisson (1993) or Christensen et al.
pullout test is applied here with electric jack.
(1991a & 1991b).
The differences observed between the results may
be due to the large uncertainties on the parameters
which characterize wall friction and reverse bearing
capacity. Generally, those parameters have been deter-
mined experimentally for an aspect ratio of 2, and that
may explain the better results obtained.

4 CONCLUSIONS

Suction caissons placed in clay and submitted to ver-


tical static undrained loading have been studied on cen-
trifuge reduce scale models. The experimental results
on bearing capacity have been compared with the the-
oretical data. In the formula presented here, the max-
imum pullout vertical static resistance of a suction
caisson embedded in clay is a function of:
Figure 9. Centrifuge model caisson of a 30 m high proto-
type (Raines & Garnier 2004). total weight of the caisson
weight of the soil plug
friction at the interfaces soil/caisson
bearing capacity of the soil
The obtained profiles display an increase of su(z)
undrained shear strength of the soil.
with depth of approximately K  1 kPa/m. Those
results have been obtained with a cone penetration The comparison between experimental data obtained
rate of 2 mm/s (model rate). on centrifuge small scale models of suction caissons

277

Copyright 2005 Taylor & Francis Group plc, London, UK


and from analytical formulas gives a better agreement REFERENCES
for the smaller aspect ratio (difference less than 30%)
than for the higher one. Allersma H.G.B., Jacobse J.A. & Krabbendam R.L. 2003.
A better understanding of the skin friction devel- Centrifuge test on uplift capacity of suction caisson with
oped at the interface between the suction caisson and active suction, ISOPE, Honolulu, 2530 May, pp. 734739.
Brinch Hansen J. A general formula for bearing capacity.
the soil should improve those results. The evaluation Danish Geotech. Institute, Bull. 11, 1961.
of the bearing capacity factor with a better knowledge Brinch Hansen J. A revised and extended formula for bear-
of the failure mechanisms should also be a source of ing capacity. Danish Geotech. Inst. Bull. 28, 1970.
improvement of the results. Christensen N.H., Haahr F. & Rasmussen J.L. 1991a. Breakout
The theoretical prediction of the maximum pullout resistance of large suction piles, OMAE vol I-B, Offshore
force in the case of more realistic loading conditions Technology, pp. 617622.
such as inclined, cyclic (loop current, storm, hurricane) Christensen N.H., Haahr F. & Rasmussen J.L. 1991b. Soil
or long term constant sustained loads is more diffi- structure interaction model for a suction pile platform,
cult, and need further investigation. Rather large pro- OMAE Vol I-B, Offshore Technology, pp. 601610.
Cluckey E.C. & Morrison M.J. 1993. A centrifuge Analytical
grammes of centrifuge tests on suction piles under Study to Evaluate Suction Caissons for TLP- Applications
complex loading conditions are running in several in the Gulf of Mexico, Design and Performance of Deep
laboratories. The collected data will be useful for cal- Foundations, pp. 141156.
ibrating the theoretical or numerical models being Clukey E.C., Morrison M.J., Garnier J. & Corte J.F. 1995.
developed for these more realistic loadings. The response of suction caisson in normally consolidated
clays to cyclic TLP loading conditions, OTC, Houston
14 May, pp. 909918.
Clukey E.C., Templeton J.S., Randolph M.F. & Phillips R.
5 NOTATION 2004. Suction caisson response under sustained loop current
loads, OTC, Houston, 36 May 2004, paper OTC 16843.
A: cross section of the caisson Colliat J.-L. Caissons succion pour lancrage de structures
ptrolires en mer profonde. Revue franaise de gotech-
Al: shaft area of the caisson nique n88, pp. 1119, 1999.
B: diameter of the caisson Cort J.F. & Garnier J. 1986. Une centrifugeuse pour la
cv: soil coefficient of consolidation recherche en gotechnique. Bulletin de liaison des labo-
D: embedment of the caisson ratoires des Ponts et Chausses. 146. pp. 528. LCPC.
Fout: force, due to water pressure effect outside the nov-dc.
caisson, applied downward Deng W. & Carter J.P. 2000. A theoretical study of the verti-
Fint: force, due to water pressure effect inside the cal uplift capacity of suction caissons, Proceedings of the
caisson, applied upward 10th International Offshore and Polar Engineering con-
Fu: ultimate load ference, pp. 342349.
Dyvik R., Andersen K.H., Hansen S.B. & Christophsen H.P.
h: depth from water surface to top of caisson 1993. Field tests of anchors in clay, ASCE, Journal of
K: slope of the shear strength profile Geotechnical engineering, Vol. 119, N 10, pp. 15151548.
L: length of the caisson Garnier J. 2002a. Modles physiques en gotechnique : II-
N, Nc: bearing capacity factors Validation de la mthode et exemples dapplication, Revue
p0: effective overburden stress at the bottom of the Franaise de Gotechnique, 98, 1er trimestre, pp. 528.
caisson before the installation of the caisson Garnier J. 2002b. Properties of soil samples used in cen-
qtip: pressure at pile tip level trifuge models. ICPMG02, St Johns, Phillips et al.;
su(z): undrained shear strength profile (Eds), Balkema, pp. 520.
s : Garnier J., Rault G. & Cottineau L.-M. 1995. Applications
u mean undrained shear strength
de la modlisation physique au domaine offshore, Coll.
T: horizontal load AUGC Les modles rduits en Gnie Civil, Nantes, 18
Ti: Friction forces inside mai, pp. 129136.
Te: Friction forces outside Iskander M., El-Gharbawy S. & Olson R. 2002. Performance
Tk: adimensionnal number suction caissons in sand and clay. Can. Geotech. J. 39:
v: loading rate 576584.
Wp: total weight of the caisson Morrison M.J., Cluckey E.C. & Garnier J. 1994. Behavior of
Wp: submerged weight of the caisson suction caissons under static uplift loading, Int. Conf.
Ws: weight of the soil plug Centrifuge 94, Singapour, Aot/Sept., pp. 823828.
Ws: submerged weight of the soil plug Puech A., Iorio J.-P., Garnier J. & Foray P. 1993. Experimental
study of suction effects under mudmat type foundations
z: depth in the marine soil 4me Conf. Can. Gnie Gotech. Marin, St-Johns, Canada,
e: external skin friction factor Vol. 3, pp. 10621080.
: unit weight of submerged soil Rahman M.S., Wang J., Deng W. & Carter J.P. 2001. A
w: unit weight of water neural network model for the uplift capacity of suction
s: unit weight of soil solid particles. caissons, Computer and Geotechnics 28, pp. 269287.

278

Copyright 2005 Taylor & Francis Group plc, London, UK


Raines R.D. & Garnier J. 2004. Physical modelling of suc- Sage Engineering 2001. Projet Clarom Caisson dancrage
tion piles in clay, OMAE Conference, Vancouver, 2025 Caisson sollicit verticalement, efforts statiques, com-
June, pp. 111. paraison mthode de prdimensionnement rsultats des
Rasmusssen J.L., Christensen N.H. & Haahr F. 1991, Soil essais en centrifugeuse, Rapport C339-03-0, Jan. 2001,
structure interaction model for a suction pile platform, pp. 116.
OMAE Conference, Vol. 1-B, pp. 601610. Steensen-Bach J.O. 1992. Recent model tests with suction piles
Rault G., Thorel L., Garnier J. 2004. Physical modelling of in clay and sand, OTC, Houston, May 47, pp. 323330.
off-shore foundations and structures. Journes AUM/AFM
Brest, 23 Sept. 10p (in french).
Renzi R., Maggioni W., Smits F. & Manes V. 1991. A cen-
trifugal study on the behavior of suction piles, Centrifuge
91, pp. 169176.

279

Copyright 2005 Taylor & Francis Group plc, London, UK


Capacity of suction caissons under inclined loading in normally
consolidated clay

R.M. El-Sherbiny, R.E. Olson, R.B. Gilbert & S.K. Vanka


The University of Texas at Austin

ABSTRACT: Laboratory experiments were conducted using instrumented prototype suction caisson in tanks
of normally consolidated kaolinitic clay. The prototype caisson was 102 mm in diameter and had an aspect ratio
of 8, with a diameter-to-wall thickness ratio of 115. Instrumentation was used to measure load, displacement,
tilt, and pore water pressure for loads ranging from horizontal to vertical. The caisson was installed half way
using dead weight followed by suction insertion to full penetration. The caisson was loaded rapidly after allow-
ing for sufficient setup time. Caisson response during loading is presented and measured capacities are com-
pared with predicted values. Caisson displacements were predominantly horizontal for loading angles less than
20 from horizontal and were predominantly vertical for loading angles above 30. Good comparison was found
between measured capacities and predictions from the PLA model.

1 INTRODUCTION that the capacities measured in the centrifuge tests


were slightly lower than values from plasticity solu-
Suction caissons are used as mooring anchors for float- tions for loading angles less than 40.
ing offshore structures with imposed loading ranging In our investigation, 1-g laboratory tests were con-
from vertical to horizontal. Suction caissons are widely ducted to study the capacity of suction caissons under
used as anchors in normally consolidated and lightly various load inclinations when loads are applied at
overconsolidated clays below deep water. Typical diam- two-thirds of the penetration depth below the mud-
eters of current caissons are 3.57.5 m with length-to- line. The measurements are compared with analytical
diameter ratios (aspect ratios) approaching ten (Tjelta, predictions.
2001).
Analytical and numerical modeling of behavior of
suction caissons under lateral loading has been inves- 2 EXPERIMENTAL SETUP
tigated by Sukumaran et al. (1999), Randolph and
House (2002), Aubeny et al. (2003a, b), Maniar (2004), Tests were conducted in a 1.2 m thick deposit of nor-
and Supachawarote et al. (2004). mally consolidated clay prepared in a steel tank meas-
Limited experimental data are available to verify the uring 1.2 by 2.4 m in plan by 1.8 m deep (Fig. 1). The
analytical and numerical models. The first horizontal test bed soil was prepared by mixing kaolinite slurry
tests on suction caissons were conducted with load and allowing it to consolidate under self-weight (Olson
attached at the mud line (Hogervorst, 1980). Keaveny et al. 2003, Coffman 2003). The final soil surface was
et al. (1994) showed that lowering the load attachment almost flat except for a 50 mm zone right next to the
point to mid depth almost doubled the capacity for a tank sidewalls.
group of two cylindrical caissons attached tangentially.
Coffman et al. (2004) examined the capacity of suction
2.1 Prototype caisson
caissons under horizontal loading using a 1-g labora-
tory model and showed that the largest capacity was The prototype caisson (Fig. 2) was constructed from
achieved when the loading point was between two a 102-mm diameter anodized aluminum tube with a
thirds and three quarters of the caisson embedment. wall thickness of 0.8 mm. A Delrin top cap provided
The results agreed well with the plasticity solution by valving and connections for various transducers and
Aubeny et al. (2003b) and finite element analysis con- mechanical equipment. The maximum penetration
ducted by Maniar (2004). Clukey et al. (2003) reported possible was 876 mm but only 813 mm of penetration

281

Copyright 2005 Taylor & Francis Group plc, London, UK


was attached to the padeye at one end and to a load cell,
a Linear Variable Differential Transformer (LVDT), and
a computer-controlled stepper motor at the other end.
The pulley around which the loading cable turned
was mounted at a horizontal distance of about 216 mm
from the caisson in the clay (Fig. 1). Vertical and hor-
izontal movements of the caisson top were deter-
mined from two LVDTs connected to the top cap with
nylon lines (Coffman 2003). Tilting of the caisson
was measured with a two-axis tilt meter.
Eight pressure transducers were connected to sens-
ing tips along the inner and outer walls of the caisson
to measure pore water pressures. Tips I1, I2, I3, were
mounted on the inside and located 584, 280, and
25 mm above the tip, respectively. Tips O1 and O2
were located on the outer front side at heights of 482
and 25 mm above the tip, respectively, while tips O3
and O4 were located on the outer backside at heights
of 254 and 25 mm above the tip, respectively. The
eighth sensing tip was placed beneath the top cap to
measure the under pressure.

2.3 Soil strength


The undrained shear strength profile of the test bed
soil was used to normalize side and tip stresses.
Undrained strengths were measured in situ using a
T-bar, cone, vane, and ball.
The T-bar was a smooth acrylic rod (25 mm diam-
eter by 100 mm long) that was penetrated at a rate of
20 mm/s. The rod was connected to a universal joint at
the opposite end before attaching to the load cell in
Figure 1. Cross section through the test tank depicting an
inclined load test. order to minimize bending moments on the load cell.
The undrained shear strength (Fig. 3) was determined
using a bearing capacity factor of 10.5 (Stewart &
Randolph 1994). The cone had a diameter of 44.5 mm
Padeye bar with cable
attachment points
and a tip angle of 60. The cone was calibrated in a
water column prior to testing. It was pushed into the
Pore pressure
soil at a rate of about 20 mm/s. A cone tip resistance
sensing tip O2
factor of 12 was used (Vanka, 2004). A 44.5-mm
Top cap
diameter ball was used in the ball penetration test. The
testing procedures duplicated those for the T-bar.
Anodized aluminum tube: The ball resistance factor was 13.5 (Randolph, 2004).
thickness = 0.8 mm Pressure line
The vane was 44.5 mm in diameter and 89 mm high.
diameter = 102 mm to transducer A rotation rate of 0.3/s was applied (typical in off-
length = 902 mm
shore practice, Young et al. 1988). The results of the
different shear strength measurements yielded very
Figure 2. Prototype caisson used in the model tests. close results except for the vane, which gave lower
undrained strengths (Fig. 3b). The lower vane strengths
was scheduled to allow for plug heave and to maintain may be the result of smaller strain rates for the vane
an aspect ratio of eight. A padeye bar allowed mul- (Randolph, 2004), in addition to the different state of
tiple points of loading. stress compared with the other tests.
Shear strength tests were performed at locations
along the centerline of the test bed prior to conducting
2.2 Instrumentation
lateral load tests. The soil properties shown in Table 1
Test fixtures and instruments were mounted on a frame and plotted in Fig. 3 were used in normalizing subse-
(Fig. 1) positioned on top of the tank. The loading cable quent measurements of caisson behavior.

282

Copyright 2005 Taylor & Francis Group plc, London, UK


Undisturbed Undrained weight) at a penetration rate of about 7.5 mm/s. At
Shear Strength (Pa) about 406 mm of penetration, dead-weight insertion
0 200 400 600 800 1000 was stopped, surcharge weight was removed, and suc-
0.0 tion was applied slowly with a vacuum regulator used
Cone (1)
to maintain a caisson penetration rate of about
Cone (2)
0.5 mm/s. Suction penetration was stopped when the
Ball (1)
caisson reached an embedded depth of about 813 mm
0.2 (Rauch et al. 2003).
Ball (2)
T-bar (1)
In the first test, the caisson was locked in place
Depth from mudline (m)

T-bar (2)
with full penetration and pore water pressures were
Vane
measured on the inner and outer walls. The time for
0.4
Average
full dissipation of excess pore water pressures was 48
hours on the exterior and 96 hours on the interior.
Because the internal pore pressures influenced the
0.6 axial capacity, a setup time of 96 hours was used for
all tests. For tests with a lateral load component, the
caisson was first locked in position at full penetration
and the loading cable was stretched tightly in position
0.8 to minimize any initial catenary.
The load was applied at a constant rate of 1.7 mm/s
in all tests. Laterally loaded tests were continued until
the padeye displacement was about 50 mm. Axially
1.0 loaded tests were continued until vertical displace-
ment reached about 120 mm.
Figure 3. Undrained shear strength profiles in test bed soil.
Data collected during load tests involved vertical
and horizontal displacements of the top cap, padeye
displacements, tilt in and orthogonal to the loading
Table 1. Properties of consolidated soil in test bed. plane, applied load, and pore water pressures. Results
were corrected for frictional losses in the pulley
Soil property Property vs. depth function ( 0.5% of the load) and stretch in the load cable.
Total unit weight t (Kg/m3) t  144  z  1360
Undisturbed undrained shear cuu  920  z  48
strength, cuu (Pa) 4 TEST RESULTS
Remolded undrained shear cur  393  z
strength, cur (Pa) Test VL-1 (Table 2) was vertically loaded at the cen-
ter of the top cap while test VL-2 was vertically
z  depth below mudline (m) loaded at the padeye. Five inclined loading tests
(INC-01 through INC-45) were conducted with load-
ing angles ranging from 1 to 45.
3 TEST PROCEDURE In test INC-01, the loading cable broke at a load of
about 382 N. The cable was replaced and the subse-
One test was performed with tensile load applied at quent measured capacity was only 1.9% greater than
the top center of the caisson. Six additional tests were the horizontal component of the capacity with the
performed with loading angles ranging from 1 to 90 load inclined at 8 and the increased stiffness of the
and the load applied at the critical depth (two thirds of soil was only about 2%. The error in measured cap-
caisson penetration). In all seven tests, the caisson acity was considered insignificant and the test results
was inserted 305 mm (3 diameters) from the adjacent were used as measured.
tank wall with a center-to-center spacing of 305 mm.
For each test, the caisson was first positioned ver-
tically with the tip at mud line, and was locked in
4.1 Caisson movement
position such that it could move only vertically.
Penetration was controlled using a stepper motor with The peak load occurred at padeye displacements of 4
displacements measured using a 1270-mm travel to 8 mm (4 to 8% of the caisson diameter), followed
rotary displacement transducer. The caisson was by a loss in capacity (Fig. 4).
inserted halfway into the clay deposit using dead The tests with load inclinations less than 20
weight (self-weight supplemented with surcharge exhibited predominantly horizontal displacements

283

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 2. Test results.

Capacity
Padeye components (N)
Test ID Loading displacement
Angle (deg.) Capacity (N) at failure (mm) Horizontal Vertical

VL-1 90 167.3 6 0.0 167.3


VL-2 90 171.3 5 0.0 171.3
INC-01 1.2 404.3 4 404.2 8.5
INC-10 8.9 400.8 6 396.8 55.6
INC-20 19.2 387.0 4 365.6 127.3
INC-30 30.6 347.4 8 299.1 176.8
INC-45 43.5 256.7 4 186.1 176.9

8.0
INC-01 Positive
400 INC-10
Tilt

Caisson tilt (deg.)


INC-30 4.0 Center of
Applied Load (N)

Load
300 Rotation
INC-01
INC-45 INC-20 INC-10
INC-20
200 0.0 INC-45
VL-2
VL-1 INC-30
100
-4.0
0 10 20 30 40 50
Padeye displacement (mm)
0
0 10 20 30 40 50
Padeye displacement (mm) Figure 6. Caisson tilt during loading.

Figure 4. Load displacement response.

caisson moved up relative to the fixed pulley. As the


40
Vertical padeye displacement (mm)

load angle approached 45, no significant horizontal


movement was noticed and the motion was vertical.
INC-45 Coffman et al. (2004) reported that the center of
30 rotation of the caisson is at two thirds the embedment
depth. They found that the top of the caisson rotated
INC-30 toward the loading point (positive rotation) for hori-
20 zontal loading above the center of rotation and away
from the loading point (negative rotation) when the
INC-20 horizontal load was applied below the center of
10 INC-01 rotation. In tests INC-01 and INC-10, the load vector
INC-10 intersected the centerline of the caisson above the
center of rotation causing a positive tilt that increased
0 with continued loading (Fig. 6). As the load inclin-
0 10 20 30 40 50 ation was increased (tests INC-20, INC-30, and INC-
Horizontal padeye displacement (mm)
45), the intersection of the load vector with the
centerline dropped below the center of rotation and a
Figure 5. Caisson displacement path during loading.
negative tilt was observed. However, continued load-
ing accompanied by vertical movement of the caisson
(Fig. 5). For test INC-30, the predominant displace- caused the load inclination to decrease and eventually
ment was vertical (Fig. 5). After 25 mm of vertical the intersection of the load vector with the centerline
displacement, the horizontal displacement started to of the caisson came above the center of rotation and
increase due to the decrease in loading angle as the the direction of rotation was reversed (Fig. 6).

284

Copyright 2005 Taylor & Francis Group plc, London, UK


Normalized excess pore Normalized excess pore
0.0 0.5
Test: INC-20 I1
Normalized Excess

pressure at failure
0 I2
Pore Pressure

-0.5 I3
-0.5
I1 -1
-1.0 O1
I1
O3 I2 I2 -1.5
Failure I3
O4 I3 O2
-1.5 -2
1.0 0.6

pressure at failure
O1 O1
Normalized Excess

O2 O2
Pore Pressure

0.5 O3 0.3
O3
O4 O4
0.0 I3 0

-0.5 -0.3

-1.0 -0.6
0 10 20 30 40 90 75 60 45 30 15 0
Padeye Displacement (mm) Load inclination from Horizontal (deg.)

Figure 7. Excess pore pressures during caisson loading. Figure 8. Excess pore pressures generated at failure.

250
Vertical component of capacity (N)

4.2 Excess pore water pressures


Measured excess pore water pressures were normal- 200 45o 30o
ized by dividing by the free-field vertical effective
stress at that depth. 150
20o
In all tests, the caisson moved upwards slightly at
the beginning of loading and suction developed inside 100
the top cap, resulting in a decrease in pore water pres- Measured Values
10o
sures in the soil plug (Fig. 7). For nearly horizontal 50 SAIL: =1.0, Nc=9.0
loading, exterior excess pore water pressure was posi- SAIL: =0.8, Nc=12
tive on the front side (O1 and O2). At load inclin- 0
ations above 10, the excess pore pressures decreased 0 100 200 300 400
Horizontal component of capacity (N)
for O2 before failure because of suction developed on
the inside. Measured excess pore water pressures on
Figure 9. Measured versus predicted capacities.
the backside (O3 and O4) decreased upon loading
because the caisson movement unloaded the soil. No
gap was observed on the backside in any tests. The axial capacity measured from the two axial ten-
The excess pore water pressures at failure tended sile tests (VL-1 with the load applied in the middle of
to be most negative on the inside for tests with large the top cap and VL-2 with the load applied to the pad-
vertical movements (Figs 5, 8), where the upwards eye) were within 2.5% of each other (Table 2). This
movement of the caisson caused internal suction. difference could be due to minor variations in strength
Horizontal loading resulted in relatively small interior spatially in the tank. Further, loading at the padeye
excess pore water pressures. caused only 2 of tilt when the test was terminated.
Exterior excess pore water pressures were negative For inclined loading, changes in the angle of load-
on the backside and positive on the front side (Fig. 8), ing at the time of failure due to continued displace-
but readings near the tip were influenced by the suc- ments were less than 1. Therefore, the initial inclination
tion developed on the inside. was used in plots and calculations.
The axial component of the capacity was within 3%
of its maximum value for loading angles between 30
4.3 Measured capacity
and 90. Observed padeye displacements (Fig. 6) sup-
The components of the measured capacity are pre- port the concept that failure was reached when the
sented in the interaction diagram (Fig. 9). The max- axial soil resistance reached its limit accompanied by
imum capacity was achieved when the caisson was predominantly axial displacements. The horizontal
loaded horizontally and the smallest capacity for axial component of the measured capacity remained close to
tensile loading. its maximum value between 0 and 10 and decreased

285

Copyright 2005 Taylor & Francis Group plc, London, UK


by only 8% at a load inclination of 20. Failure within As the load inclination exceeded zero, suction
this range was limited by lateral soil resistance with developed inside the top cap and internal excess pore
predominantly horizontal motion (Fig. 5). water pressures were negative. Along the outer wall,
the pore pressures were found to increase on the front
side and decrease on the backside for nearly horizon-
4.4 Comparison with predictions tal loading. No gap formed on the backside. The pore
pressures at the front tip of the caisson were found to
The measured capacity of the caisson was compared
be most sensitive to the pore pressures generated on
with predictions from the simplified plastic limit
the inside of the caisson.
analysis (PLA) model proposed by Aubeny et al.
For axial loading, the results for loading at the top
(2003b) (Fig. 9).
center and at the padeye were essentially the same.
In the PLA analysis, the average, undisturbed,
The components of the capacity in the vertical and
undrained shear strength profile from penetration tests
horizontal directions showed that the maximum verti-
(Fig. 3 and Table 1) was used to represent the test bed
cal component of the capacity was achieved when
soil. Rotational resistance at the caisson tip, and full
loading was between 45 and 90. In addition, no sig-
suction on the back of the caisson were assumed. The
nificant loss was observed in the horizontal compon-
lateral bearing factor (NL) was assumed to be 12 near
ent of the capacity when load inclination was between
the tip with the value decreasing at shallow depths
0 and 20.
(Aubeny et al., 2003a). Experimental investigations
The results were compared with predictions from
using 1-g and centrifuge modeling in normally con-
the PLA model proposed by Aubeny et al. (2003 a,b).
solidated clay deduced a range of the side resistance
The shear strength profile was estimated using T-bar,
reduction factor () between 0.5 and 0.8 and tip bear-
ball, and cone penetration tests, which gave consistent
ing capacity factor (Nc) between 9 and 15 (Clukey &
results. The predicted values of the capacity matched
Morrison, 1993, Randolph & House, 2002, Luke et al.
the measured capacity well when NL  12, with either
2003). For an aspect ratio of eight, Luke et al. (2003)
  1.0 and Nc  9, or   0.8 and Nc  12, as long
reported that  is more likely to be 0.8 accompanied
as no reduction was applied to NL. Additional investi-
by Nc of 12 although  of 1.0 and Nc of 9 would yield
gation using a double walled caisson is underway to
the same capacity. However, Coffman et al. (2004)
examine the  and Nc factors for limit equilibrium
showed good comparison with PLA model predic-
analysis of the axial capacity.
tions, for horizontally loaded caissons, only when
 was 1.0. Therefore, two pairs of factors were applied
to predict the capacity of the caisson under inclined
loading using the PLA model (Set 1:   0.8 and Nc  ACKNOWLEDGEMENT
12, and Set 2:   1.0 and Nc  9) (Fig. 9).
Both sets of factors agree well with the measured This study was supported by the Offshore Technology
axial capacity. However, Set 1 under estimated the Research Center at the University of Texas and Texas
horizontal component of the capacity while Set A&M University, with funding from the U.S. Minerals
2 matched the capacity well at all load inclinations. Management Service and the OTRC Industry
The PLA model (Aubeny et al., 2003b) assumes a Consortium.
reduction factor (s) on the lateral bearing factor (NL)
as a function of . The results suggest that if  is 0.8,
the reduction in lateral resistance does not follow the REFERENCES
reduction in axial resistance.
Aubeny, C. P., Han, S. W., & Murff, J. D. 2003a. Inclined
Load Capacity of Suction Caissons. Intern. J. Numerical
and Analytical Methods in Geomechanics, 12351254.
5 SUMMARY AND CONCLUSIONS Aubeny, C. P., Han, S. W., & Murff, J. D. 2003b. Refined
Model for Inclined Load Capacity of Suction. Proc.
Seven tests were conducted on a prototype suction 22nd Intern. Conf. on Offshore Mechanics and Arctic
caisson having an aspect ratio of eight. Loading was Engineering, Cancun, Mexico, June, OMAE03-37502.
applied at about two-thirds the embedment depth with Clukey, E. C., & Morrison, M. J. 1993. A Centrifuge and
load inclinations ranging from horizontal to vertical. Analytical Study to Evaluate Suction Caissons for TLP
The caisson exhibited displacements to failure ranging Applications in the Gulf of Mexico. ASCE Geotech.
Special Pub. (38): 141156.
between 3.5 and 7.5% of the diameter followed by a Clukey, E. C., Aubeny, C. P., & Murff, J. D. 2003. Comparison
drop in the load to a residual value. Displacements at of Analytical and Centrifuge Model Tests for Suction
failure were mainly horizontal for loading angles Caissons Subjected to Combined Loads. Proc. 22nd
between 0 and 20, and mainly vertical between 30 Intern. Conf. on Offshore Mechanics and Arctic Engineer-
and 90. ing, Cancun, Mexico, June, OMAE03-37503.

286

Copyright 2005 Taylor & Francis Group plc, London, UK


Coffman, R. A. 2003. Horizontal Capacity of Suction Caissons Conf. on Site Characterization, ISC-2, Porto, Portugal,
in Normally Consolidated Kaolinite. M.S. Thesis, The September, 209232.
University of Texas at Austin, December. Randolph, M. F., & House, A. 2002. Analysis of Suction
Coffman, R. A., El-Sherbiny, R. M., Rauch, A. F., & Caisson Capacity in Clay. Proc. Offshore Technology
Olson R. E. 2004. Measured Horizontal Capacity of Conf., Houston, Texas, OTC 14236.
Suction Caissons. Proc. Offshore Technology Conf., Rauch, A. F., Olson, R. E., Luke, A. M., & Mecham, E. C.
Houston, OTC 16161. 2003. Measured Response during Laboratory Installation
Hogervorst, J. R. 1980. Field Trials with Large Diameter of Suction Caissons. Proc. 13th Intern. Offshore and
Suction Piles. Proc. Offshore Technology Conf., Houston, Polar Engineering Conf., Honolulu, May, 780787.
OTC 3817. Stewart, D. P., and Randolph, M. F., (1994), T-bar
Keaveny J. M., Hansen S. B., Madshus C. and Dyvik R. Penetration Testing in Soft Clay, J. Geotechnical
1994. Horizontal capacity of large scale model anchors. Engineering, ASCE, 120 (12), pp. 22302235.
Proc. Thirteenth Intern. Conf. on Soil Mechanics and Sukumaran, B., McCarron, W. O., Jeanjean, P., &
Foundation. Engineering, New Delhi, India. Abouseeda, H. 1999. Efficient Finite Element Techniques
Luke, A. M., Rauch, A. F., Olson, R. E., & Mecham, E. C. for Limit Analysis of Suction Caissons under Lateral
2003. Components of Suction Caisson Capacity Measured Loads. Computers and Geotechnique, (24): 89107.
in Axial Pullout Tests. Proc. Intern. Symp. on Deepwater Supachawarote, C., Randolph, M., & Gourvenec, S. 2004.
Mooring Systems, Concepts, Design, Analysis, and Inclined Pullout Capacity of Suction Caissons. Proc.
Materials, OTRC, October, 112. 14th Intern. Offshore and Polar Engineering Conf.,
Maniar, D. R. 2004. A Computational Procedure for Toulon, France, May, 500506.
Simulation of Suction Caisson Behavior Under Axial and Tjelta, T. I. 2001. Suction Piles: Their Position and Applications
Inclined loads. Ph.D. Dissertation, The University of Today. Proc. of the Eleventh Intern. Offshore and Polar
Texas at Austin, August. Engineering Conf., Stavanger, Norway, June.
Olson, R. E., Rauch, A. F., Mecham, E. C., & Luke, A. M. Vanka, S. K. 2004. Laboratory Tests to Estimate Strength
2003. Self-weight Consolidation of Large Laboratory Profile of Normally Consolidated Kaolinite. M.S. Thesis,
Deposits of Clay. Proc., Pan-American Conf. on Soil The University of Texas at Austin, December.
Mechanics and Geotech. Engineering, Cambridge, Young, A. G., McClelland, B., & Quiros, G.W. 1998. In-Situ
Mass., June, 703708. Vane Shear Testing at sea. Vane Shear testing in Soils:
Randolph, M. F. 2004. Characterization of Soft Sediments Field and Laboratory Studies, American Society for test-
for Offshore Applications. Proc. of the Second Intern. ing Materials, Philadelphia, ASTM STP 1014, 4667.

287

Copyright 2005 Taylor & Francis Group plc, London, UK


A failure surface for caisson foundations in undrained soils

H.A. Taiebat
University of Technology Sydney, NSW, Australia

J.P. Carter
The University of Sydney, NSW, Australia

ABSTRACT: This paper presents the results of a series of numerical analyses of caisson foundations embed-
ded in a homogeneous soil deforming under undrained conditions. The performance of a typical caisson founda-
tion under separate axial, torsional and lateral forces is investigated, followed by the interaction of these forces
with each other. The lateral force is applied at various points along the skirt of the caisson so that the effects of
overturning moments are also included in the analyses. The ultimate capacity of the caisson under combined
loading is presented in the form of failure envelopes in the axial-lateral, axial-torsional and lateral-torsional load-
ing planes and the axial-lateral-torsional loading space. The results of this study show that although the capacity
of the caisson under lateral load depends on the location of the padeye along the caisson skirt, a unique failure
envelope, in a non-dimensional form, can be presented for the caisson regardless of the location of the padeye.

1 INTRODUCTION a caisson foundation is predominantly subjected to


axial and lateral loading, transferred from a mooring
Over the last two decades caisson foundations have line to a padeye on the caisson wall. Any misalignment
been used increasingly in marine environments for tem- of the padeye or any change in the direction of the
porary or permanent mooring of floating offshore facil- mooring line also induces torsional (twisting) loads
ities, for tension leg platforms as well as for gravity on the caisson. Caisson foundations are often subjected
based structures. Recently, their use has also been pro- to overturning moments; the effects of which are often
posed for foundations for offshore wind turbines. They considered by appropriately selecting the point of
are considered as particularly reliable and cost-effective application of the horizontal load along the skirt of the
alternatives to more conventional mooring systems in foundations.
deep and ultra deep waters. They have advantages over The response of caisson foundations to axial and
other conventional offshore anchoring systems because lateral loads has been studied previously. The focus of
of their large bearing capacity and the efficiency of their some earlier research (e.g. Bransby & Randolph 1997,
installation. Caisson foundations have typically a large 1998, Bransby 2001) has been on the capacity mobil-
diameter and a wide range of length-to-diameter ratios, ized at the tip of the caisson, ignoring the resistance
so they provide relatively large lateral and axial cap- of its skirt. In other research (e.g. Sukumaran et al.
acity. A caisson can partially penetrate into the soil 1999, Zdravkovic et al. 2001, Deng & Carter 1999,
under its own weight. Further penetration is usually 2002, Deng et al. 2001) resistances of both the tip and
facilitated by pumping water out of the caisson cham- the skirt of the caissons have been considered. The
ber, thus applying suction inside the caisson. The dif- effects of torsional loads on the axial and lateral capaci-
ference between the external and internal fluid pressures ties of caissons have also been investigated by Taiebat &
acts as an external surcharge pushing the caisson into Carter (2004), where it was shown that, in general,
the soil. This simple installation procedure is probably torsional loads reduce the axial capacity and lateral
the greatest advantage of caisson foundations over pile resistance of cylindrical caisson foundations.
foundations. A suction caisson can be withdrawn later In this study a series of finite element analyses was
by applying a positive pressure inside the chamber to performed in order to obtain an insight into the inter-
help pull it out of the soil. action of axial, lateral and torsional loading and the
In the marine environment caisson foundations are capacity of caisson foundations under combinations
subjected to various combinations of axial, lateral and of these loads. Since the main aim of this study was to
torsional loading. Acting as a part of mooring system, investigate the overall interaction of these forces, the

289

Copyright 2005 Taylor & Francis Group plc, London, UK


problem was solved primarily for a typical caisson
with a length-to-diameter (aspect) ratio of 2. This is D
within the typical range encountered in practice. A
limited number of analyses was also performed for a
caisson with an aspect ratio of 4, to investigate the
effects of the aspect ratio on the overall behaviour of L
the caisson foundations.

2 FINITE ELEMENT MODEL


3.5L
The typical caisson foundation considered here has a
diameter D and a length L, and is embedded in a homo-
geneous soil that deforms under undrained conditions.
It was assumed that the soil obeys the Tresca failure
criterion. It has a uniform undrained shear strength of
su and an undrained Youngs modulus Eu  300 su. A 8D
Poissons ratio of   0.5 ( 0.49) was assumed for
the soil to approximate the constant volume response Figure 1. Finite element mesh and geometry of the problem.
under undrained conditions. The rigidity index G/su is
therefore equal to 100, where G is the elastic shear
modulus of the soil. The caisson material has a Youngs All analyses were performed under displacement-
modulus of Ec  1000 Eu. defined conditions, where a vertical or torsional dis-
The geometry of the problem under investigation is placement was applied to the foundation at the ground
axi-symmetric, but generally the loading is of course level, or a lateral displacement was applied at various
non-symmetric. The axi-symmetric nature of the geom- locations along the skirt of the foundation, above or
etry was exploited to achieve economies in obtaining below the ground level. A combination of two or three
the finite element solutions. The finite element mesh components of displacement was also applied to the
used in the analyses has 12 wedges of elements in the foundations to investigate their behaviour under com-
circumferential direction. Each wedge of the typical bined loading. It is assumed that the loads are applied
caisson consists of 304 isoparametric (20 nodes) brick at a rate sufficiently rapid that the surrounding soil
elements. A thin layer of elements has been used around deforms under undrained conditions. No provision has
and under the caisson in order to capture the effects of been made to model any separation or de-bonding that
shearing close to the foundation. A schematic represen- may occur between the soil and the foundation.
tation of six of the wedges, i.e. half of the three-
dimensional finite element mesh used in the analyses of
the typical caisson, is shown in Figure 1, which also 3 FINITE ELEMENT RESULTS
defines the overall geometry of the finite element
model. Of course, a finer mesh may be used to obtain a The resistance of the caisson foundation under each
more accurate finite element solution to the problem in individual component of loading, i.e. axial, lateral or
hand. However, as the main aim of this study was to find torsional loading, is presented in this section.
the overall interaction of the forces at foundation failure,
this finite element mesh was considered satisfactory.
The small strain formulation used for these analyses 3.1 Axial resistance
is based on the semi-analytical approach in finite The axial bearing capacity of a caisson foundation in
element modelling described by Zienkiewicz & Taylor undrained soil, Vu, can be approximated using the
(1989), which is an efficient tool for three-dimensional conventional method of bearing capacity calculation,
problems. A semi-analytical finite element method, e.g. from the equation suggested by Vesic (1975), as:
based on representation of nodal variables in terms of
discrete Fourier series, has been integrated into the
(1)
finite element code AFENA (Carter & Balaam 1995).
Application of this method in the analyses of three-
dimensional problems has shown a considerable where A is the plan area of the caisson, Nc is the bear-
reduction in computational time, compared with a ing capacity factor for a strip footing corresponding
straightforward 3-D analysis. Details of the semi- to the undrained shear strength of the soil, s and d are
analytical method used in this study may be found factors that include the effects of the shape of the foot-
in Taiebat & Carter (2001). ing and the effects of embedment of the foundation.

290

Copyright 2005 Taylor & Francis Group plc, London, UK


30
4
25
D
D
Vertical load / (A . su)

H / (D. L. su)
20 H
D L/D=4 LL
15 2
L/D=2

10 +- V
L/D=4 1
L/D=2 L
5
0
0 0 5 10 15 20
0 10 20 30 40 50 Displacement . G / (L . su)
Vertical Displacement . G / (D . su)
Figure 3. Caisson response under lateral loading.
Figure 2. Caisson response under axial loading.

reduces significantly. At a relatively large displace-


ment the increase in the resistance becomes insignifi-
The bearing capacity factor for undrained soil cant and the ultimate load is approached. The ultimate
Nc  (2  ). The shape factor for a circular footing axial resistances of the caisson predicted by the finite
is usually suggested as s  1.2. The embedment fac- element analysis, at vertical displacements of about
tor for a circular footing has been suggested as 50% of the caisson diameter, are 18.4 A su for the cais-
d  1  0.4 tan
1(L/D); for the case of L/D  2, the son with L/D  2 and 26.6 A su for L/D  4. These
embedment factor is s  1.443 and for L/D  4, values are slightly smaller than those predicted by
the embedment factor is s  1.530. Therefore, for Equation (2). If the skirt adhesion, DLsu, is added to
the special cases considered here the axial load capac- the capacity predicted by the conventional method,
ity of a buried circular footing is given by the conven- equation (1), the results will be 16.9 A su and 25.4 A su
tional method as Vu  8.9 A su for L/D  2 and for L/D of 2 and 4, respectively. These values are closer
Vu  9.4 A su for L/D  4. The effects of the adhe- to those predicted by the finite element analyses.
sion developed on the caisson skirt are not included in
this method. 3.2 Lateral resistance
Deng & Carter (1999) recognised the effect of
adhesion developed on the skirt of the foundation and The lateral resistance of a caisson foundation in an
suggested an equation for the uplift capacity of caisson undrained homogeneous soil can be given as:
foundations in an undrained homogeneous soil as:
(3)
(2)
where Hu is the ultimate lateral capacity of the cais-
son and Nh is defined as the lateral capacity factor.
where Np  9.0 is the uplift capacity factor and Deng & Carter (1999) suggested a lateral capacity
ce  1  0.4 (L/D). Equation (2) results in an uplift factor that is a function of the point where the load
capacity of 19.44 A Asu and 28.08 A su for caissons is applied. For lateral load applied at the surface
with L/D of 2 and 4, respectively. It should be noted Deng & Carter suggested Nh  4.8.
that the uplift capacity problem of a caisson in soil The results of the finite element analysis of the
deforming under undrained conditions is a reverse caisson subjected to lateral loading applied at the
bearing capacity problem and can be treated similarly ground level are presented in Figure 3. Lateral cap-
(e.g. Anderson et al. 1993). Therefore, Equation (2) acity factors of Nh  4.0 and 4.20 are obtained for
can equally be used to calculate the compressive bear- L/D of 2 and 4, respectively. Note that the lateral dis-
ing capacity of caisson foundations. The results of the placements are normalised here with respect to L, the
finite element analysis also do not indicate any sig- skirt length, indicating a larger final lateral displace-
nificant change in the axial capacity when the direc- ment for the caisson with L/D  4.
tion of loading is reversed. The resistance of a caisson against lateral load
The results of the finite element analyses of the depends on the location of the padeye, as any eccen-
caissons under axial loads are presented in Figure 2. tricity of the lateral load applies an overturning moment
The response is approximately linear at the beginning on the caisson. The variation of the lateral resistance
of loading where about 65% of the ultimate axial of caissons with the point of load application is pre-
resistance is mobilized. After a rapid bend in the load- sented in Figure 4. In this figure z represent the coord-
deflection curve, the rate of increase in the resistance inate of the load application point, with its negative

291

Copyright 2005 Taylor & Francis Group plc, London, UK


2.0 7

5 D
1.5

T / (D3. su)
4
T
3 T
L
1.0 2 L/D=4
1 L/D=2

0
L/D=2
z/L

0.5 0 1 2 3 4 5
Rotation . G / su
L/D=4

Figure 5. Caisson response under torsional loading.


0.0

are 3.40 su D3 and 6.53 su D3 for L/D of 2 and 4,


-0.5 respectively, in excellent agreement with the theoret-
ical values obtained using Equation 4.

-1.0 4 FAILURE ENVELOPES


0 2 4 6 8 10 12
H / (D. L. su) In order to evaluate the interaction between different
Figure 4. Lateral resistance versus the load application point.
components of loading, a series of finite element
analyses was performed using different ratios of the
vertical and lateral displacements and rotation of the
sign indicating that the point is below the ground caisson. For each loading case the ultimate axial cap-
level. The responses of the two caissons considered acity is reached at a vertical displacement equal to about
here are very similar; as L/D increases the depth of 50% of the diameter of the caisson, while the ultimate
the maximum lateral resistance shifts slightly upward, torsional and lateral capacities are obtained at displace-
toward the ground level. The maximum lateral resist- ments equal to approximately 20% of the diameter. The
ances occur when the lateral load is applied at about results of the finite element analyses will be presented
0.6 L below the ground level. in the form of two-dimensional failure loci in axial-
lateral, axial-torsional, lateral-torsional loading planes,
3.3 Torsional resistance followed by a non-dimensional 3-D failure surface.
Assuming the full value of the undrained shear
strength of the soil is mobilized as a shear stress at the 4.1 Axial-lateral failure plane
caisson-soil interface, the theoretical value for the Failure loci for caissons with L/D  2 subject to
ultimate torsional capacity of a caisson foundation, combinations of axial and lateral forces, applied at
Tu, can be calculated as the sum of the base resistance different points along the caisson skirt below and
and the shaft resistance, i.e. above the ground level, are presented in Figure 6. As
explained previously, the lateral capacity of the cais-
(4) son depends on the point of application of the load. A
non-dimensional form of the failure loci can be
obtained using the maximum lateral capacity at any
Equation (4) gives the ultimate torsional capacities of point, Hmax, and the maximum axial capacity, Vmax,
3.402 su D3 and 6.545 su D3 for caissons with L/D of 2 as the normalisation factors. These non-dimensional
and 4, respectively. failure loci are presented in Figure 7. Included in this
The load deflection curves predicted by the finite figure is a normalised failure locus for the caisson
element analyses of the caissons under torsional load- with L/D  4, where the lateral load is applied at the
ing are presented in Figure 5. The response of each ground level. It can be seen that all the failure loci are
caisson is virtually elastic-perfectly-plastic. The very close to each other and the discrepancies between
torsional resistance increases linearly to its ultimate different failure loci are almost insignificant. There-
value at a rotation of about 0.35 to 0.45, after which fore it may be concluded that for most practical pur-
the resistance remains constant. The ultimate torsional poses and for any caisson a unique normalised failure
resistances predicted by the finite element analyses locus in the axial-lateral loading plane can be found

292

Copyright 2005 Taylor & Francis Group plc, London, UK


20
z /L = 1.0 1.0
z /L = 0.5
16 z /L = 0.0 0.8
z / L = -0.6
+V

Vu / Vmax
-
Vu / (A.su)

12 D z / L = -1.0 0.6
T
8 - z 0.4
+V
H
L
4 0.2

0.0
0
0.0 0.2 0.4 0.6 0.8 1.0
0 2 4 6 8 10 12
Tu / Tmax
Hu / (D . L . Su)

Figure 6. Failure locus in the axial-lateral loading plane Figure 8. Non-dimensional failure locus in the axial-
for caissons with L/D  2. torsional loading plane.

1.0
1
0.8
0.8
Vu / Vmax

0.6 D z
z/L= 1.0, L/D=2
Hu / Hmax

z/L= 0.5, L/D=2 0.6 T


- z H
0.4 +V L z / L= 0.0
H z/L= 0.0, L/D=2
L 0.4 z / L=-0.6
z/L=-0.6, L/D=2
0.2 z / L=-1.0
z/L=-1.0, L/D=2
0.2
z/L= 0.0, L/D=2
0.0
0.0 0.2 0.4 0.6 0.8 1.0 0
Hu / Hmax 0 0.2 0.4 0.6 0.8 1
Tu / Tmax

Figure 7. Non-dimensional failure locus in the axial-


lateral loading plane. Figure 9. Non-dimensional failure locus in lateral-
torsional loading plane.

which is representative of all failure loci for caissons


with L/D  2. Furthermore, this failure locus may 4.3 Lateral-torsional failure plane
also be applicable to caissons with other aspect ratios, The ultimate response of the caisson foundation, with
although this point warrants further investigation. L/D  2, to different combinations of lateral and tor-
sional deformations is presented in Figure 9 as failure
loci for 3 different points of lateral load application.
4.2 Axial-torsional failure plane The ultimate lateral and torsional forces are nor-
A normalised failure locus for the caisson with malised by their maximum values, Hmax and Tmax,
L/D  2 subjected to combinations of axial and tor- which can be obtained for pure torsional and pure lat-
sional forces is presented in Figure 8. In this figure, eral loading at different points of load application.
the axial and torsional forces are normalised by Figure 9 shows that as the torsional force increases to
their maximum values, Vmax and Tmax, obtained under its maximum value, the lateral resistance of the foun-
either pure axial or pure torsional loading. It is noted dation decreases to about 60% of its maximum value.
that these are small strain results and in reality the For lateral loads lower than about 0.6 Hmax, torsional
ultimate axial capacity may increase slightly at larger displacements govern the failure mechanism of the
vertical displacements. caisson foundation.
The failure locus presented in Figure 8 shows that
torsional forces have a very significant effect on the 4.4 Axial-lateral-torsional failure surface
axial capacity of the caisson. When much of the
shearing strength around and under the caisson is A series of finite element analyses was performed on
mobilized by torsion, the axial capacity of the foun- the typical caisson with L/D  2 in order to obtain a
dation reduces significantly. For axial loads lower general failure surface for the foundation. Various
than about 0.6 Vmax, torsional displacements govern combinations of axial, lateral and torsional displace-
the failure mechanism of the caisson foundation. ments were applied to the foundation. In all the analyses

293

Copyright 2005 Taylor & Francis Group plc, London, UK


the failure surface obtained for z/L  0, and shown
1.0
in Figure 10, can be used for other values of z/L.
0.8 Furthermore, based on the results of the limited study
Tu 0.6 performed here on the effects of the aspect ratio of the
Tmax 0.4
caisson foundation, it was shown that the non-
0.2
0.0
dimensional failure planes of the caisson with
-0.8 0.8 L/D  4 are very similar to those of the caisson with
L/D  2. Therefore, the shapes of the failure envelopes
-0.4 0.4
0.0 0.0
Vu/Vmax
0.4 -0.4 Hu/Hmax for foundations with various aspect ratios are proba-
0.8 -0.8 bly very similar to the non-dimensional 3-D failure
envelope shown in Figure 10.
Figure 10. Three-dimensional failure envelope in non-
dimensional load space.
5 CONCLUSIONS

1.0 Results of a numerical study of the capacity of caisson


Tu/Tmax
0.6 0.5 0.0
foundations under axial, lateral, and torsional loads
0.8 0.7
0.8
have been presented. Failure envelopes were presented
0.6 0.9 in non-dimensional form and it has been shown that
0.95
a single non-dimensional failure envelope can be
0.4
0.99
obtained for the foundation in the different loading
0.2 planes. The limited study performed here did not indi-
Vu 1.0 cate significant changes in the shape of the failure
0.0
Vmax envelopes for caisson foundations with different aspect
-0.2 ratios. It can be tentatively concluded that the non-
dimensional failure surface obtained in this study may
-0.4
be useful for caissons with various aspect ratios. Having
-0.6 the maximum resistances of a caisson foundation
against purely axial, lateral and torsional loading, the
-0.8
resistance of the foundation against any combination of
-1.0 the axial, lateral and torsional loading can be obtained
-1.0 -0.8 -0.6 -0.4 -0.2 0.0 0.2 0.4 0.6 0.8 1.0 using the general failure envelope presented here.
Hu/Hmax

Figure 11. Failure loci in the non-dimensional VHT space. ACKNOWLEDGEMENT

The research described in this paper was conducted


the horizontal load was applied at the ground level, as part of the work of the Special Research Centre
i.e. z/L  0. The three-dimensional failure surface for Offshore Foundation Systems, established and
was constructed based on a triangulation scheme of supported under the Australian Research Councils
linear interpolation between computed data points. Research Centres Program.
A three-dimensional image of the failure envelope
for the caisson in the axial-lateral-torsional loading
space is presented in Figure 10. Only half of the fail-
REFERENCES
ure envelope is shown as the uplift capacity problem
of a caisson in clay soils deforming under undrained Andersen, K. H., Dyvik, R., Schoder, K., Hansteen, O. E., &
conditions is merely a reverse bearing capacity prob- Bysveen, S. 1993. Field test of anchors in clay, II:
lem. Representation of the failure loci in the VHT Prediction and interpretation. Journal of Geotechnical
loading space is shown in Figure 11, where contours Engineering. ASCE, 119, 5321549.
of equal torsional load are presented. In these figures Bransby, M. F. & Randolph, M. F. 1997. Finite element mod-
all loads are normalised by their maximum values, elling of skirted strip footings subjected to combined
Vmax, Hmax and Tmax, obtained under purely axial, lat- loadings. Proc. 7th Int. Offshore and Polar Engineering
Conference. Honolulu, USA, 791796.
eral and torsional loading.
Bransby, M. F. & Randolph, M. F. 1998. Combined loading
As shown in the previous sections, a single non- of skirted foundations. Gotechnique. 48(5), 637655.
dimensional failure envelope in the HT or VT loading Bransby, M. F. 2001. Finite element analysis of jacket struc-
planes can be used for the foundation regardless of the tures with bucket foundations. Proc. 10th Int. Conference
point of the application of the lateral load. Therefore, on Computer Methods and Advances in Geomechanics.

294

Copyright 2005 Taylor & Francis Group plc, London, UK


Carter J. P. & Balaam N. P. 1995. AFENA users manual. of suction caissons under lateral loads. Computers and
Version 5, Centre for Geotechnical Research, The Geotechnics. 24, 89107.
University of Sydney, Australia. Taiebat, H. A. & Carter, J. P. 2001. A semianalytical finite
Deng, W. & Carter, J. P. 1999. Analysis of suction caissons in element method for three-dimensional consolidation
uniform soils subjected to inclined uplift loading. Report analysis. Computer and Geotechnics. 28, 5578.
No. R798, Department of Civil Engineering, The Taiebat, H. A. & Cater. 2004. Effects of torsion on caisson
University of Sydney, Australia. capacity in clay. Proc. 9th Australia New Zealand
Deng, W., Carter, J. P. & Taiebat, H. A. 2001. Prediction of Conference on Geomechanics. 1, 130136.
the Lateral Capacity of Suction Caissons, Keynote Vesic, A. S. 1975. Bearing capacity of shallow foundations,
Lecture, Proc.10th Int. Conference on Computer Methods Foundation Engineering Handbook, Eds Winterkorn &
and Advances in Geomechanics, 1, 3338. Fang, Van Nostrand Reinhold, New York, 121147.
Deng, W. & Carter, J. P. 2002 A theoretical study of the ver- Zdravkovic, L., Potts, D. M. & Jardine, R. J. 2001. A para-
tical uplift capacity of suction caissons. Int. J. of Offshore metric study of the pull-out capacity of bucket founda-
and Polar Engineering., 12(2), 8997. tions in soft clay. Gotechnique. 51(1), 5567.
Sukumaran, B., McCarron, W. O., Jeanjean, P. & Abouseeda, H. Zienkiewicz, O. C. & Taylor, R. L. 1989. The Finite Element
1999. Efficient finite element techniques for limit analysis Method. 4th Edition, McGraw-Hill, New York.

295

Copyright 2005 Taylor & Francis Group plc, London, UK


Parametric finite element analyses of suction anchors

L. Zdravkovic & D.M. Potts


Imperial College London, UK

ABSTRACT: This paper presents the results from a series of three-dimensional (3D) finite element paramet-
ric studies on suction anchors in clay. The objective is to investigate the effects of a number of factors on the
pull out capacity of suction anchors, namely the position of the attachment point, inclination of the pull out
force and anchor geometry. A typical soft clay was considered as the seabed soil and it was simulated using a
form of the modified Cam clay model. The results identify the optimum position of the anchoring point, as well
as the optimum inclination of the pull out force.

1 INTRODUCTION 2 FINITE ELEMENT ANALYSIS

Suction anchors and caissons are novel offshore foun- For the analyses of suction anchors use was made
dations used for oil and gas exploration platforms in of the Fourier Series Aided Finite Element Method
deep water. These are hollow cylindrical structures, (FSAFEM), Potts & Zdravkovic (1999). This method
made either of concrete or steel, consisting of a thin of analysis can be used for a special class of problems
wall (in relation to the diameter), that penetrates into that have axi-symmetric geometry, but whose bound-
the seabed, and a top cap. Because they are thin ary conditions and/or material properties are not axi-
walled structures, they are also referred to as skirted symmetric. Its advantage is the reduction of computer
foundations. run times, when compared to conventional 3D analysis,
Suction caissons are normally used for fixed plat- as well as the finite element mesh that remains two-
forms (e.g. Christophersen et al. 1992, Tjelta 1995) dimensional (in the r-z plane, as shown in Figure 1).
where the load from the structure is transferred to the Any variation in the out of plane direction, , is
foundation via the top cap. Suction anchors, on the expressed through a number of Fourier series har-
other hand, are employed for mooring floating vessels monics. Boundary conditions are such that the bottom
(e.g. Colliat et al. 1996) and the load from the struc-
ture is transferred to the foundation somewhere along
D/2
the height of the outer wall of the cylinder, depending r
on the position of the mooring cable. In either case, T
during the installation the cylinder initially penetrates
the seabed under its own weight. The remaining pene- z
tration is then achieved by pumping water from inside
the cylinder, thus creating a differential water pres-
sure on the outside of the top cap that pushes the
cylinder to its desired embedment. One of the main
axis of geometric symmetry
design considerations for these foundations is their
short-term pull out capacity.
Zdravkovic et al. (2001) performed an extensive
parametric study, using finite element analyses, on
the pull out capacity of suction caissons. This paper
presents the results of a similar study on suction
anchors, that investigated the effects of the position of
the attachment point, inclination of the pull out force
and anchor geometry on its pull out capacity. Figure 1. Typical finite element mesh for anchor analysis.

297

Copyright 2005 Taylor & Francis Group plc, London, UK


boundary of the mesh is fixed (i.e. zero displacements) clays in triaxial compression. The undrained conditions
in all three coordinate directions, while the right hand in the ground were simulated by prescribing a high
side vertical boundary is fixed in the r-direction. The bulk compressibility of the pore fluid.
left hand side vertical boundary, although an axis of The suction anchor was simulated as a linear elastic
geometric symmetry, is not an axis of deformational concrete material, with Youngs modulus E  30 GPa
symmetry and therefore remains unrestrained. Eight and Poissons ratio   0.15. This implies that the
nodded quadrilateral elements with reduced integra- concrete has sufficient strength and reinforcement to
tion were used. All analyses were performed with the sustain the applied loads. The installation of the anchor
Imperial College Finite Element Program (ICFEP) that was not modelled in the analyses; it was assumed to
employs a modified Newton-Raphson scheme with an be wished in place.
error controlled sub-stepping stress point algorithm The interface between the top cap and the soil under-
as its non-linear solver (Potts & Zdravkovic, 1999). neath was modelled by introducing zero thickness inter-
face elements (Day & Potts 1994). These elements
represent a layer of water trapped at the interface,
2.1 Geometry and loading conditions
which can sustain a tensile pressure of 100 kPa and
The reference geometry of the suction anchor in this cavitates if this pressure is exceeded. In all the analy-
study has a diameter D of 3 m and a skirt length L of ses this water did not cavitate because the anchor was
9 m, i.e. a D/L ratio of 1/3. To investigate the effects assumed to be embedded 1 km below sea level. There-
of anchor geometry, the skirt length was varied first, fore initially this water had a compressive pressure
keeping the same diameter (i.e. L  6 m and 15 m, equivalent to 1 km head of water.
corresponding to D/L  1/2 and 1/5 respectively). In Full adhesion was assumed between the anchor
addition, the diameter was then varied, keeping the wall and the soil.
same skirt length (i.e. D  6 m, with D/L  1/1.5).
To investigate the optimum attachment point, the
2.3 Load application
pull out force was applied at four positions along the
height of the outer wall: at the top, one third down, Ideally, the pull out force should be applied as a con-
half way down and at the bottom of the anchor wall. centrated force at a certain point on the side of the
To investigate the effects of the inclination of the anchor. However, in the FSAFEM all the variables,
pull out force, this force was applied at angle  of 0, including the applied forces, have to be expressed in
30, 45 and 90 to the horizontal at each of the four terms of Fourier series. It is therefore not straight for-
positions. ward to apply a discrete force that has a discontinuous
variation in the  direction. Consequently, the force
was applied at a node in the mesh corresponding to
2.2 Material properties
the required position, but was assumed to act as a line
The seabed soil was modelled as a homogeneous, load in the  direction, from   !20 to 20
lightly overconsolidated, soft clay with an overconsol- (  40, see inset in Figure 2).
idation ratio, OCR, of 1.1. A form of the modified The ideal distribution of the applied line load is
Cam clay constitutive model (Potts & Zdravkovic 1999) noted in Figure 2 as Applied and represents a step
was employed to simulate its behaviour, with model function. The magnitude of the pull out force is then
parameters as summarized in Table 1. These are con- the resultant of the applied line load (i.e. area under
sistent with the parameters used in the suction caisson
study by Zdravkovic et al. (2001) and produce an
undrained strength profile that varies linearly with 150
Applied and fitted load (kN/m)

Applied
depth, giving Su/vN  0.33 that is typical of soft Fitted

Table 1. Parameters for the modified Cam clay model. 100


=180 =20 =0
Parameter Description Value
20 T
50
v1 Specific volume for an isotropically 3.1
consolidated sample at p  1 kPa
TCN Critical state angle of 32
shearing resistance 0
 Slope of the VCL in v-lnp space 0.2
0 20 40 60 80 100 120 140 160 180
 Slope of the swelling line 0.03
in v-lnp space Angular coordinate, (deg)
 Poissons ratio 0.3
Figure 2. Harmonic fitting of the applied force.

298

Copyright 2005 Taylor & Francis Group plc, London, UK


the applied curve). As noted above, in the analysis the half way down the wall. This is in agreement with
line load must be expressed as a Fourier series. Conse- both the theoretical analyses of Andersen & Jostad
quently, harmonic coefficients must be specified (1999) and the model tests of Keaveny et al. (1994).
that provide a fit to the Applied line load. This was The former performed their limit equilibrium ana-
achieved using the step wise method described in lyses assuming the failure modes for an anchor loaded
Potts & Zdravkovic (1999). The degree of fit depends at 45 at the top and at mid-height of the anchor wall
on the number of harmonics used in the analysis and respectively. The latter performed large-scale model
a typical result is shown as the dotted line labelled tests of anchors at the soft clay Lysaker site near Oslo
Fitted in Figure 2. A more accurate fit requires more in Norway. The tests involved both static and cyclic
harmonics. However, the more harmonics used the loading of anchors with the load point at the top and
larger the computational resources required. A small at mid-height and a load inclination of 10 to the hori-
parametric study for the analyses performed in this zontal. Both studies indicated that the load capacity
paper showed that 10 harmonics (taking account of may be increased by a factor of 2 when using the
the   0 symmetry) were enough to provide suffi- mid-height load attachment instead of that at the top.
cient accuracy of the solution. Clearly, the smaller the Comparison of the top loading and fi way loading
angle  over which the line load acts, the larger the curves in Figure 3 confirms this conclusion and indi-
number of harmonics required. cates that the load increase factor is between 2 and 3.
However, there are studies that indicate that the
optimum attachment point maybe positioned slightly
3 RESULTS below the mid-height of an anchor. For example,
Allersma et al. (1999) performed a series of cen-
3.1 Load position and inclination trifuge tests and finite element analyses in which they
The first set of analyses was performed on the refer- varied parameters similar to those presented in this
ence geometry (D  3 m, L  9 m) of the suction paper (e.g. D/L ratio, attachment point, load inclin-
anchor. The pull out force, T, was applied at four loca- ation). For a slightly stubbier geometry of suction
tions on the outside wall: top, 3 m below (L/3), 4.5 m anchor (i.e. D/L  1/1.7, compared to D/L  1/3 pre-
below (L/2) and bottom. At each location the inclination sented here) and load inclination at 15 to the hori-
of the force was varied from 0 to 90 to the horizontal. zontal, they show that the optimum attachment point
Figure 3 summarises the results from all analyses and is at about 0.4 L from the bottom of the anchor. This
shows the variation of the maximum pull out force corresponds to about 2/3rds from the top of the
with its inclination and the anchoring position. anchor according to the nomenclature in this paper.
The first conclusion that is apparent from Figure 3 Clearly, the indication from the current and previ-
is that the optimum position of the anchor point, for ous studies is that the optimum position of the attach-
all load inclinations smaller than 90, appears to be ment point depends on the geometry of the anchor.
The optimum position of the anchor point as
deduced from the analyses presented in this paper can
3000 be explained by considering the mobilised mech-
anisms of soil deformation. These are shown in Figure 4
for the 45 inclination of the pull out force and attach-
Ultimate pull out force T (kN)

2500
ments at the top, bottom and mid-height of the anchor
wall. The arrows represent the vectors of incremental
2000 displacements of the soil mass at the last stable incre-
ment of each analysis (i.e. just before failure) and are
shown for the part of the mesh in the vicinity of the
1500
anchor. It is clear from Figure 4b that the mid-height
attachment involves predominantly a translational
top
1000 mode of failure, while both the top and bottom attach-
1/3
1/2 ments result in rotational failure and hence smaller
Top loading pull out anchor capacity.
500 1/3 loading
1/2 loading
bottom With regards to the optimum inclination of the pull
Bottom loading out force, Figure 3 shows that, while the vertical load-
0 ing (i.e.   90) gives maximum capacity for top
0 10 20 30 40 50 60 70 80 90 and bottom attachment points, an inclination of 45
Loading angle, (deg) appears to be the optimum for attachment points
(with respect to horizontal) between a third and half way down the wall.
However, the aforementioned study of Allersma
Figure 3. Ultimate pull out force for different load conditions. et al. (1999), both from centrifuge tests and finite

299

Copyright 2005 Taylor & Francis Group plc, London, UK


(a)

Figure 5. Typical failure mode for   90 loading,


independent of the attachment point.

element analyses, shows that for their optimum


attachment point (2/3rds down the wall) and the same
geometry of the anchor as described above, the pull
out capacity increases as the load inclination reduces
15.0m

from 30 to 0 to the horizontal (this was the max-


imum inclination range that they considered). This
trend is similar to that shown in Figure 3 for the bot-
tom loading condition (i.e. below the anchor mid-
height), for the same variation of the loading angle
18.0m from 0 to 30.
(b)
Another interesting result from these analyses is
the independence of the attachment position when the
anchor is pulled out vertically (i.e.   90). Figure 3
indicates the same ultimate pull out force for this load
inclination. Examination of the failure modes for all
attachment points revealed that they are almost the
same and are as shown in Figure 5. The reason for this
is the length of the anchor that, due to the constraints
provided by the surrounding soil, only allows a limited
rotation of the anchor, regardless of its attachment
point. The failure modes would probably be different
for shorter anchors and hence result in different ulti-
mate forces.
It is important to note that tension cracks along the
side of the anchor did not occur in any of the analyses.
This is probably not surprising considering that the
soil conditions analysed were those of near normal con-
solidation, with undrained strength varying linearly
(c) with depth.

3.2 Changes in anchor geometry


Figure 4. Mechanisms of failure for different attachment
points and   45: (a) top, (b) middle and (c) bottom. In the next set of analyses the length of the anchor is
varied (i.e. L  6 m and 15 m), keeping the same

300

Copyright 2005 Taylor & Francis Group plc, London, UK


6000 3.5

normalised by the weight of plug


Ultimate pull out force, T (kN)

5000
3

Ultimate pull out force, T


D=3m, L=6m
4000 D=3m, L=9m
D=3m, L=15m 2.5
3000
2
2000

1.5
1000 D=3m, L=6m
D=3m, L=9m
D=3m, L=15m
0 1
0 10 20 30 40 50 60 70 80 90 0 10 20 30 40 50 60 70 80 90
Loading angle, (deg) Loading angle, (deg)
(with respect to horizontal) (with respect to horizontal)

Figure 6. Ultimate pull out force for different anchor lengths. Figure 7. Efficiency of anchors of different length.

diameter, D, of 3 m. Only one attachment point is 6000


considered and that is the point at anchor mid-height.
The same load inclinations as before are applied and
Ultimate pull out force, T (kN)

5000
the results are summarised in Figure 6.
It is clear from Figure 6 that the increase in anchor
length results in an increase of the ultimate pull out 4000
force. However, this increase is not proportional. For
the increase of anchor length from 6 m to 9 m (i.e. 3000
50%) and from 9 m to 15 m (i.e. 66%) the increase of
pull out force is 2 (i.e. 100%) and 2.5 (i.e. 150%)
times respectively. 2000
This aspect can be further considered if the ultim-
D=3m, L=9m
ate pull out forces from each geometry are nor- 1000 D=6m, L=9m
malised by the respective weight of the soil plug
inside the anchor, as shown in Figure 7. A ratio of 1
0
indicates that the contribution to the pull out capacity
is generated mainly by the weight of the soil plug, 0 10 20 30 40 50 60 70 80 90
whereas ratios greater than 1 indicate additional con- Loading angle, (deg)
tributions from the friction between the soil and the (with respect to horizontal)
anchor wall and reverse end bearing (which increase
the efficiency of the anchor). A short and stubby Figure 8. Ultimate pull out forces for different anchor
anchor is found to be a less efficient solution when diameters.
compared to a slender anchor (Sparrevik 1998, Olson
2000). Results in Figure 7 confirm this, as the nor-
malised pull out force increases with anchor length Clearly, the increase in anchor diameter results in an
for any load inclination. increase of the ultimate pull out force for any inclination
In addition to these changes in geometry, another of loading. From the presented analyses, this increase
analysis was performed where the anchor length was appears to be proportional, as doubling the diameter
kept constant (i.e. L  9 m), while the anchor diam- results in approximately double the pull out force.
eter D was increased to 6 m. Again, only one attach- In terms of anchor efficiency, a conclusion similar
ment point is considered, at mid-height of the anchor to that from the previous set of analyses can be drawn
wall, and the same load inclinations as in the previous from Figure 9. The shorter anchor of D  6 m and
analyses. The ultimate pull out forces are compared in L  9 m is less efficient than the more slender anchor
Figure 8. of D  3 m and L  9 m.

301

Copyright 2005 Taylor & Francis Group plc, London, UK


3 For short and stubby anchors the pull out capacity
D=3m, L=9m
is dominated by the weight of the soil plug inside
the anchor. Long and slender anchors benefit from
normalised by the weight of plug

D=6m, L=9m
the additional components of friction between the
Ultimate pull out force, T

2.5 soil and the anchor wall and reverse end bearing to
the pull out capacity.

2
REFERENCES

Allersma H.G.B., Kirstein A.A., Brinkgreve R.B.J. &


1.5
Simon T. 1999. Centrifuge and numerical modelling of
horizontally loaded suction piles. Proc. 9th Int. Offshore
and Polar Eng. Conf., Brest, France. ISOPE. 1: 711717.
Andersen, K.H. & Jostad, H.P. 1999. Foundation design of
1 skirted foundations and anchors in clay. Offshore Techn.
0 10 20 30 40 50 60 70 80 90 Conf. Houston, 1: 383392.
Loading angle, (deg) Christophersen, H.P., Bysveen, S. & Stove, O.J. 1992.
(with respect to horizontal) Innovative foundation systems selected for the Snorre
field development. Behaviour of Offshore Structures,
Figure 9. Efficiency of anchors of different diameter. M.H. Patel & R. Gibbins (eds), Proc. 6th Int. Conf.
London. London: BPP Technical Services Ltd. 1: 8194.
Colliat, J-L., Boissard, P., Gramet, J-C. & Sparrevik, P. 1996.
4 CONCLUSIONS Design and installation of suction anchor piles at a soft
clay site in the Gulf of Guinea. Offshore Techn. Conf.,
This paper presents the results from a series of finite Houston, 3: 325337.
element analyses aimed at investigating the pull out Day, R.A. & Potts, D.M. 1994. Zero thickness interface
elements. Int. J. Num. Anal. Methods in Geomech. 18:
capacity of suction anchors in soft clay. The following 689708.
main conclusions can be drawn from this study: Keaveny, J.M., Hansen, S.B., Madshus, C. & Dyvik, R.
For the anchor geometry and soil conditions mod- 1994. Horizontal capacity of large-scale model anchors.
elled in this study, the optimum attachment point 13th ICSMFE, New Delhi, 2: 677680.
Olson, R.E. 2000. Deepwater anchors. Offshore Techn.
of a mooring cable is identified to be at mid-height Research Centre report. Houston, Texas.
of the anchor wall. This is in agreement with some Potts, D.M. & Zdravkovic, L. 1999. Finite element analysis
earlier limit equilibrium analyses and large scale in geotechnical engineering: Theory. London: Thomas
model tests of suction anchors found in the litera- Telford.
ture. However, it is recognised that the optimum Sparrevik, P. 1998. Suction anchor piles. State of the art
position of attachment point depends on anchor report. NGI, NR. 199.
geometry. Tjelta, T-I. 1995. Geotechnical experience from the installa-
For this attachment point (i.e. at mid-height), the tion of the Europipe jacket with bucket foundations.
optimum inclination of the mooring cable, that Offshore Techn. Conf. Houston, 2: 897908.
Zdravkovic, L., Potts, D.M. & Jardine, R.J. 2001. A para-
results in the maximum pull out force, is found to metric study of the pull-out capacity of bucket founda-
be approximately 45 to the horizontal. tions in soft clay. Geotechnique 51(1): 5567.
For long and slender anchors the pull out capacity
on vertical loading (i.e.   90) is independent of
the position of the attachment point.

302

Copyright 2005 Taylor & Francis Group plc, London, UK


Centrifuge modelling of suction piles in clay

R.D. Raines & O. Ugaz


ExxonMobil Upstream Research Company, Houston, USA

Jacques Garnier
Laboratoire Central des Ponts et Chaussees, Bouguenais, France

ABSTRACT: Most of the suction pile design methodologies in use today are analytically based because
experimental data are limited. In light of this circumstance, an experimental program was undertaken to develop
a database for the purpose of calibrating analytical design methods. In this program the behavior of suction piles
subjected to monotonic, sustained and various storm loading conditions was investigated by model tests in a
large geotechnical centrifuge. The tests were carried-out in tubs of normally consolidated clay that were pre-
pared from Speswhite power, which is a low plasticity kaolinite mineral.
The centrifuge program was designed to evaluate the fundamental response of suction piles to environmen-
tal loads and to certain pile installation issues, upon which there is no consensus for design. The primary issue
addressed in this paper is should the capacity of an anchor pile that is installed by a combination of self-weight
and suction be reduced in the interval where suction is employed. Results are also presented from a test to evalu-
ate the influence of a deep-seated sand layer on pile penetration. The response of piles to environmental loads
are not included due to space limitations. With the possible exception of the test with the deep-seated sand layer,
which is a single test, the data from the test program are considered robust because of the number of tests per-
formed and the high degree of consistency of the results.
In conjunction with the centrifuge tests, a laboratory test program was performed on the Speswhite clay to
obtain basic soil parameters for comparing the test results with finite element analysis (FEA) predictions. The
FEA results are in very good agreement with the experimental results, which lends significant credibility to
both the centrifuge scaling relations and the FEA pile-soil model. This outcome also provides confidence in
using these results to update ExxonMobil design practice and to validate other numerical models.

1 INTRODUCTION Centrifuge testing was selected for investigating suc-


tion pile behavior because it is considered necessary
Currently, there appears to be no consensus or stand- to correctly model inertial forces within the soil sam-
ardization on the design of suction piles for floating ple to obtain credible results from a model test.
production and drilling vessels. Oftentimes there may Application of the centrifuge to investigate offshore
be agreement in principle and perhaps even philosoph- issues, as well as background on principles of cen-
ical agreement, but there remains disagreement on trifuge testing and its limitations, are discussed in
how to apply the philosophy to design practice. This papers by Garnier, et al (1995) or Murff (1996). This
lack of consistency can potentially lead to unconserv- paper focuses mainly on the centrifuge tests and ana-
ative or excessively overly conservative designs, nei- lytical results. Details on the testing apparatus, testing
ther of which constitutes best practice. Consequently, procedures, and sample preparation are contained in
in the absence of a standard design practice, Exxon- the paper Physical Modeling of Suction Piles in
Mobil undertook a comprehensive suction pile research Clay by Raines, et al (2004).
program that included centrifuge testing, large-scale
tests in the Gulf of Mexico (GOM) and development
2 BACKGROUND TESTING INFORMATION
of advanced analytical models to predict suction pile
performance.
2.1 Scaling principles
This paper presents results of the centrifuge program
and analytical predictions from ABAQUS models that Testing in this program was carried out at 100-g on
were developed in conjunction with this program. 1/100th reduced scale model piles. To translate these

303

Copyright 2005 Taylor & Francis Group plc, London, UK


measurements to prototype scale the appropriate scal- 2.3 Model piles and instrumentation
ing laws must be followed. For example, measure-
The model piles are fabricated from steel and have
ments must be multiplied by 100 for dimensions and
dimensions of 310 mm  64 mm  0.5 mm. A verti-
displacements, by 1002 for applied loads, and by 1002
cal shaft is attached at the top of the pile to maintain
for diffusion (pore pressure dissipation). In accord-
pile verticality during initial penetration. The pile is
ance with the scaling laws, strain, stress, and pore
reinforced by an internal cruciform at padeye level,
pressure scale 1:1 between model and prototype. The
which is near the 2/3 point of the pile.
data presented in the paper corresponds to the proto-
The model piles and soil are instrumented and
type values, unless noted.
selected measurements are recorded continuously
during pile installation and load testing. These
include the following:
2.2 Soil characteristics
Pore pressures in the free-field and at locations
The clay bed is formed by consolidating a soil slurry inside and outside the pile.
consisting of Speswhite power and water, which has a Soil settlement.
water content of approximately 90%. The mixture is Pile displacements and rotation.
placed into cylindrical containers and consolidated in Plug displacements and water levels.
five successive layers by hydraulic press. The con- Load transferred to pile during installation.
tainer is then placed in the centrifuge and water is Load applied to pile during testing.
added to a height of about 500 mm above the soil bed.
Final consolidation is achieved in the centrifuge under
100-g artificial gravity for a period of 4 to 6 hours. 3 SUCTION VERSUS SELF-WEIGHT
At the end of centrifuge consolidation several in- INSTALLATION
flight CPT are performed at different locations to
monitor the shear strength of the soil. Several other It has been suggested by Andersen et al (2002) that
CPT are carried out in-flight throughout the course of installation by suction reduces the external skin fric-
testing, including the end of testing, to record any tion of a pile because most of the soil near the pile tip
changes in soil shear strength. A typical shear strength displaces to the inside of the pile, instead of moving
profile is shown in Figure 1. to the outside. As a result, the octahedral shear stresses
near the wall of the pile are reduced, which ultimately
leads to a reduction in skin friction. In order to account
for this effect, Andersen suggests a reduction in skin
CPT Tip Resistance, MPa
friction of 35% in the interval where suction was used
0.0 0.2 0.4 0.6 0.8 to install the pile. To investigate this claim, a series of
tests were conducted to test the effects of suction
-1.00 installation on skin friction.
-3.00 In developing the protocol for these tests, a consen-
-5.00 sus was reached that the piles should be loaded from
-7.00 the top in tension instead of along the shaft at an incline.
-9.00 Loading the piles in tension reduces the complica-
-11.00 tions inherent with inclined loading and permits test-
Penetration, m

-13.00 ing two piles in the same tub. Thus two sets of pile
-15.00 installation and loading devices were fabricated so
-17.00 that testing could occur in the same bucket without
-19.00 stopping the centrifuge.
-21.00 Suction installation. In this series of tests, the pile
-23.00 weight is counter-balanced so it does not contribute to
-25.00 installing the pile, but only serves to push the pile a few
-27.00 centimeters into the soil to form a seal, so suction can
-29.00
be applied. After the pile is seated, suction is applied by
pumping water from inside the pile, thereby using the
-31.00
differential water head to push the pile into the soil.
-33.00
Self-weight installation. Self-weight installation is
0 10 20 30 40 achieved with a two step process. First, weight is
Undrained Shear Strength, kPa added to the pile head to overcome the soil resistance.
Optimizing this weight is a non-trivial task, since too
Figure 1. Typical soil shear strength profile. little weight leaves the pile at less than full penetration

304

Copyright 2005 Taylor & Francis Group plc, London, UK


and too much weight causes the pile to penetrate vertical displacement (4 mm) and at the same rate
below the mudline. Second, it is necessary to vent the of loading (0.25 mm/sec).
water inside the pile in order to balance the internal Variations in the installation procedure were incor-
and external water pressure so the pile moves porated between buckets, however, for the suction-
unabated into the soil. installed pile. These consisted of: (1) using both
Test procedure. Typically the self-weight pile was self-weight and suction installation techniques on
installed and tested first, in order to establish the tar- some of the suction installed piles; and (2) varying
get penetration for the suction pile. Cone penetrome- the amount of suction pressure applied to the suction-
ter tests (CPT) were run in the free-field before and installed pile. These variations were incorporated to
after load testing each pile in order to monitor soil aid in generalizing the results for design purposes.
shear strength. In planning the test program the object- The details of the installation procedure for the
ive of minimizing the interaction between the piles suction-installed pile follow:
was taken into account in establishing both the pile
spacing and the test protocol. Suction installation of Test 1 Combination of self-weight (70% total
the second pile followed immediately after load test- penetration) and suction penetration.
ing and CPT in order to minimize any potential for Test 2 Differential pressure ramped up propor-
uncertainties with regard to soil shear strength. tionally to 250 kPa to evaluate effect of applying
Considerable efforts were made to be sure consis- significant suction pressure relative to soil shear
tency existed during installation and load testing of strength profile.
the two piles within the same bucket. Consequently, Test 3 Differential pressure applied immediately
emphasis was placed on achieving comparable instal- after seating but limited to maximum suction
lation times for the piles. This is critical since pore pressure required for installation, about 50 kPa.
pressure dissipation scales at 1002 in prototype dimen- Test 4 Repeat of Test 1, but with less self-weight
sions. For the same reasons, load testing of each pile penetration (50%).
occurred at essentially the same time after installa- Test results. The results of the tests are plotted as load
tion, which was typically 30 to 35 days in prototype versus displacement in Figure 2. These results repre-
time. Finally the piles were load tested to the same sent the gross ultimate capacity of the pile, since a

200 180
180 160
160 140
140 TUB 1 120 TUB 3
120
Gross (194)
100 Gross (164)
100 Self Wt
Self Wt 80
80 Net (143) Net (120)
60 60
Gross (151) Gross (133)
40 40 Suction
Suction
20 +Self-wt Net (143) 20 Net (123)
0 0
Load, daN

2 4 6 8 12 14 16 18 2 4 6 8 12 14 16 18
0 10 20 0 10 20

180 180
160 160
140 140
120 TUB 2 120 TUB 4
100 100 Gross (155)
Gross (169)
80 Self Wt 80 Self Wt
Net (117)
Net (123) 60
60
Gross (146)
40 Gross (130) 40 Suction*
Suction
20 Net (124)
20 +Self-wt Net (140)
0 0

2 4 6 8 12 14 16 18 2 4 6 8 12 14 16 18
0 10 20 0 10 20
Displacement, mm

Figure 2. Results of suction versus self-weight tests model scale.

305

Copyright 2005 Taylor & Francis Group plc, London, UK


20
Table 1. Speswhite laboratory test results.
Hand Calculation
18
Speswhite kaolin parameters
16
Vertical Load, MN

14 Tub 1 - Logarithmic Bulk Modulus 0.01


12 Tub 2 - Submerged Unit Weight 6.3 kN/m3
10 Tub 3 - Shear Modulus at 45% at Shear 316 Sutc
8 Tub 4 - Strength
ABAQUS \\\ Triaxial Compression Undrained 1.2 kPa/m line
6
Shear Strength (Sutc)
4
Triaxial Extension Undrained Shear 0.96 kPa/m line
2 Strength (Stc
u)
0

0.2 0.4 0.6 0.8 1.2 1.4 1.6 1.8


0 1.0 2.0 are shown in Table 1. The results of the ABAQUS
Vertical Displacement, m analyses, shown in Figure 3, not only correctly pre-
dict the measured capacity, but the calculated load-
Figure 3. Measured versus predicted results. displacement curves agree with the measured responses
for both elastic and inelastic displacements.
Conversely, as shown in Figure 3 the hand calcula-
portion of the applied load is used to overcome the tion overestimates the measured capacity by about
weight of the pile. The net ultimate capacity, the gross 25%. This is a concern because the final suction pile
load minus the pile weight, is a better measure for design penetration is sometimes calculated on this
comparison of results and is also shown in the figure. basis, when the vertical capacity governs the design.
By inspection, it is apparent that the piles in a given Consequently, a more detailed evaluation of the data
bucket have essentially the same net ultimate capacity was made to gain insight into this apparent anomaly.
even though they were installed by different processes. It was decided to investigate this discrepancy with
Likewise, the load-displacement curves are consistent ABAQUS to see if anything unusual occurs to the skin
between tests and no unusual behaviour was observed friction and reverse end bearing (REB) in the model
out to displacements where testing was halted. (In tub as the load is being applied. To do this, it is necessary
4 the suction pile took significantly longer to install, to isolate the skin friction and the REB.
which likely accounts for the differences in measured The skin friction can be estimated from the
capacity.) ABAQUS analysis by extrapolating the element shear
In support of these findings are soil strength data stresses to the nodes, and then integrating along the
gathered after the piles were excavated. First, upon shear surface of the neighbouring soil elements. Then
extraction of both piles, it was noted that a layer of by subtracting this value from the total calculated
clay had adhered to the outside wall of the piles, the capacity results with an estimate of REB resistance.
thickest part being near the toe. Testing of this clay Since this calculation is made for each load increment,
revealed that it was stronger than the soil in the free- load-displacement curves can be plotted illustrating
field. This occurrence was also observed on some off- the development of skin friction and REB resistance
shore piles that were installed and then extracted. In relative to each other and to the measured load-
order for the soil adjacent to the pile wall to gain displacement curve.
shear strength, the octahedral shear stresses would Figure 4 is a plot of REB versus displacement from
have to increase as a result of installation, not decrease the ABAQUS analysis superimposed on a plot of pile
as has been suggested. tip suction versus displacement. It is likely that much
Evaluation of test results. Figure 3 presents the of the REB capacity is attributable to suction, thereby
load test results from the suction-installed piles and making this a useful comparison. By observation it is
the predicted behaviour from ABAQUS analyses made seen that the REB versus displacement curve from the
without prior knowledge of the centrifuge results. Also ABAQUS analysis is very similar in both the magnitude
shown are results of a hand calculation using the pro- and the slope of the plot of suction versus displace-
cedure recommended by API RP2A for tension piles ment at the pile tip. These results tend to corroborate
in cohesive soil. this hypothesis. In addition, the back-calculated bearing
The ABAQUS analyses were made using a model capacity factor, Nc, from the ABAQUS analysis is 8,
developed in 1999 by Sage Engineering for verifying which is also a credible value.
the design of the Diana DDCV suction piles. The soil This plot of REB versus displacement from the
input parameters were based on the CPT data and ABAQUS results also helps explain why the hand cal-
strength and stiffness data from laboratory tests at MIT culation over-estimates the pile capacity. At the dis-
on reconstituted samples of Speswhite. The MIT results placement where maximum skin friction is mobilized,

306

Copyright 2005 Taylor & Francis Group plc, London, UK


Pore Pressure Measurements 25
and Gage Locations Along Pile
Axis
20

Inclined Load, MN
200
Pressure Variation, kPa

100 15 Legend
2Pi92
0 2P103 Tub 7
10 Tub 8
-100 2P106 2P109 Tub 9
FEA-EB 2P117 Tub 10
2Pi92
2P106
5
-200 ABAQUS
2P117 2P103
2P109 0
-300

0 0.5 1.0 1.5 0 0.1 0.2 0.3 0.4 0.5


Vertical Displacement, m
Vertical Displacement, mt
Figure 4. Comparison of ABAQUS REB with measured
Figure 6. ABAQUS results compared to inclined load
suction pressure below pile tip.
test data.

corresponding ABAQUS analysis (Fig. 6). By obser-


20
vation, the match quality of the ABAQUS results is
Load vs. Displacement at Pile Top
16 excellent.
Load, MN

12
5 INSTALLATION THROUGH SAND
8

4 Shortly before the installation of the suction piles for


the Diana DDCV, a cone penetrometer survey was
0 performed at the anchor locations (4  3 group) to
0 0.5 1.0 1.5 confirm site stratigraphy. During this survey it was
discovered that a 1.5 m thick sand layer existed within
Vertical Displacement, m
3.0 m of final penetration. Calculations indicated that
it might not be possible to penetrate the piles by suction,
Figure 5. Vertical load versus displacement for suction
installed pile. and therefore it was decided to move one of the piles
to avoid the sand. Relocating a pile during installation
can be costly due to schedule delays.
the REB is only about 45% of its maximum calcu- To investigate this issue, a thin (1.5 m) medium
lated value. The plot of load versus displacement in dense sand layer was introduced into the clay bed about
Figure 5 shows that as the pile continues to displace 3.0 m above final pile penetration. This thickness was
under load, the load carrying capacity only increases selected for two reasons. First, the contractor that per-
by about 15%, which indicates that the frictional formed the geophysical survey advised us that it may
resistance is decreasing. This is not surprising since it not always be possible to identify a layer of this thick-
is necessary to displace the pile in excess of one ness using current geophysical techniques. Second,
meter to fully mobilize the REB. Thus, care should be there is too much uncertainty in calculation proced-
taken when evaluating pile capacity by decoupling ures to rely solely on analytical methods for making a
the skin friction and REB components. decision on whether to relocate an anchor pile.
The results of the test showed that the pile could
not penetrate the sand using the maximum available
4 INCLINED LOAD TESTS differential pressure of 260 kPa. This pressure is
approximately 80% of the theoretical pressure to fail
A number of inclined load tests were performed as the clay plug, which is a significantly greater percent-
part of the centrifuge program including monotonic, age than the maximum pressure generally permitted
sustained, and cyclic loading due to Gulf of Mexico during an offshore installation. Thus, practically speak-
loop current, and hurricane storms, and West Africa ing, a refusal condition was encountered. Subsequent
storms. Space limitations permit presenting only results testing of the pile showed that very little REB was
from one of the monotonic loading tests and the mobilized during loading, as would be expected. Upon

307

Copyright 2005 Taylor & Francis Group plc, London, UK


completion of the test, it was observed that the pile 4. The match quality between the experimental
had penetrated only about 0.2 m into the sand. results and ABAQUS comprise to some degree a
Only one test was run with a sand layer, and it is cross validation of the centrifuge scaling laws and
suggested that additional tests be run to verify these the soil model in ABAQUS.
results and to investigate the implications of having
sand layers at different locations within the formation.
Also it would be important to establish how close the ACKNOWLEDGEMENT
tip of a pile can be to a sand layer and not signifi-
cantly influence the REB resistance. The authors wish to acknowledge Gerald Rault and
his staff for their outstanding work in this program.
6 CONCLUSIONS

The results from the test program support the follow- REFERENCES
ing conclusions:
Garnier J., Rault G. & Cottineau L.M. 1995. Applications de
1. There is no significant difference in capacity la modelisation physique au domaine de loffshore,
between piles installed by self-weight and those AUGC, Nantes, pp. 129136.
installed by suction. Murff J. D. 1996. The Geotechnical Centrifuge in Offshore
2. Depending on the geometry of the pile, it may not Engineering, Offshore Technology Conference, OTC
8286, pp. 115.
be prudent to de-couple skin friction and end bear- Raines R. D. & Garnier J. 2004. Physical Modeling of
ing when calculating capacity by hand. Suction Piles in Clay, OMAE, 2004-51343.
3. Alternative installation plans should be in place if Andersen K. H. & Jostad H. P. 2002. Shear strength along
there is a likelihood of encountering a deep sand outside wall of suction anchor in clay after installation,
layer during installation. Proc. XII ISOPE Conference, Kyushu, Japan, May.

308

Copyright 2005 Taylor & Francis Group plc, London, UK


Establishing a model testing capability for deep water foundation systems

K.S. Prakasha
Oil and Natural Gas Corporation Ltd., Nazira, India

H.A. Joer
Advanced Geomechanics, Perth, Australia (formerly COFS, University of Western Australia)

M.F. Randolph
COFS, University of Western Australia, Perth, Australia

ABSTRACT: Foundation engineering is known to be a semi-empirical science where quantitative estimates


often involve empirical coefficients based on observations. These observations may be based on monitoring of
instrumented prototype behaviour but are more often based on data from model tests designed to explore new
technology. As exploitation of hydrocarbon resources moves into ever greater water depths, new foundation and
anchoring systems, such as suction caissons and bucket foundations are being adopted, with an increasing num-
ber installed around the world. Since experience is still limited and instrumented prototypes are rare, it is neces-
sary to undertake model tests to validate and fine-tune analytical models or design calculations. With this
motive, a modern laboratory testing facility has been developed at the Oil and Natural Gas Corporation Ltd
(ONGC) by the Centre for Offshore Foundation Systems (COFS). The principal objective of the facility was to
investigate the behaviour of deep foundation systems such as suction caissons and bucket foundations in clay
and sand materials. A complex loading mechanism that allows a combination of vertical (V), horizontal (H) and
moment (M) loading to be applied to model foundations was designed. In addition, all peripheral apparatus neces-
sary for the preparation of 1 m diameter sand or clay samples were developed, including a compression frame
for the preparation of clay samples, a pluviation device for the preparation of sand samples, site characteriza-
tion instruments (cone and T-bar penetrometers) and a range of model foundations and associated load cells. A
syringe pump for suction installation of caissons and a chain loading device were also developed. The paper
describes the various components of the facility. Preliminary test results obtained during the commissioning of
the facility are presented.

1 INTRODUCTION analysis has been used, but there are still difficulties
in assessing key parameters such as interface friction.
Offshore oil exploration and production requires con- Given the paucity of data from field installations, it is
struction and installation of structures to support and necessary to carry out model tests to arrive at proper
house production facilities and drilling rigs. Fixed design procedures.
structures such as conventional piled jacket or gravity- Key design uncertainties include the vertical load
base platforms are no longer feasible for water depths capacity of bucket foundations and the load capaci-
exceeding 200 m. Jacket platforms supported on bucket ties of suction anchors under typical angles of chain
foundations are found to be viable up to depths of loading and pad eye locations. The behaviour during
150 m, but for water depths beyond 200 m some form installation should also be studied in detail especially
of floating or tethered structure is generally optimal, in sands owing to possibility of cavitation or quick
and suction anchors appear to be one of the most sand conditions.
attractive anchoring alternatives. The requirements for conducting model tests on
The design basis for suction caissons has so far been deep water foundations are: availability of represen-
based on limit equilibrium approaches either through tative soil material in required quantities; character-
upper bound solutions (Randolph & House 2002) or isation of laboratory soil; a plan for the model tests to
through trial failure surfaces with consideration of side be conducted; model testing equipment and neces-
friction (Andersen & Jostad 2002). Finite element sary instrumentation and control software; model

309

Copyright 2005 Taylor & Francis Group plc, London, UK


foundations; etc. Details of the model testing labora-
tory established in Oil and Natural Gas Corporation Consolidation
Ltd. by the Centre for Offshore Foundation Systems frame
Slurry mixer
are presented in this paper.

2 SOIL MATERIAL

Every attempt should be made to collect drag samples


of representative material from the sea floor. However,
it is often difficult and costly to get the in situ soil for
model testing owing to the large quantity required (1 to
2 tonnes). In such cases reference onshore materials
such as clay with similar plasticity and normalized
shear strength properties or commercially available
kaolin can be used. Beach sand with similar angularity,
Sand rainer
Cc, Cu and Dr - relationships, can provide a suitable
reference material for marine sand.
Extensive laboratory investigations are required
containers
including static and cyclic triaxial and simple shear
tests, small strain stiffness assessment from resonant
column or bender element tests, consolidation tests, Figure 1. General arrangement of model testing laboratory.
index and classification tests, fall cone tests to estimate
strength and sensitivity relationships etc. A particular
challenge is relating the model soil to the in situ soil to 9 months. However, the spacing between tests, and
conditions, and gauging how to adjust results from between the foundation and walls of the container,
model foundation tests accordingly. should be such as to eliminate interaction. This will
generally necessitate a minimum of 4 diameters (cen-
tre to centre) spacing between adjacent tests, and
3 PLANNING MODEL TESTS 2-diameter spacing between the container wall and the
centre of the foundation for clays. A larger spacing is
Model tests are generally planned on the basis of either required for sands depending on friction angle. The
(a) a direct geometric model of the prototype problem, depth from tip to base of container should preferably
with measured results scaled appropriately for design, be twice the diameter.
or (b) a simplified model that is used to validate or
calibrate software, which is then used for design. It is
important to have a clear objective for any model test, 4 MODEL TESTING SET UP DIFFERENT
whether to observe a failure mechanism or to quantify EQUIPMENTS AND THEIR USES
some particular aspect of performance. Where pos-
sible, the entire construction or installation sequence for The model testing facility has been established in a
an offshore facility should be followed, monitoring spacious laboratory equipped with an overhead electric
relevant data at each stage. As the 1 g models do not crane for moving heavy equipment. The set-up includes
simulate prototype geostatic stresses (especially in equipment for sample preparation, consolidation, load-
sands), the results obtained form 1 g model tests can- ing, data acquisition and instrumentation, foundation
not be scaled directly to estimate prototype perform- models and specialist equipment required for per-
ance. However, results observed from these model tests forming specific tests.
can be compared with those obtained from numerical
or analytical solutions for the model test conditions to
4.1 Sample preparation
calibrate and develop the analytical models for design
as envisaged in (b) above. Geostatic stresses can only Samples are prepared in cylindrical containers of 1 m
be simulated using techniques such as a centrifuge. diameter and 0.4 m high. The height can be increased
Compromises will generally be necessary, both in by stacking container rings above each other.
terms of use of instrumentation and in the number of Clay samples are prepared by preparing the clay
tests undertaken in a given sample. Multiple tests in one slurry and pumping it into containers by means of a
sample allows comparison between tests and saves screw type slurry pump. The slurry mixer consists of
time, especially for tests in clay where consolidation a cylindrical hopper with a blade mounted on a verti-
of the sample is a very slow process and may take up cal shaft for mixing (Fig. 1). The speed of the blade

310

Copyright 2005 Taylor & Francis Group plc, London, UK


can be controlled. Clay slurry, free from stones and
vegetation, is prepared outside and poured into a hop-
per and mixed thoroughly for 24 hours at a water con- Vertical
tent in excess of twice the liquid limit. The salinity of actuator
sea water is maintained. Vacuum is applied while mix-
ing to remove the dissolved air and for proper mixing.
Sand samples are prepared by raining the sand
under free fall from a predetermined height to obtain
the desired density. The sand rainer is a basket fitted
with a 3.5 mm sieve at the bottom that fits inside the
bins (Fig. 1) The required density can be achieved by
changing the height of fall. A minimum relative density
of 30% can be achieved. Carbon-dioxide is passed
through the sand sample from the bottom of the con-
Horizontal actuator
tainer before infiltrating water from the bottom to
achieve saturation.

5 CONSOLIDATION EQUIPMENT
Figure 2. Complete test set up with actuator and foundation
Consolidation is carried out by applying a vertical
model in position.
consolidation pressure to the clay slurry. The pressure
is applied by means of a hydraulically operated piston
fitted with O rings (Fig. 1). The hydraulic system is The rate of loading should be slow enough to cap-
mounted on a reaction frame with a maximum cap- ture sufficient data points, and should endeavour to
acity of 900 kN; this corresponds to a pressure of just match the prototype situation. For tests in clay,
over 1 MPa, which can be regulated. The friction is undrained conditions can be achieved for displace-
minimised by perfectly machining the container insides ment rates typically 0.1 mm/s for model sizes of 50 to
and greasing the O rings. The consolidation pressure 100 mm diameter, or around 1 mm/s for a penetrome-
is applied in stages, with a waiting period to ensure ter of 5 to 10 mm diameter. The key criterion is to
95% consolidation has been achieved at the final pres- maintain the non-dimensional quantity vd/cv (where v
sure. Suction can be applied by means of a suction is velocity, d the diameter and cv the consolidation
pump at the bottom of the sample to achieve a strength coefficient) greater than about 30 (Finnie & Randolph
gradient. 1994). In sand, most loading will be drained, although
cyclic loading may lead to some accumulation of
excess pore pressures.
In cyclic loading, the control program reverses the
6 LOADING DEVICE horizontal or vertical displacement direction when the
targeted load or displacement is detected. A typical
The load is applied on the model foundations by means loading frequency of 0.1 Hz simulating storm loading
of a bi-axial actuator. The actuator is a displacement- is recommended for cyclic tests, although this can be
type machine (electrical rather than hydraulic) but runs varied to suit individual project needs. For the tests in
under feedback control (Fig. 2). The maximum strokes Indian marine clays, this rate of loading is sufficient
are 700 mm in the vertical direction and 500 mm in to produce undrained response, as the consolidation
the horizontal direction. The displacements are meas- at the tip of the foundation would be less than 10%
ured using encoders. Feedback control allows cyclic percent at the end of a one hour storm. However, in
or monotonic load or displacement, with independent sand, this rate of loading is not sufficient to produce
control of the vertical and horizontal axes. The actuator undrained response. Pore fluid of high viscosity (sil-
is designed so that the vertical tower can be uncoupled icon oil or Methyl cellulose) can be used to create
and then both axes may be used separately. Two undrained response.
motors and gearboxes (one for each direction) allow
the rate of displacement to be controlled in both
directions. In addition, the vertical tower can be tilted, 7 DATA ACQUISITION AND CONTROL
at increments of 5 angles up to 30, to apply inclined
loading to the foundation. Constant rate of loading can LABVIEW software has been used for data acquisi-
be achieved by feedback from the computer-controlled tion and control. It has provision for conducting the
system. tests under displacement or load control. It allows

311

Copyright 2005 Taylor & Francis Group plc, London, UK


specification of limiting displacements and loads profile occur as the testing proceeds. Figure 3 shows
with provision for alarms (or automatic stopping to a comparison of three T-bar tests conducted during
displacement) if these limits are exceeded. In load- the commissioning tests on clay. There is good con-
controlled tests, actuation is achieved through a feed- sistency between the three tests.
back system. The frequency of data acquisition can The peak in T-bar resistance at a depth of just over
be prescribed and the data is stored on disk every 15 100 mm is due to slight drying out of the surface of
seconds. The data acquisition system (DAS) consists the sample that occurred, prior to an additional batch
of a 32 channel data logger. of slurry being added to form the final upper 100 mm
Each of the three axes (x, z and syringe pump for of the sample. Over the main depth penetrated by the
application of suction) may be driven in either mono- caisson, the average T-bar resistance is around 90 kPa,
tonic or cyclic mode. In cyclic mode, the user enters corresponding to a shear strength of 8.5 kPa.
the displacement limits to drive between, in mm or
load limits, as well as specifying the number of cycles
9.2 Installation tests
to drive. Cyclic mode has the ability to drive the actu-
ator in either velocity mode or period mode. Velocity The models are aligned on the soil bed (Fig. 4) and
mode is simply mm/s, whilst period mode is input in the water or air is then extracted from the top of the
seconds and load in load units/s. model using a syringe pump. The rate of displace-
ment of the syringe pump and hence the applied suc-
tion or rate of installation can be controlled during the
8 MODEL FOUNDATIONS test. The suction pressure inside the model is meas-
ured using the pressure transducer fitted at the top of
Three types of model foundations are provided: fixed the model.
connection foundations; foundations loaded by a chain; Difficulties may be encountered while installing
and multiple foundations (simulating a structure on foundations in sand as seepage through the sand may
bucket foundations). Foundations with different length
to diameter ratios (say 1, 2 and 5) are possible. The T-bar resistance (kPa)
foundations are designed with a cap on top, which -150 -100 -50 0 50 100 150 200
allows connection of the load cells, the pump inlet 0
and pressure transducer. A specially designed pressure
transducer fitting has been made so that a hypodermic 50
needle can be fitted inside the foundation. 100
Penetration (mm)

In situ tests are conducted using miniature cone


and T-bar penetrometers in order to measure the 150 T-bar 1
actual strength profile for subsequent analysis of the 200 T-bar 3
test data and correlation with prototype conditions T-bar 4
(Stewart & Randolph 1994). 250 T-bar size
300 d = 10 mm
L = 50 mm
9 SPECIFIC TESTS AND EXAMPLE 350
RESULTS 400

Specific model tests that are frequently carried out Figure 3. Profiles of T-bar resistance.
include installation, vertical load tests, horizontal load
tests, inclined tension tests by means of chains, inclined
load tests with the actuator, multiple foundation tests
and cyclic tests. Testing procedures for such tests and
results from the commissioning tests in clay for a
model foundation of 100 mm diameter, 100 mm long
Model foundation
(L/d ratio of 1) are presented below.

9.1 Sample strength characterisation. Consolidated clay bed


The strength profile through the sample is determined
by T-bar penetration tests conducted periodically
through the programme of model tests. It is important to
assess whether any consistent changes in the strength Figure 4. Model foundation being inserted into the clay bed.

312

Copyright 2005 Taylor & Francis Group plc, London, UK


exceed the available volume of the syringe pump. The 9.3 Vertical load tests
pore fluid may be changed to overcome this. Care
After the model has been installed into the clay, it may
must also be exercised in selecting the rate of syringe
be pushed further into the soil bed and then pulled
displacement, so as not to apply too much suction
out. This allows determination of the capacity of the
which may lead to cavitation and upheaval.
foundation and also the vertical stiffness (from unload-
The results from commissioning tests in clay are
reload loops) for use in analysis. Results from the
shown in Figure 5. As can be seen, it is necessary to
commissioning test in clay are shown in Figure 6. As
regulate the suction pressure at the beginning of the
can be seen, the capacity of just under 1250 N is
test until smooth penetration is achieved. The vertical
mobilized at a small displacement and is maintained
load on the model recorded by the multi-axial Stroud
over a displacement of 20 mm. The pull out capacity
load cell during installation is also presented in the
is much less than the capacity in compression, prob-
figure. As can be seen very high suction was neces-
ably due to tension failure at the skirt base. These load
sary to install the model, although in the field the
tests used the multi-axial Stroud load cell to measure
depth of water will ensure positive total pressures.
the vertical load.
Note that a high vertical load was applied in the final
stages of installation, once the soil plug had contacted
the top cap of the foundation, and the maximum load 9.4 Horizontal load tests (side-swipe tests)
is equal to the bearing capacity of the foundation
Horizontal side-swipe tests (Tan 1990) are conducted
(see Fig. 6).
by pushing the installed foundation horizontally while
keeping the vertical load constant (Fig. 7). These tests
are carried out to determine the horizontal capacity of
0 1400 the suction anchors, and also to determine the approx-
-10 1200 imate yield envelope in horizontal-vertical load
-20 1000 space. Horizontal side-swipe tests at low vertical
Suction pressure (kPa)

load, and without rotation, simulate the field condi-


-30
Vertical load (N)

800 tion and can be carried out in a controlled manner.


-40 Suction pressure 600 Multi-axial Stroud load cells, which have three pairs
-50 Vertical load 400 of strain-gauged webs for measuring vertical load,
-60 200 horizontal load and moment applied to the founda-
tions are used. The results from the commissioning
-70 0
test in clay are shown in Figure 8. As can be seen from
-80 -200 the figure the horizontal capacity mobilised is insen-
-90 -400 sitive to the applied vertical load.
-100 -600
0 20 40 60 80 100
Penetration (mm)

Figure 5. Installation of model foundation.

1400
1200
1000
Rigid rod
Vertical load (N)

800
600
400
200
0
-200 Stroud load cell
-400
-600
0 5 10 15 20 25 30
Vertical displacement (mm)
Figure 7. Horizontal load test arrangement in sand with
Figure 6. Results from vertical load test on model foundation. stroud load cell.

313

Copyright 2005 Taylor & Francis Group plc, London, UK


600 700 14
Horizontal loading

Inclination of model (degrees)


500 & unloading loop up to 600 12
60 mm displacement,
Horizontal load (N)

for vertical load of 1100 N 500 10

Chain load (N)


400 Loss of control after
400 excessive inclination 8
300 of model
300 6
Horizontal loading & unloading loops Chain load
200 up to 15 mm displacement, for vertical 200 4
loads of 400, 600, 800 and 1000 N
100
100 2
Inclination
0 0 0
0 10 20 30 40 50 60 0 20 40 60 80
Horizontal displacement (mm) Extent of chain pulled (mm)

Figure 8. Results from horizontal side-swipe test. Figure 10. Results from monotonic chain load test.

800
700
600
Initial load of
Chain load (N)

500 300 N
400
300
200
Horizontal
Chain frame 100
0
0 20 40 60 80
Figure 9. Arrangement for chain load test.
Extent of chain pulled (mm)

9.5 Chain load tests Figure 11. Results from cyclic chain load test.
Inclined loading tests on foundations that are free to
rotate are carried out by pulling on a chain. The chain an initial tensile load of 300 N was applied before
is attached to the model foundation prior to installa- cycling so as to keep the chain always under tension.
tion and is tensioned through a pulley (Fig. 9). These Then cyclic loads with a 10 second period and of dif-
tests simulate the actual field conditions although it is ferent amplitudes were applied. The results from these
difficult to apply loads at very small angles due to the commissioning tests are shown in Figure 11. It can be
model tank dimensions and other limitations. The inclin- observed by comparison with Figure 10 that the capa-
ation of the model is measured continuously using city of the foundation remains nearly the same even
inclinometers. In addition, a miniature load cell is after cycling.
incorporated within the chain. Results from commis-
sioning tests in clay (chain load and inclination vs. the 9.7 Multiple foundation tests
extent of chain pulled) are shown in Figure 10. As can
be seen from the figure, inclination becomes exces- Bucket foundations generally operate as a group, with
sive soon after achieving the maximum capacity. one below each jacket leg. It would be preferable to
test these at model scale with multiple model founda-
tions rigidly attached to one another. This requires a
9.6 Cyclic load tests
suitable combination of foundation models and clay
Cyclic loading tests can be carried out by applying sample; even with the 1 m diameter container here,
varying horizontal loads or chain loads. They require only one test could be conducted within the sample.
sophisticated data acquisition control systems as At this stage, no tests have been conducted on multiple
discussed earlier. In the commissioning tests in clay, foundations.

314

Copyright 2005 Taylor & Francis Group plc, London, UK


10 CONCLUSIONS REFERENCES

The data presented in the paper demonstrate that Andersen, K.H. & Jostad, H.-P. 1999. Foundation design of
model tests provide an excellent opportunity for mak- skirted foundations and anchors in clay. Proc. Offshore
ing observations on deep water foundation systems. It Technology Conf., Houston, Paper OTC 10824.
Finnie, I.M.S. & Randolph, M.F. 1994. Punch-through and
requires elaborate planning, soil data and sophisti- liquefaction induced failure of shallow foundations on
cated equipment to carry out model tests. Since, these calcareous sediments. Proc. Int. Conf. on Behaviour of
1 g models have limitations in simulating in-situ Offshore Structures, BOSS 94, Boston, 217230.
stresses, their results cannot be used directly for esti- Randolph, M.F. & House, A.R. 2002. Analysis of suction
mating prototype performance by scaling. However, caisson capacity in clay. Proc. Annual Offshore Technology
their results can be used for calibrating and improving Conference, Houston, Paper OTC 14236.
analytical models for design. Since both analytical Stewart, D.P. & Randolph, M.F. 1994. T-bar Penetration testing
and physical models have their own advantages and in soft clay, J. Geot. Eng.Div., ASCE, 120(12), 223035.
limitations, their role is complementary in arriving at Tan, F.S. 1990. Centrifuge and theoretical modelling of conical
footings on sand. PhD Thesis, Cambridge University, UK.
suitable designs.

315

Copyright 2005 Taylor & Francis Group plc, London, UK


Reliability-based design considerations for deepwater mooring system
foundations

R.B. Gilbert, Y.J. Choi, S. Dangayach & S.S. Najjar


The University of Texas at Austin

ABSTRACT: Foundations for floating production systems in deep water are designed using methods adopted
from those used for jacket platforms in shallow water. While significant efforts have been undertaken to quan-
tify and understand the reliability of pile foundations for jackets, comparatively little effort has been made to do
the same for foundations in floating production systems. In this paper, a reliability analysis is conducted for a
study spar that is representative of existing technology and design practice in the Gulf of Mexico. Biases and
uncertainties that are inherent in the design loads and capacities are addressed. The effect of a lower-bound
value on the foundation capacity is considered. Finally, the reliability for the foundation in the study spar is
compared with that for a pile foundation in a typical jacket platform and with an industry-recommended target
level for mooring systems. The major conclusions are (1) foundation designs for floating production systems
may be excessively conservative; and (2) a lower-bound or minimum available foundation capacity can have a
significant effect on the reliability and is useful design input.

1 INTRODUCTION level that has been recommended by industry for


mooring systems. The paper concludes with a discus-
Foundations for floating production systems in deep sion of the implication of these results on current
water are designed using methods that were adopted design practice.
from those used for jacket platforms in shallow water
(e.g., API 1993). However, while significant efforts
have been undertaken to quantify and understand the 2 STUDY SPAR
reliability of pile foundations for jackets (e.g., Tang
1990; Hamilton & Murff 1992; Tang & Gilbert 1993; The design for the study spar was developed by an
Horsnell & Toolan 1996; and Bea et al. 1999), com- industry consortium named Deepstar. It is intended to
paratively little effort has been made to do the same be representative of existing technology for deepwa-
for foundations in floating production systems. Two ter oil production in the Gulf of Mexico. For the pur-
exceptions are Ronold (1990), who analyzed the reli- poses of this paper, this study spar is also intended to
ability of a pile in a tension leg platform, and Clukey be representative of a generic floating production sys-
et al. (2000), who analyzed a suction caisson for a tem in deep water.
variety of mooring system configurations. In both The study spar has a classic hull with a diameter of
cases, a very high-degree of reliability was calculated 37 m and a length of 215 m. It was designed with a
for current design practice. mooring system for three different water depths, nom-
This paper addresses reliability-based design con- inally 1,000 m, 2,000 m and 3,000 m. The mooring sys-
siderations for foundations of floating production tem consists of fourteen lines in an omni-directional
systems in deep water. A reliability analysis is con- pattern. The spar with a 1,000-m water depth has a
ducted for a study spar that is representative of exist- semi-taut system with wire rope lines; the spars with a
ing technology and design practice in the Gulf of 2,000-m and 3,000-m water depth have a taut mooring
Mexico. Biases and uncertainties that are inherent in system with polyester lines. The foundation for each
the design loads and capacities are addressed. The mooring line is a suction caisson with a length that is
effect of a lower-bound value on the capacity is con- six times the diameter and a padeye that is located at
sidered. Finally, the reliability for the foundation in two-thirds of the penetration depth below the mudline.
the study spar is compared with that for a pile foun- The required design capacities for individual lines
dation in a typical jacket platform and with a target and anchors in the study mooring systems were

317

Copyright 2005 Taylor & Francis Group plc, London, UK


determined using a 100-year hurricane event with FSmedian
both an intact mooring system (intact case) and with 1 2 3 4 5 6 7 8 9 10
the most-heavily loaded line missing (damage case). 1

Probability of Foundation Failure in Lifetime


When typical factors of safety from practice are Typical
applied to the study systems, the damage case gener- for
ally governs the design of the lines and foundations. 0.1 Jackets
Even in instances where the intact case governs the
design, the required design capacities from the dam- Total c.o.v. = 0.7
0.01
age case are very close to those from the intact case.
Therefore, for simplicity in presenting and explaining
the results in this paper, the damage case will be
0.001
assumed to govern in all instances. 0.6
The soil profile was assumed to be a normally con-
0.5
solidated clay that is representative of typical condi- 0.0001 0.4
tions in the Gulf of Mexico. In all three water depths, 0.3
the axial capacity of the suction caisson governs its
design. 0.00001

Figure 1. Results from a conventional reliability analysis


on a pile in a typical jacket platform.
3 RELIABILITY FRAMEWORK

A conventional reliability analysis for an offshore


design values represent biases between the median
foundation provides a useful framework as a starting
or most likely value in the design life and the value
point to consider the reliability of the suction caissons
that is used in the design check with the factor of
for the study spar. A conventional reliability analysis
safety. For context, the median factor of safety is
can be generalized in a convenient mathematical form
between three and five for a pile in a typical jacket
as follows:
platform.
The coefficients of variation in Equation 1 represent
uncertainty in the load and the capacity. For an off-
(1) shore foundation, the uncertainty in the load is gener-
ally due to variations in the occurrence and strength
of hurricanes at the platform site over the design life.
where P(Load  Capacity) is the probability that the The uncertainty in the capacity is due primarily to
load exceeds the capacity in the design life, which is variations between the actual capacity in a storm load
also referred to as the lifetime probability of failure; compared to the capacity predicted using the design
FSmedian is the median factor of safety, which is method. The denominator in Equation 1 is referred to
defined as the ratio of the median capacity to the as the total coefficient of variation:
median load; and  is the coefficient of variation
(c.o.v.), which is defined as the standard deviation (3)
divided by the mean value for that variable. Equation
1 assumes that the load and capacity each follow log-
normal distributions, a common assumption in typ- As an example, typical c.o.v. values for a jacket plat-
ical reliability analyses for offshore foundations (e.g., form range from 0.3 to 0.5 for the load, 0.3 to 0.5 for
Tang & Gilbert 1993). the capacity, and 0.5 to 0.7 for the total.
The median factor of safety in Equation 1 can be The relationship between the probability of failure
related to the factor of safety used in design: and the median factor of safety and the total c.o.v. is
shown on Figure 1. An increase in the median factor
of safety and a decrease in the total c.o.v. both reduce
the probability of failure. For context, the lifetime
(2) failure probabilities for a pile in a typical jacket foun-
dation range from 0.005 to 0.05 (Fig. 1). Note that the
event of foundation failure, i.e. axial overload of a
single pile in the foundation, does not necessarily
lead to collapse of a jacket. Failure probabilities for
where the subscript design indicates the value used the foundation system are ten to 100 times smaller
to design the foundation. The ratios of the median to than those for a single pile (Tang & Gilbert 1993).

318

Copyright 2005 Taylor & Francis Group plc, London, UK


4 BIAS AND C.O.V. VALUES FOR SPAR Table 1. Bias and c.o.v. values for foundation in study spar.
FOUNDATIONS
Water depth (m)
The reliability for a foundation in the study spar will
1,000 2,000 3,000
depend on appropriate values for the bias and c.o.v.
values for the load and capacity (Fig. 1). loadmedian/loaddesign(damage) 0.41 0.70 0.71
capacitymedian/capacitydesign 1.21.4 1.21.4 1.21.4
4.1 Foundation load FSdesign(damage) 1.52.5 1.52.5 1.52.5
FSmedian 48 35 35
A numerical model (Ding et al. 2003) was used to
load 0.30 0.14 0.11
simulate mooring line loads during different sea capacity 0.3 0.3 0.3
states for the study spar. From the resulting time his- total 0.4 0.3 0.3
tories and a probabilistic description of hurricanes for
the Gulf of Mexico (Winterstein & Kumar 1995), a Notes: Design life  20 years; Damage case governs
probability distribution was developed for the max- design; Axial loading governs design capacity.
imum load in a line during a storm (Dangayach 2004).
The vertical load at the anchor was then determined
2.0
from the maximum load and corresponding angle in Luke et al 2003

Measured/Predicted Capacity
Clukey and Morrison 1993
the mooring line at the mudline using the analytical Clukey et al 2003
model developed by Neubecker & Randolph (1995). Randolph and House 2002
1.5 Cho et al 2003
The results are summarized in Table 1 for the most House and Randolph 2001
heavily loaded line during a storm.
The conservative bias in the median load versus 1.0
the design load is greater for these spar foundations
than for a pile in a typical jacket platform, where the 0.5
ratio of the median to the design load is between 0.7
and 0.8 (Tang & Gilbert 1993). This conservative bias
is especially significant for the semi-taut mooring 0.0
system (1,000-m water depth) due to the effect of 0.0 0.1 1.0 10 100 1000 1E04 1E05 1E06
removing a line in establishing the design load. The Measured Capacity (kN)
loads are shared more evenly between the lines in the
taut mooring systems, which minimizes the impact on Figure 2. Measured versus predicted axial capacity for
each line when one line is removed. model tests on suction caissons in normally consolidated
clays.
Also, the coefficients of variation in the spar foun-
dation load are smaller than for a pile in a typical jacket
in models for predicting the uplift capacity of suction
platform, where the c.o.v. values are generally between
caissons in normally consolidated clays. This data-
0.3 and 0.5 (Tang & Gilbert 1993). There are several
base of 25 tests includes the following: laboratory-
reasons for smaller uncertainty in the foundation loads
scale model tests (Luke et al. 2003), centrifuge tests
on the spar. First, the line loads are less sensitive to
(Clukey & Morrison 1993; House & Randolph 2001;
wave height for a spar mooring system in deep water
Randolph & House 2002; Clukey & Phillips 2002;
compared to a fixed jacket in shallow water (e.g.,
Clukey et al. 2003), and full-scale field tests (Cho et al.
Banon & Harding 1989). Therefore, variations in the
2003).
sea states over the design life are less significant for the
For the purposes of a preliminary analysis, the pre-
spar mooring system. Second, the mooring system is
dicted capacities in each test were calculated using
simpler to model than a jacket, meaning that there is
the model developed by Aubeny et al. (2003 a,b) with
less uncertainty in the loads predicted by the model.
an alpha value of 1.0 for side friction and a bearing
Finally, the spar line loads are dominated by pre-
capacity factor of 9.0 for the reverse end bearing.
tension versus environmental loads; variations in the
Shear strengths reported by each investigator were
load due to variations in the sea states therefore have a
used directly. Refer to Najjar (2005) for details of this
smaller effect on the total line load. This effect of pre-
analysis. Ratios of measured to predicted capacities
tension is particularly significant for the taut mooring
for the 25 tests in the database are plotted on Figure 2.
systems (2,000-m and 3,000-m water depths), which
The average value for the ratio of measured to pre-
consequently have the smallest c.o.v. values (Table 1).
dicted capacity on Figure 2 is 0.99, indicating an
unbiased prediction model. A more careful analysis
4.2 Foundation capacity
was performed on a set of laboratory-scale model
A database comprised of published load tests was tests that were conducted on a suction caisson with a
assembled and used to evaluate biases and uncertainties range of loading angles (El-Sherbiny et al. 2005). The

319

Copyright 2005 Taylor & Francis Group plc, London, UK


results of this analysis also support the use of an alpha FSmedian
of 1.0 and a bearing capacity factor of 9.0 to produce 1 2 3 4 5 6 7 8 9 10
unbiased predictions for capacity. 1

Probability of Foundation Failure in Lifetime


While there are variations in design practice with Typical
suction caissons, typical values for alpha and the bear- for
ing capacity factor in practice are 0.6 to 0.8 and 7 to 9, 0.1 Jackets
respectively. Therefore, there is a conservative bias that
is introduced with these nominal values. To quantify Total c.o.v. = 0.7
this bias, the suction caisson design was considered for 0.01
the study spar. The side friction contributes about 50
percent of the total capacity, and the reverse end bearing
0.001
contributes about 35 percent of the total capacity. As an 0.6
example, if there is a bias of 1.0/0.7 on the side friction
0.5
and a bias of 9.0/7.0 on the end bearing, the composite 0.0001 0.4
bias on the total capacity is 1.3. For typical designs, the 0.3 Study Spars
bias will range from about 1.2 to 1.4. Coincidentally,
the design bias for pile foundations on jacket platforms 0.00001
is also about 1.3 (Tang & Gilbert 1993).
The coefficient of variation in the ratio of meas- Figure 3. Results from a conventional reliability analysis
ured to predicted capacity about the average on on study spar foundations.
Figure 2 is 0.28. This value is very similar to the value
of 0.3 that is typically used for pile foundations on 5 LOWER-BOUND FOUNDATION CAPACITY
jacket platforms (Tang & Gilbert 1993).
One consequence of the relatively large median fac-
tors of safety and small coefficients of variation in the
4.3 Median factor of safety, total c.o.v. and applied loads for spar foundations is that the presence
reliability of a minimum or lower-bound capacity can have a
The biases in the load and the capacity are combined large effect on the reliability (Gilbert et al. 2005). A
together with the design factor of safety through simple estimate of the lower-bound capacity for a
Equation 2 to determine the median factor of safety, suction caisson in normally consolidated clay can be
and the results are summarized in Table 1. Typical obtained using the remolded strength of the clay to
factors of safety being used in practice for the damage calculate side friction and end bearing.
case, which governs design for the study spar, range The database of load tests (Fig. 2) was used to
between 1.5 and 2.5. In addition, the total c.o.v. val- investigate the existence of a lower-bound capacity.
ues are obtained from Equation 3 (Table 1). For each test, a predicted lower-bound capacity was
The relationship between the reliability and calculated using the remolded undrained shear
FSmedian and total is re-plotted on Figure 3. For the strength with an alpha of 1.0 and a bearing capacity
spar foundation, the probability that the axial load factor of 9.0. Details for how the lower-bound cap-
will exceed the capacity of the suction caisson during acity was calculated in each test and a discussion of
a 20-year design life is on the order of 0.0001 or relevant assumptions are provided in Najjar (2005).
smaller. For comparison, this failure probability is The ratio of the calculated lower-bound capacity to
more than two orders of magnitude smaller than that the measured capacity is shown on Figure 4.
for piles in typical jacket structures (Fig. 3). For all tests analyzed, the ratio of the calculated
There are several reasons for the difference in reli- lower-bound capacity to the measured capacity is less
ability on Figure 3 between anchors for floating pro- than or equal to 1.0, providing compelling evidence
duction systems and piles in jacket platforms. A new for the existence of a lower-bound axial capacity. The
source of conservatism was introduced for floating ratio of lower-bound capacities to measured capacities
production systems with the damage case, where the ranges from 0.25 to 1.0 with an average value of 0.6.
factor of safety is applied to a load corresponding to The effect of a lower-bound capacity on the reli-
one line missing. In addition, due to concerns with ability of the foundation is shown on Figure 5 for the
using new types of foundations such as suction cais- study spar in 2,000 m of water. In these calculations,
sons, the factors of safety were generally increased. the lognormal distribution for the capacity is assumed
Finally, there tends to be less uncertainty in the loads to be truncated at the lower-bound value using a mixed
on mooring system anchors compared with jacket piles. lognormal distribution (Najjar 2005). The probability
The result of increased conservatism and decreased of failure is calculated through numerical integration.
uncertainty for anchors compared to jacket piles is The results on Figure 5 show the significant role
reflected on Figure 3. that a lower-bound capacity can have on the reliability.

320

Copyright 2005 Taylor & Francis Group plc, London, UK


2.0 1E+00
Lower Bound/Measured Capacity

Luke et al 2003
Clukey and Morrison 1993

Probability of Failure in Lifetime


Clukey et al 2003 1E-01 1,000-m Water Depth
Randolph and House 2002
1.5 Cho et al 2003
House and Randolph 2001 1E-02 Target

1E-03 2,000-m Water Depth


1.0
1E-04 3,000-m Water Depth
0.5 1E-05

1E-06
0.0
0.0 0.1 1.0 10 100 1000 1E04 1E05 1E06 1E-07
Measured Capacity (kN) 1 1.5 2 2.5
Design FS (Damage Case)
Figure 4. Calculated lower-bound versus measured axial
capacity for suction caissons in normally consolidated clays. Figure 6. Reliability for study spar foundations versus
design factor of safety (20-year design life).
1E+00
FSdesign(damage) = 1.5 annual probabilities of failure represent the rate of
Probability of Failure in Lifetime

1E-01 occurrence for high-consequence events.


FSdesign(damage) = 2.0
In contrast to an event that is dominated by a time
1E-02
FSdesign(damage) = 2.5 varying load, the uncertainty in the failure of an off-
1E-03 shore foundation is dominated by uncertainty in the
capacity. This capacity does not vary randomly with
1E-04 time. Therefore, it is not appropriate to consider the
probability of failure as a rate of failure. If the actual
1E-05
capacity is higher than expected, then the annual rate
1E-06 of failure due to storm loading may be very small. If
the actual capacity is lower than expected, then the
1E-07 annual rate of failure may be larger.
0 0.2 0.4 0.6 0.8 A more appropriate measure of the reliability for a
Ratio of Lower-Bound to Median Capacity foundation is the probability of failure during the life-
time of the structure. This probability was calculated
Figure 5. Effect of lower-bound capacity on probability of in Figures 3 and 5 by considering the time-varying
failure (study spar in 2,000-m water depth).
component of the load to determine the distribution
of the maximum load applied to the foundation over
its lifetime.
For the average lower-bound capacity from Figure 5, In order to compare failure probabilities in a
0.6 times the median capacity, the probability of fail- design life with target probabilities of failure that are
ure is more than 1,000 times smaller with the lower- expressed as annual rates, the target probabilities
bound than without it for a design factor of safety of should be converted to a probability of failure in a
1.5 in the damage case. Furthermore, for design fac- lifetime. Since it is implicit in published failure rates
tors of safety of 2 or 2.5 in the damage case, the prob- that event occurrences are statistically independent
ability of failure for a lower-bound capacity that is 0.6 with time, the probability of failure in a lifetime, T,
times the median capacity is essentially zero. can be obtained from the following:

6 TARGET RELIABILITY
(4)
Results from reliability analyses are generally pre-
sented in terms of the annual probability of an event where pannual is the annual failure rate. Therefore, the
in offshore applications. For example, Goodwin et al. target failure rate of 2  10
4 per year for a single
(2000) recommend a target probability of failure of mooring line recommended by Goodwin et al. (2000)
2  10
4 per year for a single mooring line. The motiv- corresponds to a target probability of failure of 0.004
ation for using annual probabilities is that many in a 20-year design life.
events in offshore applications, such as hurricanes The probability of failure in a lifetime is shown on
and explosions, occur randomly with time. These Figure 6 for the study spar foundation in the three

321

Copyright 2005 Taylor & Francis Group plc, London, UK


water depths. For these foundations, the lower-bound Arctic Engineering, Cancun, Mexico, June, OMAE03-
capacity was calculated to be about 0.43 times the 37502.
median capacity. The target probability of failure is Banon, H. and Hardin, S. J. 1989. Methodology for Assessing
also shown on Figure 6. A factor of safety of 1.3 for Reliability of Tension Leg Platform Tethers. Journal of
Structural Engineering, Vol. 115, No. 9: 22432259.
the damage case would achieve the target reliability Bea, R. G., Jin, Z., Valle, C. and Ramos, R. 1999. Evaluation
for all three water depths. In comparison, current of Reliability of Platform Pile Foundations. J. Geotech.
practice has this factor of safety between 1.5 and 2.5. and Geoenvir. Engrg., ASCE, 125(8): 696704.
Cho, Y., Lee, T. H., Chung, E. S. and Bang, S. 2003. Field
7 CONCLUSIONS AND DISCUSSION Tests on Pullout Loading Capacity of Suction Piles in
Clay. Proc. of the Intl. Conference on Offshore
Mechanics and Arctic Engineering, OMAE: pages-7.
The first major conclusion from this work is that Clukey, E. C., Banon, H. and Kulhawy, F. H. 2000.
foundation designs for floating production systems Reliability Assessment of Deepwater Suction Caissons.
may be excessively conservative. The probability of Proc. Offshore Technology Conf., Houston, OTC 12192:
failure values being achieved are several orders of 777785.
magnitude smaller than industry-recommended tar- Clukey, E. C. and Phillips, R. 2002. Centrifuge Model Tests
gets for components (single line or foundation) in a to Verify Suction Caisson Capacities for Taut and Semi-
mooring system. The practical concern with such Taut Legged Mooring Systems. Proc. Intl. Conference on
excessive conservatism is that installation can become Deepwater Offshore Technology.
Clukey, E. C. and Morrison, M. J. 1993. A Centrifuge and
unnecessarily costly and problematic. In fact, if one suc- Analytical Study to Evaluate Suction Caissons for TLP
tion caisson in the mooring spread cannot be installed, Applications in the Gulf of Mexico. ASCE Geotech.
then the risk actually increases because the system is Special Pub. (38): 141156.
in a damaged state from the start. Clukey, E. C., Aubeny, C. P. and Murff, J. D. 2003.
The second major conclusion from this work is Comparison of Analytical and Centrifuge Model Tests for
that a lower-bound or minimum available capacity Suction Caissons Subjected to Combined Loads. Proc.
can have a significant effect on the foundation reli- 22nd Intl. Conf. on Offshore Mechanics and Arctic
ability. One practical implication of this effect is to Engineering, Cancun, Mexico, June, OMAE03-37503.
include a design check using the lower-bound cap- Dangayach, S. 2004. Reliability Analysis of Mooring
System for Spar. M. S. Thesis, University of Texas at
acity in addition to the nominal or design capacity. A Austin: 84 pp.
second practical implication is that installation data Ding, Y., Kim, M., Chen, X. and Zhang, J. 2003. Coupled
can be used to update the reliability even when signif- Analysis of Floating Production System. Proc. Intern.
icant set-up is expected. Symp. on Deepwater Mooring Systems, Concepts, Design,
Ongoing work in this area is addressing the reli- Analysis, and Materials, Houston, Texas: 152167.
ability of the system of foundations, chains and lines. El-Sherbiny, R. M., Olson, R. E., Gilbert, R. B. and Vanka, S. K.
2005. Capacity of Suction Caissons Under Inclined
Loading in Normally Consolidated Clay, Proc. Frontiers in
ACKNOWLEDGEMENT Offshore Geotechnics, Perth, Western Australia: in review.
Gilbert, R. B., Najjar, S. S. and Choi, Y. J. 2005. Incorporating
This study was supported by the Offshore Technology Lower-Bound Capacities into LRFD Codes for Pile
Foundations. Proc. Geo-Frontiers Congress, Austin, Texas:
Research Center at The University of Texas at Austin in press.
and Texas A&M University. Prof. Jun Zhang and his Goodwin, P., Ahilan, R. V., Kavanagh, K. and Connaire, A.
research group at Texas A&M performed the hydro- 2000. Integrated Mooring and Riser Design: Target
dynamic analyses to predict mooring-line loads. Reliabilities and Safety Factors. Proc. of OMAE: 185792.
Profs. Rick Mercier and Skip Ward of Texas A&M Hamilton, J. M. and Murff, J. D. 1992. Selection of LRFD
provided technical guidance. Resistance Factors for Pile Foundation Design. Proc. of
Structural Congress 92, ASCE, San Antonio, Texas:
788795.
REFERENCES Horsnell, M. R. and Toolan, F. E. 1996. Risk of Foundation
Failure of Offshore Jacket Piles. Proc. of Offshore Tech.
API 1993. API RP 2A-LRFD: Recommended Practice for Conf., OTC 7997: 381392.
Planning, Designing and Construction Fixed Offshore House, A. R. and Randolph, M. F. 2001. Installation and
Platforms Load and Resistance Factor Design, First Pullout Capacity of Stiffened Suction Caissons in
Ed., American Petroleum Institute, Washington, D. C. Cohesive Sediments. Proc. 11th Intl. Offshore and Polar
Aubeny, C. P., Han, S. W. and Murff, J. D. 2003a. Inclined Engineering Conference.
Load Capacity of Suction Caissons. Intl. J. Numerical Luke, A. M., Rauch, A. F., Olson, R. E. and Meacham, E. C.
and Analytical Methods in Geomechanics, 12351254. 2003. Components of Suction Caisson Capacity
Aubeny, C. P., Han, S. W. and Murff, J. D. 2003b. Refined Measured in Axial Pullout Tests. Proc. Intl. Symp. on
Model for Inclined Load Capacity of Suction Caissons. Deepwater Mooring Systems, Concepts, Design,
Proc. 22nd Intern. Conf. on Offshore Mechanics and Analysis, and Materials, OTRC: 112.

322

Copyright 2005 Taylor & Francis Group plc, London, UK


Najjar, S. S. (2005), Importance of Lower-Bound Capacities Tang, W. H. 1990. Performance Reliability of Offshore
in Geotechnical Reliability Assessments, Ph.D. Piles. Proc. of Offshore Tech. Conf., OTC 6379: 299308.
Dissertation, The University of Texas at Austin: 317 pp. Tang, W. H. and Gilbert, R. B. 1993. Case Study of Offshore
Neubecker, S. R. and Randolph, M. F. 1995. Performance of Pile System Reliability. Proc. of Offshore Tech. Conf.,
Embedded Anchor Chains and Consequences for Anchor OTC 7196: 677686.
Design. Proc. Offshore Technology Conf., OTC 7712, Winterstein, S. R. and Kumar, S. 1995. Reliability of
Houston, Texas: 191200. Floating Structures: Extreme Response and Load Factor
Randolph, M. F. and House, A. 2002. Analysis of Suction Design, Proc. Offshore Technology Conf., OTC 7758,
Caisson Capacity in Clay. Proc. Offshore Technology Houston, Texas: 569578.
Conf., Houston, Texas, OTC 14236.
Ronold, K. O. 1990. Reliability Analysis of Tension Pile.
Journal of Geotech. Engineering, Vol. 116, No. 5:
760773.

323

Copyright 2005 Taylor & Francis Group plc, London, UK


Validation of the use of beam-column method for suction caisson design

J. Cao, Y. Li, K.-M. Tjok & J.M.E. Audibert


Fugro-McClelland Marine Geosciences, Inc., Houston, Texas, USA

ABSTRACT: This paper presents data to validate the application of the Beam-column (BMCOL) method as
an efficient tool for suction caisson design. First, methods to develop t-z, Q-z, and p-y curves are discussed and
are compared with those by conventional methods recommended by API RP 2A (2000). Then, the application
of the BMCOL method in suction caisson design is presented, including investigation of axial and lateral load
vs. displacement curves, prediction of caisson lateral behavior, padeye location optimization, and estimation of
shear force and bending moment distributions along the caissons wall. Finally, the applicability of the BMCOL
method for suction caisson design is discussed and compared to rigid rotation theory for piles with various slen-
derness ratios.

1 INTRODUCTION suction caissons have much larger diameters and are


shorter than conventional piles, a critical step in using
The BMCOL method has been used by the authors to the BMCOL method for suction caisson design is to
evaluate suction caissons axial and lateral behavior. develop soil springs that can appropriately represent
This method was originally developed to investigate the caisson/soil lateral interaction. Details on the selec-
the behavior of slender driven pipe piles (Bogard & tion of soil springs, especially for the p-y curves, are
Matlock 1977). It treats the pile foundation as a series discussed in the following section.
of discrete elements, represented by linear springs,
which are acted upon by nonlinear springs represent-
ing the surrounding soil. 2 T-Z, Q-Z AND P-Y SOIL REACTION CURVES
Suction caissons differ from piles in that: 1) they
are shorter and stiffer and 2) the load is applied some- 2.1 T-z curves
where between mudline and the caissons tip instead
Several factors can affect the caissons capacity (Huang
of at the top, as it is generally the case for piles.
et al. 2003), including set-up effect, padeye trenching
Although some concern exists within the geotech-
effect, long-term sustained creep effect, etc. These
nical community about the applicability of the BMCOL
factors can be incorporated in the unit skin friction as
method to suction caissons, our design experience
follows:
over the past 8 years indicates that the BMCOL method
can be successfully and effectively used for suction
caisson design. (1)
This paper presents how the BMCOL method has
been applied in suction caisson design to help investi- where setup is the set-up factor, trench is the trench
gate the axial and lateral load vs. displacement behav- factor, creep is the long-term creep factor, twist is the
iors, padeye location optimization, and shear force and out-of-plane load factor, NGI is the active suction
moment distributions in the caisson wall. The BMCOL effect factor, post is the post-peak reduction factor
method can be a very useful time-saving and cost- (see Fig. 1), API is the API RP 2A (2000) shear trans-
effective tool, especially during the preliminary design fer factor, and Su is the soil undrained shear strength
phase. at the depth under consideration.
The soil-caisson interaction is implemented in the The t-z (unit skin friction vs. pile displacement)
BMCOL method through t-z curves for the axial dis- curve for cohesive soils recommended by API RP 2A
placements, a Q-z curve for the reverse end bearing (2000) can be used to describe the skin friction devel-
(REB) and p-y curves for the horizontal (lateral) dis- opment for suction caissons in the BMCOL analysis.
placements. The caisson/soil interaction depends on For example, according to the API RP 2A (2000)
both the soil behavior and the caissons stiffness. Since approach shown on Figure 1, the maximum skin friction

325

Copyright 2005 Taylor & Francis Group plc, London, UK


z/D (for Q/Qmax) 1.2
0 0.1 0.2 0.3 0.4 0.5 1.0
1.2
0.8
t/tmax or Q/Qmax

1.0

p/pu
0.8 0.6
0.6 API RP 2A (2000)
0.4
Q-z curve
0.4 Range of tRES for clay Stevens & Audibert (1979)
(tRES=0.7~0.9 tmax) 0.2
0.2
0.0
0.0 0 3 6 9 12 15
0 0.02 0.04 0.06 0.08 0.1 y/yc
z/D (for t/tmax)
Figure 2. Lateral load vs. deflection curves.
Figure 1. t-z and Q-z Curves as recommended by API RP
2A (2000).
where Su is the undrained shear strength for undis-
is mobilized at a displacement of one percent of caisson turbed clay, D is the pile diameter,  is the soil sub-
diameter (0.01D). The skin friction decreases to a value merged unit weight, J is a dimensionless empirical
equal to 70 to 90 percent of its maximum value at a dis- constant with values ranging from 0.25 to 0.5, and D
placement equal to two percent of the caisson diam- is the pile diameter.
eter (0.02D), and remains constant beyond this point. For Z  ZR:

2.2 Q-z curves (4)


The reverse end bearing (REB) can be calculated The critical depth ZR is the depth below seafloor
using the following equation: where the soil failure mechanism changes from a
(2) shallow wedge to a deep flow around the pile.
P-y curves from API RP 2A (2000) may not be
entirely appropriate for large diameter caissons. Stevens
where Nc is a reverse end bearing factor, Su,tip is the and Audibert (1979) recommended a modified method
undrained shear strength of soil at the level of the for piles with large diameters, which is believed to
caisson tip or 1/4D to 1/2D below the caisson tip, and be more suitable for suction caissons. The modified
Abase is the full cross section area of the caisson. method differs from the API RP 2A (Matlock 1970)
Studies by various researchers (Cluckey & Phillips formulation in two points:
2002, Cao 2003, Randolph & House 2002) indicate
that the reverse end bearing factor Nc ranged from 6 1. As indicated on Figure 2, the p/pu vs. y/yc curve is
to 15, depending on how test results are analyzed and stiffer for y/yc values from 0 to 1; and
interpreted. A value of 9 is usually used for suction 2. yc is determined by:
caisson design.
The Q-z (tip resistance vs. tip movement) curve (5)
recommended by API RP 2A (2000) for piles can be
adopted for suction caissons. As indicated on Figure 1, where 50 is the axial strain at 50% of the peak
a displacement of ten percent of the caissons diame- deviator stress for cohesive soil. Comparing with the
ter (0.1D) is required to fully mobilize the reverse end yc  2.550D given in API RP 2A (2000), the modi-
bearing. fied yc is much smaller, however, the overall shape of
the curves is very much the same as in API RP 2A
2.3 P-y curves (Fig. 2).
The API RP 2A (2000) recommends the use of the An example is given here to illustrate the difference
Matlock (1970) approach to develop lateral load- between the two methods. Two BMCOL analyses were
displacement (p-y) curves for soft clays. Based on carried out, one using the API RP 2A p-y curve criter-
this criterion, the ultimate unit lateral soil resistance (pu) ion and one using the Stevens and Audibert (1979)
increases from 3SuD to 9SuD, as the depth (z) increases method. The caisson configuration, soil and loading
from 0 to the critical depth (ZR) according to: conditions are the same for both analyses. The results
of the BMCOL analyses are presented on Figure 3. It
(3) can be seen from this figure that, using the API RP 2A

326

Copyright 2005 Taylor & Francis Group plc, London, UK


Lateral Displacement, [m] 21000 Finite element analyses
-0.2 -0.1 0 0.1 0.2 0.3
18000

Mooring Line Load, [kN]


-6
Mudline 15000
Distance Along Caisson From Mudline, [m]

0 Stevens and
12000 API RP 2A (2000)
Audibert (1979)
Beam-Column analyses
6 Stevens and API RP 2A 9000
Audibert method (2000)
(1979) 6000
12 Thin line: load angle = 30 from horizontal
3000 Thick line: load angle = 18 from horizontal

18 0
0 0.3 0.6 0.9 1.2 1.5
24 Total Displacement at Padeye, [m]

Figure 4. Comparison of load vs. total displacement


30
curves Obtained Using BMCOL and FEA Methods.
Padeye located at 18.3 m BML
36
3.1 Load vs. displacement curves
Figure 3. Comparison of caisson lateral behavior using 3.1.1 Load vs. total displacement
p-y curves based on Stevens and Audibert (1979) and API
RP 2A (2000).
Figure 4 presents the load vs. total displacement
curves obtained from BMCOL analyses, using both
the API RP 2A (2000) p-y curves and the Stevens and
Audibert (1979) p-y curves, for two padeye load
p-y curves, the caisson has a displacement of about angles (18 and 30), in comparison to the finite elem-
0.1 m toward the mooring line. Using the Stevens and ent analysis results. As shown on this figure, based on
Audibert p-y curves, the caisson has a slight back- the BMCOL analyses, the caisson responses using the
ward rotation, and the maximum lateral displacement Stevens and Audibert p-y curves are stiffer than those
is only on the order of 0.01 m. using the API RP 2A p-y curves. The finite element
A lateral load test was carried out on one of the results indicate that the caisson response is signifi-
anchor caissons for a Gulf of Mexico project. The test cantly stiffer as compared to the BMCOL solutions
results indicated that the BMCOL results using the using the API RP 2A p-y curves, but only slightly
Stevens and Audibert p-y curves were in reasonable stiffer than the BMCOL solutions using the p-y
agreement with the field test results for both the curves developed using the Stevens and Audibert
axial and the lateral displacements at the mudline. approach. Both the API RP 2A and the Stevens and
Unfortunately, these data cannot be released to the Audibert methods overestimate the caissons ultimate
public domain. capacity and the caisson deflections as compared
to the finite element results, with the Stevens and
Audibert method being much closer to the finite elem-
ent solutions.
3 BEAM COLUMN METHOD VALIDATION
FOR SUCTION CAISSON DESIGN
3.1.2 Load vs. axial displacement
The BMCOL method has been used for suction cais- Axial load vs. padeye displacement curves obtained
son design in predicting load vs. displacement behav- by BMCOL analyses and by finite element analyses
ior, checking capacity, optimizing the padeye are compared on Figure 5. As can be seen from this
location, and finding the shear force and bending figure, up to about 16000 kN or 90% of the total
moment distributions along the caisson wall. To valid- capacity, the axial load vs. displacement curve from
ate the use of this method for suction caissons, the the BMCOL analyses is in very good agreement with
results from the BMCOL analyses are compared with that from the finite element analyses. Beyond this
the finite element analysis (FEA) results in the fol- point and for a given load, the BMCOL analyses indi-
lowing sections. The development of the FEA model, cate a slightly smaller displacement than the finite elem-
including the FE mesh and boundaries, the constitu- ent analysis. The difference between the BMCOL
tive models for soil and caisson wall materials and the results and the FEA results is less than 5%, however.
simulation of loading are discussed in a companion This difference is considered to be quite acceptable in
paper for this conference (Cao et al., 2005). geotechnical engineering.

327

Copyright 2005 Taylor & Francis Group plc, London, UK


Vertical Displacement, [m] 15

Normalized Lateral Capacity, Nk


14 L/D = 6 FEA
0 0.1 0.2 0.3 0.4 0.5 13
21000 Vertical 12 L/D = 6 BMCOL
Stevens and Audibert (1979)
Mooring Line Load, [kN]

18000 11
Lateral 10
15000 9
API RP 2A (2000) 8
12000 7
6
9000 5
4
6000 3
Thin line: Finite element analysis 2
3000 Thick line: BMCOL analysis 1
0
0 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
0 0.01 0.02 0.03 0.04 0.05
Padeye Position (Zp/L)
Lateral Displacement, [m]
Figure 6. Lateral load vs. padeye location.
Figure 5. Load vs. displacement curves comparison
between BMCOL and FEA methods.
the embedment depth below mudline. It can also be
seen from this figure that, up to a Zp/L value of 0.6,
3.1.3 Load vs. lateral displacement
the lateral capacity obtained by the BMCOL analysis
Load vs. padeye lateral displacement curves obtained
is slightly lesser than that from the finite element
by BMCOL analyses and by finite element analyses
analysis. Beyond this point, a smaller capacity is cal-
are compared on Figure 5. As shown on this figure,
culated by the BMCOL analysis. In general, the lat-
for a given load, the lateral displacement obtained by
eral capacity results from the BMCOL analysis are
the BMCOL analyses using API RP 2A p-y curves is
very close to those obtained by the finite element
about two to three times that from the finite element
method, and the optimal padeye position is found to
analysis. However, the lateral displacement obtained by
be essentially the same by both analysis methods.
the BMCOL analyses using the Audibert and Stevens
p-y curves is in very good agreement with the finite 3.2.2 Lateral behavior optimization
element analysis results. Hence, it can be concluded As indicated above and by previous studies (Cao et al.
that the BMCOL method can reasonably simulate the 2005), there exists a critical padeye depth for a given
behavior of a suction caisson under a static load, pro- caisson embedment depth, which is usually at a depth
vided that stiffener p-y curves (e.g., Stevens & Audibert with a Zp/L ratio of 0.66 or larger. The caissons lateral
method 1979) are used in the BMCOL analysis. behavior is different when the padeye is set above, at
or below this point. The BMCOL method can be used
3.2 Padeye location optimization by suction caisson designers to find the padeye criti-
cal depth for a given caisson configuration in order to
3.2.1 Lateral capacity optimization optimize its lateral behavior.
Based on previous studies (Deng & Carter 2000; Cao As an example, the results of the BMCOL analyses
et al. 2005), a suction caissons maximum lateral cap- for different padeye depths are presented on Figures 7
acity is obtained when the loading point is set at about through 9. It can be seen from Figure 7 that the cais-
63 to 67% of its embedment depth. This loading point son exhibits an unfavorable rotational failure mode
is called the optimal padeye position. It is very impor- (toward the mooring line) when the padeye is located
tant for designers and operators to determine the opti- above the critical depth. Figure 8 shows the ideal
mal padeye position for a given caisson configuration, translational (or plowing) failure mode. Figure 9
as it results in the most cost effective caisson design. presents the caisson behavior with a padeye below the
We have used the BMCOL method extensively to critical padeye depth, for which the caisson translates
determine the optimal padeye position in suction and rotates backward. This backward-translational
caisson designs. Figure 6 presents the variation of the failure mode is preferred, as any possible tendency for
lateral capacity vs. padeye location obtained by the a separation between soil and caisson on the backside
BMCOL and the finite element analyses. The lateral of the caisson would be mitigated.

capacity was normalized by DLSu. As can be seen
from Figure 6, the normalized lateral capacity (Nk)
3.3 Shear force and bending moment distribution
reached a maximum value when Zp/L (ratio of padeye
along caisson wall
depth to total embedment depth) reaches 0.67. In other
words, the optimal padeye location is at a Zp/L value The shear force and bending moment distributions
of about 0.67, which is approximately two-thirds of along the caisson wall are required to perform the

328

Copyright 2005 Taylor & Francis Group plc, London, UK


Lateral Displacement, [m] Lateral Displacement, [m]
-0.2 -0.1 0 0.1 0.2 0.3 -0.2 -0.1 0 0.1 0.2 0.3
-6 -6

Distance Along Caisson From Mudline, [m]


SPAR
Distance Along Caisson From Mudline, [m]

Mudline SPAR Mudline


0 0

Original
6 Original Forward rotation. 6 caisson
caisson Not preferred. position
position
12 12
Backward rotation.
Preferred and
18 18 conservative.

24 24

30 30
Padeye located at 15.3 m BML Padeye located at 18.3 m BML
36 36

Figure 7. Padeye located above critical depth. Figure 9. Padeye located below critical depth.

thick wall section and internal stiffeners at the padeye


should be used in the structural design.
Lateral Displacement, [m]
-0.2 -0.1 0 0.1 0.2 0.3
-6 4 DISCUSSIONS
Distance Along Caisson From Mudline, [m]

Mudline SPAR
0 A common concern about the applicability of the
BMCOL method in suction caisson analyses is that
Original this method was developed based on the elementary
6 caisson Plowing. beam theory and primarily to be used for slender
position Ideal but difficult (long) piles. Suction caissons are usually stout, and
to guarantee.
12 they behave more like a rigid body under horizontal
loads and moments. Therefore, it is more appropriate
to treat the caisson as an ideal rigid rotation body by
18 assuming that the movement of the caisson is linearly
changing from the top to the bottom, especially when
24 the caisson is very stout (i.e., L/D 2).
The solutions using the BMCOL method are com-
pared with those using the rigid rotation theory for
30 caissons of various L/D ratios. Figures 10 and 11 pre-
Padeye located at 17.3 m BML sent the deflections, moments and shear forces for a
36 caisson with L/D ratio of 5. It is shown that the two
solutions almost coincide, except for the local area
Figure 8. Padeye located at critical depth. where the concentrated load is applied.
The second example is for a caisson with L/D ratio
of 2.5. The results are presented in Figures 12 and 13.
As shown in Figure 12, the rigid rotation theory pre-
structural design (i.e., wall thickness and stiffeners) dicts a much larger rotation than the BMCOL method.
during the design phase. These can be obtained from Also, the neutral depth (0 displacement) is deeper.
the BMCOL analysis results. Both the shear force Correspondingly, there is a noticeable difference in
distribution and the bending moment distribution the moment and shear force diagrams (Fig. 13).
(i.e., Fig. 11) in the wall along the caisson indicate The above analyses indicate that the BMCOL
that the maximum shear force and bending moment method has enough accuracy for relatively long cais-
occur in the wall at the padeye location. Hence, a son, when L/D is larger than 5. However, when the

329

Copyright 2005 Taylor & Francis Group plc, London, UK


Caisson Deflection (mm) Pile Deflection (mm)
-5 0 5 10 15 -160 -80 0 80 160 240
0 0
Rigid Rotation

Distance Along Caisson from Mudline, (m)


BMCOL
Distance Along Caisson from Mudline, (m)

2
5
4

10 Padeye Elevation
6
Padeye Elevation

15 8

10
20
12
Rigid Rotation
25 BMCOL 14

16
30
Figure 12. Comparison of deflection (L/D  2.5).
Figure 10. Comparison of deflection (L/D  5).

Moment (MN-m) Moment (MN-m)


-80 -40 0 40 80 -20 -15 -10 -5 0 5 10
0 0
Distance Along Caisson from Mudline, (m)
Distance Along Caisson from Mudline, (m)

2
5
Moment-Rigid Rotation
4 Moment-BMCOL
Moment-Rigid Rotation Shear-Rigid Rotation
Moment-BMCOL Shear--BMCOL
10 Shear-Rigid Rotation
Shear--BMCOL 6

15 8

10
20
12

25 14

16
30 -12 -9 -6 -3 0 3 6
-8 -4 0 4 8
Shear Force (MN)
Shear Force (MN)
Figure 13. Comparison of moment/shear diagram
Figure 11. Comparison of moment/shear diagram
(L/D  2.5).
(L/D  5).

L/D ratio is 2.5 or smaller, the end condition (restric- 5 CONCLUSIONS


tion) effect becomes prominent, and the results from
the BMCOL method are prone to noticeable bias The BMCOL method is a time-saving and economical
compared to those using rigid rotation theory. tool, which can be used very effectively and efficiently

330

Copyright 2005 Taylor & Francis Group plc, London, UK


to perform sensitivity studies of suction caissons (at Lateral Loads. Proceedings of 9th Annual Offshore
least at the preliminary design phase) to predict the Technology Conference, Houston, Paper 2953.
load vs. displacement, check the capacity, optimize Cao, J. 2003. Centrifuge Modelling and Numerical Analysis
the padeye location, and find the shear force and bend- of the Behaviour of Suction Caissons in Clay. Ph. D
thesis, Memorial University of Newfoundland, St, Johns,
ing moment distributions along the caisson wall. In NF, Canada.
order to effectively simulate the behavior of suction Cao, J., Audibert, J.M.E., Tjok, K-M. & Hossain, M. K.
caissons in the BMCOL analyses, the t-z and Q-z 2005. Validation of the Use of Finite Element Method for
curves can be developed using the API RP 2A (2000) Suction Caisson Design. International Symposium on
approach, but the p-y curves should be developed Frontiers in Offshore Geotechnics, September 1925,
according to the Stevens and Audibert method (1979). Perth, Western Australia.
The two major limitations of the BMCOL method Clukey, E.C. & Phillips, R. 2002. Centrifuge Model Tests
for suction caissons are: 1) the physical size of the to Verify Suction Caisson Capacities for Taut and Semi-
caisson cannot be taken into consideration, and 2) Taut Legged Mooring Systems. Proceedings, International
Conference on Deepwater Offshore Technology.
only in-plane load can be analyzed. It is suggested Deng, W. & Carter, J.P. 2000. Uplift Capacity of Suction
that a three-dimensional finite element analyses be Caissons in Uniform Soil. GeoEng 2000; Melbourne,
carried out to investigate the behavior of a suction Australia, papers on CD.
caisson at the final design phase. Huang, J., Cao, J. & Audibert, J.M.E. 2003. Geotechnical
Design of Suction Caisson in Clay. Proceedings of 13th
International Offshore and Polar Engineering Confer-
ence, Vol. 2, pp 770779.
REFERENCES Matlock, H. 1970. Correlations for Design of Laterally
Loaded Piles in Soft Clay. Offshore Technology
American Petroleum Institute (API) 2000. Recommended Conference, OTC 1204.
Practice for Planning, Designing, and Constructing Fixed Randolph, M.F. & House, A.R. 2002. Analysis of Suction
Offshore Platforms Working Stress Design, API Caisson Capacity in Clay, Offshore Technology Confer-
Recommended Practice 2A WSD (RP 2A-WSD), 21th ence, OTC 14236.
Ed., API, Washington, D.C. Stevens, J.B. & Audibert, J.M.E. 1979. Re-Examination of
Bogard., J.D. & Matlock, H. 1977. A Computer Program for P-Y Curve Formations. Proceedings of Offshore Technol-
the Analysis of Beam-Columns under Static Axial and ogy Conference, Houston, Paper OTC 3402.

331

Copyright 2005 Taylor & Francis Group plc, London, UK


Validation of the use of finite element method for suction caisson design

J. Cao, J.M.E. Audibert, K.-M. Tjok & M.K. Hossain


Fugro-McClelland Marine Geosciences, Inc., Houston, Texas, USA

ABSTRACT: This paper presents the development of a three-dimensional (3-D) finite element analysis
(FEA) model for caisson design, using the finite element code ABAQUS/Standard, to investigate the behavior
of suction caissons in clays. The 3-D FEA model can be used to effectively simulate the behavior of a suction
caisson under long-term sustained (loop current) loading conditions or under hurricane loading conditions. The
capabilities of this model are briefly introduced in this paper, in association with capacity check, optimization
of the padeye position, prediction of the axial load vs. displacement and lateral load vs. displacement curves,
investigation of out-of-plane loading (twist) effects, and determination of soil reactions (pressures) on the cais-
sons wall.

1 INTRODUCTION use of FEA and the accuracy of the FEA results. As


such, we felt there was a need to justify and validate
Suction caissons have proven to have several advan- the use of the FEA method for suction caisson design.
tages over other mooring/anchoring system for float- This paper presents the development of a three-
ing structures in both economical and technical aspects. dimensional model, using the finite element code
During the last decade, suction caissons have been ABAQUS/Standard, to investigate the behavior of
widely used as foundations for many types of off- suction caissons. The model was calibrated using load
shore structures (Huang et al. 2003), and for more than vs. displacement curves recommended by API RP 2A
36 fields in water depths as deep as 1650 m (5413 ft) (2000); i.e., a displacement of one percent of the cais-
(Andersen & Jostad 2002). As hydrocarbon explor- son diameter (0.01D) is needed to fully mobilize wall
ation and production head out into progressively skin friction and a displacement of ten percent of the
deeper waters, the oil and gas industry is making diameter (0.1D) is needed to fully mobilize the reverse
greater use of floating platforms (e.g., SPARs, semi- end bearing (REB). The capabilities of this model,
submersibles, deep draft caisson vessels (DDCVs), such as capacity check, optimization of the padeye
tension leg platforms (TLPs), etc.) for both drilling and position, prediction of the axial load vs. displacement
production units. Hence, suction caissons are expected and the lateral load vs. displacement curves, investi-
to play an important role in the future (Cao 2003). gation of out-of-plane loading (twist) effects, and soil
The behavior of suction caissons depends not only reactions (pressures) on the caissons wall, are dis-
on the specific site soil and environmental conditions, cussed below.
but also on the caissons structural configurations,
such as internal stiffeners (e.g., padeye stiffeners, lon-
gitudinal and ring stiffeners) and location of the
2 DEVELOPMENT OF A 3-D FINITE
padeye. It is important for designers and operators to
ELEMENT MODEL
understand the behavior of suction caissons (i.e.,
foundation strength and deformation parameters)
2.1 Finite element mesh and boundaries
during the design period. Therefore, a sensitivity
study is always required to evaluate the effects of dif- A 3-D finite element mesh (shown in Fig. 1) was con-
ferent structural configurations and various environ- structed to investigate the behavior of suction cais-
mental conditions on the caissons behavior. Although sons in clay. The mesh included 15 main soil layers
laboratory and centrifuge model tests and field tests with thicknesses that could be easily varied to charac-
can be used, sensitivity studies are usually carried out terize the various soil strata. For the soil in the vicin-
using 2-D or 3-D FEA, because FEA offer definite time ity of the caissons external wall and within the soil
and cost advantages. However, in some instances, plug, each main layer was further divided into four
some of the regulatory agencies have questioned the sub-layers to obtain a relatively fine mesh around the

333

Copyright 2005 Taylor & Francis Group plc, London, UK


the model was set to correspond to the undrained
shear strength of the clay. The clay was assumed to be
normally consolidated and total-stress based analyses
were performed. The undrained behavior of the clay
was simulated by setting the Poissons ratio (v) close
to 0.5.

2.3 Model for caisson wall material


The caisson wall material was modeled as an elastic
material, with Youngs modulus and Poissons ratio
representative of steel, i.e., E  2.0E8 kPa (2.9E7 psi)
and v  0.3. These parameters can be easily reassigned
when the caisson is made of concrete or other materials.

2.4 Simulation of loop current and hurricane


loading
The FEA model can be used to study the behavior of
suction caissons under loop current loading condi-
Figure 1. 3-D Finite element mesh and boundary tions or under hurricane loading conditions.
conditions. For a loop current (long-term sustained) loading
condition: the caisson is conservatively modeled as
caisson. The padeye could be located on the caissons a friction pile, as the reverse end bearing (REB)
wall at the designed elevation (i.e., any place between could be lost over time and, thus, cannot be counted
mudline and caisson tip). The soil mass was extended on. The caissons capacity is the sum of the external
to 20 times the caissons diameter below the caissons and internal friction resistances and the submerged
tip. Laterally, the far boundary was set at a distance weight of the caisson. Only a small displacement is
equal to 15 times the caissons diameter from the cen- required to fully mobilized skin friction. This behav-
ter of the caisson. Based on our experience with FEA, ior can be simulated by separating the soil plug top
the mesh boundaries were far enough from the caisson from the caisson cap in the finite element mesh.
to minimize any boundary effects. The side bound- For a hurricane (relatively short-term) loading
aries of the FEA mesh were assumed to be frictionless condition: the caisson capacity is the sum of the
surfaces in the vertical direction, and constrained external skin friction, the REB at the caisson tip and
from moving in the horizontal direction. The base of the submerged weight of the caisson. The REB com-
the mesh was fixed in both the vertical and horizontal ponent (about 40 to 60% of the total caisson capacity)
directions. For the soil mass, 20-node reduced inte- can be effectively mobilized when a suction caisson is
gration solid elements (C3D20R) were used. For the subjected to hurricane loading. This behavior is simu-
caisson, 8-node reduced integration shell elements lated using a hard contact between the soil plug and
(S8R) were used. As shown in Figure 1, thin elements the caisson cap in the finite element mesh. The REB
were used for the soil in direct contact with the out- is progressively mobilized as the caisson is pulled
side surface of the caisson in order to better model a out of the seabed. A relatively large displacement is
very narrow zone of soil that is disturbed during cais- required to fully mobilize the REB.
son installation and is highly sheared during loading.
A special feature (rigid bar elements) was used to
model the off-wall load application (i.e., padeye and 3 VALIDATION OF THE FINITE ELEMENT
shackle connection), so that out-of-plane loads could MODEL
be applied in the analyses. A total of 4419 biquadratic
elements and 16089 nodes were required to construct The standard t-z (friction vs. displacement) and Q-z
the mesh, generating 51255 solution variables. (REB vs. displacement) curves recommended by API
RP 2A (2000) are presently adopted in the suction
caisson design.
2.2 Constitutive model for soil material
For example, according to API RP 2A, the devel-
The well-known Drucker-Prager elastic-plastic mater- opment of the ultimate friction capacity along the
ial model available in ABAQUS/Standard (Hibbitt, internal/external skin of the caisson requires an axial
et al. 1998) was used for the clay soils, with the friction displacement equivalent to 0.01D (D being the caissons
angle (f) set equal to zero. Yielding of the material in diameter). On the other hand, the development of the

334

Copyright 2005 Taylor & Francis Group plc, London, UK


ultimate end bearing or REB may require a displace-
ment as large as 0.1D. These equivalent axial dis-
placements do not include the influence of the piles
deformation, which is negligible as the Youngs mod-
ulus of pile/caissons wall (steel) is about 7000 times
larger than that of soil. Therefore, these load vs. dis-
placement curves were used to calibrate the 3-D FEA
model.
A typical Gulf of Mexico clay material was used to
calibrate the FEA model. The undrained shear strength
profile was assumed to linearly increase from 2.4 kPa
(50 psf) at the mudline, at a rate of 1.5 kPa/m depth
(10 psf/ft). The submerged unit weight profile linearly
increases from 4 kN/m3 (25pcf) at mudline to 8 kN/m3
(50 pcf) at a depth of about 36 m (120 ft) below mud-
line (BML), and a constant value of 8 kN/m3 (50 pcf)
was used below this depth. The Youngs modulus of
the soil was assumed as 300 Su, where Su is the
undrained shear strength. A Poissons ratio of 0.45
was used for the soil.

3.1 Skin friction verification


In order to investigate the mobilization of the external Figure 2. Friction vs. displacement curve comparison
and internal skin friction resistance alone, the top of between FEA and API PR 2A (2000) results.
the soil plug was not connected to the caissons top at
mudline in the FEA mesh. The caisson was modeled
as a friction pile during the analysis, with negligible
(if any) reverse end bearing being mobilized.
The results from the finite element analyses are
presented in Figure 2 and indicate that the friction
resistance increases with the caissons displacement
until the ultimate friction resistance is fully mobilized
at a displacement of about 1.7% of the caissons diam-
eter. It can be noted that the load vs. displacement
curve from the FEA is slightly softer than that recom-
mended by API RP 2A (2000). Considering that the
required displacement to mobilize the ultimate fric-
tion is very small, the results from the FEA and from
the API RP 2A (2000) method are considered to be in
very good agreement.

3.2 Reverse End Bearing (REB) verification


In order to investigate the mobilization of the reverse
end bearing resistance alone, a hard contact between
the top of the soil plug and the caissons top at mud-
line was set in the FEA mesh. In addition, the soil elem-
ents around the caissons outside wall were assigned
negligible shear strength, so that the contribution of
the skin friction resistance to the total capacity would Figure 3. REB vs. displacement curve comparison
be negligible. between FEA and API PR 2A (2000) results.
The FEA results are presented in Figure 3. As
expected, much larger displacements are required to ratio of 50%, the load vs. displacement results from
develop the ultimate REB component, as compared to the FEA are in very good agreement with those rec-
those required to develop the ultimate skin friction. ommended by API RP 2A (2000). Beyond this
As can be seen in Figure 3, up to a REBmob/REBmax point, the FEA results show a larger displacement

335

Copyright 2005 Taylor & Francis Group plc, London, UK


than would be indicated by the API RP 2A (2000) out-of-plane twist. The mooring line load vs. vertical
criterion. However, the difference between the FEA displacement and the mooring line load vs. lateral
results and the values recommended by API RP 2A displacement curves can be obtained, and the pullout
(2000) is generally less than 15%, and is on the con- capacity of the caisson can be determined based on
servative side. Thus, the results are considered quite the load vs. displacement curves. Usually, the caisson
acceptable for geotechnical engineering. capacity obtained by the FEA is slightly less than that
obtained by the limit equilibrium analysis.
3.3 Summary
4.2 Padeye depth optimization and lateral behavior
The load vs. displacement curves obtained from the
3-D finite element analyses for both skin friction and An important advantage of suction caissons is their
reverse end bearing indicate slightly softer responses ability to resist substantial lateral loads, common in
than those recommended in API RP 2A (2000). Since the offshore environment. In general, a suction cais-
the results obtained by the FEA model are on the con- son will achieve its maximum holding capacity when
servative side, they are considered reasonable and it exhibits a desired translational (plowing) failure
acceptable. mode. Several studies (Randolph & Houlsby 1984,
Murff & Hamilton 1993) were carried out regarding
the behavior of piles subjected to lateral loading applied
4 APPLICATION OF THE 3-D FINITE at the top of the pile. However, these results may not
ELEMENT MODEL IN SUCTION CAISSON be directly applicable to suction caisson anchors for
DESIGN two reasons: 1) the padeye location for suction cais-
son anchors is usually BML, and 2) suction caissons
4.1 Capacity confirmation are shorter and stiffer than piles, resulting in a differ-
ent behavior. A finite element study on the lateral
The behavior of a suction caisson varies with differ- behavior of suction caissons was performed (Deng
ent loading conditions (i.e., loop current vs. hurri- and Carter 2000) using an aspect ratio (ratio of
cane). For the present state-of-practice, the axial embedment to diameter, L/D) ranging from 1 to 2.5
pullout capacity of a suction caisson is estimated and a ratio of padeye depth BML to embedment
using the limit equilibrium method, adapted from API (Zp/L) varying from 0 to 1. They found that the
RP2A (2000), as follows: normalized lateral capacity Nk (normalized by
For loop current loading conditions: the available
L  D  S u) did not change much with the aspect
caisson resistance to uplift is the sum of the external ratio, and the optimal padeye location was roughly at
and internal skin friction and the caissons submerged Zp/L  0.63.
weight: The behavior of suction caissons with larger aspect
ratio values (i.e., L/D ranging from 6 to 30) was
(1) investigated by the 3-D FEA model and the results are
presented in Figure 4. It can be seen from this figure
where, Qtotal is the total axial pullout capacity, Qexternal that the optimal padeye position is at a Zp/L value
and Qinternal are the external and internal skin friction, of about 0.67, or at approximately two-thirds of the

respectively, W caisson is the submerged weight of the embedment depth.
caisson.
For hurricane loading conditions: the available
caisson resistance to uplift is the sum of the external
skin friction, the REB at the caisson tip and the cais-
sons submerged weight:

(2)

As mentioned earlier, several factors could reduce


the caissons capacity (Huang et al. 2003). Some of
them, including set-up effect, trench effect, NGI fac-
tor (NGI 1999), and post-peak reduction factor, can
be evaluated and estimated from the caisson geome-
tries and various laboratory soil tests.
A 3-D finite element analysis can incorporate all
the factors mentioned above, as well as the effect of Figure 4. Lateral load vs. padeye location curve.

336

Copyright 2005 Taylor & Francis Group plc, London, UK


Based on a series of FEA runs with various Zp/L caisson being pulled out should the applied loads
ratio values, three typical caisson failure modes were exceed the capacity of the caisson, and is the pre-
identified, as presented in Figure 5. As can be seen, ferred behavior.
there exists a critical padeye depth. When the padeye
load is applied at a point above the critical padeye
4.3 Twist effect on axial capacity
depth, as shown in Figure 5 (b), the top part of the
caisson moves towards the SPAR (or floating struc- Often, a suction caisson is subjected to an out-of-
ture) and the caisson bottom moves backward away plane mooring load, and its holding capacity could be
from the SPAR. The caisson exhibits a somewhat slightly reduced due to twist effect. The reduction in
unfavorable (i.e., forward rotation, and not translation capacity depends on the site soil conditions and on
as would be desired) failure mode. Such a padeye the twist angle (equivalent to the rotation angle of the
location is considered unacceptable. When the load is caisson during its installation). The twist effect on
applied at the critical depth, the caisson uniformly caisson capacity was investigated effectively using a
translate without rotation (Fig. 5 (c)), and there is 3-D finite element analysis.
concern that a gap might develop at the back of the Figure 6 presents the load vs. displacement curves
caisson, resulting in a possible decrease in capacity. from a 3-D finite element analysis for three twist
When the padeye location is slightly deeper than the angles under a loop current loading condition (i.e.,
critical padeye depth, as shown in Figure 5 (d), the cais- the caisson behaves as a friction pile). The maximum
son rotates, with its top moving away from the SPAR capacity reduction is about 3 and 8% for twist angles
and the caisson bottom moving forward toward the of 5 and 10 degrees, respectively.
SPAR. The caisson exhibits a slightly counter-rotational Additional FEA studies indicated that, when a suc-
failure mode. This mode reduces the possibility of a tion caisson is subjected to a hurricane loading condi-
tion, the caissons holding capacity is reduced by
about 2 and 5% for twist angles of 5 and 10 degrees,
respectively. The caisson capacity reduction for a hur-
ricane loading condition is less than that for a loop
current loading condition because the reverse end
bearing (up to 60% of the overall capacity) is almost
not influenced at all by the twist effect.

Figure 5. Caisson lateral behavior vs. padeye location. Figure 6. Twist effect on caisson axial capacity.

337

Copyright 2005 Taylor & Francis Group plc, London, UK


4.4 Soil reactions on caisson wall for shown in Figures 7 (c) and (d), the reduction in both
operation phase outside circumferential pressure and the outside
downward friction from about 16.5 m to the caissons
Soil reactions (pressures) against the suction caissons
tip was due to the NGI reduction factor of 0.65
wall are required for the caisson structural design
applied to the external skin friction for the suction
(i.e., installation phase, removal phase and operation
assisted penetration portion. Likewise, the slight
phase). To find the soil reactions on the caissons wall,
increase in the outside circumferential pressure and
limit equilibrium based methods are often used for the
the outside downward friction at and below 18.5 m
caisson installation and removal phases. However,
BML was due to the effect of a reduction coefficient of
such methods are quite limited and may not capture
0.92 (trenching factor) applied to the external skin fric-
the complexities of the boundary conditions. For the
tion for the portion from mudline to the padeye depth.
operation phase, a 3-D finite element analysis is highly
recommended.
As discussed above, many factors influencing the
behavior of suction caissons could be incorporated in 5 CONCLUSIONS
the 3-D finite element analyses. The FEA analyses
can provide the soil reactions (pressures) on the cais- This paper has presented the development and appli-
sons wall, including radial inward pressure on the cation of a three-dimensional model using the finite
external wall, radial outward pressure on the internal element code ABAQUS/Standard for simulating the
wall, downward friction and circumferential friction behavior of suction caissons. The applications of this
on the external wall and the internal wall, and down- model for suction caisson designs was also discussed
ward pressure on the caisson top. These pressures can in the paper. The following conclusions can be drawn:
be used for the structural design of the caisson.
Examples of the soil reactions obtained through The 3-D FEA model can be used effectively to
FEAs are shown in Figure 7. accurately simulate the behavior of suction caissons
As mentioned earlier, various capacity reduction under various loading conditions (e.g., loop current
factors can be incorporated in the FEA model. As loading and hurricane loading), including pre-
diction of load vs. displacement curves, capacity
check, optimization of the padeye location, investi-
gation on the out-of-plane loading (twist) effect,
and soil reactions (pressures) on the caisson wall.
The load vs. displacement curves for both friction
and reverse end bearing by the FEA are in good
agreement with the methods adapted from API RP
2A (2000).
The capacity obtained by the finite element analysis
is slightly smaller than that by limit equilibrium
methods because several factors affecting the capac-
ity can be incorporated in the 3-D FEA model. This
is important when evaluating factors of safety.
For twist angles of 5 and 10 degrees, the caisson
capacity is reduced by about 3 and 8% under loop
current loading, and by about 2 and 5% under hur-
ricane loading.

ACKNOWLEDGEMENT

The authors gratefully appreciate fugros manage-


ment support in the publication and presentation of
this paper.

REFERENCES

American Petroleum Institute (API) 2000. Recommended


Figure 7. Soil reactions along caissons wall by FEAs. Practice for Planning, Designing, and Constructing Fixed

338

Copyright 2005 Taylor & Francis Group plc, London, UK


Offshore Platforms Working Stress Design. API Hibbitt, Karlsson & Sorensen, Inc. 1998. ABAQUS/Standard
Recommended Practice 2A-WSD (RP 2A-WSD), 21th Users Manual. Volume II, Version 5.8.
Ed., API, Washington, D.C. Huang J., Cao J. & Audibert J.M.E. 2003. Geotechnical
Andersen K.H. & Jostad H.P. 2002. Shear Strength Along Design of Suction Caisson in Clay. Proceedings of 13th
Outside Wall of Suction Anchors in Clay After International Offshore and Polar Engineering
Installation. Proceedings of 12th International Offshore Conference, Vol. 2, pp. 770779.
and Polar Engineering Conference, Kitayushu, Japan, Murff J.D. & Hamilton J.M. 1993. P-Ultimate for Undrained
ISOPE, Vol. 2, pp. 785794. Analysis of Laterally Loaded Piles. Journal of Geo-
Cao J. 2003. Centrifuge Modelling and Numerical Analysis technical Engineering, ASCE, Vol. 119(1), pp. 91107.
of the Behaviour of Suction Caissons in Clay. Ph.D Norwegian Geotechnical Institute (NGI) 1999. Set-up
thesis, Memorial University of Newfoundland, St. Johns, Effects Outside Skirt Wall Skirted Foundations and
NF, Canada. Anchors in Clay. Joint Industry Sponsored, May 11.
Deng W. & Carter J.P. 2000. Uplift Capacity of Suction Randolph M.F. & Houlsby G.T. 1984. The Limiting Pressure
Caissons in Uniform Soil. GeoEng 2000; Melbourne, on A Circular Pile Loaded Laterally in Cohesive Soil.
Australia, papers on CD. Gotechnique. 34(4), pp. 613623.

339

Copyright 2005 Taylor & Francis Group plc, London, UK


Developments in the Australian frontiers

Copyright 2005 Taylor & Francis Group plc, London, UK


Geotechnical interpretation for the Yolla A Platform

P.G. Watson & C. Humpheson


Arup, Houston

ABSTRACT: The Yolla A Platform was installed in the Bass Strait (Australia) in March 2004. The platform
is a steel gravity based structure, and is founded on a complex soil strata comprising primarily calcareous silt.
Similar material has not been encountered previously, and an intensive soil characterization program was under-
taken during the design process. This paper presents aspects of the soil testing undertaken, focusing on testing
of the calcareous silt. A comparison of insitu and laboratory testing performed using equipment and samples
from different investigation contractors is also presented.

1 INTRODUCTION

The Yolla A Platform was installed in the central Bass


Strait (Australia) in March 2004 as part of the BassGas
Project, for client/operator Origin Energy. The site loca-
tion is shown in Figure 1.
The Yolla A Platform is shown in Figure 2. The
platform is a self installing steel gravity base platform,
and is one of Arups ACE range of offshore platforms.
The foundation comprises a skirted foundation, with
skirts penetrating 5.4 m into the seabed (forming 12
individual compartments in the base).
This paper provides:
A brief overview of the site investigations per- Figure 1. Site location.
formed at the site;
An overview of the insitu soil testing, focusing on
the CPT testing performed by different operators,
as well as T-bar testing performed in the final
investigation;
A summary of some of the laboratory test results,
focusing on simple shear testing on samples
recovered using different soil sampling techniques.
Full details of the site investigation and soil charac-
terisation performed as part of the Yolla A Platform
design are to be documented in a journal paper cur-
rently in preparation. (a) Artist impression (b) Prior to sailout, Batam (Indonesia)

Figure 2. The Yolla A platform.


2 OVERVIEW OF SOIL INVESTIGATIONS
characterise the soil, and provide geotechnical parame-
PERFORMED
ters for design.
The following insitu testing and soil sampling was
Soils encountered at the Yolla A Platform location con-
carried out at the platform location:
sist of interbedded calcareous sandy silts, sands and
clay. Other (intermediate) soil types are also encoun- Thales Geosolutions [in August 2001]: Miniature
tered. Numerous soil studies have been undertaken to cone penetrometer tests (MCPTs) with tip

343

Copyright 2005 Taylor & Francis Group plc, London, UK


area  1.75 cm2. Nine tests performed to 10 m qnet (MPa) qnet (MPa)
0 0.5 1 1.5 2 0 2 4 6
depth. 0 0
Benthic Geotech [in October 2001]: Cone penetrom-
eter tests (CPTs) using a standard cone with tip Benthic
area  10 cm2. Two tests were performed from 2 10 Thales
Fugro
018 m depth, and one test from 3045 m Depth.
Limited soil sampling undertaken to 100 m depth,

Depth (m)

Depth (m)
4 20
with samples of 44 mm diameter (typical) recovered.
Fugro [in October 2002]: CPT testing using a large 6 30
cone with area  15 cm2, with five tests performed
to 30 m depth. CPT testing included dissipation test-
8 40
ing. T-bar testing (4 tests) up to 30 m, including
cyclic testing. Recovery of 90 mm diameter undis-
turbed samples using a gravity piston corer to 10 50
approximately 13 m.
Figure 3. Summary of PCPT results.
Conventional characterization testing was carried out
by Douglas Partners and Western Geotechnics, whilst Although good agreement was observed, it was
specialist laboratory testing was completed at the noted that the qnet results obtained by Thales Geosolu-
Centre for Offshore Foundation Systems (COFS) at tions are higher than those obtained by either Benthic
the University of Western Australia. Geotech or Fugro. In addition, the Fugro results are
marginally lower than those obtained by Benthic
Geotech. Potential reasons for the variations include
3 SUMMARY OF SOIL TYPES (but are not limited to):
ENCOUNTERED
Possible influence of the large shells on the CPT
Four primary (interlayered) soil types have been iden- results. This is expected to influence the mini
tified at Yolla: penetrometer more than either of the larger
penetrometers.
Sandy SILT, such as found at skirt tip level (5.4 m). Influence of thin layers/stratification. Lunne et al.
Very Sandy SILT, dominant from skirt tip level to (1997) discusses the fact that smaller diameter
around 10 m. probes attain the maximum cone resistance for a sin-
CLAY, such as encountered at 1.63.2 m below gle layer with less penetration distance. In stratified
seabed. soils, this implies smaller probes may respond
SAND, encountered at 3.2 m, and again at various quicker to changes in stratigraphy than larger probes.
depths through the stratigraphy. The layer at 3.2 m The result is likely to be greater extremes of qnet in
includes significant large (oyster) shells and other the Thales results, which values are expected to be
molluscs. closer to the actual strength of the individual layers.
Other (intermediate) soil types are also encountered The effect of drainage. This is believed to be the
at Yolla, although focus of the interpretation was on most significant factor in the results observed, and
the above soil types. is now discussed in more detail.
Of the above soil types, it is the Sandy SILT and The degree of drainage may be assessed using the non-
Very Sandy SILT that are most critical to foundation dimensional ratio vD/cv, where v is the penetration
design these soil types being (predominantly) found rate, D is the probe diameter and cv is the coefficient of
from skirt tip depth and through the zone of influence consolidation. Unpublished centrifuge test data from
for foundation design. Select test results in these Watson (1999) suggests that, for the range of vD/cv
materials are presented in this paper. expected for the different cone sizes, some difference
in cone resistance may be expected. The centrifuge
data, performed on calcareous silt using a model size
4 SUMMARY OF IN-SITU TESTING
cone penetrometer, suggests that the likely ratio
between fully undrained (excluding viscous effects)
4.1 Cone penetrometer testing
and fully drained cone response is in the range 6 to 8.
All CPTs were performed at a constant rate of 2 cm/s. This range in soil response is comparable to results
A summary of the average results obtained from each published by Hight et al. (1994), who presented results
operator is shown in Figure 3. Note that individual from CPTs on clayey sands. Data from this paper, and
test results showed very repeatable results, with little from testing at Yolla, is plotted on Figure 4 which
variability. shows normalized cone resistance versus normalized

344

Copyright 2005 Taylor & Francis Group plc, London, UK


80 Nk
CP1 (3.5 - 10 m) 0 10 20 30
0
70
Data from
60 Hight et al. 5

Nk = 20
50
10

Depth (m)
qt/'v

40

30

Nk = 14
15

20
20
10

0 25
-0.2 0 0.2 0.4 0.6 0.8 1
Bq Figure 6. Interpreted Nk from CPT data.

Figure 4. Effect of drainage of tip resistance.


The measured resistance to penetrate the T-bar is
converted to soil strength simply by dividing the resist-
su (kPa) qtbar E / qtbar P ance by a T-bar factor (Nt). Figure 5 shows the undrained
0 50 100 150 200 0 0.5 1 shear strength profile derived from the T-bar results
0 0 assuming a value of Nt  10.5. This is a traditional
value supported by testing performed in the laboratory
4 4 and at (controlled) onshore sites, and is documented
TB1 in Stewart and Randolph (1994) and Watson (1999).
8 TB2 8 Figure 6 shows the interpreted Nk from the Yolla
TB3
data, determined by comparing the T-bar and CPT
Depth (m)

Depth (m)

TB4
12 Average 12 data. The range in values is consistent with the data
reported in Lunne et al. (1997).
16 16

20 5 LABORATORY TESTING
20

24
5.1 Monotonic simple shear testing (DSS)
24
This section presents selected results from monotonic
Figure 5. Summary of T-bar results. simple shear tests.
Stressstrain responses from monotonic DSS test-
ing on the Very Sandy SILT are presented in Figure 7.
pore pressure, Bq. The results are (surprisingly) simi- Similar results were determined for the Sandy SILT,
lar, and show a potential increase in cone resistance although less dilation was observed for this material.
up to 8 fold. Figure 7 also shows the design line adopted.
The influence of drainage on CPT results was a Key observations from the monotonic DSS testing
major consideration in the foundation design, as this include:
was taken to be an indicator of the potential for
drainage to influence the skirt penetration resistance. The responses collapse into a narrow band when
normalised by the insitu vertical stress (vo).
The stressstrain response comprises an initial stiff
4.2 T-bar testing response at low strain, followed by a gradual
The T-bar is a tool specifically designed for investiga- increase in the observed shear stress as strain
tion of soft sediments, and previous examples detail- increases. The increasing shear strain is associated
ing its use offshore are provided in Randolph et al. with dilation, with peak shear strength generally
(1998) and Hefer and Neubecker (1999). observed at shear strains of greater than 30%.

345

Copyright 2005 Taylor & Francis Group plc, London, UK


1.4 su (kPa)
0 50 100 150 200
1.2 0
Tests at = 30% (DSS)
lower 'vo
Stress ratio (/'vo)

1.0

CIU tests 5
0.8

0.6 CPT (Nk = 16)


Tbar (Nt = 10.5)

0.4 10 DSS

Benthic samples

Depth (m)
0.2 Fugro samples
Design line 15
0.0
0 5 10 15 20 25 30
Strain (),%
(a) Normalised stress strain response
20
0.8

0.6
25
Stress ratio (u/'vo)

0.4

0.2
30
(a) Maximum strain level ( = 30 %)
0.0
0 5 10 15 20 25 30
su (kPa)
-0.2
0 50 100 150 200
-0.4 0
Benthic samples = 12 % (DSS)
Fugro samples
-0.6
Strain (), %
5
(b) Normalised pore pressure strain response

Figure 7. Monotonic DSS: very sandy SILT. CPT (Nk = 16)


Tbar (Nt = 10.5)
10
DSS

The influence of dilation is confirmed in the pore


pressure response, which demonstrates a rapid
Depth (m)

increase in pore pressure at low strain, followed by 15


a gradual reduction with increasing strain.
Although not shown, the amount of dilation is less
in the Sandy SILT samples.
There appears to be little difference between the 20
test results on Benthic and Fugro soil samples.
This is despite differences in the sampling and soil
handling procedures, and the DSS sample test size.
The Very Sandy SILT tests show a potential stress 25
effect. This is broadly consistent with recent test-
ing and interpretation which has demonstrated that
vertical stress plays a significant role in determin-
30
ing the stressstrain response of dilatant calcar-
eous sands, as reported in Finnie and Hefer (2002). (b) Intermediate strain level ( = 12 %)

Figure 8 compares the measured T-bar and CPT soil Figure 8. Comparison between insitu and laboratory
strength profiles with the strength derived from the strength measurement.

346

Copyright 2005 Taylor & Francis Group plc, London, UK


monotonic soil strength (a) at the maximum strain in
the DSS (  30%); and (b) at a strain level of B =ave/ cyc
max = cyc + ave
  12%. The comparison has been made for all mater- 0.5

Applied shear stress (t)


ial types encountered (not just the Very Sandy SILT 0.45
discussed above). cyc
0.4
It is clear that (for the silt material) the DSS results
at high strain result in a higher undrained soil strength 0.35 ave

than the insitu testing, and that improved comparison is 0.3


Time

observed using a   12%. Note that reasonable com-

cyc/'vo
parison was observed for the clays for both strain 0.25

levels, as the simple shear results (performed on clay 0.2


samples) do not show any significant dilation, and the
strength at   12% is close to the maximum obtained. 0.15 Design line

The reasons for obtained good agreement between 0.1


insitu and DSS results at   12% are poorly under- Benthic sample
0.05 Fugro sample
stood. However, in reviewing these results, the fol- Monotonic (design)

lowing is noted: 0
1 10 100 1000
Although the T-bar and CPT results shown are N
assumed to represent undrained penetration (a) Direct comparison of results
response in the various silt and clay layers, there is 0.5
Error bars indicate range of
potential for some drainage effects to influence 0.45 observed monotonic response
these results. All DSS results are undrained. (from Benthic and Fugro samples)

The laboratory strength of the SAND layers (such 0.4


as at 4 m) is derived from the response of the Very 0.35
Sandy Silt, since no DSS tests was performed on
this material. It is therefore not expected that the 0.3
cyc/'vo

laboratory and insitu strength profiles will agree 0.25


for these soil layers.
0.2
Although plotted against the maximum strain
achieved in the simple shear apparatus (  30%), 0.15
this is not necessarily the peak obtainable shear
0.1
strength. Laboratory results at high strain become
somewhat questionable, as the apparatus can influ- 0.05
ence the results obtained.
0
Calcareous soil deposits generally exist (insitu) in 1 10 100 1000
the form of an open (high void ratio) matrix, pri- N
marily resulting from the depositional environ- (b) Including reduction factor on Benthic results
ment and grain type. At high strain (greater than
  30%), it is possible that this open soil struc- Figure 9. Comparison between Benthic and Fugro DSS
ture may collapse, leading to very low remoulded results, Very Sandy SILT (data points are number of cycles
strengths. At Yolla, this conclusion is supported by at constant cyc/vo to reach cyc  5%).
the cyclic T-bar tests (not reported in this paper),
which showed the remoulded strength at high 5.2.1 Comparison between Benthic Geotech and
strain to be very low in the calcareous silt layers. Fugro soil response
In collapsing soils, it is therefore possible that the Figure 9 presents select results from cyclic DSS test-
CPT and T-bar measure soil strengths somewhat ing on the Very Sandy SILT.
lower than the peak obtained in simple shear. For Results are presented in the form of S-N curves, plot-
the Yolla silts, this appears to coincide with the ting observed number of cycles (N) required to reach a
stress measured in the simple shear apparatus at certain amount of cyclic strain (cyc) at different cyclic
  12%. shear stress ratio (cyc/vo). The results in Figure 9 are
for cyc  5%. Also plotted is the monotonic strength at
Further work is required to explore the above issues.
N  1, taken from the design line derived for mono-
tonic simple shear loading. The following is noted:
5.2 Cyclic simple shear testing
Although some scatter is evident in the data, a
This section presents select results from cyclic simple fairly consistent response is observed, enabling
shear testing. design lines to be drawn as shown.

347

Copyright 2005 Taylor & Francis Group plc, London, UK


Unlike the monotonic DSS testing, where samples 0.5
Monotonic (Insitu)
recovered from the Benthic Geotech and Fugro Benthic (Insitu)
investigations showed similar strength, the Benthic 0.45
Fugro (Insitu)
Geotech samples seem to indicate slightly higher 0.4 Monotonic (Insitu + Platform)
cyclic soil strength than the Fugro samples. Figure Benthic (Insitu + Platform)
Fugro (Insitu + Platform)
9(b) show the data including a constant factor of 0.85 0.35
applied to all the Benthic data (i.e. 0.85  cyc/vo).
0.3

cyc/'vo
As shown, there is good agreement between the two
data sets when the factor is included. 0.25
The reason for the (relatively small) difference 0.2
between the Benthic and Fugro cyclic data is unclear.
Potential reasons include: 0.15

Problems were encountered with vertical stress con- 0.1


trol in the Benthic Geotech cyclic testing, due to the
0.05
small sample size. It was observed that a greater Range of insitu stress = 30 - 38 kPa
fluctuation in total vertical stress occurred during Range of insitu + platform stress = 55 - 72 kPa
0
the Benthic Geotech cyclic tests, as the apparatus 1 10 100 1000
struggled to maintain a constant vertical stress at the N
(high) strain rates used. Better control was achieved
in the Fugro tests, associated with the larger sample Figure 10. Influence of consolidation stress, Very Sandy
size. It is unclear whether this would necessarily SILT.
result in the Benthic Geotech samples overestimat-
ing the cyclic strength, but it is nevertheless one
potential difference between the two data sets. 0.50
Potential effect of the end platens. All tests were Error bars indicate range of
performed using end platens with skirts only (no 0.45 observed monotonic response

pins or ridges). Previous experience has shown that 0.40


inadequate platen support in cyclic tests may result Benthic (B < 0.1)
Benthic (0.1 < B < 0.3)
in premature failure of the sample along the inter- 0.35
Fugro (B < 0.1)
face between the platens and the sample, rather
0.30 Fugro (0.1 < B < 0.3)
than within the soil sample itself.
cyc/'vo

Fugro (0.3 < B < 0.5)


In the current study, all samples were tested using 0.25 Monotonic
skirted platens without additional pin supports. It
is possible that, due to the lower aspect ratio of the 0.20
skirts, the larger Fugro samples may be influenced
0.15
by potential platen effects to a greater extend than
the Benthic Geotech samples. This would imply 0.10
that the true cyclic strength is underestimated with
the Fugro Survey samples. 0.05

0.00
5.2.2 The effect of vertical consolidation stress 1 10 100 1000
Figure 10 illustrates the effect of consolidation stress on N
the cyclic tests performed on Very Sandy SILT (Benthic
Geotech test results include the 0.85 reduction factor). Figure 11. Influence of load bias on cyclic response, Very
The results appear consistent with the monotonic Sandy SILT.
results, showing a distinct influence of vertical stress on
soil response, with tests performed at the higher vertical
stress giving lower (normalized) soil strength. influence on the observed cyclic stress strain
response. This is broadly consistent with studies
5.2.3 The effect of load bias undertaken by NGI, and reported in various papers
Figure 11 examines the effect of load bias on the including Andersen (1992); Andersen et al. (1988).
observed cyclic DSS response for the Very Sandy SILT The cyclic DSS tests have been used to derive S-N
(B  0 represents pure 2-way loading and B  1 rep- contours for cyclic strains in the range 0.5 to 15%.
resents pure 1-way loading: see definition in Figure 9). The results are presented in Figure 12 for the Very
The data demonstrates that, for the range of bias Sandy SILT. A similar design curve has been derived
considered, application of average stress has little for the Sandy SILT material.

348

Copyright 2005 Taylor & Francis Group plc, London, UK


0.6 Yolla A Platform location, Bass Strait (Australia).
10% These materials have not been encountered, or inves-
tigated, elsewhere.
Results presented include insitu CPT and T-bar
0.5
testing performed by different investigation contract-
ors, and monotonic and cyclic simple shear testing.

0.4 5%
REFERENCES
cyc/'vo

3%
Andersen, K.H. 1992. Foundation design of offshore gravity
0.3
2% structures. Cyclic Loading of Soils, M.P. OReilly & S.F.
1% Brown (eds), Blackie and Son Ltd.
Andersen, K.H., Kleven, A. and Heien, D. 1988. Cyclic soil
0.2 0.50% data for design of gravity structures. ASCE, J. of
Geotech. Eng., 114(5) 517539.
0.25% Finnie, I.M.S. and Hefer, P. 2002. The cyclic resilience of
calcareous sediments. Draft copy provided for review.
0.20%
0.1 Hefer, P. and Neubecker, S. 1999. A recent development in
0.15% offshore site investigation tools: The T-bar. Proc. 1999
Australiasian Oil and Gas Conference, Perth.
Hight, D.W., Georgiannou, V.N. and Ford, C.J. 1994.
0 Characterisation of clayey sands. Proc. 7th Int. Conf.
1 10 100 1000 Behaviour of Offshore Structures, BOSS94, Pergamon,
N
321340.
Figure 12. Cyclic strain contour diagram, Very Sandy SILT. Lunne, T., Robertson, P.K. and Powell, J.J.M. 1997. Cone
Penetration Testing. Blackie Academic and Professional.
Randolph, M.F, Hefer, P.A., Geise, J. and Watson, P.G. 1998.
Improved seabed strength profiling using T-bar pen-
Although not reported here, the S-N curves have etrometer, Proc. Int Conf Offshore Site Investigation and
been used to determine cyclic soil strength for use in Foundation Behaviour New Frontier, SUT, London.
traditional foundation design calculations. Stewart, D.P. and Randolph, M.F. 1994. T-bar penetration
testing in soft clay. J. of Geotechnical Eng. Div, ASCE,
120 (12), 22302235.
Watson, P.G. Unpublished centrifuge data, 1999. The
6 CONCLUSIONS Univeristy of Western Australia.
Watson, P.G. 1999. Performance of skirted foundations for
This paper presents select results from a soil charac- offshore structures, PhD Thesis, The University of
terisation study performed on material found at the Western Australia.

349

Copyright 2005 Taylor & Francis Group plc, London, UK


Preloading of drag anchors in carbonate sediments

S.R. Neubecker, M.P. ONeill & C.T. Erbrich


Advanced Geomechanics, Perth, Western Australia

ABSTRACT: Drag anchors are often used to provide a fixed anchorage for offshore floating facilities. As part
of the mooring installation, drag anchors are required to be preloaded. The preloading embeds the anchor
beneath the seabed and also defines a threshold load, such that the anchor will not move for applied loads below
this level. In the past, simple guidelines have been used to define the level of preload required. However, with
the recent development of a more detailed analysis method by DNV, the implications for anchors in carbonate
silty soils are significant. This paper discusses how the DNV method may be applied to preloading drag anchors
in carbonate sediments, and some of the implications that arise.

1 INTRODUCTION to the quasi-static load, or 80% of the intact dynamic


load as specified by Lloyds (1999).
For permanently moored offshore oil and gas facil- However, when applied to carbonate soils, the
ities there is often a strict requirement on the amount DNV method may lead to much higher preload levels,
of movement of the central mooring point that can be due to the greater susceptibility of the soil to cyclic
sustained by the system. This is usually because there degradation.
are flexible flowlines that rise from the seabed to the This paper will discuss the philosophy behind the
floating structure, and to ensure that these risers are DNV method, and how it may be applied to drag
not overstressed the watch-circle of the vessel needs anchor design and preload in carbonate soils.
to be limited.
Consequently, fixed anchors are provided for the
anchor lines, and these must generally be designed so
that there is no movement (or slippage) during the life 2 METHODOLOGY
of the facility. The consequences of anchor movement
are potentially disastrous. In the case of piles, gravity The philosophy outlined in RP-E301 is that the drag
boxes, and suction caissons, the anchor point can be anchor should not move under the design loads and
safely designed so as not to move significant distances that this should be guaranteed by the adoption of
under the design loads. However, when using drag appropriate factors of safety and preload levels. The
anchors, there is the potential that some movement may UHC (or characteristic anchor resistance), Rc, of a
occur, particularly under cyclic loading. To ensure drag anchor comprises the sum of the installation
that this does not happen, drag anchors are required to anchor resistance, Ri and the predicted post-installation
be preloaded to a high enough level so that in-service effects of consolidation (Rcons) and cyclic loading
movements are limited to an acceptable level. (Rcy);
The design procedure outlined in the DNV Recom-
mended Practice RP-E301 Design and Installation of (1)
Fluke Anchors in Clay (DNV 2000) presents a model
to assess the performance of drag anchors in cohesive
seabeds which is based on sound theoretical princi- The full equation for the UHC given in RP-E301
ples. Specifically it enables the determination of the also includes an additional term due to the possible
amount of preload that will give the same anchor sys- seabed friction, but for simplicity, this component in
tem reliability after explicitly catering for cyclic the above equation has been ignored.
degradation of the soil at the site. RP-E301 is based on a limit state design approach,
When applied to traditional clay soils, the DNV and since the above equation for the UHC does not
method typically results in similar preload recommen- include any factor of safety it is recast including a
dations as other simpler approaches, such as preloading material factor (m) as;

351

Copyright 2005 Taylor & Francis Group plc, London, UK


(2) set-up of the remoulded soil adjacent to the surface of
the anchor will occur, and the full shear strength will
ultimately be reinstated.
Note that in this equation m is only applied to the However, evaluating the effect of consolidation is not
post-installation effects, since it is assumed that Ri as simple as saying that the anchor will set-up by an
has been accurately measured during installation and amount equal to the soil sensitivity. This is because the
hence has no uncertainty associated with it. anchor resistance is divided into two geotechnical com-
The design line tension for the drag anchor is ponents; forces acting normal to the fluke, and forces
defined as; acting parallel to it (or shear forces). It is only the shear
forces acting on the anchor surfaces that will set up.
(3) The normal forces will essentially remain unchanged
as they are not a result of full soil remoulding.
where Tc-mean is the characteristic mean line tension, RP-E301 recommends values for Ucons for drag anchors
Tc-dyn is the characteristic dynamic line tension and as a function of sensitivity, St, as shown in Table 1.
mean and dyn are load factors. For the example considered here, RP-E301 has a
A satisfactory anchor design which meets the no default value for Ucons of 1.45 for a typical normal
movement criteria is obtained when Rc exceeds Td, clay with a sensitivity (St) of 2.5. This value will be
which is achieved by varying the anchor installation adopted in this case study.
resistance, Ri. This in turn leads to an equation for the
minimum preload during anchor installation (Tmin) 3.2 Normal clay Ucy
which is defined as;
For the range of clays given in RP-E301, a reasonable
value for Ucy assuming 10 equivalent cycles of load-
(4) ing is 1.3 (full 1-way cycling, i.e. a/suD  0.5). These
recommendations are based on results from a data-
base of cyclic testing on clays, as shown on Figure 1.
3 EXAMPLE 1 NORMAL CLAY
3.3 Results normal clay
In order to demonstrate how the DNV code may be
used to assess preload requirements, an example is Inserting the parameters for cyclic loading and con-
presented for a conventional clay soil. solidation derived in the previous sections into the
For this example we will assume a Consequence equation for minimum preload, we get;
Class 1 scenario which implies no major consequences
such as loss of life or major structural or environmen- Table 1. Recommendations for consolidation factor Ucons.
tal damage. For a Quasi-static analysis the load is
not divided into mean and dynamic components and a Soil sensitivity DNV recommendation
single overall load factor of 1.7 is recommended St Ucons
(Td  1.7.Tc-qs, where Tc-qs is the quasi-static load). A
1.0 1.00
material factor m of 1.3 is recommended for this
2.0 1.30  0.05
load case (RP-E301, Table 5-1). 2.5 1.45  0.10
We also need to determine the critical parameters
Rcons and Rcy for which guidance may also be sought
from RP-E301;

(5)

where Ucons is the consolidation factor and Ucy is the


combined cyclic loading and load rate factor. The
selection of these parameters is discussed in the fol-
lowing sections.

3.1 Normal clay Ucons


It is generally considered that during installation of a
drag anchor the shear strength on the sliding surfaces
of the anchor is su, where  is the inverse of the soil
sensitivity, St. Following installation, consolidation or Figure 1. Recommendations for Ucy from DNV.

352

Copyright 2005 Taylor & Francis Group plc, London, UK


it is possible to develop an expression for the consol-
idated shear strength of a shear surface, su-con, as a
(6) function of the inclination of the surface to the hori-
zontal as follows:
Hence it may be seen that for this example, the
minimum required preload works outs to be very simi- (7)
lar to the quasi-static load (Tc-qs) in normal clay,
which is consistent with most current practice.
where  is the inclination of the frictional anchor sur-
face to the horizontal. The expression results in full
4 EXAMPLE 2 CARBONATE SOIL consolidation (back to su-mon) for horizontal surfaces,
and no consolidation for vertical surfaces.
Now we will look at the same example, but applied to For this illustrative example a value for Ucons of 1.7
a typical uncemented silty carbonate soil. For this has been assumed, which is actually a better result
case, RP-E301 gives no guidance on appropriate val- than for the normal clay. This reflects the much
ues for Ucons and Ucy so we will have to consider these higher sensitivities expected in many uncemented
from a more fundamental site specific basis. silty carbonate soils.

4.1 Carbonate soil Ucons 4.2 Carbonate soil Ucy


For traditional clays, the DNV method has provided When we consider Ucy, a different story is evident.
guidance on the amount of increase in the overall For normal clays, Ucy exceeds 1 since the (viscous)
anchor holding capacity due to consolidation, as a rate dependent strength gain more than offsets the true
function of the sensitivity of the soil. RP-E301 would degradation due to cyclic loading. However there is no
suggest that increasing St would lead to increasing evidence of such rate dependency with some carbon-
Ucons, although no guidance is given for St exceeding ate soils and cyclic degradation can be significant.
2.5 (see Table 1). These estimates have been made on An illustration of the response of carbonate sedi-
the basis that the anchor resistance is comprised of ments to cyclic loading is presented on Figure 2. This
both sliding and bearing forces on the various anchor figure shows that as the apparent overconsolidation
elements. The shear forces (parallel to anchor sur- of the soil increases (su-mon/vo), the relative
faces) are assumed to setup according to the inverse of undrained strength under cyclic loading (su-cyc/su-mon)
the sensitivity, whereas no strength increase is attributed decreases. This aspect of carbonate soils is further
to the bearing forces (normal to the anchor surfaces). discussed in Erbrich (2005).
This approach yields typical values for the increase in It can be seen that typical values for the cyclic
holding capacity (Ucons) as presented earlier in Table 1. strength of these soil types can be as low as 50% of
As shown in this table, the increase in anchor capacity the monotonic strength. For this example, the value of
is always less than the soil sensitivity because there is Ucy has therefore been taken as 0.5.
a significant bearing component of anchor resistance,
which does not experience setup.
It is apparent from site specific testing that sensi- 1.6
tivities of typical carbonate sediments can be much 1.4
higher than for typical clays, with typical values for St
1.2
ranging between 5 and 20.
Another aspect where carbonate soils differ from 1
su-cyc/su-mon

typical cohesive soils is that setup on some shear elem-


0.8
ents may not be fully realised. This is true of vertical
surfaces (such as pile walls or skirts), where penetra- 0.6
tion and shearing leads to a loss of normal stress due 0.4
to the soil compressibility and arching mechanisms,
and recovery of strength is generally less than indi- 0.2
cated by the sensitivity of the soil. 0
However, for horizontal surfaces which are over- 0 0.5 1 1.5 2 2.5 3
lain by a given depth of soil, the normal stress should su-mon/v'
eventually revert to the in situ conditions. Loss of nor-
mal stress due to soil compressibility can not occur in Figure 2. Typical cyclic strength data for carbonate
the same way as is possible for vertical surfaces. Hence sediments.

353

Copyright 2005 Taylor & Francis Group plc, London, UK


4.3 Results carbonate soil above the quasi-static) can be sustained with only
minimal anchor movement in-service.
Hence in this case;
If the philosophy of preloading to ensure no move-
ment under the intact total loads and allowing some
(8) movement under ultimate conditions is adopted for
carbonate soils, the preloading requirements will be
It is important to realise here that we have a minus less stringent than illustrated in the previous example.
sign in the result which is contrary to the inherent Re-writing the final expression for the preload level
assumption built into the RP-E301 formula for deter- in carbonate soils, without the load and material fac-
mining Tmin. For consistency in this case it is there- tors results in the following equation:
fore necessary to recast the formula for Tmin slightly,
with the material factor enhancing the cyclic and con- (11)
solidation effects;
This philosophy results in preload levels that are
(9) still higher than traditional approaches would require,
but not as stringent as if the full DNV approach of no
Hence for this illustrative case we now have; movement under ultimate loads was imposed. This is
an approach that has been successfully adopted to
define preload levels for drag anchors in carbonate
(10) soils on several projects in Australia.

This means that, for the various parameters selected 6 FUTURE DIRECTIONS
for this simple example, the appropriate preload for
the carbonate soil example is about twice the quasi- In this paper we have examined the preload require-
static load. ments for drag anchors in traditional clays and in car-
bonate sediments based on the philosophy outlined by
DNV. We have also considered the effect of preloading
5 ADJUSTED APPROACH FOR to ensure no movement under the required ultimate
CARBONATE SOILS capacity, versus preloading to a lower level where some
redistribution of anchor load may occur under the
The case study presented in the previous sections of extreme damaged condition.
this paper demonstrates that there may be significant The DNV method is very useful in that it provides
differences between the level of preload required for a rational framework for assessing the components of
drag anchors in carbonate sediments, and drag anchors anchor resistance over time (due to consolidation) and
in traditional clays, in order to achieve the same in- under cyclic loading. This enables the anchor preload
service reliability. This difference is primarily due to to be defined for various levels of reliability.
the large cyclic degradation of carbonate materials. One aspect of anchor behaviour that may be exam-
It is noted that the DNV method has been con- ined in more detail in the future is the response of the
structed to ensure that there is no anchor movement anchor to dynamic loading over and above the anchor
under the factored loads (and also using a material fac- resistance (taking into account consolidation and
tor on the soil response). This is equivalent to ensur- cyclic loading). If it is assumed that the anchor resist-
ing that the anchor will not move at all under any ance defines a threshold to anchor movement, then a
extreme event, and that the required ultimate capacity dynamic load above this threshold will simply accel-
can be reached without further anchor movement. erate the anchor under the net load (being the applied
This is a somewhat more stringent philosophy than load minus the threshold resistance). The anchor will
typically adopted in mooring design. It has generally then decelerate when the load drops below the thresh-
been accepted in the past that anchors should be pre- old until it becomes stationary. A cycle by cycle
loaded to ensure no movement under the intact condi- analysis could be carried out to determine the accu-
tion, but for a damaged (one line broken) condition, mulated anchor movement over the design life. This
some small amount of movement may be permitted would require detailed time-histories of the applied
resulting in redistribution of load to other anchors in anchor loads.
the mooring arrangement. Unfortunately, there is a potential difficulty with
Furthermore, previous practice of preloading the this approach. In practice the cyclic strength does
anchor to at least the quasi-static load implicitly more than define a threshold to movement; in reality
acknowledges that higher dynamic loads (over and it defines a trigger point for a cyclic failure to be

354

Copyright 2005 Taylor & Francis Group plc, London, UK


initiated. Once this occurs, even cyclic loading of At the present time it is unfortunate that no full
smaller amplitude than that which triggered the fail- scale or even model scale tests have been performed
ure in the first place may still give rise to progressive that address cyclic loading of drag anchors in carbon-
movements (see Erbrich 2005 and Bye et al. 1995). ate soils. Hence the enhanced preload requirements
Nevertheless, with a drag anchor it may be possible to for carbonate sediments discussed in this paper could
move out of the zone of damaged soil into a new be considered somewhat speculative, and it is possible
undamaged zone with only relatively modest extra that our analysis overstates (or even understates) the
displacements (ie. potentially a few metres). In some true magnitude of the problem at hand. Nevertheless,
cases displacements of this magnitude might still prove an interesting and highly relevant analog to the problem
to be acceptable. of a drag anchor preloaded in cyclically degradable
This is an area where further research is required if soils was the recent preloading of spudcans at the
this philosophy is to be pursued in future. Trefoil prospect in the Bass Strait (Erbrich 2005).
If it could be demonstrated that the mooring arrange- This case history involved the application of cyclic
ment is able to withstand a certain amount of anchor preloading to the spudcans resulting in the triggering
movement over the life of the field, and if a model for of a cyclic foundation failure, accompanied by large
anchor movement under dynamic loading can be devel- foundation settlements. Importantly, the cyclic pre-
oped with high reliability, then the preload require- load required to trigger failure was only around 65%
ments would be able to be reduced from the level of the static preload that had previously been success-
required for a no-movement philosophy. fully applied to the spudcans.
The authors believe that this case history alone is
enough to raise the alarm wherever cyclically degrade-
7 SUMMARY AND CONCLUSIONS able soils (particularly carbonate sandy silts/silty
sands) are encountered and where consideration is
The simple numerical examples presented in this paper being given to any kind of foundation where static
show quite clearly the potential differences in drag preloading is normally used to prove the foundation
anchor behaviour between normal clay and carbon- capacity. A wise and prudent approach to engineering
ate soils, if all other things are considered equal. The in such soils is to adopt a cautious approach, such as
clear conclusion is that the potential for high cyclic that outlined in this paper for drag anchors, until and
degradation of carbonate soils means that a much unless a better approach can be demonstrated through
higher pre-load is likely to be required in order to obtain a future testing programme.
the same reliability for a drag anchor compared to
elsewhere in normal clay, assuming that the anchor
were not permitted to move after installation.
REFERENCES
However, if movement of the drag anchor could be
catered for in the design of the mooring system then it Bye A., Erbrich C.T. Rognlien B. & Tjelta T.I. 1995.
may not be necessary to impose higher preloads. The Geotechnical Design of Bucket Foundations. Proc.
drag anchor could then be sized according to the max- Offshore Technology Conference, OTC 7793, Houston,
imum allowable movement (larger anchors required Texas.
for smaller movement). For this case it would be neces- DNV 2000. Design and Installation of Fluke Anchors in Clay.
sary to perform some kind of dynamic analysis and Det Norske Veritas Recommended Practice RP-E301.
balance the energy expended in dragging the anchor Erbrich C.T. 2005. Australian Frontiers Spudcans on the
through (cyclically degraded) soil compared to the Edge. Proc. International Symposium on Frontiers in
Offshore Geotechnics, Perth, Western Australia. Balkema:
energy imposed from the mooring system during the Rotterdam.
design storm. Also, if re-tensioning of the mooring Lloyds 1999, Rules & Regulations for the Classification of
system post-storm was considered unacceptable then a Floating Offshore Installation at a Fixed Location;
it would be necessary to consider a lifetime of storms Part 3 Unit Types & Special Features. Lloyds Register
for this assessment. of Shipping, London.

355

Copyright 2005 Taylor & Francis Group plc, London, UK


The geotechnical performance of Deep Penetrating Anchors in
calcareous sand

M.D. Richardson, C.D. OLoughlin & M.F. Randolph


The University of Western Australia, Perth, Western Australia

ABSTRACT: Deep Penetrating Anchors have been shown to be a feasible anchoring system in normally con-
solidated clay. However, substantial oil and gas deposits exist in areas of the world that are characterised by cal-
careous ocean floor sediments. This paper examines the viability of the use of such anchors in calcareous sand
through a series of centrifuge model tests and associated analytical studies. It is shown that whilst the embed-
ments and vertical capacities in calcareous sand are lower for similar impact velocities than in normally consoli-
dated clay, the results compare favourably with field test data reported for torpedo shaped anchors in uncemented
calcareous sand in the Campos Basin, Brazil. In addition a prediction model developed for calculating expected
embedments in clay has been adapted to calcareous soil conditions and gives results that agree well with the
measured data.

1 INTRODUCTION

The Deep Penetrating Anchor (DPA) is a deepwater


offshore foundation concept originally proposed by
Lieng et al. (1999) as a cost effective anchoring system
for floating oil and gas production facilities. Exhaustion
of shallow water hydrocarbon deposits is increasingly
forcing the oil and gas industry to exploit deepwater
resources. As such suitable mooring techniques are
required to ensure the viability of these operations. The
high cost of anchor handling vessels and increased
installation durations mean that installation costs for
conventional offshore anchoring systems increase dra-
matically with increasing water depth. The DPA is seen Figure 1. Deep Penetrating Anchor: (a) Schematic (after
as beneficial in this regard in that installation costs Lieng et al. 1999) (b) Model anchor 0 flukes.
are largely independent of water depth as no external
energy sources or mechanical operations are required
during installation. Medeiros (2001) report cost sav- Extensive centrifuge testing of model DPAs in
ings of approximately 30% associated with the use of normally consolidated clay has been reported by
conceptually similar torpedo anchors in the Campos OLoughlin et al. (2004). The results of these tests
Basin, Brazil. show that tip penetrations of 1.5 to 3 times the anchor
DPAs are rocket shaped anchors (see Figure 1a), length are possible resulting in vertical holding capaci-
which when released from a specified height above the ties of between 3 and 5 times the anchors dry weight.
seabed, free-fall through the water column and pene- Watson & Randolph (1998) report that oil and gas
trate the ocean floor by means of the anchors self operations on the North West Shelf of Australia are
weight and the kinetic energy gained during free-fall. increasingly moving into water depths exceeding 900 m.
Once installed, uplift forces due to environmental load- With this increase in water depth comes an associated
ing are mainly resisted by the friction developed at the need for alternative foundation design, which is espe-
anchor-soil interface. Detailed discussion of DPA con- cially important considering the problematic nature of
struction, installation and feasibility is given by Lieng the calcareous sediments present in the area. Calcareous
et al. (1999, 2000). sands originate from biological processes such as

357

Copyright 2005 Taylor & Francis Group plc, London, UK


sedimentation of skeletal debris and coral reef forma- the dried material recovered from the seabed down to
tion and are characterised by highly angular and a maximum particle size of 0.3 mm. The sieved
brittle particles and the presence of varying degrees material was subsequently mechanically mixed for
of cementation (Murff 1987), often resulting in trou- several hours to ensure uniform particle distribution
blesome foundation design. The requirement for throughout the sample. Upon completion of the mixing,
effective deepwater foundations in these conditions the sand was placed loosely in a centrifuge strongbox,
motivated the investigation of DPA performance in with internal dimensions measuring 650 mm long,
calcareous sand. 390 mm wide and 325 mm deep, and saturated via
Medeiros (2001, 2002) report the results of field drainage holes in the bottom of the box. The strong-
tests conducted using a torpedo anchor in the Campos box was then placed on a vibrating table at a low speed
Basin. The study focused on the likely embedment, in setting for approximately 1 hour. Drainage at the bot-
various soil conditions, of 30 inch (762 mm) diam- tom of the sample was provided via a woven, felt
eter, 12 m long cylindrical piles filled with scrap metal drainage blanket overlying a 10 mm deep layer of
and concrete and fitted with a conical tip. Torpedo coarse sand. The overall sample height after preparation
anchor installations conducted in 200 m of water in the (including the drainage layer) was 225 mm for Box 1
Corvina field examined anchor performance in unce- and 210 mm for Box 3. This preparation technique
mented calcareous sand. For drop heights of 30 m above resulted in an average saturated unit weight of
the seabed, average tip penetrations of approximately 15.2 kN/m3 and a relative density of 46  5% repre-
15 m were achieved. senting a medium dense sample.
This paper presents the results of centrifuge tests In order to save time, Box 2 was prepared by recon-
conducted on model DPAs in calcareous sand recov- stituting the sample in Box 1. The original sample was
ered from the seabed in the area surrounding the reconstituted by carefully hand mixing the sample
North Rankin platform on the North West Shelf of under water to avoid ingress of air. The sample was then
Australia. The results of these tests are compared with placed back on the vibrating table for approximately
DPA model tests conducted in normally consolidated 30 minutes before being mounted on the centrifuge
clay and full scale field trial results reported by for further testing. A sample height of 205 mm was
Medeiros (2001, 2002). achieved for Box 2.
Soil characterisation tests were conducted to meas-
ure tip penetration resistance and to ensure consist-
2 CENTRIFUGE TESTING ency between individual samples. A standard model
cone penetrometer (diameter 10 mm, cone angle 60),
2.1 Soil properties installed at a rate of 1 mm/s was utilised in each box.
Average tip resistance profiles for each box are pre-
Calcareous sand obtained from the seabed near the
sented in Figure 2.
North Rankin platform was used throughout the test
Apart from the top 2040 mm, average CPT profiles
series. The soil properties in the vicinity of this plat-
in Box 1 and Box 3 show a relatively uniform increase
form are widely reported (Angemeer et al. 1975,
in tip resistance with sample depth. Comparison of
Renfrey et al. 1988) whilst further studies on the gen-
average tip resistances prior to, and at the conclusion
eral engineering characteristics of calcareous sands
of testing in Box 3 indicate an increase in strength
are presented by Murff (1987) and Jewell (1993).
during the course of testing. This may be attributed to
Some key properties of the soil used in these tests are
summarised in Table 1.
Tip Resistance (MPa)
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0
2.2 Sample preparation 0 0

Three separate samples were prepared for this test 20 4


programme. Box 1 and Box 3 were prepared by sieving 40 8
Prototype Depth (m)

60 12
Model Depth (mm)

80 16
Table 1. Summary of soil properties.
100 20
Specific gravity, Gs 2.73 120 24
Min. dry unit weight 7.46 kN/m3 140 28
Max. dry unit weight 10.1 kN/m3 160
Box 1
32
Box 2
Min. void ratio, emin 1.65 Box 3 - before testing
180 Box 3 - after testing 36
Max. void ratio, emax 2.59
Porosity, n 6272% 200 40
Friction angle,  40
Figure 2. Average CPT tip resistance profiles.

358

Copyright 2005 Taylor & Francis Group plc, London, UK


the sample settling to a certain degree due to ramp- to prevent lateral movement during free-fall. The
ing up and down of the centrifuge, leading to a max- impact velocity of the model anchors was measured
imum density and strength condition after several using a photoemitter-receiver pair located at the base of
ramp up/down cycles. the anchor guide. The impact velocity was determined
The average CPT profile for Box 2 indicates rela- by recording the time taken for the anchor to enter and
tively high tip resistances within the top 40 mm of the pass through the emitters beam. Release of the model
sample compared with those tests conducted in Box 1 anchor in flight was achieved via the use of a resistor
and Box 3. At greater depths the measured tip resist- which when supplied with current, heated up and sub-
ance remains relatively constant with depth. This sequently burnt through a release cord attached to the
behaviour tends to suggest some form of segregation anchor padeye. The anchor guide, release mechanism
or non-uniformity within the sample and implies that and photoemitter-receiver pair used throughout testing
the reconstitution method adopted was not effective are the same as those adopted by OLoughlin et al.
in producing a sample consistent with those of Box 1 (2004). By varying the initial vertical position of the
and 3. As a result of this inconsistency, anchor instal- anchor within the guide, the release height and ensuing
lation and pullout tests conducted in Box 2 are not impact velocity can be controlled.
considered in this paper. Once installed, a vertical pullout test is conducted.
Typically several minutes elapsed between installa-
tion of the anchor and commencement of the pullout,
2.3 Model anchors allowing any excess pore pressures developed during
The model anchor employed throughout the test pro- DPA installation to dissipate. The mooring line passes
gramme was a 1:200 scale model, fabricated from from the anchor padeye to the load cell, which in turn
brass and designed in accordance with the Type I is attached to the actuator. By driving the actuator ver-
Torpedo Anchor (no flukes) reported by Medeiros tically upward, the load-displacement response of the
(2001) (see Figure 1b). The anchor had a mass of anchor during pullout was measured via the load cell
14.9 g (119,200 kg prototype) and was fitted with an and vertical displacement transducer on the actuator.
ellipsoidal shaped tip. The anchor dimensions are Each of the vertical pullout tests were carried out
given in Table 2. monotonically at a rate of 0.3 mm/s, the standard rate
The mooring chain utilised was a nylon coated adopted for DPA tests in clay (OLoughlin et al. 2004).
stainless steel wire trace of diameter 0.45 mm, which Tests conducted in normally consolidated clay
satisfied the basic scaling law of centrifuge modelling and reported by OLoughlin et al. (2004) employed a
(Taylor 1995), whilst providing sufficient tensile spacing of 5 anchor diameters between tests to avoid
strength and minimising adverse effects such as interaction effects. These effects however were expected
stretching and unravelling. to be of far greater significance in calcareous sedi-
ments. Therefore a standard minimum spacing of 10
anchor diameters was adopted for testing reported
2.4 Test procedure here. A separation of 20 diameters was employed near
All tests were performed at 200 g in the fixed beam
geotechnical centrifuge at the University of Western
Australia, which is described in detail by Fahey et al.
(1990).
The test setup in the centrifuge is shown in Figure 3,
with the anchor positioned at the desired drop height
above the sample surface and attached to the actuator
via the mooring chain and anchor padeye (OLoughlin
et al. 2004).The rotational nature of the acceleration
field and Coriolis effects in the centrifuge necessi-
tates the installation of model anchors down a guide,

Table 2. Anchor dimensions.

Property Symbol Model (mm) Prototype (m)

Length L 75 15
Tip length Ltip 11.4 2.28
Diameter D 6 1.2 Figure 3. Centrifuge test setup (after OLoughlin et al.
2004).

359

Copyright 2005 Taylor & Francis Group plc, London, UK


the walls of the strongbox to avoid any boundary 3.2 Vertical pullouts
effects during installation. Additionally, the need to
The peak vertical holding capacities achieved in this
avoid lateral components of acceleration, other than
test series range from 3055 N (1.22.2 MN prototype),
due to Coriolis effects, restricted DPA installation and
representing efficiencies of between 1 and 2 times the
pullout tests to the centreline of the strongbox. These
anchors dry weight. The dependence of ultimate vertical
limitations resulted in a maximum of 7 installations
capacity on embedment depth is demonstrated in
and subsequent pullout tests in each sample.
Figure 5. Evidently the shallower embedment in cal-
In Box 3, two static push (slow rate embedment)
careous sand results in lower vertical pullout capacities
tests were conducted to measure the static resistance
than in clay. However OLoughlin et al. (2004) indicate
of the model DPA during penetration. These tests
that optimisation of anchor geometry could lead to
were conducted at a rate of 1 mm/s, in order to match
much higher impact velocities, which in turn would
the installation rate of the CPTs.
result in higher anchor embedments and subsequent
holding capacities.
3 TEST RESULTS Figure 5 also shows the theoretical capacities cal-
culated for DPA tests in both calcareous sand and nor-
3.1 Model anchor installations mally consolidated clay. Idealising the model DPA as
a closed ended cylindrical pile, the vertical pullout
Drop heights ranging from 0300 mm (060 m proto- capacity in calcareous sand may be formulated as:
type) were considered during the course of this study
with the aim of examining the relationship between
anchor impact velocity and embedment depth. These (1)
drop heights resulted in impact velocities of up to
approximately 24 m/s. where Fv is the vertical pullout capacity, Ws is the
Over the range of impact velocities considered in submerged anchor weight, Nq is the bearing capacity
Box 1 and Box 3, tip embedments ranging from approxi- factor, v is the vertical effective stress, Ap is the pro-
mately 50110 mm (1022 m prototype) were meas- jected area of the anchor, is the ratio of shaft fric-
ured. These correspond to embedments of between 0.7 tion to effective overburden stress and Ashaft is the
and 1.5 times the anchor length. These results are sum- shaft area of the anchor. It is important to note that the
marised in Figure 4 together with equivalent results for bearing capacity during the vertical pullout tests
normally consolidated clay (OLoughlin et al. 2004). It relates to the bearing resistance at the anchor padeye,
can be seen that at similar impact velocities, the tip not the tip. Hence for those tests in which the anchor
embedments in calcareous sand are on average half was not fully embedded, no bearing resistance is
those for installations in normally consolidated clay. mobilised and the entire capacity is provided by the
These results also show the same trend of increasing shaft friction and anchor weight.
embedment depth with increasing impact velocity. The Selection of values for Nq and in calcareous sand
centrifuge embedment results are in good agreement is somewhat uncertain. During pullout, the low padeye
with relevant field data reported by Medeiros (2001, embedments permit the analysis of the DPA padeye
2002) who measured tip embedments of 1.25 times the
anchor length.
Prototype Capacity (MN)
0 1 2 3 4 5 6
Impact Velocity (m/s) 0 0
Box 1
-5 0 5 10 15 20 25 30 35 25 Box 3 5
0 0 Clay
Box 1 50 Theoretical (calcareous sand) 10
Prototype Embedment (m)

25 Box 3 5 Theoretical (clay)


Model Embedment (mm)

Clay 75 15
50 10
Prototype Embedment (m)
Model Embedment (mm)

100 20
75 15
125 25
100 20
150 30
125 25
175 35
150 30
200 40
175 35
225 45
200 40 STRONGBOX LIMIT
250 50
225 STRONGBOX LIMIT 45 0 30 60 90 120 150
250 50 Model Capacity (N)

Figure 4. Embedment depth data in calcareous sand and Figure 5. Measured and predicted vertical pullout capacity
normally consolidated clay. in calcareous sand and normally consolidated clay.

360

Copyright 2005 Taylor & Francis Group plc, London, UK


as a vertically loaded plate anchor, which fails the soil sensitivity of to selection of appropriate bearing
by shearing through to the surface. Rowe & Davis capacity factors. These parameter values result in
(1982) present bearing capacity (anchor capacity) friction ratios of approximately 1%, which are lower
factors for vertically loaded plate anchors in sand of than friction ratios for model jacked piles in unce-
17, for embedments normalised by the anchor diam- mented North Rankin calcareous sand (23%)
eter of up to 8. Normalised DPA padeye embedments reported by White (2005), but are comparable to typ-
of less than 6 were recorded during the test series and ical values for silica sand of 0.51.5%.
as such an average Nq  4 was selected based on the
plate anchor research. The value of was then
selected to fit the measured pullout capacity data. Best 4 EMBEDMENT DEPTH PREDICTION
agreement was achieved with Nq  4 and  0.3.
Average values of 0.030.15 are reported by Figure 5 indicates the strong dependence of DPA
Randolph (1988) for field tests on driven piles at the capacity on embedment depth. Consequently, confi-
North Rankin site. These values are noticeably dence in calculating anchor capacity is dictated to a
lower than that used to calculate the theoretical large degree by the success of predicting DPA embed-
anchor capacity (  0.3); the differences may lie in ment depth for a given anchor drop height and hence
the method of installation or extraction. White & impact velocity. A similar model to that adopted for
Lehane (2004) report that the cyclic loading associ- predicting DPA embedment in clay, reported by
ated with pile driving causes friction fatigue. OLoughlin et al. (2004), has been adapted to predict
However, DPA installation is essentially monotonic, the embedment of model DPAs in calcareous sand.
which White & Lehane (2004) show to cause no This embedment model considers the resistant forces
reduction in the mobilised shaft friction.  0.3 is acting on the anchor during the soil penetration
therefore consistent with these observations, exceed- phase. The resulting anchor deceleration may be
ing the values reported by Randolph (1988). determined using:

3.3 Static push tests (3)


In order to assess the static resistance of the anchor,
two static push tests were conducted in Box 3. The where m is the anchor mass and Rf is a rate dependent
measured static resistance profile can be formulated term to account for velocity dependence of the soil
using an expression similar to that in Equation 1: resistance terms. A semi-logarithmic rate dependent
function (Biscontin & Pestana 2001) was utilised for
(2) Rf, given by:

where Fs is the static resistance force. The CPT tip (4)


resistance at the conclusion of testing in Box 3 was
used to deduce an average value of Nq  32, over the
measured range of DPA tip embedments. An average where  is a constant,  is the anchor velocity and s
Nq was adopted as experimental evidence shows that is the reference penetration velocity at which the
the bearing capacity factor decreases with increasing static resistance was measured (1 mm/s).
ambient stress level in silica sand (Randolph 1988). Strain-rate effects in clay are well documented,
Similar effects are likely to be observed in calcareous with typical  values of 520% per log cycle of shear-
sand and can be attributed to decreasing peak friction ing rate (Randolph 2004).  values were back-calculated
angles with increasing stress and increased soil from the embedment depth measurements taken in
compressibility. Box 3 using in the first instance the measured static
Although Nq  32 is considerably larger than the resistance profile and in the second instance those
bearing capacity factor employed for the vertical pull- formulated in Equation 2, with Nq  32,  0.42
out tests (Nq  4), the difference is due to the shallow and Nq  35,  0.3. An average value of   0.006
embedment of the anchor padeye, with a failure (0.6% per log cycle) was calculated for the measured
mechanism to the free surface. Adopting Nq  32 in static resistance, which when compared with 520%
Equation 2, and calculating the shaft friction to match for clay indicates minimal rate effects during DPA
the measured static resistance gave  0.42. Back installation in calcareous sand. Using the formulated
analysis using  0.3 (from the vertical pullout tests) static resistance,  was approximately 2% per log
gives Nq  35, which is tolerably similar to Nq  32 cycle for both sets of parameters. The higher values of
calculated from the CPT tip resistance, indicating the  do not suggest the presence of any rate effects but

361

Copyright 2005 Taylor & Francis Group plc, London, UK


Impact Velocity (m/s) that the rate dependency of soil strength in calcareous
-5 0 5 10 15 20 25 30 35 sand is minimal.
0 0
Measured - Box 3
Predicted - measured
5 Predicted - Nq = 32 25
ACKNOWLEDGEMENTS
Prototype Embedment (m)

Model Embedment (mm)


Predicted - Nq = 35
10 50
The work described here forms part of the activities of
15 75 the Special Research Centre for Offshore Foundation
Systems, established and supported under the Australian
20 100 Research Councils Research Centres Program.
25 125

30 150
REFERENCES
Figure 6. Comparison of predicted and measured embed- Angemeer, J., Carlson, E.D., Stroud, S. & Kurzeme, M.
ment depths Box 3. 1975. Pile Load Tests in Calcareous Soils Conducted in
400 feet of Water from a Semi-Submersible Exploration
Rig. Proc. Offshore Technology Conference, Houston,
are indicative of the agreement between the measured Texas, USA: 657670.
and formulated static resistance. CPTs conducted Biscontin, G. & Pestana, J.M. 2001. Influence of peripheral
by Joer etal. (1998) in slightly cemented calcareous velocity on vane shear strength of an artificial clay.
sand over a small range of penetration velocities ASTM Geotechnical Testing Journal 24(4): 423429.
(0.21 mm/s) support the observation of minimal rate Dayal, U. & Allen, J.H. 1975. The effect of penetration rate
effects in calcareous soil. Additionally, a study by on the strength of remolded clay and sand samples.
Dayal & Allen (1975) indicates that the effects of rate Canadian Geotechnical Journal 12(3): 336348.
Fahey, M., Finnie, I.M.S., Hensley, P.J., Jewell, R.J.,
of shearing on the ultimate strength of granular soils Randolph, M.F., Stewart, D.P., Stone, K.J.L., Toh, S.H. &
is negligible. Windsor, C.S. 1990. Geotechnical centrifuge modelling
The measured data from installations in Box 3 are at the University of Western Australia. University of
shown in Figure 6 against the predicted embedments Western Australia. Research report No. G:1000.
calculated using the measured and formulated static Jewell, R.J. 1993. An introduction to calcareous sediments.
resistances with the appropriate  values. Reasonable University of Western Australia. Research Report No.
agreement between the experimental and theoretical G1075.
results can be observed in each case. Joer, H.A., Jewell, R.J. & Randolph, M.F. 1998. Cone pene-
trometer testing in calcareous sediments. Proc. 17th Int.
Conf. Offshore Mechanics & Arctic Engineering, Lisbon,
Portugal.
5 CONCLUSIONS Lieng, J.T., Hove, F. & Tjelta, T.I. 1999. Deep penetrating
anchor: Subseabed deepwater anchor concept for floaters
This paper has presented results from a series of cen- and other installations. Proc. 9th International Offshore
trifuge model tests carried out in North Rankin calcar- and Polar Engineering Conf., Brest, France 1: 613619.
eous sand to investigate the penetration and pullout Lieng, J.T., Kavli, A., Hove, F. & Tjelta, T.I. 2000. Deep pen-
resistance of Deep Penetrating Anchors. Prototype etrating anchor: Further development, optimisation and
embedment depths of up to 1.5 times the anchor length capacity clarification, Proc. 10th Int. Offshore & Polar
were achieved for velocities approaching 24 m/s. These Engineering Conf., Seattle, USA 2: 410416.
Medeiros, C.J. 2001. Torpedo anchor for deep water. Proc.
embedments agree well with field test data reported by
Deepwater Offshore Technology Conf., Rio de Janeiro,
Medeiros (2001, 2002) in uncemented calcareous sand. Brazil.
Maximum vertical capacities of approximately 2 times Medeiros, C.J. 2002. Low cost anchor system for flexible
the anchor dry weight were recorded. Embedments and risers in deep waters. Proc. Offshore Technology Conf.,
subsequent capacities in calcareous sand were on aver- Houston, Texas, USA.
age 50% of those reported for normally consolidated Murff, J.D. 1987. Pile Capacity in Calcareous Sands: State
clays at similar impact velocities (OLoughlin etal. of the Art. Journal of Geotechnical Engineering 113(5):
2004). The potential for higher embedment depths and 490507.
subsequent capacities with higher impact velocities OLoughlin, C.D., Randolph, M. F. & Richardson, M.D.
2004. Experimental and Theoretical Studies of Deep
suggests that the DPA has potential as an anchoring sys- Penetrating Anchors. Proc. Offshore Technology Conf.,
tem in calcareous sediments. Houston, Texas, USA.
It is possible to predict the embedment depth of Randolph, M.F. 1988. The axial capacity of deep founda-
DPAs in calcareous sand from the static resistance tions in calcareous soil. Proc. Int. Conf. on Calcareous
profile. Low back-calculated  values demonstrate Sediments, Perth, Australia 2: 837857.

362

Copyright 2005 Taylor & Francis Group plc, London, UK


Randolph, M.F. 2004. Characterisation of soft sediments for Watson, P.G. & Randolph, M.F. 1998. Failure Envelopes for
offshore applications. Proc. 2nd Int. Conf. on Site Caisson Foundations in Calcareous Sediments. Applied
Characterisation, Porto, Portugal. Ocean Research 20(1-2): 8394.
Renfrey, G.E., Waterton, C.A. & van Goudoever, P. 1988. White, D.J. 2005. A general framework for shaft resistance
Geotechnical data used for the design of the North on displacement piles in sand. Int. Symp. Frontiers in
Rankin A platform foundation. Proc. Int. Conf. on Offshore Geotechnics, Perth, Australia. Submitted.
Calcareous Sediments, Perth, Australia 2: 343355. White, D.J. & Lehane 2004. Friction fatigue on displace-
Rowe, R.K. & Davis, E.H. 1982. The behaviour of anchor ment piles in sand. Gotechnique 54(10): 645658.
plates in sand. Gotechnique 32(1): 2541.
Taylor, R.N. 1995. Centrifuges in modelling: principles and
scale effects. Geotechnical Centrifuge Technology:
1933. Routledge mot E.F. & N Spon.

363

Copyright 2005 Taylor & Francis Group plc, London, UK


Seabed geotechnical characterisation with a ball penetrometer deployed
from the Portable Remotely Operated Drill

P.J. Kelleher
Benthic Geotech Pty Ltd., Sydney, NSW, Australia

M.F. Randolph
The University of Western Australia, Perth, Western Australia

ABSTRACT: This paper describes the methodologies and results of a recent marine geotechnical site inves-
tigation undertaken in Bass Strait, Australia. The investigation was undertaken to characterise the shallow
seabed sediments at an offshore site, for prediction of spudcan foundation performance and stability, prior to the
arrival of a jackup drilling rig. A range of geotechnical data was acquired. Soil sampling, piezocone testing and
spherical ball penetrometer testing were undertaken continuously to 28 m below seabed, using PROD (the
Portable Remotely Operated Drill), a multi-purpose seabed drilling and testing platform. This paper describes
the sampling and in situ testing undertaken, with particular emphasis on the results achieved using a relatively
new type of soft soil penetrometer, the spherical ball penetrometer.

1 INTRODUCTION fine-grained sediments. The ball penetrometer is the


axisymmetric equivalent of the cylindrical T-bar pen-
A geotechnical site investigation was recently completed etrometer (Watson et al. 1998, Randolph et al. 1998).
at Trefoil, location T/18P, Bass Strait, Victoria. The work The previous seabed investigations at Yolla revealed
was undertaken by Benthic Geotech on behalf of Origin a highly variable shallow geology comprising interbed-
Energy Resources Ltd and the T/18P joint venture. ded carbonate clays, silt and sands. Accurate estimation
The Trefoil site is located in Bass Strait, Australia. of spudcan penetration in these deposits has proved dif-
Very limited geotechnical data are available in this ficult, owing to the susceptibility for the large (15 to
region of Bass Strait, other than information acquired 20 m diameter) foundations to punch through thin sand
during the Yolla Platform site investigation (located layers, particularly if the load on the spudcan is cycled.
35 km to the East of Trefoil). In general, the spudcan response is undrained in the
The project was undertaken to investigate the com- clay and silt zones (compared with partially drained
position and strength of the shallow seabed sediments conditions around penetrometers) and the fine-grained
at the site. The data were used to predict the penetration sediments can also show significant softening. This
and stability of spudcan foundations, prior to arrival characteristic is important in assessing the operational
and jacking up of the Ensco 102 mobile drilling rig. stability of jack-up rigs during design storm conditions.
The field work was carried out using the Portable
Remotely Operated Drill (PROD). PROD is a self-
contained seabed drilling platform capable of under- 2 SITE INVESTIGATION OBJECTIVES
taking soil and rock sampling, together with a range
of in situ tests, to depths of 100 m below seabed, in The key objective was to acquire adequate geotech-
water depths of up to 2,000 m (Carter et al. 1999). nical information for the development of an appropriate
Piston and rotary core sampling, together with piezo- spudcan installation procedure, and for the prediction
cone penetrometer testing (CPT) and spherical ball of spudcan stability during the passage of a design
penetrometer (SBP) testing were completed. The tech- storm through the site. The required data included:
niques used to acquire the data, together with a descrip- continuous sampling over the expected zone of
tion and discussion of the data is presented. Particular influence of the spud cans;
emphasis is placed on the results achieved with the CPT testing at multiple locations within the pro-
SBP, a relatively new type of full-flow penetrometer posed rig footprint to characterise site stratigraphy
that is well suited to strength characterisation for soft and spatial variability;

365

Copyright 2005 Taylor & Francis Group plc, London, UK


piezocone dissipation testing to define consolida- Penetrometer is thrust
tion characteristics of identified soil layers; into ground using
ball penetrometer tests, to define the strength and PROD drill string
sensitivity of the finer-grained soil layers.

Instrumentation,
3 PROD CAPABILITIES data storage and
transmission
PROD is capable of undertaking sampling and in situ assembly
testing within a single deployment to the seabed. Once
landed on the seabed and isolated from ship move-
ment, it is powered and operated remotely via a single
umbilical. Push rod and
A suite of sampling and in situ testing tools are car- anti-friction
ried to the seabed on each deployment. The operator sleeve
can switch between rotary coring, piston sampling and
thin-walled sampling via selection of the appropriate
tool from the drills magazines. In addition to sam- Spherical ball
pling, piezocone and ball penetrometer testing can be
carried out to characterise the stratigraphy in situ. Pore water
The ability to switch between rotary coring, sam- pressure filter
pling or penetrometer testing is advantageous in com-
Figure 1. Spherical ball penetrometer for PROD.
parison with conventional seabed reaction frames and
drop corers. Where hard layers are encountered, which The SBP has equivalent testing capabilities to the T-Bar
might otherwise prevent sampling or penetrometer penetrometer (Stewart & Randolph 1994, Randolph
testing at greater depths, the operator simply switches et al. 1988). Compared to the T-Bar, however, the ball
to rotary core barrels from the magazine until the hard has a more suitable geometry for deployment down
layer is penetrated. a borehole, with less risk of snagging the edge of the
hole. The SBP geometry (see Figure 1) comprises a
3.1 Sampling 60 mm diameter, hardened smooth spherical ball,
Three types of sampling barrels may be deployed: attached to the end of a 20 mm diameter, 200 mm long,
high tensile push shaft. The push shaft incorporates a
rotary core barrels, for recovery of well consoli- friction reducing outer sleeve. The ball includes a pore
dated sediments and rock cores; water filter at mid-height, which is internally con-
piston sampling barrels, for recovery of continuous nected to a pore water pressure transducer.
sediment cores; Ball penetration resistance (qb) and pore water
thin walled sampling tubes, for the recovery of pressure (ub) can be measured in real time during pen-
high quality clay cores. etration, with the penetration resistance expressed as
The design of the sampling tools has taken account of
the detailed recommendations of engineering codes (1)
of practice, such as Eurocode 7 (2000).
where Pb is the load recorded by the load cell and Ab
3.2 Piezocone testing is the projected area of the ball. Further, correction
Standard piezocone penetrometer equipment, provid- for pressure effects is negligible. The load, and hence
ing data to the operator in real time, may also be resistance, qb, is monitored during both penetration
deployed. The equipment and testing methodologies and extraction stages.
comply with the requirements of the international As for the T-Bar, closely bracketed plasticity
reference test procedure for the cone penetration test solutions are available for the ratio, Nball, between pene-
(ISSMGE 1999), and the Swedish Recommended tration resistance and the undrained shear strength
Standard for CPT (1992). (Randolph et al. 2000). These solutions suggest that
the ball resistance should be some 20 to 25% higher
than for a cylindrical (T-bar) penetrometer. However,
3.3 Spherical ball penetrometer testing
in practice the ball and T-bar resistances are found to
A new spherical ball penetrometer (SBP) has recently be closely similar with, if anything, the ball resistance
been incorporated on PROD, with the aim of improv- marginally lower than that for the T-bar (Chung &
ing characterisation of soft fine-grained soil layers. Randolph 2004, DeJong et al. 2004). This has been

366

Copyright 2005 Taylor & Francis Group plc, London, UK


attributed to soil characteristics such as strength

Stratum Penetration, m
anisotropy, strain rate effects and softening, none of

Penetration, m

Soil Profile
which is accounted for in the plasticity solutions. SOIL DESCRIPTION
qc, CONE TIP RESISTANCE (MPa)
0 4 8 12 16
Full-flow penetrometers such as the T-bar and ball TORVANE STRENGTH, (kPa)
have advantages relative to the cone, including: min- Greenish gray silty fine calcareous SAND
0 20 40 60 80

imal correction for pore water pressure effects and with shell fragments

overburden pressure; more accurate (and versatile) ana-


4.0
lytical solutions for the corresponding resistance fac- 5 Soft greenish gray sandy calcareous CLAY
with scattered shell fragments & H2S odour
tors; and the ability to conduct meaningful cyclic
penetration tests to evaluate remoulded strengths in situ. 7.3
8.0
Light gray clayey calcareous SILT with
shell fragments
The larger projected area of the ball mobilises sig- 9.0 Dark greenish gray silty fine calcareous
SAND with scattered shell fragments
nificantly higher resistance loads compared to the 10
10.6
Pale olive gray sandy carbonate SILT
intermixed with shell fragments
cone. This enables shear strength to be measured to a 12.1
Pale olive gray silty fine carbonate SAND
intermixed with shell fragments

resolution of better than 0.2 kPa using PRODs SBP. 13.0 Greenish gray silty fine to medium
calcareous SAND intermixed with fine
to medium sized shell fragments
The low-resolution capability is advantageous for 15 Light gray silty fine to medium
calcareous SAND intermixed with
strength measurement in very soft sediments. fine to medium sized shell fragments

The SBP monitors pore water pressure at the mid-


height of the ball. In principle, the excess pore pres- 18.5
Greenish gray silty fine calcareous SAND
sure at this position reflects pore pressure developed 20
20.5
- intermixed with shell fragments below 18.9m

due to shearing (as opposed to changes in total Firm greenish gray calcareous CLAY
- with silt partings below 21.3m

stress). As is shown later, this provides an extremely 23.2


24.2 Light gray siliceous carbonate sandy SILT
sensitive measure of the compactive or dilatant char- 25 intermixed with fine to course shell fragments
Greenish gray silty fine siliceous carbonate
acteristics of the soil being penetrated. Dissipation 26.7
SAND intermixed with fine to medium
shell fragments
testing can also potentially be carried out, although at 27.4 Greenish gray silty fine calcareous SAND
with scattered cemented calcareous
present no analytical or numerical solutions have siltstone and shell fragments
End of Borehole at 27.4m
30
been developed for the interpretation of dissipation
tests around full-flow penetrometers.
Sophisticated control software allows a wide range Figure 2. Preliminary borehole log for the Trefoil site.
of manoeuvres of the down-hole tools. Positioning is
accurate to approximately 1 mm. A variety of cyclic
displacement routines are available to evaluate the mudline and are 4 m, 1.6 m and 3.8 m thick, respect-
remoulded characteristics of the soil. Tools can be pene- ively. Predominantly fine-grained sands overlie the
trated and extracted over a range of velocities, from cohesive layers. With respect to soil bearing capacity,
approximately 10 mm/sec to 100 mm/sec, which can this stratigraphy would be expected to provide a highly
provide useful data for estimating rate effects and par- variable, often brittle, bearing response as the foot-
tial consolidation during sounding (Randolph 2004). ing penetrates adjacent strong and weak soil layers.
For the Trefoil site a number of large displacement Torvane testing was completed on core recovered
cyclic tests were completed in the fine-grained soil from the clay layers at the site; average vane strengths
layers. The data were used to provide a preliminary of about 12 kPa and 38 kPa were measured respect-
estimate of soil sensitivity at the Trefoil site. ively in the two clay layers commencing at 4 m and
20.4 m below mudline. An average submerged unit
weight of about 8 kN/m3 was also determined.
4 INVESTIGATION RESULTS

4.1 Site geology 4.2 Cone penetrometer testing


The majority of samples were extruded and classified Superimposed plots of the CPT data, net cone resist-
on the vessel. Visual logging, Torvane testing and soil ance (qcnet), and pore pressure parameter (Bq), are pre-
unit weight testing was completed, enabling the devel- sented on Figure 3. Borehole TRE1 was completed
opment of a preliminary bore log for the site (see adjacent to the sampling borehole to about 28 m depth.
Figure 2). No further laboratory testing was under- Three shallow boreholes (TRE4, TRE5 and TRE6)
taken on the vessel. were completed to 13 m depth within the proposed
The geology comprises interbedded carbonate footprints of the jackup rigs spud cans.
sands, silts and clays, with fine to medium shell frag- The virtually identical traces indicate very low
ments throughout. At approximately 27 m a hard layer spatial variability across the site. The soft cohesive
was identified, which was not explored further. layers are clearly defined in the plots, characterised
Three cohesive soil layers were identified at the by relatively low qcnet and high Bq values. Elsewhere,
site; these commence at 4 m, 9 m and 20.4 m below the data are typical of fine carbonate silts and sands,

367

Copyright 2005 Taylor & Francis Group plc, London, UK


qcnet (MPa) Excess pore pressure ratio, Bq Table 1. Ball penetrometer testing sequence.
0 2 4 6 8 10 -0.2 0 0.2 0.4 0.6 0.8
0 Soil Start Target No. of
Phase Activity type* depth (m) depth (m) cycles

5 1a Install sa-cl 0 7.02


b Cycle cl 7.02 6.5 7
c Extract cl-sa 6.5 0
10 2a Reinstall all 0 27.2
b Extract sa-si-cl 27.2 21.04
Depth (m)

Hole-TRE1 c Cycle cl 21.04 20.52 7


15 Hole TRE4 d Extract cl-sa/si 20.52 17.45
Hole-TRE5 e Cycle sa/si 17.45 16.95 3
Hole-TRE6
f Extract sa/si 16.95 15.0
20 g Cycle sa/si 15.0 14.5 2.5
h Extract sa/si-si 14.5 10.0
i Cycle si 10.0 9.48 8
25 j Extract all 9.48 0

* cl  clay; si  silt; sa  sand; e.g. sa/si  sand and silt;


30 cl-sa  clay transitioning to sand; all  all types (clay, silt
and sand).
Figure 3. Overlay of all CPT data.

with occasional spikes in the data believed to be Penetration Resistance qcnet & qb (MPa)
mainly attributable to shell fragments encountered 0 2 4 6 8 10 12
throughout the profile. The variation in Bq values is 0
due to varying degrees of consolidation and dilational
characteristics of the different layers. A hard layer 5
was intercepted at approximately 27 m below seabed.
With CPT refusal occurring during repeated attempts
to penetrate this layer, it was decided that further 10
exploration would not be required below this depth.
Depth (m)

15 Cone
4.3 Ball penetrometer testing Ball top of fine-grained layer
bottom of fine-grained layer
4.3.1 SBP testing program
20
A single borehole (TRE3) was completed using the
ball penetrometer to a maximum depth of 27 m, where
cone refusal had previously occurred. The borehole 25
was completed in two phases. Initially, the SBP was
penetrated to a depth of 6.5 m within the upper soft
clay layer that commences at about 4 m. The SBP was 30
then cycled over the depth interval 6.5 m to 7 m (a
cycling amplitude of about 8 ball diameters); there- Figure 4. Comparison of cone and SBP resistance.
after the ball was extracted from the ground. In the
second phase the ball was pushed monotonically to
refusal at approximately 27 m. It was then extracted and qb) and Figure 5 (excess pore pressure ratios, Bq
from the ground to four separate layers of interest (see and Bball). The excess pore pressure ratios are defined
Table 1) and cycled until the ball resistance reduced to in the conventional way as the ratio of excess pore
an approximately residual value. The layers targeted pressure to net penetration resistance.
for cyclic SBP testing had previously been defined Also included on Figures 4 and 5 are the approxi-
from the CPT, and corresponded to layers with moder- mate boundaries of the fine grained (clay and silt)
ately high Bq values. All testing was undertaken at the layers identified at the site. The qb data follows the
standard cone penetration velocity of about 20 mm/s. qcnet data closely in the top 10 m of the seabed,
although with slightly lower peak resistance in the
4.3.2 SBP monotonic test data sandy layers. At greater depths the trends in penetra-
Monotonic CPT and SBP penetration data are pre- tion resistance are very similar; however, the ball
sented on Figure 4 (net penetration resistance, qcnet resistance tends to plot lower than the CPT resistance.

368

Copyright 2005 Taylor & Francis Group plc, London, UK


Excess pore pressure ratios, Bq and Bball qb (kPa)
-0.4 -0.2 0 0.2 0.4 0.6 0.8 1 -200 0 200 400 600
0 6.4

Top of fine-grained layer


5 6.5

Bottom of fine-grained layer


6.6
10
CYCLE #1
Depth (m)

Depth (m)
Ball
Cone
6.7 CYCLE #2
15
CYCLE #3
6.8 CYCLE #4
20 CYCLE #5
6.9 CYCLE #6
25 CYCLE #7
7

30 7.1
Figure 5. Comparison of cone and SBP pore pressure data. Figure 6. Cyclic SBP penetration data from upper clay layer.

In the softer layers at 15 m and 20 to 24 m, the ball


resistance is only 70 to 80% of the cone resistance.
Similar trends are observed for the cone and SBP an approximately zero mean. The negative Bball values
excess pore pressure data (Figure 5). However, as imply only partial drainage as the ball passes, in spite
expected, the maximum Bball values are much less of high consolidation coefficients in those layers.
than the corresponding Bq values, since the excess
pore pressure measured in the SBP is that due to 4.3.3 SBP cyclic test data
shearing, rather than total stress change. This also An example of a cyclic test plot is presented on
leads to the Bball values being much more sensitive to Figure 6. In this example the ball bearing pressure is
variations in dilational characteristics of the soil, or in monitored in both tension and compression as the ball
this case to changes in soil grading. is cycled over an amplitude of 0.5 m between 6.5 and
In the clay layers (commencing at 4 m and 20.4 m 7 m (see Table 1).
below seabed) the pore pressure response on the ball is Degradation in penetration and extraction resist-
clearly positive but fluctuates to a greater degree than ance with increasing cycle number was observed in all
the Bq data. Localised negative spikes in the ball data, the cyclic tests. For the test in Figure 6, the penetra-
such as at a depth of 6.6 m, may be affected by shell tion resistance decreases from approximately 190 kPa
fragments in the vicinity of the ball, but could also (initial installation), to about 40 kPa after 7 cycles (rep-
reflect thin coarse-grained layers. No similar reduction resenting a remoulded strength of about 20% of the
in Bq is seen at the depth of 6.6 m, although a negative peak, or a sensitivity of 5).
spike was measured in one cone profile (TRE6 see The data from three cyclic tests undertaken in the
Figure 3). finer-grained soil layers are summarised in Figure 7.
In the predominantly silty layer (approximately 9 to In this figure the ball resistance has been normalised
10.6 m below mudline) the pore pressure response on against the approximate effective overburden pres-
the ball becomes positive, though to a much lower sure. All three layers rapidly asymptote to a residual
magnitude than measured in the clay layers. The lower qb/vo ratio of less than 1, and in the case of the silt
pore pressure response in this layer can be attributed to layer about 0.3. The remoulded strengths in the clay
the moderate dilatancy typically exhibited by normally layers are about 20% of the initial monotonic strength,
consolidated carbonate silts under shear. while that for the silt layer is less than 5%.
In the sand layers the Bball data is generally negative. It is the authors understanding that a similar high
An exception to this trend is evident over the depth sensitivity to cyclic loading was observed for silty
interval 14 m to 16 m, where the Bball data plots as layers at the Yolla site. However, it is possible that the
moderately positive. This trend could represent a apparent reduction in full-flow penetrometer resist-
localised increased in fines content, which was not ance (T-bar at Yolla, ball at Trefoil) may be partly
identified during preliminary logging of the cores. In attributable to an open slot forming above the pen-
the sand layers the cone Bq parameter fluctuates around etrometer during partially drained penetration. That

369

Copyright 2005 Taylor & Francis Group plc, London, UK


Cycle Number tended to give progressively lower resistance with
-2 increasing depth. The excess pore pressure data from
the ball indicated a much greater sensitivity of this tool
to changes in soil grading with depth, and thus an
0
enhanced capacity to define changes in soil stratigraphy
with depth.
2 The SBP was used to explore the sensitivity of the
fine-grained soil layers to cyclic loading, making use
of PRODs ability to apply precise two-way depth and
qb/'vo

4 velocity control of the penetrometer. The cyclic test


data indicated a high sensitivity of the soft soil layers
6
Cycle set #1 - 6.5m to 7m (CLAY) - Penetration at the site to cyclic loading. Degradation in ball pene-
Cycle set #1 - 6.5m to 7m (CLAY) - Extraction tration resistance (which has a direct correlation with
Cycle set #2 - 20.5m to 21m (CLAY) - Penetration undrained shear strength) was found to occur very
8 Cycle set #2 - 20.5m to 21m (CLAY) - Extraction rapidly in the fine grained soil layers at the site.
Cycle set #3 - 9.5m to 10m (SILT) - Penetration Residual penetration resistances of less than 25% of
Cycle set #3 - 9.5m to 10m (SILT) - Extraction the initial resistance were estimated from the data.
10
1 2 3 4 5 6 7 8

Figure 7. Summary of cyclic test data in fine soil layers.


ACKNOWLEDGEMENTS

would account for the very rapid initial drop in cyclic The authors would like to thank the T/18P Joint ven-
resistance in such materials (such as for the cyclic test ture partners for the permission to publish the
between 9.5 and 10 m in Figure 7). Where partial or methodology and results of this site investigation. The
full slotting occurs the reduction in penetration resist- Joint Venture comprises Origin Energy Resources Ltd
ance of the ball (or T-bar) may be considerably (operator), AWE Petroleum Pty Ltd, CalEnergy Gas
reduced, as bearing resistance is mobilised only on a (Australia) Ltd and Wandoo Petroleum Pty Ltd.
reduced portion of the gross area of the penetrometer. The spherical ball penetrometer described above is
Confirmation of the actual sensitivity of the soil is the subject of international patent applications.
needed from cyclic shear testing in the laboratory.

5 CONCLUSIONS REFERENCES

Carter, J.P., Davies, P.J. & Krasnostein, P. 1999. The future of


This paper has presented the results of a geotechnical offshore site investigation robotic drilling on the
site investigation undertaken with the Portable seabed. Australian Geomechanics, 34(3), 7784.
Remotely Operated Drill. The reason for the site Eurocode 7. 2000. Geotechnical Design, Part 3: Design
investigation was to obtain high quality geotechnical assisted by field testing. European Pre-Standard ICS
data for the prediction of spudcan behaviour and sta- 91.080.01; 93.020.
bility, prior to the arrival of the Ensco 102 jackup rig Chung, S.F. & Randolph, M.F. 2004. Penetration resistance
at the Trefoil site, Bass Strait, Australia. in soft clay for different shaped penetrometers. Proc. 2nd
The site geology from recovered samples was found Int. Conf. on Site Characterisation, Porto, 1: 671678.
DeJong, J.T., Yafrate, N.J., DeGroot, D.J. & Jakubowski, J.
to comprise interbedded calcareous sands, silts and 2004. Evaluation of the undrained shear strength profile
clays (Figure 2). in soft layered clay using full-flow probes. Proc. 2nd Int.
The CPT data, obtained from 4 separate locations Conf. on Site Characterisation, Porto, 1: 679686.
across the proposed footprint of the jackup, indicated ISSMGE. 1999. International reference test procedure for
very low spatial variability. The stratigraphy deduced the cone penetration test (CPT) and the cone penetration
from the profile of cone resistance was found to be test with pore pressure (CPTU). Report of ISSMGE
in close agreement with the results of the sampling TC16 on Ground Property Characterisation from In-situ
borehole. Testing, Proc. 12th Eur. Conf. on Soil Mech. and Geot.
A spherical ball penetrometer (SBP) was pushed to a Eng., Amsterdam, Balkema. 3: 21952222.
Randolph, M.F. 2004. Characterisation of soft sediments for
maximum depth of 27.2 m below seabed at the site. offshore applications, Keynote Lecture, Proc. 2nd Int.
Penetration resistance as well as pore pressure meas- Conf. on Site Characterisation, Porto, 1: 209231.
ured at the mid-height of the ball were monitored in real Randolph, M.F., Hefer, P.A., Geise, J.M. & Watson, P.G.
time. The measured penetration resistances of the ball 1998. Improved seabed strength profiling using T-bar
followed the same trends as the cone data, although penetrometer. Proc Int. Conf. Offshore Site Investigation

370

Copyright 2005 Taylor & Francis Group plc, London, UK


and Foundation Behaviour New Frontiers, Society Stewart, D.P. & Randolph, M.F. 1994. T-Bar Penetration test-
for Underwater Technology, London, 221235. ing in Soft Clay. Journal of Geotechnical Engineering.
Randolph, M.F., Martin, C.M. & Hu, Y. 2000. Limiting Volume Number 120 (12): 22302235.
Resistance of a Spherical penetrometer in cohesive mater- Watson, P.G., Newson, T.A. & Randolph, M.F. 1998. Strength
ial. Geotechnique 50 (5): 573582. profiling in soft offshore soils. Proc. 1st Int. Conf. On Site
Swedish Geotechnical Society. 1992. Recommended Characterisation ISC 98, Atlanta, 2: 13891394.
Standard for Cone Penetration tests. Established by the
Swedish Geotechnical Society, June 15 1992, report num-
ber 1:92 E.

371

Copyright 2005 Taylor & Francis Group plc, London, UK


Static and cyclic behavior of laterally loaded piles in calcareous sand

W.D. Guo & B.T. Zhu


School of Engineering, Griffith University, QLD, Australia

ABSTRACT: Elastic-plastic closed-form solutions for laterally loaded free-head piles using idealized p-y curves
have been established, and implemented into a spreadsheet program GASLFP by the first author. Based on a generic
limiting force profile (LFP), the solutions can accommodate various effects including group effect, soil type and
loading regime through determining appropriate values of three parameters for the LFP. In this paper, the idealized
p-y curves are discussed against existing p-y curves exclusively for calcareous sediments. Responses of single piles
under static and/or cyclic loading are investigated using GASLFP, focusing on examining the LFP mobilized along
the piles. Based on six case studies, determination of LFP for piles in calcareous sediments is presented for design.

1 INTRODUCTION Mt= Pt e
Pt y

Calcareous sediments are encountered in offshore

Soil resistance,p (kN/m)


Plastic
engineering in many parts around the world. Piles x Plastic zone
zone p
installed in such an environment are subjected to static pu
as well as cyclic lateral loading transferred from super- Spring, k Slider, pu
Elastic
structures, or wave. Design of such piles has been zone L-xp
based on p-y curve method ( p  soil reaction and k
Elastic zone
y  local pile deflection). Small scale field tests, and 1
centrifuge tests on piles were undertaken previously to yu y
obtain p-y curves (Wesselink et al. 1988, Novello 1999, Membrane, Np
Dyson & Randolph 2001). However, it seems that there x
is little research on the difference in predicted pile (a) Load transfer model (b) Load transfer (p-y) curve
response using different shapes of p-y curves together
with the effect of gap developed between the pile and Figure 1. Load transfer model for free-head piles (Guo 2003).
soil under cyclic loading (Randolph et al. 1988).
The problem here refers to a free-head pile under a
lateral load Pt applied at an eccentricity, e above the where pu  limiting force per unit length [FL
1];
groundline. The pile-head is free to rotate and translate AL  gradient of the LFP [FL
1
n]; 0  a constant
with no constraints along the effective pile length except to include the force at ground surface [L]; x  depth
for the soil resistance. As shown in Figure 1, the pile-soil below ground surface [L]; n power of the sum of
interaction at each depth is depicted by a spring in series 0 and x. For piles in sand, it follows: AL  Ngsd 2
n,
with a slider, which represents an idealized elastic- in which s  effective unit weight of the sand;
plastic p-y curve. The spring has a subgrade modulus, k, d  diameter of the pile; Ng  the limiting force factor.
and the slider has a limiting force per unit length, pu. The solutions are for responses of pile deflection,
Once the pile deflection arrives at the yield displace- rotation, bending moment, shear force and soil resist-
ment yu  pu/k, the soil resistance reaches the pu, which ance, which are presented in explicitly expressions
extends to a depth called slip (gapping) depth, xp. Below of the xp, and the LFP etc. The solutions have also
the slip depth, the soil is in elastic state that is simulated been implemented into a spreadsheet program called
by the spring and a fictitious membrane with the latter GASLFP (Guo 2003). The program (also referred to as
having a constant fictitious tension, Np. For this model, CF) allows response of piles to be readily predicted
closed form solutions have been developed by Guo if values of k (Np), AL, 0, and n are known (Note
(2003), in terms of a generic limiting force profile that k(Np) can be estimated using shear modulus Gs,
(LFP), which is Guo & Lee 2001). Should responses of laterally loaded
piles be available, values of Gs, AL, 0, and n can be
(1) back-estimated.

373

Copyright 2005 Taylor & Francis Group plc, London, UK


30
Wesselink et al. (1988)
Novello (1999)
Dyson & Randolph (2001) 600 200 kPa

Deviator stress, kPa


Guo (2003)

20
100 kPa
400
p / s d2

50 kPa
200
10 c = 25 kPa

0
0 2 4 6 8 10
0 Axial strain, %
0 5 10 15
y/d Figure 3. Stress vs. axial strain of Kingfish B sand (after
Hudson et al. 1988).
Figure 2. p-y curves at x  2d (d  2.08 m).

Table 1. Static hardening p-y curves for Kingfish B sand. Kingfish B sand shown in Figure 3 (Hudson et al.
1988). Such a similarity was examined previously by
References p-y curve McClelland & Focht (1958). However, the normalized
displacement for lateral loading, y/d is about 2.5 times
Wesselink et al. p  Rd(x/x0)n (y/d)m the axial strain upon reaching maximum pressure pu, or
(1988) x0  1 m, n  0.7, m  0.65, deviator stress, as elaborated late on. At each cell pres-
R  650
sure, c, the deviator stress increases with axial strain,
Novello (1999) p  Rd
v0n qc1
n (y/d)m qcd and reaches a constant once the strain is beyond the
R  2, n  0.33, m  0.5, dashed line. The intersection of the dashed line with

v0  sx
each stress-strain curve may be viewed as the yield
Dyson & Randolph p  Rsd2 (qc/sd)n (y/d)m point. For instance, at c of 25100 kPa, the yield strain
(2001) R  2.7, n  0.72, m  0.6 is about 3%8%. Similar stress-strain relationship, and
yield strain level of 2%7% for c  5100 kPa was
noted as well for the calcareous sands from Leighton
In this paper, back-estimation will be carried out
buzzard, Dogs Bay, Ballyconneely and Bombay Mix
to investigate variation of the LFP with cyclic load-
(Golightly & Hyde 1988).
ing along with the gapping (slip depth) effect on piles
embedded in calcareous sand.
2.2 Difference in predicted pile response
Using suitable parameters, the three hardening models
2 IDEALIZED AND HARDENING P-Y
can produce similar p-y curves thus pile response. For
CURVES
this reason, only the p-y curve proposed by Wesselink
et al. (1988) was employed to predict the response of the
2.1 Comparison of p-y curves
fictitious pile using program COM624P (FHWA 1993).
A fictitious pile was installed in calcareous, Kingfish This is shown in Figure 4 together with those pre-
sand (Kingfish B platform site, Bass Strait). The pile dicted using GASLFP. Using Ep  3.0  104 MPa and
had a diameter, d of 2.08 m, and Youngs modulus, Ep Gs 5.0 MPa, the critical pile length Lcr (Guo & Lee
of 3.0  104 MPa. The sand had the properties of 2001) was estimated to be 9.24d (19.22 m). Thus the
s  8.1 kN/m2,   31, Gs 5.0 MPa, and had a pile deflection and soil reaction mainly take place within
cone tip resistance, qc linearly increasing with depth at this depth as shown in Figures 4(c) and (d).
a gradient of 400 kPa/m. The limiting force may be Figure 4 indicates that: (1) p-y curves from the two
described by n  1.7, 0  0 and Ng  0.33 K p2. An models offer similar prediction in pile-head deflection,
example p-y curve at x  2d is plotted in Figure 2, yt (Fig. 4a), maximum bending moment, Mmax (Fig. 4b),
together with those produced by the three p-y curve distributions of normalized deflection, y/d (Fig. 4c), and
models provided in Table 1. The latter three p-y curves normalized bending moment, M/(sd4) (Fig. 4e); (2)
all increase with the pile deflection without limit, thus The p-y models affects remarkably the distribution
they are referred to as hardening models. of soil reaction, p/(sd2) (Fig. 4d), especially within
The idealized p-y curve by Guo (2003) is quite simi- a depth of 4d. Using the elastic-plastic model, the
lar to the stress-strain relationships of the (near surface) normalized soil resistance increases, following the LFP,

374

Copyright 2005 Taylor & Francis Group plc, London, UK


2.5 2.5 selected to represent the bending moment. Same results
COM624P_Wesselink et al
CF
COM624P_Wesselink et al
CF
from one instrumented pile may result in different p-y
2.0 2.0 curves. Overall, a true p-y curve should also offer a con-
xp/d =2.36 sistent prediction against measured distribution of soil
1.5 1.5 reaction with depth.
Pt (MN)

Pt (MN)
xp/d =1.58
1.0 1.0

0.5
3 STATIC AND CYCLIC RESPONSE
0.5
OF PILES
0.0 0.0
0 20 40 60 80 0 4 8 12 Using GASLFP, static and cyclic responses of four dif-
yt (mm) Mmax (MN-m)
(a) (b) ferent steel pipe piles in calcareous sand were investi-
y/d (%) 2
gated against measured data. The properties of the piles
p / sd
0 1 2 3
0 4 8 12 16
and soils were tabulated in Tables 2 and 3, respectively.
0
xp/d =1.58
0 Each case study is presented below.
Plastic
2 1.58 d
2
xp/d =2.36 2.36 d 3.1 Kingfish B onshore tests
4 4
Two tubular steel piles (piles A and B) were of 356 mm
x/d

Elastic
x/d

6
OD, 4.8 mm wall thickness, 6.27 m in length, and
6
Wesselink et al. (Pt = 1000 kN)
24 MN-m2 in flexural stiffness. They were driven into
8 Wesselink et al.(Pt = 2000 kN)
8
Wesselink et al. (P = 1000 kN)
t an onshore test pit that was filled with saturated, unce-
Wesselink et al. (P = 2000 kN)
CF (Pt = 1000 kN)
CF (P = 1000 kN)
t
t
mented calcareous sand. The sand was dredged from
CF (Pt = 2000 kN)
10 10
CF (P = 2000 kN)
t the vicinity of Kingfish B, Bass Strait (Williams et al.
(c) (d) 1988). Laterally loaded tests were performed on both
M/s d
4
pu/sd
2 piles with bending moments, head rotation, head dis-
0 25 50 75 0 10 20 30 40 placement, displacement of the sand surface, and pore
0 0
pressure being recorded. Pile A was initially pushed
monotonically (one load-unload loop at each load level)
5
1 to 106 kN at a free-length of 0.37 m above groundline,
xp/d = 2.36
and at loading rates of 1 kN/min (virgin loading), and
x/d

5 kN/min (unloading and reloading), respectively. The


x/d

2
pile was then pulled monotonically to failure in the
10
opposite direction. Only the response for the push test
Wesselink et al.(P = 1000 kN)
t 3
Wesselink et al.(P = 2000 kN)
t
was examined herein (Case 1). Pile B was subjected to
CF (P = 1000 kN)
t
CF (P = 2000 kN)
t
a series of two-way cyclic loadings with 250 sec period
15 4 in the initial cycle, 60 sec period up to 100 cycles and
(e) (f) 250 sec period in the 101th cycle.
Figure 4. Responses predicted using p-y curves proposed by 3.1.1 Case 1: Test A (static loading)
Wesselink et al. (1988) and Guo (2003): (a) pile head deflec-
tion; (b) maximum bending moment; (c) deflection; (d) soil
In the test pit for Test A, the void ratio of the sand placed
reaction; (e) bending moment: (f) LFP. was 1.21, thus the dry unit weight was interpolated to
be 12.4 kN/m3, as the near surface Kingfish B sand
(Hudson et al. 1988) has a maximum and minimum
from zero at groundline to the maximum value at the dry density (dmax and dmin) of 15.4 and 10.9 kN/m3,
slip depth, xp (1.58d, and 2.36d at Pt  1.0, and respectively, corresponding to a void ratio of 1.07 and
2.0 MN, respectively) (Fig. 4d). Afterwards, the resist- 1.48. The saturated unit weight was estimated as
ance decreases with depth in the elastic zone. The soil 17.85 kN/m3 using a specific gravity of the soil par-
reaction peaked at the slip depth is due to ignorance of ticles of 2.72. A submerged unit weight of 8.04 kN/m3
the transition zone (Guo 2003); and (3) Depth of max- was used as the water level in the test pit was main-
imum bending moment, xm at Pt  2.0 MN reduces tained 50 mm below the sand surface. The friction angle
from 9.3 m (4.47d) to 8.06 m (3.88d), using the harden- may be less than 35 as it was a very loose to loose sand
ing and idealized models, respectively. (Kulhawy & Mayne 1990); and the cone resistance was
Normally, the soil reaction is double differentiated typically 1.5 to 3 MPa. A post peak friction angle of 31
from discrete bending moment data that were obtained for loose sand from Kingfish B was used here. Youngs
from instrumented piles. Hence, it is very sensitive to modulus was taken as the secant one at 50% ultimate
the accuracy of measured moment and the function deviator stress, E50 i.e. E50  5.6 MPa (at c  50 kPa),

375

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 2. Properties of piles.

Case Loading L(m) d(m) e(m) EI(MN


m2) Ep(MPa)

1/2 S/C* 5.9 0.356 0.37 24.0 3.044  104


3 static 32.1 2.137 2.4 79506 7.77  104
4 static 6.0 0.37 0.45 98.0 1.065  105
5/6 S/C* 5.9 0.356 0.254 1.035  10
3 5.065  104

S/C*: S  static loading for cases 1 and 5; C  cyclic loading for cases 2 and 6.

Table 3. Soil parameters used for predictions.

s  Gs 0
Case Sand (kN/m3) (deg.) (MPa) (m) n Ng Lcr/d

1/2 uncemented 8.04 31 2.2 0 1.7 0.9/1.4* 11.4


3 uncemented 8.04 31 3.45 0 1.7 1.2 12.9
4 uncemented 6.4 30 2.0 0 1.7 1.0 15.8
5/6 cemented 8.45 28 3.4 0.15 1.7 2.4/1.5* 11.6

0.9/1.4* Ng  0.9 for static loading, and  1.4 cyclic loading at terminal cycle.

150 M (kN-m) This indicates that the Gs used is lower than the actual
Measured
Wesselink et al. (1988)
0 50 100 150 200 value around the tip.
0
CF At a maximum load of Pt  106 kN, the slip depth
100 xp was calculated as 1.689 m (4.74d). At the depth,
(1) pile deflection was 22.68 mm (i.e. yu/d  6.37%);
Pt (kN)

2
40 kN (2) the effective overburden stress was 13.6 kPa
(1.689  8.04). Under this stress, Figure 3 indicates
x (m)

50 61 kN
106 kN that the deviator stress reaches its peak value at an axial
xp/d = 4.74
4 strain at about 2.5%, thus about 40% the value of yu/d.
Measured The soil within the depth xp must have yielded, thus use
0 Wesselink et al.
0 40 80 120 CF
of the idealized p-y curves is more suitable than that
yt(mm) 6 of the hardening model proposed by Wesselink et al.
(1988), although the latter can also give a good predic-
Figure 5. Predicted vs. measured pile responses for onshore tion on pile response.
Test A, Kingfish B.
3.1.2 Case 2: Test B (Cyclic loading)
Identical properties to those for Test A were adopted
which was obtained from consolidated drained tri- for pile B. Ng was back calculated as 2.5 Kp2 and 1.4 K2p,
axial tests (Hudson et al. 1988). Assuming s  0.3, respectively, for cycle 1 and the terminal cycle against
Gs was obtained as 2.2 MPa. This Gs and Ep of measured pile deflection at 0.11 m height above ground-
3.044  104 MPa (24/(  0.3564/64)) allowed Lcr line as shown in Figure 6. They were about 2.78 and
to be calculated (Guo & Lee 2001) as 11.4d 1.56 times that of Test A under static loading, respect-
(4.05 m embedment length of 5.9 m). ively, which may be partly contributed to local densifi-
For uncemented sand, 0 may be taken as 0 and n as cation of the soil during cyclic loading (Wesselink et al.
1.7. Using GASLFP, Ng was then back calculated as 1988). The corresponding LFPs were plotted in Figure 7
0.9 Kp2 by matching the predicted with measured Ptyt along with that for Test A. The value of Ng at the ter-
(at groundline) relationship, and bending moment distri- minal cycle is about 0.56 times that of cycle 1. This
butions as shown in Figure 5. Except that the maximum degradation may be attributed to increase in the depth of
bending moment at Pt  106 kN was 5.9% higher than gap, xp. For instance, at Pt  110 kN, it increased from
the measured value, pile head displacements and bend- 2.73d at cycle 1 to 4.04 d at the terminal cycle.
ing moments at other load levels were equally well pre-
dicted against those by Wesselink et al. (1988) (Fig. 5),
3.2 Case 3: Kingfish B centrifuge tests
irrespective of the difference in the p-y curves (e.g.
Fig. 2). It should be noted that the predicted bending A series of centrifuge tests were reported by Wesselink
moment near the pile tip was higher than the measured. et al. (1988). Among them, one prototype pile was of

376

Copyright 2005 Taylor & Francis Group plc, London, UK


150 20 M (MN-m)
Measured_cycle 1 Measured
Measured_terminal cycle Wesselink et al. (1988) 0 50 100 150
CF_cycle 1 CF 0
CF_terminal cycle 15
5
xp/d = 3.08 3.3 MN

Pt(MN)
100 xp/d = 2.92 10
10
Pt (kN)

(a) 15

x (m)
6.2 MN
5 20
xp/d = 4.04 (b) 15 MN
50 25
0 30 Measured
0 200 400 600 Wesselink et al. (1988)
CF
yt(mm) 35

0
0 20 40 60 Figure 8. Response for Kingfish B centrifuge test: (a) Ptyt;
yt (mm) (b) Moment.

Figure 6. Predicted vs. measured pile response for onshore


Test B, Kingfish B. agreement was noted between the predicted and meas-
ured Pt-yt (at groundline) relationships (Fig. 8a) and
bending moments (Fig. 8b) as did by Wesselink et al.
pu/sd2 (1988). Similar to Test A, the higher bending moment
0 100 200 300 400
around the pile tip (Fig. 8b) was due to a higher
0 actual value of Gs than the average value used in the
Pile B_cycle 1 calculation.
Pile B_terminal cycle
1 Pile A
At a maximum load Pt of 16.03 MN, the slip depth xp
was calculated as 6.59 m (3.08 d). At the depth, (1)
the pile deflection was 323.78 mm (i.e. yu/d  15.15%);
2 xp/d = 2.92
(2) The effective overburden stress was 53.0 kPa
x/d

(6.59  8.04); and (3) The deviator stress reached its


3 xp/d = 4.04 peak value at an axial strain of around 6% (Fig. 3), i.e.
40% of yu/d. It should be noted that for large diameter
piles, (1) the value of xp/d was smaller than that for small
4
xp/d = diameter, onshore test piles. (2) Ng increased about 30%,
4.74 as also noted previously (Stevens & Audibert 1979).
5 Nevertheless, (3) the onshore and centrifuge tests indi-
cate that piles in Kingfish B sand can be well predicted
Figure 7. LFPs for onshore Test B, Kingfish B. using GASLFP and a LFP generated by using n 1.7,
0  0 and Ng  (0.91.2)K2p.

2.137 m in diameter, 34.5 m in length. Load was applied


3.3 Case 4: North Rankin B in-situ test
at 2.4 m above the sand surface. Having a EpIp of
79506.0 MN-m2 (Ep  7.77  104 MPa), the pile was An in-situ static lateral load test on a free head pile was
installed in a submerged uncemented Kingfish B cal- undertaken in North Rankin B on the north west Shelf
careous sand. of Western Australia (Renfrey et al. 1988). The pile was
Similar to that of the onshore pit tests (Wesselink penetrated into the sand about 6.0 m and subjected to
et al. 1988), the submerged unit weight was taken as loading at 0.45 m above groundline. It consisted of an
8.04 kN/m3; the post peak friction angle as 31. An aver- upper 3.5 m long, 370 mm OD and 30 mm thick pipe
age value of Gs  3.45 MPa was obtained, and Lcr was pile, and a lower 2.35 m long pipe pile of 340 mm OD
calculated to be 12.9d (27.48 m). In particular, the and 12.5 mm thickness that was threaded with the upper
secant moduli at half of ultimate deviator stress, E50 part. The use of an enlarged diameter in the upper layer
were reported as 8.0, 5.6, 12.3, and 10.0 MPa at c of 25, was because of its dominant effect on the pile response.
50, 100, and 200 kPa, respectively, the corresponding Taking the Youngs modulus of the pile material as
Gs was 3.1, 2.2, 4.7, 3.8 MPa (s  0.3). This allows the 2.1  105 MPa, the EpIp of the pile was calculated to be
above-mentioned average Gs to be estimated. 98.0 MN-m2 thus Ep  1.065  105 MPa.
Using GASLFP and a LFP with n 1.7, 0  0 and At the site of North Rankin B Platform, the soil
Ng  1.2 K2p, the pile head displacement and bend- was uncemented to weakly cemented calcareous silt
ing moment were predicted as shown in Figure 8. Close and sand sediments to a depth of 113 m below seabed.

377

Copyright 2005 Taylor & Francis Group plc, London, UK


250 0.6
Measured_Lower jack level
Measured_Upper jack level xp/d = 7.68
CF_Lower jack level Measured
CF_Upper jack level
200 CF

0.4
150 Static

Pt (kN)
Pt (kN)

xp/d = 5.32
100
0.2 xp/d = 8.32

50 Cyclic

0.0
0 0 2 4 6 8
0 50 100 150 200
yt (mm)
yt (mm)

Figure 9. Predicted vs. measured in-situ pile test, North Figure 11. Comparison of measured pile deflection with
Rankin B. predicted.

3.4 Cases 5 and 6: Bombay High model tests


Golait & Katti (1988) presented static (Case 5) and
pu / sd2
cyclic (Case 6) model tests on a pile with L/d  50.
0 50 100 150 200 Relevant properties were tabulated in Tables 2 and
0
3. The stainless steel pipe pile was 25.4 mm OD
and had a flexural rigidity of 1.035 kN-m2 (thus
Ep  5.065  104 MPa). It was placed in artificially
prepared calcareous mix made with 40% beach sand,
2
56% calcium carbonate and 2.5% sodium-meta-silicate.
The sand was equivalent to Bombay High cemented cal-
x/d

careous sand in respect of strength, plasticity, and stress-


xp/d = 5.32 strain relationship. In the 1.4  1.0 m  2.0 m (height)
4
model tank, the sand was compacted to a void ratio of
1.0  0.05 and then saturated. Taking specific gravity
as 2.72 and moisture content as 37% (interpolated from
29%41% for void ratio of 0.71.1), submerged unit
6 weight s was obtained as 8.45 kN/m3. From the hyper-
bolic type of stress-strain relationships obtained from
Figure 10. LFP for in-situ test, North Rankin B. CU tests (Golait & Katti 1986), the secant Youngs modu-
lus at 50% ultimate stress and a confine pressure c
of 50 kPa was found as 8.8 MPa and  as 28. Taking
s  0.3, the value of Gs was obtained as 3.4 MPa. With
Within the upper 30 m, the submerged unit weight s the Ep and Gs, Lcr was obtained as 11.6d (0.295 m L).
was 6.4 kN/m3. It was muddy and very loose near the Using GASLFP with n 1.7, through matching
surface, the  was thus taken as 30 (Reese et al. 1988). the predicted with the measured pile displacements at
Using 0  0, Ng  1.1 K2p and Gs  2.0 MPa, the groundline (Fig. 11), it followed that 0  0.15 m,
pile displacements at locations of the upper and lower Ng  2.4 K2p and 1.54 K2p, respectively for static, and
jacks were predicted using GASLFP, Figure 9, which cyclic loading. The cemented sand did have adhesion
compare quite well with the measured data. Lcr was (Guo 2003), thus 0  0. The values of Ng were quite
obtained as 5.9 m (15.95d). The value of k was calcu- similar to those for the Kingfish B onshore tests (Cases
lated as 5.56 MPa (Guo & Lee 2001). This value agrees 1 & 2). Within the slip depths of 7.68d and 8.32d at
well with 5.72 MPa used by Reese et al. (1988), and the maximum static and cyclic loads, respectively, the
is higher than 34 MPa recommended by Renfrey Ng for cyclic loading was about 0.64 times that for
et al. (1988). At a maximum load Pt of 185 kN, the static loading. The corresponding LFPs are shown in
slip depth xp was calculated as 1.97 m (5.32d ) and the Figure 12.
pile deflection at groundline was about 67.1 mm The good comparisons between the measured and
(18.1%d). predicted pile responses at cycle 1 and the terminal

378

Copyright 2005 Taylor & Francis Group plc, London, UK


Golait, Y.S. & Katti, R.K. 1987. Stress-strain idealization of
pu / s d2 Bombay High calcareous soils. Proc. 8th Asian Reg. Conf.,
0 300 600 900 ICC 86, 2: 173177.
0 Golait, Y.S. & Katti, R.K. 1988. Some aspects of behavior of
piles in calcareous soil media under offshore loading condi-
Static tions. In Jewell and Andrews (eds), Proc. of Engineering for
cyclic calcareous sediments: 199207. Balkema, Rotterdam.
3 Golightly, C.R. & Hyde, A.F.L. 1988. Some fundamental prop-
erties of carbonate sands. In Jewell and Andrews (eds),
x/d

Proc. of Engineering for calcareous sediments: 6978.


Balkema, Rotterdam.
xp/d = 7.68 Guo, W.D. & Lee, F.H. 2001. Theoretical load transfer
6
approach for laterally loaded piles. International Journal
xp/d = of Numerical and Analytical Methods in Geomechanics,
8.32 25(11): 11011129.
Guo, W.D. 2003. A simplified approach for piles due to
9 soil movement. In Culligan, Einstein and Whittle (eds),
Proc. 12th Pan-American Conf. on Soil Mechanics and
Figure 12. LFPs for Bombay High model tests. Geotechnical Engrg., 2: 22152220. Verlag Glckauf
GmbH.
Hudson, M.J., Mostyn, G., Wiltsie, E.A. & Hyden, A.M. 1988.
cycle in Cases 2 and 5 indicate the gapping effect can Properties of near surface Bass Strait soils. In Jewell and
be simulated by reducing Ng to 0.560.64 Ng. Andrews (eds), Proc. of Engineering for calcareous sedi-
ments: 2534. Balkema, Rotterdam.
Kulhawy, F.H. & Mayne, P.W. 1990. Manual on estimating
4 CONCLUSIONS soil properties for foundation design. Report No. EL-6800
Electric Power Research Institute, Cornell Univ. Ithaca,
In this paper, the spreadsheet program GASLFP based N.Y., 250 pp.
on elastic-plastic closed-form solutions for free head McClelland, B. & Focht, J.A. 1958. Soil Modulus for Laterally
piles (Guo 2003) was employed to investigate the static Loaded Piles. Transactions, American Society of Civil
and cyclic responses of laterally loaded piles in cal- Engineers, Paper No. 2954: 10491063.
careous sand. Six case studies demonstrate that, with Novello, E.A. 1999. From static to cyclic p-y data in calcar-
idealized p-y curves, the pile responses can be well pre- eous sediments. In Al-Shafei (ed.), Proc. of Engineering
dicted. For the limiting force profile, it is noted that: for calcareous sediments: 1727. Balkema, Rotterdam.
Randolph, M.F., Jewell, R.J. & Poulos, H.G. 1988. Evaluation
(1) n 1.7; 0  0, and 0.15 m for uncemented, and of pile lateral load performance. In Jewell and Andrews
cemented sands, respectively; Ng  (0.92.5)K2p for (eds), Proc. of Engineering for calcareous sediments:
static loading that is based on post peak friction angle. 639645. Balkema, Rotterdam.
(2) Ng for cyclic loading is (0.560.64) times that for Reese, L.C., Wright, S.G., Roesset, J.M., Hayes, L.H.,
static loading; (3) Maximum xp is (38.3)d, within Dobry, R. & Vallabhan, C.V.G. 1988. Analysis of piles
which reduction of limiting force mainly occurs under subjected to lateral loading by storm-generated waves. In
cyclic loading, indicating the gapping effect. Jewell and Andrews (eds), Proc. of Engineering for calcar-
eous sediments: 647654. Balkema, Rotterdam.
Renfrey, G.E., Waterton, C.A. & Goudoever, Van, P. 1988.
ACKNOWLEDGEMENT Geotechnical data used for the design of the North Rankin
A platform foundation. In Jewell and Andrews (eds),
The work reported herein is currently supported Proc. of Engineering for calcareous sediments: 343355.
by Australian Research Council Discovery grant Balkema, Rotterdam.
DP0209027 awarded to the first author. This financial Stevens, J. B. & Audibert, J.M.E. 1979. Re-examination of p-y
curve formulations. 11th Annual Offshore Technology
assistance is gratefully acknowledged. Conference, Dallas.
Wesselink, B.D., Murff, J.D., Randolph, M.F., Nunez, I.L. &
Hyden, A.M. 1988. Analysis of centrifuge model test data
REFERENCES from laterally loaded piles in calcareous sand. In Jewell and
Andrews (eds), Proc. of Engineering for calcareous sedi-
Dyson, G.J. & Randolph, M. F. 2001. Monotonic lateral load- ments: 261270. Balkema, Rotterdam.
ing of piles in calcareous sand. J. of Geotechnical and Williams, A.F., Dunnavant, T.W., Anderson, S., Equid, D.W. &
Geoenvironmental Engineering, ASCE, 127(4): 346352. Hyden, A.M. 1988. The performance and analysis of lateral
FHWA 1993. COM624P-laterally loaded pile analysis pro- load tests on 356 mm dia piles in reconstituted calcareous
gram for the microcomputer. Report No. FHWA-SA-91-048, sand. In Jewell and Andrews (eds), Proc. of Engineering for
Washington, D.C., USA. calcareous sediments: 271280. Balkema, Rotterdam.

379

Copyright 2005 Taylor & Francis Group plc, London, UK


Foundation solutions for offshore wind turbines

Copyright 2005 Taylor & Francis Group plc, London, UK


Design aspects of monopiles in German offshore wind farms

K. Lesny & J. Wiemann


Institute of Soil Mechanics and Foundation Engineering, University of Duisburg-Essen, Germany

ABSTRACT: As a result of the German energy policy aimed at a broad expansion of renewable energies, a
number of wind farms in the North Sea and the Baltic Sea consisting of several hundred wind energy convert-
ers of up to 5 MW rated power each are now in planning. The present paper provides an overview of the current
plans and the specific site conditions in German waters. Using the example of a monopile foundation important
geotechnical design aspects are discussed, reviewing the application of state-of-the-art offshore design guide-
lines. The influence of large monopile diameters on pile-soil-interaction is discussed by comparing the results
of a finite-element-modeling to standard p-y lateral loading analysis procedures.

1 OFFSHORE WIND FARM PROJECTS Table 1. Approved offshore wind farm projects in the
Exclusive Economic Zone (BSH 2004).
1.1 Political background
Distance Water depth
The ambitious goals of the federal government in Wind farm Installed power [km] [m]
Germany call for the percentage of renewable energies
to be increased from todays 3.5% to 12.5% in 2010 and Borkum 180  35 MW 34 2329
20% in 2020. Being especially suitable to Germanys Riffgrund
renewable energy conditions, offshore wind energy Borkum
plays a major role in attaining these goals. Taking into Riffgrund
West 458  2.5/5 MW 4050 3035
account the high risk and the huge investments associ- Borkum 208  3.55 MW 4350 30
ated with offshore wind energy the renewable energy West
sources act (2004) guaranties special market prices of Butendiek 80  3 MW 3040 1620
9.1 ct./kWh to be paid over a period of 12 years for off- Sandbank24 980  35 MW 100 3040
shore wind energy converters being installed before Amrumbank 80  3.55 MW 35 2125
2010. Additionally, the energy suppliers are bound to West
collect the power from renewable energy sources. Nordsee Ost 250  45 MW 30 1924

1.2 Current situation


the Exclusive Economic Zone which have been
Realizing the above mentioned goals means that at approved up to now are summarized.
least 5000 offshore wind energy converters with a
rated power of 5 MW each need to be installed within
the next two decades. Due to the attractive funding a
2 SITE CONDITIONS
lot of wind farm projects with 20 to 1000 wind energy
converters per wind farm have been requested up to
2.1 Environmental conditions
now. They are located mainly in the Exclusive
Economic Zone, which is in wide parts identical with The environmental conditions in the southern North
the German continental shelf. Only few have been Sea are characterized by high maximum wave heights
requested within the 12-sm-Zone, because large parts and high wind speeds, as shown in Table 2.
of this area are reserved for environmental protection, The wind climate in the Baltic Sea is considered to
shipping routes, cable or pipeline routes. Hence, the be similar, whereas the wave climate with maximum
distance of the requested sites to shore is up to wave heights of about 9 m and a storm surge current
120 km, and the water depth ranges between 20 and surface velocity of 0.10.3 m/s is moderate. In con-
50 m. In Table 1, some of the wind farm projects in trast to the North Sea ice loading has to be taken into

383

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 2. Environmental conditions in the southern
North Sea.

Wind:
Hub-height 50-year extreme 10 min mean wind 50.0 m/s
Hub-height 50-year extreme 5 s gust 60.0 m/s
Water depth:
Mean water depth 35.0 m
50-year extreme water depth 41.0 m
Wave & currents:
50-year maximum wave height Hmax 22.3 m
Related wave period T 14.5 s
50-year tidal current surface velocity 1.71 m/s
50-year storm surge current surface velocity 0.43 m/s

Table 3. Quasi-static loading at mud line for a 5-MW


turbine.

North Sea Baltic Sea

Vertical Load V [MN] 35 35 Figure 1. Examples of typical marine strata in the southern
Horizontal Load H [MN] 16 6 North Sea (left) and the Baltic Sea north of Ruegen (right).
Bending Moment M [MNm] 562 280
Torsional Moment MT [MNm] 4 4
streams during the warm intervals are a typical feature
of the southern North Sea. These channels are filled
with various sediments of usually low bearing capacity.
account. However, reliable data on the environmental The marine soil strata in the Baltic Sea is more
conditions is not available yet. complex. The upper Holocene layers mainly consist of
Assuming an offshore wind energy converter with mud (sandy, silty soils with organic contents, sapro-
a rated power of 5 MW, a hub-height of about 95 m pelic mud) (Gigawind 2004). These layers may have a
above still-water-level and a rotor of 125 m diameter thickness of up to 20 m and more and have a very low
these environmental conditions lead to an approxi- bearing capacity. The Pleistocene soil layers contain
mate quasi-static loading at mud line as shown in highly over-consolidated glacial till over cemented
Table 3 (Richwien & Lesny 2003). The loading refers clays with stones and boulders. North of Ruegen,
to load case E2.2 according to the regulations of the large chalk deposits are to be found below the glacial
German Lloyd (GL 1999). In E2.2 the 50-year extreme till which is covered only by thin sandy sediment
wave loading is combined with a reduced 50-year wind layers. Typical examples of the marine strata in the
gust and stand-by conditions for the turbine. This leads southern North Sea and the Baltic Sea north of Ruegen
to extremely high bending moments, which control the are depicted in Figure 1. The characteristic values of the
foundation design. relevant soil parameters are based on current experi-
ences (Gigawind 2004).
2.2 Soil conditions
The submarine strata in the southern North Sea is char-
3 FOUNDATION DESIGN
acterized predominantly by non-cohesive soil layers of
varying density (Wiemann et al. 2002). Holocene sedi-
3.1 Overview
ments (silty sands or sandy silt, tidal mud) of low
bearing capacity overlay Pleistocene sediments. The Most of the foundation concepts for offshore wind
thickness of the Holocene sediments is usually less than energy converters currently under discussion are
3 m, but may reach 15 m in some areas. The Pleistocene adopted from traditional offshore engineering (Fig. 2),
sediments consist of medium dense to dense fine to such as the gravity foundation, the tripod or jacket
medium sands or coarse sands. However, the density foundation. The bucket foundation is a relatively new
especially of the deeper layers can be very high. Basin design concept and still under development. A first
deposits of highly over-consolidated glacial till or clay prototype with a 3 MW turbine has been installed
can be found as well. Boulders up to 10 m3 are located 2003 in Frederikshavn (Ibsen 2004).
especially along the Pleistocene-Holocene boundary. Monopiles frequently have been installed as foun-
Huge channel systems formed by the large melt water dations for offshore wind energy converters, e.g. at

384

Copyright 2005 Taylor & Francis Group plc, London, UK


The spring characteristics define the non-linear
relation between the subgrade reaction per unit length
of the pile p and the horizontal pile deflection y tak-
ing into account the material behavior of the soil and
the pile stiffness.
The p-y-curves are mostly based on theoretical
failure models for the determination of the maximum
subgrade reaction pu(z), whereas their shape and
Figure 2. Foundation concepts for offshore wind energy parameters are derived from field test results. They have
converters. been established for different types of soil, e. g. for sand,
stiff or soft clay. Reese et al. (1974) defined com-
posed and therefore non-steady p-y-curves for sand
as shown in Figure 3. In API (2000) these p-y-curves
are substituted by a closed hyperbolic formula:

(1)

In Equation 1 ks,0 represents the initial modulus of


subgrade reaction (see Fig. 3) depending on the angle
of internal friction or the relative density of the soil.
The quantity pu(z) is the maximum subgrade reaction:
Figure 3. Beam-on-elastic-springs-model and p-y-curves
according to Reese et al. (1974) for sand.
(2)

Horns Rev in Denmark or Arklow Bank in Ireland. The coefficients C1, C2 and C3 are functions of the
Beside its simplicity the major advantage of this angle of internal friction and can be determined acc.
design concept is that the loading due to wave, cur- to API (2000). The parameter A is a correction factor
rents and ice can be clearly defined. Another aspect is for the theoretical value of pu(z) in Equation 2. It has
the limited consumption of the ground, which is been derived from field tests with piles of approxi-
favourable for the ecological acceptance of an off- mately 0.60 m diameter (Reese et al. 1974) to:
shore wind farm. However, the magnitude and the
combination of loads as shown in Table 3 call for pile
diameters of approximately 6 m and more which are far
(3)
beyond any experiences regarding design and installa-
tion. Pile driving in the dense North Sea sands or the
hard sediments of the Baltic Sea will be limited, thus
calling for other installation methods. Furthermore, A thorough description of the p-y-method and an
there is no sufficient experience up to now regarding overview of different p-y-curves and new develop-
a reliable design of large diameter monopiles. ments is given in Reese & van Impe (2001).
The standard p-y procedure is a well established
design method, which has been verified through the
3.2 Design aspects of monopiles experiences gained over the past 30 years. However,
Monopile foundations transfer bending moments and these experiences are limited to piles up to 12 m
horizontal loads into the subsoil by horizontal sub- diameter. Thus, it has to be investigated if this design
grade reaction. The standard design procedure for method thoroughly reflects the soil-pile interaction of
these structures according to the relevant guidelines large diameter monopiles.
in offshore engineering (API 2000, GL 1999, DNV
2004) is the well-known p-y-method. Within this
4 APPLICABILITY OF THE P-Y-METHOD TO
approach the pile-soil system is modeled as a beam
LARGE DIAMETER MONOPILES
resting on elastic springs (Fig. 3). According to the
Winkler-hypothesis the springs act independently of
4.1 Numerical model and input parameters
each other. Hence, the subgrade reaction at the pile in
a certain depth is not influenced by the pile displace- The applicability of the standard p-y-method to
ments at other depths. large diameter monopiles has been investigated by

385

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 4. Cross-sectional parameters of the analysed pipe
piles.

Diameter, Thickness, Cross-section, Moment of inertia,


d [m] t [m] A [m2] I [m4]

1 0.02 0.062 0.0074


2 0.03 0.186 0.0901
4 0.05 0.621 1.2103
6 0.07 1.304 5.7330

Table 5. Soil parameters for Essen Sand. Figure 4. Numerical model of a laterally loaded pile
distribution of the void ratio.
Relative density ID 0.55
Void ratio e 0.629
Angle of internal friction  40.5
Oedometric modulus Es (mean stress range) 5080 MN/m2
Weight/submerged weight / 17/10 kN/m3
Initial modulus of subgrade reaction ks,0 19,000 kN/m3

Finite-Element modeling of the pile-soil behavior. In


the FE-analyses a steel monopile was modeled as a
half pipe pile assuming linear-elastic material behav-
ior. The pile parameters are given in Table 4. The pile-
soil contact was modeled by the Coulomb friction law
excluding tension stresses.
A homogenous non-cohesive soil was assumed and
its material behavior was described by a linear elastic
ideal plastic constitutive model with the Mohr-Coulomb
failure criterion. Additionally, a hypoplastic constitu-
tive model as formulated by von Wolffersdorff (1996)
was used. The basic soil parameters for the Essen
Sand used in all analyses are summarized in Table 5.
As an example, the numerical model of a laterally Figure 5. Pile resistance acc. to the p-y-method (API
loaded pipe pile showing the distribution of the void 2000) and results of the FE-analysis.
ratio is depicted in Figure 4.
FE-analysis using the hypoplastic material model for
4.2 Analysis of the maximum subgrade reaction different contact friction angles  between 0 and 
are shown.
Within this analysis pile slices of constant length
Apparently, the results according to the standard
z  L  2.5 m with different pile diameters of 1, 2, 4
p-y-method represent the outer boundaries of the FE-
and 6 m have been investigated. For this length the
results. Equation 2 refers to the solution by Reese
ratio z/d 2.625 and therefore the correction factor
et al. (1974), in which the friction angle was assumed
A acc. to Equation 3 is A  0.9. The ultimate bearing
to be  0. This leads to a lower boundary of the
capacity results from the maximum subgrade reaction
pile resistance. For d  1 m this solution coincides
pu(z) using Equations 2 and 3 integrated over the
with the FE-result for   0, whereas for larger
depth z of the pile slice:
diameters it is a conservative approach.
Modifying the theoretical value of pu(z) in
Equation 2 with the correction factor A acc. to
(4) Equation 3 results in an upper boundary of the pile
resistance (Fig. 5). As mentioned above, the factor A
was derived from field test results, and thus includes
In Figure 5 the pile resistance acc. to Equation 4 is different influences from the installation process,
depicted with and without consideration of the possible inhomogenities of the soil or the friction
correction factor A. Additionally, the results of the between pile and soil according to the conditions on

386

Copyright 2005 Taylor & Francis Group plc, London, UK


the test site. However, these influences cannot be Table 6. Pre-design of the piles according to the standard
specified separately. p-y-method (API 2000) for a pile head distortion of
The FE-results in Figure 5 show a large influence   0,7.
of the friction angle  which apparently increases in
Diameter, Length, Moment, Displacements,
correspondence to an increasing diameter. For d  1 m
d [m] L [m] M [MNm] yhead [m]
the results of the FE-analysis for    coincide
with the results of the p-y-method. However, for a 1 10.6 3.98 0.031
more realistic contact friction angle of   2/3 the 2 17.1 30.2 0.049
pile resistance according to the p-y-method is about 4 28.5 245.0 0.081
65% larger than the FE-results for all investigated 6 38.9 855.0 0.109
diameters. Hence, it can be concluded that the theo-
retical solution acc. to Equation 2 leads to a conser-
vative estimation of the maximum subgrade reaction,
but the correction factor A acc. to Equation 3 should
be applied with great care.

4.3 Analysis of the pile-soil-interaction


For the analysis of the pile-soil-interaction all investi-
gated piles have been pre-designed using the p-y-
method (API 2000) for the parameters displayed in
Tables 4 and 5. Table 6 shows the pile lengths together
with the bending moments and the pile head displace-
ments for a constant pile head distortion of   0.7.
This value has been provided by the serviceability
conditions of a typical 5-MW turbine.
In the FE-analyses the elasto-plastic material model
assuming an oedometric modulus parabolically increas-
ing with depth z and the hypoplastic material model
have been used. The resultant deflection lines are
depicted in Figure 6.
Figure 6 shows that the deflection line for a pile
diameter of d  1 m according to the p-y-method Figure 6. Deflection lines for moment loading according
coincides with the deflection line of the FE-calcula- to the standard p-y-method (API, 2000) and FE-results.
tion with elasto-plastic material behavior. However,
the deflection line of the 6 m pile resulting from the
FE-analysis shows no rigid fixation of the pile in con- In Equation 5 the modulus of subgrade reaction is
trast to the result of the p-y-method, which leads to assumed to increase linearly with depth: ks(z)  ks,max .
much greater pile head displacements. z/L. According to Terzaghi ks can be estimated to:
In both cases, the numerical results based on
hypoplasticity show a smaller soil resistance against
the pile compared to the elasto-plastic material model (6)
resulting in larger pile head displacement. This results
from the stress dependent formulation of the
hypoplastic material model. Especially in zones near For the pile lengths LLc in Table 6 calculated
the subsurface, which contribute predominantly to the with the p-y-method Equations 5 and 6 result in an
subgrade reaction, the low stress level leads to very unrealistically high stiffness of the soil. E.g. for the
large strains (see also von Wolffersdorff 2004). 6 m pile the stiffness is ES,max  538.4 MN/m2 at the
However, the pile length determined with the p-y- pile tip. The initial slope of the p-y-curves ks,0  z
method is apparently not sufficient. This is due to the according to Figure 3 and Table 5 also implies a linear
stiffness relations implicated by the p-y-method. increase of the modulus of subgrade reaction with
Titze (1970) calculated the critical length Lc of the depth, which leads to a similar magnitude at the pile
pile ensuring a rigid fixation to: tip of ES,max  739.1 MN/m2. If these high stiffness
values are considered in the FE-calculations, the
resulting deflection lines show a rigid fixation of the
(5) pile as well (Fig. 7). However, the pile head displace-
ments for a linear distribution of ES in the FE-analysis

387

Copyright 2005 Taylor & Francis Group plc, London, UK


larger diameter, however, need a greater critical
length. This leads to different pile-soil-stiffness rela-
tions, which are not covered by the test results.
To take these effects into account a simple procedure
has been derived to modify the initial modulus of
subgrade reaction in the p-y-curve acc. to Equation 1
as follows:

(7)

In Equation 7 a diameter of 1 m was chosen as ref-


erence quantity dref. It represents the diameter up to
which the p-y-method has been verified by experi-
ence. The exponent a has been determined to 0.5 for
dense Sand and 0.6 for medium dense Sand. More
details are described in Wiemann & Lesny (2004).
The results of the modified p-y-method for a pile
Figure 7. Deflection lines according to the p-y-method and of 6 m diameter are depicted in Figure 8. The modifi-
FE-results with oedometric moduli derived from the p-y- cation acc. to Equation 7 leads to greater pile length
method. to ensure rigid fixation of the pile. For this increased
pile length the FE-results show a fixed pile, too.
Furthermore, the difference between the pile head
displacements are significantly smaller as compared
to the results of the standard p-y-method. However,
they are still larger as the deflections calculated with
the modified p-y-method. This effect may result from
the linear increasing stiffness assumed in the calcula-
tions as shown already in Figure 7.

5 CONCLUSION

The offshore wind farm projects currently planned in


Germany are very specific due to difficult site condi-
tions and the great number of wind energy converters
to be erected in a limit period of time. This especially
emphasizes the need for innovative foundation con-
cepts to guarantee the economic benefit of the pro-
jects. The monopile is one of the foundation concepts
under discussion. Monopiles are designed using the
well-known p-y-method. However, it has been shown
based on FE-analyses that the standard p-y-method
Figure 8. Deflection line for moment loading with modi- overestimates the pile-soil-stiffness of large diameter
fied initial slope of the p-y-curves. monopiles at great depths which may result in an
insufficient pile length. A simple procedure to modify
the initial stiffness of the p-y-curves has been pre-
as assumed for the p-y-curves are larger than the dis- sented which better accounts for these effects. The
placements resulting from the p-y-method. Only if a pile length can be properly calculated using this
power law distribution of the stiffness is assumed, the method, whereas the absolute value of the pile head
deflection lines resulting from the FE-analysis coin- displacements should be used with great care.
cide with the results of the p-y-method. Finally, it should be mentioned that the analyses
The reason for these differences may be found in presented here are limited to quasi-static loading of
the derivation of the p-y-curves from field test results. the monopile. The effects of cyclic loading on the
These field tests represent certain pile-soil-stiffness monopile behavior are not well understood up to now
relations for pile diameters smaller than 1 m. Piles of and are a major task of future research.

388

Copyright 2005 Taylor & Francis Group plc, London, UK


ACKNOWLEDGEMENTS Lesny, K. Richwien, W. & Wiemann, J. 2002. Gruendungst-
echnische Randbedingungen fr den Bau von Offshore-
The work presented within this paper is part of the Windenergieanlagen in der Deutschen Bucht.
GIGAWIND and GIGAWIND research projects Bauingenieur Band 77.
Reese, L. C. & van Impe, W. F. 2001. Single Piles and
which are funded by the Federal Ministry for the Pile Groups Under Lateral Loading. Rotterdam:
Environment, Nature Conservation and Nuclear Balkema.
Safety. Its support is gratefully acknowledged. Reese, L. C.; Cox, W. R.; Koop, F. D. 1974. Analysis of
Laterally Loaded Piles in Sand. Proceedings of the VI
Annual Offshore Technology Conference. Dallas.
REFERENCES Richwien, W. & Lesny, K. 2003. Machbarkeitsstudie fuer
die Monopile-Gruendung einer Offshore-Windenerg-
API 2000. American Petroleum Institute. Recommended ieanlage Windpark Arkona-Becken Suedost. Universitaet
Practice for Planning, Design and Constructing Fixed Duisburg-Essen.
Offshore Platforms - Working Stress Design. Washington Titze, E. 1970. Ueber den seitlichen Bodenwiderstand bei
D.C.: API Publishing Services. Pfahlgruendungen. Bauingenieur-Praxis Heft 77. Berlin:
BSH 2004. Bundesamt fr Seeschifffahrt und Hydrographie, Verlag Ernst u. Sohn.
Hamburg & Rostock. http://www.bsh.de Wiemann, J., Lesny, K. & Richwien, W. 2002. Mitteilungen
DNV 2004. Det Norske Veritas. Design of Offshore Wind aus dem Fachgebiet Grundbau und Bodenmechanik Heft
Turbine Structures. Offshore Standard DNV-OS-J101: 29. Gruendung von Offshore-Windenergieanlagen
DNV. Gruendungskonzepte und geotechnische Grundlagen.
Gesetz zur Neuregelung des Rechts der Erneuerbaren Essen: Prof. Dr.-Ing. W. Richwien.
Energien im Strombereich 2004. Bundesgesetzblatt Wiemann, J. & Lesny, K. 2004. Evaluation of the Pile
Jahrgang 2004 Teil I Nr. 40. Bonn. Diameter Effects on Soil-Pile Stiffness. Proceedings of
Gigawind 2004. Abschlussbericht 20002003. Bau the 7th German Wind Energy Conference (DEWEK),
und umwelttechnische Aspekte von Off-shore 2021 October 2004. Wilhelmshaven.
Windenergieanlagen. Hannover: Gigawind. von Wolffersdorff, P.-A. 1996. A Hypoplastic Relation for
GL 1999. Rules and Regulations. Offshore Windenergy Granular Materials with a Predefined Limit State
Converters. Hamburg: Germanischer Lloyd. Surface. Mechanics of Cohesive-Frictional Materials
Ibsen, L. B. 2004. Development of the Bucket Foundation Vol. 1: 251271.
for Offshore Wind Turbines, a Novel Principle. 3. von Wolffersdorff, P.-A. 2004. Implementation of
Symposium Offshore-Windenergie, Bau- und umwelt- Hypoplasticity. Plaxis User-Defined Soil Models
technische Aspekte. Hannover: Gigawind. Meeting. Bundesanstalt fuer Wasserbau. Karlsruhe.

389

Copyright 2005 Taylor & Francis Group plc, London, UK


Finite element modelling of horizontally loaded monopile foundations for
offshore wind energy converters in Germany

K. Abdel-Rahman & M. Achmus


Institute of Soil Mechanics, Foundation Engineering and Waterpower Engineering,
University of Hannover, Germany

ABSTRACT: Offshore wind energy offers a huge potential for the expansion of renewable energies in Germany.
High demands have to be made concerning foundation design and construction methods to find an economically
and technically optimal solution. A promising foundation concept which can be used in this field is the monopile,
which in principle is a prolongation of the tower shaft into the ground. The behaviour of large-diameter monopiles
under the horizontal loading induced by wind and wave loads is analysed using a three-dimensional finite elem-
ent model. Non-linear elastoplastic soil behaviour is taken into account. The results of the numerical simula-
tions are presented and compared with the results of the p-y method commonly used for the design of laterally
loaded piles.

1 INTRODUCTION Currently, artificial competitiveness is maintained in


Germany by a law forcing energy suppliers to buy wind
Offshore wind farms promise to become an important current at a fixed price.
source of energy in the near future. It is expected that However, the costs of wind turbines are falling and
within 10 years, wind parks with a total capacity of thou- are expected to continue doing so over the coming
sands of megawatts will be installed in European seas. decade. Also, experience will be gained in building
Onshore wind energy has grown enormously over offshore wind farms, so that the offshore construction
the last decade to the point where it generates more industry will likely find other cost-savings. Hence it
than 10% of all electricity in certain regions (such as is hoped that offshore wind energy will become really
Denmark, Schleswig-Holstein in Germany). However, competitive in time.
due to noise and visual pollution, further expansion of
onshore wind energy is limited.
These problems in the development of wind energy 2 FOUNDATION REQUIREMENTS
can be resolved by the installation of wind energy farms
in offshore regions, e.g. in the North Sea and in the A number of different foundation types can be used
Baltic Sea. This solution has the following advantages: for offshore wind energy converters (OWECs). The
major types under discussion are the monopile, the
1. availability of large continuous areas, suitable for gravity and the tripod foundations.
major projects, A monopile foundation consists of a large-diameter
2. higher wind speeds, which generally increase with steel pile, which is in principle simply a prolongation
distance from the shore, of the tower shaft into the ground. The monopile must
3. less wind turbulence, which allows the turbines to be able to transfer both lateral and axial loads from
harvest the energy more effectively and reduces the the structure into the seabed. The steel piles are of
fatigue loads on the turbine. simple tubular construction which is inexpensive to
produce and provide a low cost fabrication option.
Of course, also problems arise in moving offshore. The gravity foundation, unlike the piled founda-
Offshore installation costs are much higher than tion, is designed with the objective of avoiding tensile
onshore, and the integration of the offshore wind farm loads between the support structure and the seabed.
into the electricity network is much more expensive. This is achieved by providing sufficient dead loads to
Economic construction and design methods are indis- stabilise the structure under the overturning moments
pensable to make offshore wind energy competitive. which result from wind and wave action.

391

Copyright 2005 Taylor & Francis Group plc, London, UK


+ 190 m Soil structure The analysis of the behaviour of large monopiles
0m
under monotonic horizontal loading is the objective of
S,u
this paper.
-1 m
Assuming a water depth of 30 m and a maximum
+ 130 m
design wave height of 14.5 m, the design horizontal
load for a monopile with a diameter of 7.5 m amounts to
mS - fS about 8 MN, the resultant horizontal force acting about
30 m above sea level, i.e. nearly at still water level.
Thus, the corresponding bending moment at sea bed
+ 30 m
level is about 240 MNm.
-7 m
Additionally a vertical load of 10 MN representing
_ 0,0 m
+ the own weight of the turbine, the blades and the tower
mS,g
was assumed. Such loads have to be considered analyz-
ing the behaviour of monopiles.
D = 7,5 m

3 COMMON DESIGN OF HORIZONTALLY


Figure 1. Monopile foundation of a wind energy plant. LOADED PILES

The design procedure for OWEC foundations is in


The tripod consists of a spatial steel frame trans- Germany given in the Germanische Lloyd rules and
ferring the forces from the tower to primarily tension regulations (GL 1999). In this regulation concerning
and compression forces in three hollow steel piles the behaviour of piles under horizontal loading refer-
driven into the seabed, located in the corners of a tri- ence is made to the regulation code of the American
angle. In contrast to the monopile, the steel piles used Petroleum Institute (API 2000). The Norwegian guide-
are of smaller diameters (less than 2 m). lines (DNV 2004) also refer to the API code. In the API
The soil conditions in the North Sea and the Baltic code the p-y method is recommended for the design
Sea are characterized by pleistocene and holocene sedi- of horizontally loaded piles.
ments. The pleistocene soils were strongly preloaded In principle, the p-y method is a subgrade modulus
during ice ages and are thus highly overconsolidated method with non-linear and depth-dependent load-
(e. g. boulder clays) or in dense state (sands and deformation (p-y) characteristics of the soil springs. In
gravels). They are overlayed by holocene soils of the API code, the following procedure is given to con-
varying thickness like loose or medium dense silty sands struct p-y curves for sandy soils:
or at some locations peat or mud (Lesny et al. 2002). 1) The ultimate lateral resistance per unit length pu
In the past OWECs established in the North Sea is taken as the minimum of two expressions. The first
and in the Baltic Sea have rated power outputs of max- one is valid for shallow depths, whereas the second is
imum 2 MW and are located at small distances from valid for greater depths:
the coast at moderate water depths of up to about 8 m.
Most of these structures are founded on monopiles. (1)
The OWECs now planned in the German offshore
areas shall have rated outputs of up to 5 MW and are thus
much larger than the converters already installed. The (2)
planned plants are to have hub heights of approximately
100 m and rotor diameters of approximately 120 m Herein z is the given depth in metres, D is the average
(Fig. 1). Moreover, the wind farms shall be installed in pile diameter in metres,  is the effective unit weight
areas far away from the coast. At these locations, water of soil (kN/m3). The coefficients c1, c2, c3 are depend-
depths from 20 m to 40 m (or even 50 m) are expected. ent on the friction angle of the soil.
For the large depths the tripod foundation is the 2) The p-y curve is given at a specific depth by the
most promising foundation type. But, for depths of up following expression:
to about 30 m the monopile foundation is thought to
be an alternative. This foundation type is simple and
elegant and has advantages concerning the damage (3)
risks in case of a ship collision. However, for mono-
piles with diameters in the range of 6 to 8 m, which
are necessary for OWEC sizes and water depths des- where A  3.0
0.8 z/D 0.9 for static loading and
cribed above, no experience exists concerning the A  0.9 for cyclic loading, p is the soil resistance per
load-deformation behaviour. unit length, y is the actual lateral deflection and k is the

392

Copyright 2005 Taylor & Francis Group plc, London, UK


initial modulus of subgrade reaction determined as a
function of the friction angle.
This method is verified only for piles with diameters
of up to about 2 m. The question is to be answered,
whether the method can be used also for the design of
large-diameter piles.
In the following, results of numerical calculations of
the load-deformation behaviour of monopiles are pre-
sented and compared with the results of the API p-y
method. The calculations with the API method were car-
ried out by means of the LPILE program (Lpile 2000).

4 FINITE ELEMENT MODELLING

4.1 Model features


For the investigation of the behaviour of laterally loaded
monopiles with large diameters, a three-dimensional
(3-D) numerical model was established. Figure 2. Finite element mesh.
The computations were done using the finite elem-
ent program system ABAQUS (Abaqus 2004). In
Table 1. Material parameters used for dense sand.
order to carry out many calculations for varying bound-
ary and loading conditions, a large computer system Unit buoyant weight  11.0 kN/m3
with parallel processor technology was used to min- Oedometric stiffness parameter  600
imize the time effort. Oedometric stiffness parameter  0.55
The aim of the investigation was to analyse the Poissons ratio  0.25
behaviour of a large monopile in principle and to check Internal friction angle  35.0
whether the API method can be used for such large Dilation angle  5.0
piles. For that, an idealized homogeneous soil consist- Cohesion c 0.0 kN/m2
ing of dense sand was considered. A monopile diam-
eter of D  7.5 m and a wall thickness of 9 cm was
assumed. The loading consists of a resultant horizontal The material behaviour of sand and soil in general
force acting at a given height h above the sea bed level. is very complex. In the case of monotonic loading con-
The bending moment at sea level is thus M  H  h. By sidered here, essential requirements on the material law
variation of H and h any water depths or load combin- are the consideration of the non-linear, stress-dependent
ations can be considered. Additionally, a vertical load soil stiffness and the consideration of possible shear fail-
was applied to take the structures weight into account. ure. An elasto-plastic material law with Mohr-Coulomb
Due to the symmetric loading condition only a half- failure criterion was used. The soil stiffness is repre-
cylinder representing the sub-soil and the monopile sented by a stiffness modulus for oedometric compres-
was considered. The discretized model had a diameter sion ES and a Poissons ratio . A stress dependency of
of 90 m, which is twelve times the pile diameter. The the stiffness modulus was accounted as follows:
bottom boundary of the model was taken 15 m below
the base of the monopile. With these model lengths
the calculated behaviour of the pile is not influenced (4)
by the boundaries. A view of the discretized model
area is given in Figure 2.
For the soil as well as for the pile 8-node continuum Herein at  100 kN/m2 is a reference (atmospheric)
elements were used. The interaction behaviour between stress and  is the current mean principal stress in the
the monopile and the sand soil is simulated using contact considered soil element. The parameter  determines
elements. The maximum shear stress in the contact the soil stiffness at the reference stress state and the
area is determined by a friction coefficient. For the parameter  rules the stress dependency of the soil
calculations presented herein this coefficient was set stiffness.
to   0.4. The advantage of the material law used is that it can
The material behaviour of the monopile was assumed be generally used for non-cohesive as well as for cohe-
to be linear elastic with the parameters E  2.1  sive soils. The parameters used for the calculations pre-
105 MN/m2 (Youngs modulus) and   0.2 (Poissons sented here are typical for a dense sand (EAU 1996) and
ratio) for steel. are given in Table 1.

393

Copyright 2005 Taylor & Francis Group plc, London, UK


The finite element calculation is executed stepwised.
At first, for the generation of the initial stress state the
whole model area is discretized using soil elements
only. Subsequently, the monopile is generated by
replacing the soil elements located at the pile position
by steel elements and activating the contact conditions
between pile and soil. Then the vertical load is applied,
and finally the horizontal load is applied and increased
step by step. The monopile elements were extended
above the ground surface of the model in order to real-
ize different load combinations (horizontal forces and
bending moments).

4.2 Model results


A monopile with a diameter of D  7.5 m and an
embedded length below sea bed of L  30 km is taken
as the basic case. With an assumed water depth of 30 m
the resultant design wave load acts at h  30 m above Figure 3. Mobilised bedding pressure for a monopile with
sea bed level, i. e. h/L  1.0. D  7.5 m, L  30 m in dense sand, H  8 MN, M 
For a resultant horizontal load of 8 MN, which is a H  h  240 MNm.
possible design load for the considered water depth,
the horizontal (bedding) stresses acting on the pile in
the symmetry plane are shown in Figure 3. The char-
acteristic loading behaviour of the pile with bedding
stresses of opposite sign above and below a point of
rotation can be seen clearly. For the considered case
the point of rotation lies about 22 m below sea bed.
For the horizontal force of H  8 MN and the
bending moment at sea bed level of M  240 MNm
the pile displacement at sea bed level amounts to
about w  4 cm and the angle of rotation to about
  0.17.
From a practical point of view, the design proced-
ure for a monopile foundation depends mainly on Figure 4. Comparison between finite element results and
API method results regarding the horizontal displacement at
the calculation of the horizontal displacement and
sea bed level (D  7.5 m, h  L  30.0 m).
the pile rotation with respect to the applied loading
conditions.
In Figures 4 and 5 the calculated force-displacement
and the force-rotation relationships for the basic case
considered are given. Also the results obtained with
the API method are shown for comparison. With this
method, the pile deformations are smaller and may
thus be underestimated for large horizontal forces.
However, for horizontal forces less than about 6 MN
the results of the API method and the numerical cal-
culations are in fairly good agreement.

4.3 Variation of pile length and height of loading Figure 5. Comparison between finite element results and
point API method results regarding the rotation angle at sea bed
In the scope of a parametric study the height of the level (D  7.5 m, h  L  30.0 m).
horizontal load above sea bed and with that the bend-
ing moment was varied in order to take different The results are shown in Figures 6 and 7 as load-
water depths or load combinations into consideration. deformation and load-rotation curves, respectively.
Additionally, a pile with a smaller embedded length The expected influence of the pile length and the
of L  20 m was analysed. height of the horizontal load is evident. Regarding

394

Copyright 2005 Taylor & Francis Group plc, London, UK


Figure 8. Equivalent lateral stiffness curves for a monopile
Figure 6. Force-displacement curves for a monopile D 
D  7.5 m, L  30 m embedded in dense sand.
7.5 m, L  30 m/L  20 m embedded in dense sand.

Figure 9. Equivalent rotational stiffness curves for a


Figure 7. Force-rotation curves for a monopile D  7.5 m, monopile D  7.5 m, L  30 m embedded in dense sand.
L  30 m/L  20 m embedded in dense sand.
Thus, such results also need verification. However, in
this aspect it should be considered that with lower order to investigate the influence of the monopile diam-
h/L-values also the design load is generally lower, eter, the embedded length and the type of loading, the
because shallower water normally corresponds with numerical model presented seems to be more reliable
lower design wave heights. than the API method.
For preliminary design steps diagrams can be help-
ful, which allow a simple determination of the approxi-
5 CONCLUSIONS mate pile deformations to be expected for a specific
case.
According to the results of the numerical calculations As integral stiffness values of a monopile founda-
carried out for monopiles of large diameter for high tion two spring stiffnesses can be defined as follows:
design loads, the p-y curve method given in API (2000)
understimates pile deformations. The main reason for
these results is probably an overestimation of the ini-
tial soil stiffness in large depths by the API method. (5)
Moreover, for a large-diameter pile the shearing resist-
ance in the pile tip area may play an important role
compared to a small-diameter piles. Since this method
is not verified by measurements at large-diameter piles,
it should in general not be used for the design of Herein H and M  H  h are the load values, and w and
monopile foundations.  are the pile displacement and rotation at sea bed level.
Thus, for the time being the execution of numerical From the results of numerical calculations diagrams
investigations is recommended for the design of the can be developed, from which these integral stiffness
large monopiles planned in the German offshore areas. values can be determined. For the basic case considered
Of course it must be mentioned that also the validity herein such diagrams are presented in the Figures 8
of numerical results are limited with respect to the accur- and 9. For comparison also the respective curves
acy of the material law and the parameters assumed. obtained with the API method are given.

395

Copyright 2005 Taylor & Francis Group plc, London, UK


With the API method nearly constant stiffness val- REFERENCES
ues are obtained, which means that the monopile foun-
dation stiffness is almost independent of the loading Abaqus 2004. Users manual, Version 6.4.
level. In contrast, with the numerical calculations a Achmus, M. & Abdel-Rahman, K. 2003. Numerische
decreasing stiffness is obtained with increasing load Untersuchung zum Tragverhalten horizontal belasteter
level. The curves in Figures 8 and 9 show, that the Monopile-Grndungen fr Offshore-Windenergieanlagen.
load-deformation relationship obtained by the API 19th Christian Veder Kolloquium, Graz/Austria.
API 2000. American Petroleum Institute. Recommended
method is nearly linear. Nonlinearities, which are to be Practice for Planning, Designing and Construction Fixed
expected for high load levels, are thus not covered by Offshore Platforms- Working Stress Design, API Recom-
the method. The reason for this is that with the small mended Practice 2A-WSD (RP2A-WSD), 21st edition,
displacements in relation to the diameter of the mono- Dalllas.
pile only the nearly linear beginning portion of the DNV 2004. Det Norske Veritas. Design of Offshore Wind
API p-y curve becomes relevant (see also Achmus & Turbine Structures. Offshore Standard, Norway.
Abdel-Rahman 2003). EAU 1996. Recommendations of the Committee for Waterfront
Further investigations of the authors will be the Structures, Harbours and Waterways, 7th edition, Ernst &
derivation of diagrams for preliminary design of Sohn, Berlin.
GL 1999. Germanischer LIoyd. Rules and Regulations, IV
monopiles for other soil profiles typically for the North Non-Marine Technology, 2 Offshore Wind Energy
Sea and the Baltic Sea and also for varying pile diam- Converters, 9 Foundations, Hamburg, Germany.
eters and embedded lengths. Moreover, the behaviour Lesny, K. & Richwien, W. & Wiemann, J. 2002. Grndung-
of monopile foundations under cyclic loading must stechnische Randbedingungen fr den Bau von Offshore-
be and will be a subject of future research. Windenergieanlagen in der Deutschen Bucht.
Bauingenieur 77, p. 431438.
Lpile 2000. Users Manual, Version Lpile plus 4.0.
ACKNOWLEDGEMENTS Reese, L.C. & Van Impe, W.F. 2001. Single Piles and Pile
Groups under Lateral Loading. A.A. Balkema, Rotterdam/
Brookfield.
The results presented in this paper were obtained in the
framework of the FORWIND research group funded
by the Government of the federal state of Lower Saxony,
Germany. The support is gratefully acknowledged.

396

Copyright 2005 Taylor & Francis Group plc, London, UK


Tripods with suction caissons as foundations for offshore
wind turbines on sand

M. Senders
School of Civil and Resource Engineering, The University of Western Australia, Perth, Australia

ABSTRACT: This paper affirms that tripods with suction caissons can be used as foundations for offshore
wind turbines on sand. By modeling the complete structure and the occurring loads in a computer program a
better understanding of the distribution of load and footing deflections is created. Results of calculations for a
typical offshore wind turbine show the tripod primarily utilises a push-pull system between the legs to overcome
the occurring moment. The (non-dynamic) elastic behaviour of the foundation is an order stiffer than the elas-
tic behaviour of the superstructure and therefore does not contribute much to the deflection of the nacelle.
Unfortunately the exact behaviour of these suction caissons under a dynamically loaded tripod is not known and
should therefore be thoroughly investigated. With a better understanding, the tripod with suction caissons can
become a reliable and relatively cheap foundation for offshore wind turbines.

1 INTRODUCTION are design tools for wind turbine performance and


loading. They do not however, focus upon the founda-
Wind turbines are currently built offshore in Western tions although they are an essential part of the overall
European countries like Denmark, England and The design.
Netherlands. The foundation of these 100 m high It should be noted that this paper describes tripods
structures typically consist of a driven mono pile or with suction caissons only in uniform sand. In reality
gravity base structure (for water depths 25 m. Det the soil conditions in an entire wind farm (e.g.
Norske Veritas 2004). 50 km2) will vary considerable. This variability will
Other types of stiffer and cheaper foundations directly influence the behaviour of individual founda-
should be considered however because: a) the current tions in the farm if one type and size of foundation is
foundations may be unsuitable in water deeper than used. It is however believed that this soil variability
20 m, where the real potential for offshore wind energy can be taken into account during the design if the pre-
lies (Beurskens & Jensen 2003) and b) it is calculated cise behaviour of the foundation and the exact soil
that the current industry preferred mono pile founda- conditions are known.
tion contributes approximately 30% of the total
investment costs (DTI 2001).
Foundations consisting of tripods have the advan-
2 LOADS
tage that they can be designed to withstand the loads
occurring on offshore wind turbines even in relatively
2.1 General
deep water (2050 m). They can also make the struc-
ture stiffer, decreasing maximum deflection of the Loads on offshore wind turbines typically consist of
nacelle. Where suction cans are placed underneath dead load, wave and current loading and wind loading
the tripods, the installation costs will be significantly on the structure and the rotors (see Figure 1).
lower than for other foundation types. Compared with As a reference case in this paper a 3 MW turbine
quadropods these tripods use less steel and are less will be used comprising a hypothetical combination
labour intensive to fabricate. of the Vestas V90 (Vestas 2003) and the Dowec 6 MW
For an accurate foundation design it is essential to (Kooijman et al. 2003). It has a 100 m high support
understand which loads are acting on the complete structure and is placed in 25 m water depth. The foun-
structure and how it reacts to these loads. Commercial dation consists of a 15 m high tripod with three 11 m
computer programs such as BLADED (Garrad & OD suction caissons of 5.5 m height placed at 26 m
Hassan) and FLEX5 (Stig ye) are available, which centres.

397

Copyright 2005 Taylor & Francis Group plc, London, UK


2.2 Dead load Cd.air  drag coefficient in air (here: 0.7) [
]
The typical underwater weight of the reference foun- Fwind  wind load on structure [N]
.
dation is in the order of 5 MN. The dead load of the s  velocity of the structure [m]
superstructure is in the order of 3 MN (Vestas 2003). uwind  velocity of the wind [m/s]
The turbine which is placed on top (nacelle  rotors) air  density of air (here 1.25) [kg/m3]
gives an extra load in the order of 1 MN. This load The induction factor a can be found by combin-
will act eccentrically (see Figure 1) creating a moment ing blade element theory with momentum theory
of about 1 MNm. (Burton et al. 2001) and has a theoretical maximum
value of 1/3. In pitch controlled wind turbines this
2.3 Wave and current value will be lowered considerably beyond the optimal
Wave and current forces acting on the underwater wind speed, to allow generation of a constant power.
structural members can be calculated using the exten- Typical wind loading on the support structure
ded Morisons equation. A typical wave and current above 25 m water depth with a wind velocity of
load on the complete support structure in 25 m depth 15 m/s at hub height is in the order of 0.1 MN. The
water with 7 m waves having a period of 12 s and a force on rotors, 45 m in length is in the order 0.3 MN.
current of 2 m/s is in the order of 1 MN and can create In total this creates a moment at foundation level of
a moment at foundation level of about 10 MNm. about 35 MNm. If an extreme gust of twice the aver-
age wind speed is considered this force on the rotors
2.4 Wind can increase to approximately 1.2 MN and conse-
quently create a moment at foundation level of about
Wind load on the superstructure and the rotors can be
140 MNm. This large moment should be counteracted
calculated by the general formula (Burton et al. 2001):
by the foundation.

2.5 Extra loading


In addition to these loading conditions it may be neces-
(1) sary to consider other loads such as earthquake load-
ing, ship impacts and collision, snow and ice loads,
transient loading (shutdown or fault events) and the
Where: influence of scour (Det Norske Veritas 2004). These
a  axial flow induction factor [
] conditions will however not be covered in this paper.
Ad  rotating area of rotors or support [m2]
pile area
3 INSTALLATION IN SAND

Costs involved with the installation of a tripod with


suction caissons (pump and simple supply vessel)
will be lower than costs involved with installation of
(mono) driven piles (underwater hammer and large
barge with crane). Furthermore, the installation of the
tripod foundation can immediately be followed by the
installation of the remaining parts of the wind turbine,
without expensive equipment becoming idle. This
means that the whole operation can be performed
with one spread of vessels, which requires only one
mobilisation and one weather window.
Suction caissons have been installed in (very)
dense sand with success (Erbrich & Tjelta 1999,
Ibsen et al. 2003). However, it is observed that the soil
properties inside the caisson may be subject to slight
change. During the installation process heave inside
the caisson can also occur. Experiments have been
carried out to get a better understanding of these two
phenomena (Allersma et al. 1997, Tran et al. 2005),
but more research is needed to quantify the influence
Figure 1. Loading on offshore wind turbine. on the overall behaviour.

398

Copyright 2005 Taylor & Francis Group plc, London, UK


4 COMPUTER PROGRAM SOS_3D acting upon the complete structure. It uses rotor
depending parameters such as pitch angle, rotor speed
A computer program, named SOS_3D, has been used and optimum power to calculate the force acting on
to investigate the distribution of the load and the the rotor for a given average wind speed.
dynamic behaviour of this typical offshore wind tur- On a pitch controlled wind turbine the turbines elec-
bine with tripod foundation. tronic controller checks the power output of the tur-
SOS_3D has been developed to evaluate the bine several times per second. When the power output
response of wave-structure-soil interaction problems becomes too high, it sends an order to the blade pitch
(for details see Bienen & Cassidy 2005). Its inte- mechanism which immediately pitches (turns) the
grated approach to shallow foundation plasticity fol- rotor blades slightly out of the wind. For the max-
lows analysis of plane-frame jack-ups by Martin imum force calculations it is however assumed that the
(1994), Thompson (1996) and Cassidy (1999). blades will have a fixed pitch angle based on the aver-
Foundation models for sand include amongst age wind. The force is thereafter calculated by using the
others the sophisticated Model C. This model has applied wind speed instead of the reference wind speed.
been developed by Cassidy (1999) and Houlsby & Either a random wind speed in time is given based
Cassidy (2002) for flat circular footings on dense on a Von Karman spectrum or a wind cycle based on
sand subject to a combination of vertical, moment IEC 61400-1 to incorporate a maximum 50 year gust
and horizontal loading (V, M, H). By combining a yield (Burton et al. 2001). Examples of both can be seen in
surface (Fig. 2) with internal elastic behaviour, a plas- Figure 3.
tic potential and a hardening law, the model describes
the load-displacement relationship for footings on
dense sand. Because the plasticity framework allows
the response of the foundation to be expressed purely 5 INPUT FOR RUNS
in terms of force resultants on the footing, it can
be coupled directly to the numerical analysis of a 5.1 Support structure dimensions
structure. The support structure in this calculation consists of a
SOS_3D has been extended with a 2D wind com- pile increasing from 4.28 m OD at 5 m above seabed
ponent, making it possible to calculate wind forces to 2.3 m OD at 100 m. The 15 m high tripod, with
bracings under 45, 18.435 and 0, is attached to this
support pile (see Fig. 4). The tripod is modeled in 2D
by replacing two caissons by one which can take twice
the load and by making the connections twice the
original strength. Details can be found in Table 1.

5.2 Turbine specifications


The dimension and specifications of the reference
turbine are presented in Table 2. Blade specifications
(shape, drag and lift coefficients) are scaled from the
Dowec 6 MW (Kooijman et al. 2003).

Figure 2. Model C yield surface shape.


z-axis
y-axis

35 Support
x-axis pile
Wind speed [m/s]

30
25
20
15
10 Vref = 15 m/s von Karman
5
TI = 0.1 IEC 61400-1 x or y -axis
0
0 5 10 15 20 25 30 35 40 45 50 TOP VIEW SIDE VIEW
Time [s]
Figure 4. Tripod with suction caissons (dotted  3D;
Figure 3. Schematisation of wind speed in time. solid  2D).

399

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 1. Input parameters for SOS-3D.

Begin End
[m;m] [m;m] D A I E W
[x;z] [x;z] [m] [m2] [m4] [N/m2] [kg]

15; 5.5 0; 10.5 0.75 0.057 0.004 2.1E11 6361


15; 5.5 0; 20.5 0.75 0.057 0.004 2.1E11 9482

7.5; 5.5 0; 10.5 1.5 0.114 0.007 2.1E11 12722

7.5; 5.5 0; 20.5 1.5 0.114 0.007 2.1E11 18964
15; 5.5
7.5;5.5 1.5 0.114 0.007 2.1E11 55935
0; 10.5 0; 20.5 4.28 0.671 1.503 2.1E11 52694
0; 20.5 0; 55.5 3.785 0.478 0.857 2.1E11 131354
0; 55.5 0; 80.5 3.125 0.263 0.322 2.1E11 51646
0; 80.5 0; 105.4 2.576 0.127 0.108 2.1E11 24910
0; 105.4 0; 105.5 2.301 0.072 0.047 2.1E11 56

Table 2. Turbine specifications. Table 3. Input parameters for foundation Model-C.

Number of [m] 3 vin [m/s] 4 Suction can  soil:


Blades Radius [m] 5.5 WT [m] 0.055
Rblade [m] 45 vout [m/s] 25 Height [m] 5.5 Top Cap [m] 0.11
Eccentricity [m] 0.5 nominal []
2.4 Surface [
] rough G [MPa] 7.5
Maximum [MW] 3 nominal [rad/s] 2.56  [] 35  [kN/m3] 10
Power
Weight [kg] 104000 Kp [
] 0.9 Failure envelope  strain hardening:
a1 [
]
0.2 1 [
] 0.99
a0 [
] 0.0 2 [
] 0.99
m0 [
] 0.15 h0 [
] 0.337
5.3 Footing q0 [
] 0.1 pm [
] 0.0316
fv0m [
] 0.99 fkinit [
] 3.3
Two footing models were used for the calculations:
Elasticity:
a) fixed footing with no displacement and rotation k1 [
] 3.270 k2 [
] 1.005
b) flexible caissons schematised in a Model C footing. k3 [
] 3.468 k4 [
]
0.660
The Model C footing in sand is combined with an k5 [
] 12.636
elastic response for caissons with flexible skirts as
calculated by Doherty & Deeks (2003). To be able to
have tensile capacity in the model C footing an extra second natural frequency (0.315 Hz and 1.742 Hz
load was placed on each footing equal to the expected respectively).
friction plus the weight of the suction caissons (7 MN
calculated according DNV, 1992). Parameters for the
6 RESULTS AND DISCUSSION
Model C used can be found in Table 3 (further infor-
mation on the meaning of the parameters is presented
6.1 Wind turbine
by Cassidy 1999).
Power production, blade pitch, axial force in the rotor
5.4 Loading dimensions shaft and induction factor a are calculated for wind
speeds between 0.5 and 35 m/s with interval steps of
External loads are arbitrarily chosen and do not rep- 0.5 m/s. Results can be found in Figure 5.
resent a maximum load case. The structure is placed It can be seen that the maximum force on the rotors
in 25 m water depth with regular Airy waves of 7 m does not occur with the highest wind speed, but at
height and a period of 12 s. A regular current of 2 m/s around 12 m/s where the optimal power output of
is assumed to be present at the wind turbine location. 3 MW is reached. After this point the induction factor
The wind has an average wind speed of 15 m/s and is a decreases significantly because of a change in
subjected to a 50 year gust loading as described by pitch angle of the blades.
IEC 61400-1. However, in this example it does not
occur once every 50 years but is not realistically
6.2 Nacelle displacement
repetitive. An extra load is added to the dead load as
described in 5.3 to schematise the tensile capacity. The displacement of the nacelle is an essential com-
Structural damping is set as 0.02 at the first and ponent in the design of the complete structure.

400

Copyright 2005 Taylor & Francis Group plc, London, UK


3.5 20 Time (s)
3

Pitch angle, Q []
15 0 50 100 150 200
Power [MW]

2.5
10
30 V total
2
Caisson1

V load [MN]
1.5 5
20 Caisson 2 and 3
1
0.5 0
0 -5 10
0 10 20 30 40 0 10 20 30 40
Wind Speed [m/s] Wind Speed [m/s]
0
500 0.4
Time (s)
Induction factor, a [-]

400 0.3
0 50 100 150 200
Force [kN]

300 0.2 3
H total
200 0.1 Caisson1

H load [MN]
2 Caisson 2 and 3
100 0
0
1
-0.1
0 10 20 30 40 0 10 20 30 40
0
Wind Speed [m/s] Wind Speed [m/s]
-1
Figure 5. Wind turbine characteristics. Time (s)
0 50 100 150 200
1
0 25 50 75 100 125 150 175 200 Caisson1
M load [MNm]
4 Caisson 2 and 3
y displacement [m]

3 0.5
2
1 0
0 Caisson
Fixed footing
-1
-0.5

Figure 6. Nacelle displacement. Figure 7. Footing forces.

Because the period of the regular wave (12 s) and regu-


Time (s)
lar wind gust (14 s) are not exactly similar, loading
0 50 100 150 200
will differ in time. Figure 6 presents the displacement 5.56
of the nacelle over a period of 200 s.
y displacement [m]

The results for the fixed footing and the foundation 5.58
with caissons display only a very slight variation (the
maximum displacement between the two systems 5.60
Caisson 1
varies by only 4%). The main reason for this is that Caisson 2 and 3
the super structure is an order of magnitude more 5.62

flexible than the foundation. Time (s)


The nacelle displaces nearly 3 m, which is compar- 0 50 100 150 200
0.006
able with an average tilt of 1.7. This may appear to
x displacement [m]

be excessive given that 0.5 is commonly quoted as 0.004


the maximum allowed by the operating company.
0.002

0.000 Caisson 1
6.3 Footing Caisson 2 and 3
-0.002
The forces acting on the three suction caissons are Time (s)
presented in Figure 7, the displacement and rotation 0 50 100 150 200
in Figure 8. Caisson 1 represents the left caisson (Fig. 4) 0.0004

and caisson 2 and 3 are the two on the right.


[degrees]

The figure showing vertical load demonstrates that 0.0002

the structure uses a push-pull system between the legs


0.0000
to overcome the occurring moment.
Caisson 1
By comparing the calculated displacements and Caisson 2 and 3
-0.0002
rotations with hand calculations making use of
Dohertys elastic coefficients (Doherty & Deeks Figure 8. Footing displacements and rotations.
2003) it can be concluded that the whole system

401

Copyright 2005 Taylor & Francis Group plc, London, UK


exhibits elastic behaviour. This means that the applied Table 4. Natural frequencies for different footings.
loads fall within the failure envelope of the foundation.
Nevertheless the vertical load for caisson 1 reaches Natural frequency
a low value of 1.1 MN at the peak gust. It therefore
Footing G 1st 2nd 3rd
uses 5.9 MN of the total frictional capacity of 7 MN
[
] [MPa] [Hz] [Hz] [Hz]
and, at the same time, overcomes a horizontal load of
1.1 MN and a moment loading of 0.7 MNm. Fixed infinite 0.323 1.866 5.525
The y-displacement starts at 5.585 m, due to the fact Caisson 7.5 0.315 1.742 4.902
that the caisson is firstly installed by means of a virtual Caisson 3.0 0.301 1.595 4.444
preload of 400 MN. With the chosen elasticity assump-
tion it takes around 20 mm of displacement of caisson 1 stiffness as in capacity) can be important, but degra-
to mobilize enough resistance to overcome the moment. dation is not implemented in the footing model and is
The total horizontal force consists of the combina- therefore not seen back in the results.
tion of the periodic gust and wave cycles and is dis-
tributed evenly over the caissons. The maximum
horizontal force occurs at around 106 s. The load taken 7 CONCLUSIONS AND RECOMMENDATIONS
by caissons 2 and 3 varies between 0.0 and 0.7 MN
whereas for caisson 1, it varies between
0.4 and A tripod with suction caissons is a good alternative
1.1 MN. The accompanying horizontal displacement for foundations of offshore wind turbines on sand.
is in between 0 and 3 mm for caisson 2 and 3 and in They can be designed to withstand the large occurring
between
2 and 5 mm for caisson 1. forces, and are comparatively easy to install. The rel-
Finally the results show that very little moment ative high moment is counteracted by a push-pull sys-
is taken by the suction caissons (maximum of tem between the caissons. Maximum pull force is
0.7 MNm). The rotation () of the caissons is very critical for failure of this foundation. Displacement of
small and both caissons have a similar behaviour in the nacelle shows greater dependence on the super-
their rotation. This means that the tripod construction structure (96%) than on the foundation (4%). Never-
is stiff and rotates relatively uniformly. theless natural frequencies are influenced by the found-
ation and failure of the total system can occur when
tensile capacity is not as large as calculated.
6.4 General interpretation
In the near future SOS_3D will be adjusted so that it
Comparison of Figure 6 and 7 demonstrates that in can also correctly calculate the hydrodynamic forces
this calculation the nacelle responds quite directly to on diagonally tripod members. Furthermore a foun-
the gusts but less so to the passing waves. It should be dation model based on API py-curves will be imple-
noted however that for waves with a period around the mented, allowing comparisons with monopiles. First
natural frequency (3 s) the dynamics would begin to calculations shows deflections of typical monopiles
play an important role and totally different results are an order higher than tripod constructions.
may be obtained. Although several researchers (both experimental as
Although the foundation seems not to contribute analytical) have been investigating the behaviour of
much to the deflection of the nacelle, the superstruc- suction caissons (Byrne 2000, Doherty & Deeks 2003),
ture and foundation can not be treated separately in not much research has been performed regarding the
the overall design. The natural frequencies of the hypothesis that degradation of frictional capacity and
complete structure vary for instance in case the foot- stiffness response during cyclic loading in sand can
ing is fixed or consist of a caisson. In the latter case occur. However this frictional degradation might lead
the stiffness of the soil (here described with an uni- to an unconservative design. Centrifuge tests on scale
form G over depth) has an influence as well. Table 4 models are planned to be performed in 2005 and can be
summarises the first 3 natural frequencies calculated used as a first check of this hypothesis (Senders 2004).
by SOS_3D according the Rayleigh method. Although For large wind parks it will be an important (cost)
it seems not to make a significant difference, these factor that similar foundations are used over the
variations in natural frequencies can play an impor- whole area. Therefore the influence of the variability
tant role in the design of the complete structure and of the soil upon different types of offshore wind tur-
have to be calculated accurately. bine foundations should be investigated.
If the cyclic behaviour of the suction caisson
would be different than anticipated (e.g. less tensile
capacity or less stiffness) the whole wind turbine ACKNOWLEDGEMENTS
might fall over or tilt even more. Design codes like
API (2000) and DNV (1992) state in a remark that The author would like to thank the International
degradation of sand due to cyclic loading (both in Postgraduate Research Scholarship (IPRS) and Fugro

402

Copyright 2005 Taylor & Francis Group plc, London, UK


for their financial support. Furthermore he would like Doherty, J.P. & Deeks, A.J. 2003. Elastic response of circu-
to thank Messrs. M. Liingaard, J. v/d Tempel and J.v. lar footings embedded in a non-homogeneous half-space.
Iperen for their useful comments. Geotechnique, 53(8), 703714.
DTI. 2001. Monitoring & Evaluation of Blyth Offshore
Wind Farm. DTI/Pub URN 01/1524, DTI.
Erbrich, C.T. & Tjelta, T.I. 1999. Installation of bucket foun-
dations and suction caissons in sand geotechnical per-
REFERENCES formance. Proceedings of the Annual Offshore Technology
Conference, May 3-May 6 1999, OTC10990, 725735.
Allersma, H. G. B., Plenevaux, F. J. A. and Wintgens, J.-F. P. C. Houlsby, G.T. & Cassidy, M.J. 2002. A plasticity model for
M. E. 1997. Simulation of suction pile installation in sand in the behaviour of footings on sand under combined load-
a geocentrifuge Proceedings of the 1997 7th International ing Geotechnique, 52(2), 117129.
Offshore and Polar Engineering Conference. Part 1 (of Ibsen, L.B., Schakenda, B. & Nielsen, S.A. 2003. Devel-
4), May 2530 1997, Honolulu, HI, USA, 761766. opment of the Bucket Foundation for Offshore Wind
API. 2000. Recommended Practice for Planning, Designing Turbines: a Novel Principle. Proceedings of the US Wind
and Constructing Fixed Offshore Platforms Working Energy Conference 2628 august 2003. p.12.
Stress Design. Kooijman, H.J.T., Lindenburg, C., Winkelaar, D. & Hooft, v/d
Beurskens, J. & Jensen, H.P. 2003. Wind Energy R&D pri- E.L. 2003. Aero-elastic modelling of the DOWEC 6 MW
orities. EWEA. pre-design in PHATAS. DOWEC-F1W2-HJK-01-046/9.
Bienen, B. & Cassidy, M.J. 2005. Simulation of the soil- Martin, C.M. 1994. Physical and numerical modelling of off-
structure-fluid interaction of offshore jack-up structures. shore foundations under combined loads, Oxford, London.
submitted to 11th International Conference of IACMAG, Senders, M. 2004. Foundations of offshore wind turbines by
2005, Turin, Italy. suction cans; assessment of cyclic loading, Research
Burton, T., Sharpe, D., Jenkins, N. & Bossanyi, E. 2001. Wind Proposal
Energy Handbook, John Wiley & Sons Ltd, Chichester. Thompson, R.S.G. 1996. Development of non-linear numer-
Byrne, B.W. 2000. Investigations of Suction Caissons in ical models appropriate for the analysis of jack-up units,
Dense Sand, Oxford, London. Oxford, London.
Cassidy, M.J. 1999. Non-Linear Analysis of Jack-Up Tran, M.N., Randolph, M.F., & Airy, D.W. 2005 in prep.
Structures Subjected to Random Waves, Oxford, London. Study of seepage flow and sand plug loosening in instal-
Det Norske Veritas. 1992. Foundations, Classification Notes lation of suction caissons in sand. Fifteenth International
No 30.4, Hovik. Offshore and Polar Engineering (ISOPE) Conference
Det Norske Veritas, 2004. Design of Offshore Wind Turbine Seoul, Korea, June 1924, 2005.
Structures. Offshore Standard DNV-OS-J101. Vestas. 2003. V90 3.0 MW An efficient way to more power.

403

Copyright 2005 Taylor & Francis Group plc, London, UK


The tensile capacity of suction caissons in sand under rapid loading

Guy T. Houlsby, Richard B. Kelly & Byron W. Byrne


Department of Engineering Science, Oxford University

ABSTRACT: We develop here a simplified theory for predicting the capacity of a suction caisson in sand,
when it is subjected to rapid tensile loading. The capacity is found to be determined principally by the rate of
pullout (relative to the permeability of the sand), and by the ambient pore pressure (which determines whether
or not the water cavitates beneath the caisson). The calculation procedure depends on first predicting the suc-
tion beneath the caisson lid, and then further calculating the tensile load. The method is based on similar prin-
ciples to a previously published method for suction-assisted caisson installation (Houlsby & Byrne 2005). In the
analysis a number of different cases are identified, and successful comparisons with experimental data are
achieved for cases in which the pore water either does or does not cavitate.

1 INTRODUCTION applicable for calculating the capacity. For the purposes


of calculation an idealised case of a foundation on a
Suction caissons are an option for the foundations for homogeneous deposit of sand is considered here.
offshore structures. Under large environmental loads The resistance on the caisson is calculated as the sum
the upwind foundations of a multiple-caisson founda- of friction on the outside and the inside of the skirt. The
tion might be subjected to tensile loads. Recent research effective stresses on the annular rim are likely to be
indicates that serviceability requirements will often dic- sufficiently small that they can be neglected, and it is
tate that, under working and frequently encountered assumed that the soil breaks contact with the lid of the
storm loads, tensile loads on caissons should be avoided, caisson. The frictional terms are calculated in the same
as they are accompanied by large displacements. How- way as for the installation calculation (Houlsby & Byrne
ever, it may be appropriate to design structures so that 2005), by calculating the vertical effective stress adja-
under certain extreme conditions the caissons are cent to the caisson, then assuming that horizontal effec-
allowed to undergo tension. It is therefore necessary to tive stress is a factor K times the vertical effective stress.
have a means of estimating the tensile capacity of a Assuming that the mobilised angle of friction between
caisson foundation, whilst recognising that large dis- the caisson wall and the soil is  then we obtain the
placements may be necessary to mobilise this capacity. result that the shear stress acting on the caisson is vK
The calculations are also relevant to the holding capac- tan . Note that in the subsequent analysis the values of
ity of caisson anchors subjected to pure vertical load, K and  never appear separately, but only in the combi-
and to calculation of forces necessary to extract a cais- nation K tan , so it is not possible to separate out the
son rapidly (for whatever reason). effects of these two variables. Allowance is made, how-
Under rapid tensile loading, a suction caisson in ever, for the possibility of different values of K tan  act-
sand will exhibit a limiting load which will typically ing on the outside and inside of the caisson. A difference
consist of a suction developed within the caisson, and between this analysis and conventional pile design is
friction on the outer wall. However, a number of dif- that the contribution of friction in reducing the vertical
ferent possible modes of failure exist. The purpose of stress further down the caisson is taken into account.
this paper note is to set out simple calculations for capa- If, as a preliminary, no account is taken of the
cities under various failure modes, and to compare these reduction of vertical stress close to the caisson due to
with experimental results. the frictional forces further up the caisson, then the
tensile vertical load on the caisson for penetration to
2 TENSILE CAPACITY CALCULATIONS depth h is given by:

2.1 Drained capacity


If the tensile load is applied very slowly, then pore pres- (1)
sures will be small, and a fully drained calculation is

405

Copyright 2005 Taylor & Francis Group plc, London, UK


A similar analysis follows for the stress on the out-
side of the caisson. We assume that (a) there is a zone
V' Mudline
between diameters Do and Dm  mDo in which the ver-
tical stress is reduced through the action of the upward
z h friction from the caisson, (b) within this zone the verti-
cal stress does not vary with radial coordinate and (c)
Di there is no shear stress on vertical planes at diameter
Dm. We then obtain the same results as for the inside
Do of the caisson, but with Zi replaced by Zo  Do
(m2
1)/(4(K tan )o).
Figure 1. Caisson geometry. Alternative assumptions could be made for the
variation of Dm with depth, but at present there is little
evidence to justify any more sophisticated approach.
If Dm is taken as a variable, then the differential equa-
tion for vertical stress will usually need to be integrated
numerically.
Accounting for the effects of stress enhancement,
Eq. (1) becomes modified to:

Figure 2. Vertical equilibrium of a slice of soil within the


caisson.

(3)
where the dimensions are as in Figure 1, and  is the
effective weight of the soil. V is the buoyant weight
of the caisson and structure.
A check should always be made that the friction
calculated inside the caisson does not exceed the In the special case where m is taken as a constant
weight of the trapped soil plug h D2i /4. and uniform stress is assumed within the caisson this
Ignoring the reduction of the stress in this case becomes:
proves unconservative (i.e. it overestimates the force
that can be developed), so we develop here a theory
which takes this effect into account. Consider first the
soil within the caisson. Assuming that the vertical effec-
tive stress is constant across the section of the caisson, (4)
the vertical equilibrium equation for a disc of soil
within the caisson (Fig. 2) leads to:

(2) The calculation accounting for stress reduction


obviates the need to check that the internal friction
does not exceed the soil plug weight, as the capacity
asymptotically approaches that value at large h.
Writing Di/(4(K tan )i)  Zi, Eq. (2) becomes
dv /dz v /Zi  , which has the solution v  
Zi
(1
exp(
z/Zi)) for v  0 at z  0. The total fric- 2.2 Tensile capacity in the presence of suction
tional terms depend on the integral of the vertical If the caisson is extracted more rapidly, then transient
effective stress with depth, and we can also obtain. For excess pore pressures will occur, and the suction
within the caisson will need to be taken into account.
For small We return later to the calculation of the relationship
between the rate of movement and the suction, but
h/Zi the integral simplifies to h2/2 as in Eq. (1). For first address the calculation of load in terms of the
brevity in the following we shall write the function suction. If the pressure in the caisson is s with respect
to the ambient seabed water pressure, i.e. the absolute
y(x)  (exp(
x)
1  x), so that in the above pressure in the caisson is pa  w hw
s (where pa is
atmospheric pressure, w is the unit weight of water
and hw the water depth), then we at first assume that
. the excess pore pressure at the tip of the caisson is as,

406

Copyright 2005 Taylor & Francis Group plc, London, UK


i.e. the absolute pressure is pa  w(hw  h)
as. D2/4  A):
There is therefore an average downward hydraulic gra-
dient of as/w h on the outside of the caisson and
upward hydraulic gradient of (1
a)s/w h on the
inside.
(8)
We assume that the distribution of pore pressure on
the inside and outside of the caisson is linear with depth.
A detailed flow net analysis shows that this approxima-
tion is reasonable. The solutions for the vertical stresses
inside and outside the caisson are exactly as before, where Z  D(m2
1)/(4K tan ). Neglecting the
except that  is replaced by   as/h outside the cais- effects of stress reduction would give:
son and by 
(1
a)s/h inside the caisson. The
capacity, accounting for the pressure differential across
the top of the caisson and pore pressure on the rim (only (9)
relevant for a thick caisson), is again calculated as the
sum of the external and internal frictional terms:
which means that the capacity is simply calculated by
applying a linearly varying factor to the suction force
beneath the lid.

2.3 Undrained failure


(5)
A further condition should be considered: that of
undrained failure of the sand. In any dilative sand,
however, the pore pressures developed under undrained
In the special case of m constant and a uniform conditions are potentially so large that invariable
stress assumed within the caisson, this gives: (except in very deep water) the cavitation mechanism
would intervene first. Since the undrained strength of
sand is in any case very difficult to determine, we do
not pursue this case here.

3 RELATIONSHIP BETWEEN SUCTION


AND DISPLACEMENT RATE
(6) At low displacement rates, the rate of influx of water
q to the caisson can be calculated by Darcys law, and
equated to the rate of displacement times the area of
We can often make a further simplifying assump- the caisson. Flow calculations were presented by
tion, that the suction is sufficiently large that the Houlsby & Byrne (2005), and yield:
soil within the caisson liquefies and therefore

(1
a)s/h  0. For a large suction this means that
a  1 and almost all of the suction appears at the cais- (10)
son tip. The above rearranges to give as/h  s/h,
and Eq (6) can be simplified to:
where F is a dimensionless factor as determined by
the procedures in Houlsby & Byrne (2005), which
may be fitted approximately by the equation F 
3.6/(1  5h/D) for 0.1  h/D  0.8.
If the displacement rate is increased, the above con-
(7) dition is interrupted by one of two conditions (a) the
suction becomes large enough for liquefaction of the
sand within the caisson to occur or (b) cavitation occurs
In the case either that the thickness of the caisson within the caisson.
is small, or that a  1 this simplifies to the following When liquefaction occurs, the permeability of the
(writing the outer diameter as D, and the caisson area liquefied sand increases to a large value, with the

407

Copyright 2005 Taylor & Francis Group plc, London, UK


3.5
Dimensionless flow factor FL

3.0
2.5
2.0
1.5
1.0
Note that this will imply a sudden jump in s and V at
the onset of liquefaction.
0.5
0.0 (c) Cavitation without liquefaction
0.0 0.2 0.4 0.6 0.8 1.0
Aspect ratio h/D Onset of cavitation occurs at s  pa(1
f )  whw.
After that
dh/dt is unbounded, s is constant and:
Figure 3. Dimensionless flow factor for liquefaction case.

result that the a factor in the calculation of the load


changes (as noted above) to near unity. The displace-
ment rate may still be estimated from a flow calcula- as in case (a).
tion, but the appropriate boundary condition now
becomes one of the suction applied at the base rather (d) Cavitation with liquefaction
than top of the caisson. Modified values of F (termed Since s is constant once cavitation occurs, this condition
FL for this case) are given in Figure 3, and may be can only occur when liquefaction occurs before cavita-
fitted approximately by the equation F  1.75  1.9 tion. Onset of cavitation is at s  pa(1
f )  w hw,
exp(
5h/D) for 0.1  h/D  1.0. after which
dh/dt is unbounded, s is constant and:
When cavitation occurs, either before or after lique-
faction, the displacement rate becomes unlimited and
(assuming that cavitation occurs at an absolute pres-
sure fpa where f is a constant), the suction will be con-
stant and determined by pa  w hw
s  fpa, or Note that the above cases only occur in order (a), (b),
s  pa(1
f )  w hw. In practice it appears that the (d) or (a), (c). When several possibilities exist for calcu-
factor f is near zero. lating load capacity it is often true that the correct case
is simply found by calculating all cases and then taking
the lowest value. Note in this analysis that this simple
4 SUMMARY OF ANALYSIS CASES approach cannot be adopted as the onset of some states
can preclude other cases occurring, and the calculated
The following summary presents equations for the load is not necessarily the lowest of the cases.
above cases, for a thin-walled caisson. To simplify the
equations we neglecting here the stress reduction
effect, although this should be included in more accur- 5 COMPARISONS WITH DATA
ate calculations:
We present here a number of pullout tests conducted
two sands and at different pullout rates. The tests were
conducted in a pressure chamber: some tests at an ambi-
ent (mudline) water pressure equal to atmospheric, and
some at atmospheric plus 200 kPa. The model caisson
was 280 mm diameter, 180 mm skirt length. In the fol-
lowing the loads presented include the caisson weight.
The first test reported here (Test 9) was conducted
on Redhill Sand, at a pullout rate of 100 mm/s and
atmospheric pressure. Figure 4 shows the record of
(for
dh/dt  0, s  0 and these reduce to the equa- suction developed beneath the lid of the caisson
tions for the fully drained case). against time, and Figure 5 shows the corresponding
vertical load. It can be seen that (with a minor initial
fluctuation) the suction rapidly approaches 100 kPa,
(b) Liquefaction without cavitation
at which stage cavitation occurs. At around 3044.5 s
Onset of liquefaction occurs at h/(1
a), after that there is a sudden loss of both suction and vertical

408

Copyright 2005 Taylor & Francis Group plc, London, UK


20.0 0
4257 4257.5 4258 4258.5 4259 4259.5
0.0 -50
3042.8 3043.3 3043.8 3044.3 3044.8 3045.3
-20.0
-100
pressure (kPa)

Experiment

pressure (kPa)
-40.0 Theory
Experiment
Theory -150
-60.0

-200
-80.0

-100.0 -250

-120.0
-300
t (s) t (s)

Figure 4. Pressure v. time for test 9. Figure 7. Pressure v. time for test 10.

2.0
5.0
0.0
3042.8 3043.3 3043.8 3044.3 3044.8 3045.3 0.0
-2.0 Experiment 4257 4257.5 4258 4258.5 4259 4259.5
Theory
-5.0
-4.0 Experiment
V (kN)

Theory
-10.0
V (kN)

-6.0

-8.0 -15.0

-10.0 -20.0

-12.0 -25.0
t (s)
-30.0
t (s)
Figure 5. Vertical load v. time for test 9.
Figure 8. Vertical load v. time for test 10.

2.0
on each of Figures 4 to 6, and it is clear that the theory
1.5
(whilst not capturing some of the detail at the begin-
ning of the pullout) predicts the broad trends of the
test correctly.
V / sA

1.0
Figures 7 and 8 show corresponding results for
Experiment Test 10 (at the same pullout rate) but at an ambient
Theory
0.5 pressure of atmospheric plus 200 kPa. The suctions
developed at this rate of loading are insufficient to
0.0 cause cavitation, which would occur at
300 kPa rela-
3042.8 3043.3 3043.8 3044.3 3044.8 3045.3
tive to ambient. It can be seen that again the theory
t (s)
predicts the overall pattern of behaviour well. This
time it is case (b) that applies. The fluctuations in pre-
Figure 6. V/sA v. time for test 9.
dicted suction (and hence load) are due to minor vari-
ations in the calculated velocity of extraction.
load, but this is of little practical interest since by Figures 9 and 10 show the results from Test 11,
then the displacements are enormous and about three- which is directly comparable to Test 9, but this time at
quarters of the caisson had been pulled out of the soil. a pullout rate of only 5 mm/s. Although the suctions
Figure 6 shows the ratio V/sA, showing that this are sufficient to cause liquefaction, the pullout rate is
ratio remains approximately constant during most of such that the suction is sufficiently small so that cavi-
the pullout. tation does not occur, and the vertical loads are cor-
It can readily be shown that the suction in this case respondingly lower too. The predicted suction and
rapidly increased to sufficient value to cause lique- load are also shown on the Figures. The match to the
faction (which would occur at a suction of only about data could be improved by adjusting the permeability,
3 kPa), and that the relevant case for analysis here but the value used in the predictions were deliberately
is case (d). The predicted values from the theory des- kept the same for all three tests discussed. The per-
cribed above (including stress reduction) are also shown meability value used was k  0.5  10
3 m/s, which

409

Copyright 2005 Taylor & Francis Group plc, London, UK


5 10

0 0
2980 2990 3000 3010 3020 3030 8890 8892 8894 8896 8898 8900 8902
pressure (kPa)

-5 -10

V (kN)
-10
-20

-15 Experiment Experiment


Theory -30 Theory

-20
t (s) -40
t (s)

Figure 9. Pressure v. time for test 11. Figure 12. Vertical load v. time for test 23.

1.0
Table 1. Predicted and measured tensile loads.
0.5

0.0 Max tensile load (kN)


2980 2990 3000 3010 3020 3030
-0.5
Test Predicted Measured
V (kN)

-1.0

-1.5 Test 11 (5 mm/s, 0 kPa) 1.1 2.4


Experiment
-2.0 Theory Test 9 (100 mm/s, 0 kPa) 10.1 11.1
Test 10 (100 mm/s, 200 kPa) 25.6 24.2
-2.5
HP5 sand: 30.6 33.2
-3.0 Test 23 (25 mm/s, 200 kPa)
t (s)

Figure 10. Vertical load v. time for test 11.


The predicted and measured values of maximum
tensile load for the three tests on Redhill sand and one
on HP5 are shown in Table 1. The order of magnitude
50
of the tensile load is correctly predicted in all cases,
0
8890 8892 8894 8896 8898 8900 8902
even though the actual capacity of the caisson varies
-50 greatly in the different tests.
pressure (kPa)

-100

-150 Experiment
Theory
-200 6 CONCLUSIONS
-250

-300
In this paper we develop a simplified theory for pre-
dicting the maximum tensile capacity of a caisson
-350
t (s) foundation in sand. The calculated capacity depends
critically on the rate of pullout (in relation to the per-
Figure 11. Pressure v. time for test 23. meability) and the ambient water pressure (which
determines whether cavitation occurs). The theory is
used successfully to explain widely differing experi-
mental results for caissons pulled out under different
is somewhat higher than estimated previously for this conditions.
sand (Kelly et al. 2004). The other parameters used
are K tan   0.7 and m  1.5.
Finally, Figures 11 and 12 present equivalent data
for a test on HP5 sand, which is much finer that Red- REFERENCES
hill Sand, and has an estimated permeability of
Houlsby, G.T. and Byrne, B.W. 2005. Design procedures for
k  0.2  10
4 m/s. The extraction rate was 25 mm/s, installation of suction caissons in sand, Proc. ICE, Geo-
and in this case, although the extraction rate is lower, technical Engineering, in press.
the pore pressures are sufficient to cause cavitation Kelly, R.B., Byrne, B.W., Houlsby, G.T. and Martin, C.M.
even with the ambient pressure of atmospheric plus 2004. Tensile loading of model caisson foundations for
200 kPa. structures on sand, Proc. ISOPE, Toulon, Vol. 2, 638641

410

Copyright 2005 Taylor & Francis Group plc, London, UK


Moment loading of caissons installed in saturated sand

Felipe A. Villalobos, Byron W. Byrne & Guy T. Houlsby


Department of Engineering Science, Oxford University

ABSTRACT: A series of moment capacity tests have been carried out at model scale, to investigate the effects
of different installation procedures on the response of suction caisson foundations in sand. Two caissons of
different diameters and wall thicknesses, but similar skirt length to diameter ratio, have been tested in water-
saturated dense sand. The caissons were installed either by pushing or by using suction. It was found that the
moment resistance depends on the method of installation.

1 INTRODUCTION

Suction caisson foundations are increasingly being


used in offshore applications. They have been used for
fixed structure applications, as described by Bye et al.
(1995), and also for floating facilities (House 2002).
More recently they are being considered as founda- 100 m
tions for offshore wind turbines (Byrne & Houlsby
2003). The wind turbine structures may be founded on
single or multiple caissons. The multiple caisson prob-
lem is addressed by Kelly et al. (2004), so in this paper
we concentrate on the single caisson problem. Typical
dimensions and loads for this problem are shown in
Figure 1. Byrne & Houlsby (2003) describe this prob- 90 m
6 MN
lem in detail, but the main differences in loads on the
foundations for offshore wind turbines as compared to
typical oil and gas structures are that: (a) the vertical 4 MN
load is much smaller, (b) the horizontal and moment
loads are proportionately larger. New design methods
h = 30 m
must be developed to allow safe designs to be engin-
eered for this regime of loading. As a result Byrne et al.
(2002) describe a research project aimed at developing
such design guidelines. This paper outlines the results
from a part of that project.
Initial studies of the moment capacity of caisson
foundations in the laboratory were carried out in
drained sand. Preliminary results from these experi- Figure 1. Dimensions and magnitude of loads for a 3.5 MW
ments are described by Byrne et al. (2003). As the sand turbine structure founded on a monopod suction (adapted
used during the tests was dry, the caissons were from Byrne & Houlsby 2003).
installed into the prepared sand bed by applying a ver-
tical load. The advantage of using dry sand is that the process. The different installation techniques may
test bed can be prepared quickly, and a large number of impose different stress paths on elements of soil around
tests can be carried out at specified densities. To miti- the caisson, which in turn may affect the response of the
gate the effects of scale, the tests beds were chosen to caisson to the applied loads. Therefore it is necessary to
be relatively loose. Clearly using installation by apply- carry out experiments similar to those in the dry sand,
ing vertical loads is different from the procedure that but on caissons installed by suction, to observe if there
has to be used in the field i.e. the suction installation are any fundamental differences in behaviour.

411

Copyright 2005 Taylor & Francis Group plc, London, UK


2 EQUIPMENT AND MATERIALS Table 1. Redhill 110 properties (Kelly et al. 2004).

2.1 Sand samples D10, D30, D50, D60 (mm) 0.08, 0.10, 0.12, 0.13
Coefficients of uniformity, Cu 1.63, 0.96
The sand used during the experiments was a commer- and curvature Cc
cially produced sand called Redhill 110. The proper- Specific gravity, Gs 2.65
ties of this sand are given in Table 1. Minimum dry density, min (kN/m3) 12.76
The sand samples were saturated with water inside Maximum dry density, max (kN/m3) 16.80
a tank of diameter 1100 mm and depth 400 mm. Pre- Critical state friction angle, cs 36
paration of the test bed involved an initial phase of
fluidisation by an upward hydraulic gradient induced
in the sand bed. The sample was then densified by
vibration under a small confining stress. The density
was determined by measuring the weight and the vol-
ume of the sample. The preparation process was
halted once a target density was reached. The peak
triaxial angle of friction was estimated as 44.1 to
45.2 from the correlation of Bolton (1986), for the
range of relative densities tested (see Table 3).

2.2 Testing procedure


Tests were performed using a three degree-of-freedom
loading rig (3DOF) designed by Martin (1994). This
rig, shown in Figure 2, can apply any combination of
vertical, rotational and horizontal displacement (w,
Figure 2. 3DOF-loading rig.
2R, u) to a footing by means of computer- controlled
stepper motors (R is the radius of the footing). Byrne
(2000) has installed a software control program, so
Table 2. Geometry of the model caissons tested.
that any combination of vertical, moment or horizon-
tal load (V, M/2R, H) can also be applied to the foot- Diameter, 2R (mm) 293 200
ing. All displacements and loads are monitored and Length of skirt, L (mm) 146.5 100
recorded using appropriate data-acquisition routines Thickness of the skirt wall, t (mm) 3.4 1.0
as well as being used within feedback control rou- Aspect ratio, L/2R 0.5 0.5
tines. It is possible to apply loads and displacements Thickness ratio, 2R/t 86 200
to the footing which represent the offshore environ-
ment loads of gravity, wind, waves and currents. The
geometry of the model suction caissons used in the connector arm to the 3DOF
experiments is given in Table 2. The model caissons loading rig
were fabricated from aluminium alloy, with a rela-
tively smooth (but not polished) surface. air valve fluid valve
The loading apparatus was modified to allow the
footings to be suction installed. Previous experiments L -w hf
had only used caissons forced into the ground by ver- l=w+L
tical load. To enable the suction installation phase to
be carried out, the equipment was modified as shown
250 mm reservoir
in Figure 3. The suction caisson, attached to the 3DOF
loading rig, was pushed into the ground about 30 mm 1100 mm
with the air valve open. This allowed the pressure
inside the caisson to equilibrate to the outside pres-
sure. On reaching a penetration of 30 mm the air valve Figure 3. Suction device.
was closed, and the fluid valve opened. The fluid from
inside of the caisson was connected to a reservoir,
which was slowly lowered to increase the head differ- the footing was kept constant using feedback control.
ence, hf, between the inside and outside of the cais- The reservoir was connected to the suction caisson by
son. The head difference was increased to a maximum a pipe of 6 mm internal diameter, chosen to allow suf-
of 300 mm (3 kPa), whilst the vertical load applied to ficient water flow with minimal head loss.

412

Copyright 2005 Taylor & Francis Group plc, London, UK


This procedure allowed the caisson to be installed 2000 Theory (Houlsby and Byrne, 2005)
by suction whilst connected to the loading rig. Once 1800 Experiments, D = 293mm
the suction phase was complete, it was possible to 1600 Experiments, D = 200mm

Vertical Load, V (N)


carry out experiments similar to those carried out on 1400
the dry sand as described by Byrne et al. (2003). 1200
1000
2.3 Comments on installation methods 800
The two different installation methods have been 600
described by Houlsby & Byrne (2005). Installation by 400
vertical load involves pushing the caisson into the 200
ground. The resistance to penetration is given by 0
the friction on the inside and outside of the wall, 0 20 40 60 80 100 120 140
Vertical Displacement (skirt penetration), h (mm)
and the bearing resistance on the skirt tip. Due to
silo effects the stresses around the skirt, and at the
Figure 4. Pushing installation tests and theoretical calcula-
tip, are enhanced, leading to larger resistances than tions for both caissons.
may be given by a more conventional pile calculation.
Houlsby & Byrne (2005) developed expressions for
predicting the resistance to penetration for caissons,
Table 3. Installation tests (Suction and Pushing).
taking account of these silo effects. The equations
they developed are used below to provide a compari- .
2R h
son with the experimental results. In these calcula- Test (mm) V (N) Vo (N) Rd (%) (mm/s)
tions it will be assumed that (K tan )i,o takes a value
of 0.9, and the stress enhancement factors m and n are FV6_5_1S 200 5 398 75 0.04
taken as 1. FV7_5_1P 200 425 74 0.50
Installation by suction requires an initial penetra- FV6_2_1S 200 40 410 75 0.04
tion to create a seal at the skirt tip. Typically 10% to FV6_3_1P 200 428 75 0.50
20% of the caisson skirt penetrates into the ground FV6_8_1S 200 60 400 75 0.04
under its own weight. The seal allows the suction FV7_1_1P 200 469 74 0.50
FV8_1_1S 293 10 1700 81 0.04
process to begin and should prevent the occurrence of FV8_2_1P 293 1772 81 0.20
an unconfined flow failure (i.e. a piping failure). Once FV7_3_1S 293 60 1700 74 0.06
sealed, the caisson will penetrate into the ground under FV7_4_1P 293 1740 74 0.20
the application of suction. Typically a pump will FV7_1_3S 293 120 1500 74 0.07
remove fluid from inside the caisson, creating a pres- FV7_2_1P 293 1741 74 0.40
sure differential on the caisson lid, as well as inducing
hydraulic gradients in the soil. The hydraulic gradi-
ents lead to changes in the effective stresses around Table 3 is the Relative Density. Also shown on Figure 4
the caisson skirt that are beneficial to installation. is a theoretical prediction calculated using the methods
Houlsby & Byrne (2005) have developed expressions of Houlsby & Byrne (2005).
to calculate the required suction for installation of Figure 5 shows the load penetration curves for a
caissons. This expression is used below for compari- pushed test and a suction-installed test. In the latter
son with the experimental data. the vertical load was kept constant at 60 N during the
suction phase. For this test the curve labelled V  S
shows the net vertical load due to applied load plus
3 RESULTS OF THE INSTALLATION TESTS the pressure differential on the caisson lid. It is clear
that there is a significant difference between this net
Figure 4 shows the load-displacement results for cais- vertical load and the vertical load for the caisson
sons pushed into the ground at a constant rate. The installed by pushing. The difference between these
results are shown as vertical load V against vertical curves represents the beneficial effects of the
penetration h. The maximum values of V obtained hydraulic gradients set up within the soil due to the
during these tests were approximately 400 N for the suction.
footing of diameter 200 mm and 1700 N for the Figure 6 shows one of these tests compared to the
293 mm diameter footing. These maximum values of V theoretical predictions of Houlsby & Byrne (2005). In
(denoted by Vo in Table 3) represent preconsolidation all of the experimental tests the suction was applied
vertical loads which might be used for interpreting the after approximately 30 mm of penetration. The suc-
results within the context of a yield surface model tion force shown in Figure 6 is slightly underesti-
(Gottardi et al. 1999, Houlsby & Cassidy 2002). Rd in mated by the calculations.

413

Copyright 2005 Taylor & Francis Group plc, London, UK


2000 Table 4. Moment capacity tests.
1800
Vertical Load, V and V + S (N)

V (M/2R)y Hy . . .
1600
V .
Test M/2RH (N) (N) (N) u /2R w /2R
1400 by pushing FV7_4_1
1200 by suction FV7_3_1 V = 60N FV6_5_2S 1.03 5.5 6.7 4.8 0.391
0.397
1000 FV7_5_2P 1.02 6 14.3 12.8 0.490 0.445
800 FV6_2_2S 1.06 40 11.8 10.4 0.463
0.122
600 V+S FV6_3_2P 1.05 40 24.1 21.7 0.465
0.284
400 FV6_8_2S 1.05 60 18.3 16.4 0.501
0.051
200 V
FV7_1_2P 1.03 60 29.2 26.9 0.505
0.253
FV8_1_2S 1.04 10 14.8 15.1 0.310
0.409
0
0 20 40 60 80 100 120 140 FV8_2_2P 1.03 10 33.4 32.9 0.569
0.551
Vertical Displacement (skirt penetration), h (mm)
FV7_3_2S 1.04 60 30.9 27.7 0.404
0.299
FV7_4_2P 1.03 60 42.0 40.4 0.446
0.483
FV7_1_4S 1.04 120 39.7 40.4 0.362
0.125
Figure 5. Comparison between pushed installation and
FV7_2_2P 1.03 120 56.3 53.3 0.477
0.289
suction installation for 293 mm diameter caisson.

300 Test FV7_3_1: V


Calculated V 30
200
Vertical Load, V and S (N)

Test FV7_3_1: S
Moment Load, M/2R (N)
Calculated S Pushed installation
100
20
0
0 20 40 60 80 100 120 140 Suction installation
-100 10

-200
0
-300 0 0.25 0.5 0.75 1 1.25 1.5 1.75 2
Rotational Displacement, 2R(theta) (mm)
-400 Vertical Displacement (skirt penetration), h (mm)
Figure 7. Moment capacity tests, load-rotation response
Figure 6. V and S comparison between experimental result showing yield points (M/2R)y.
and calculation for a suction installed test of the 293 mm
diameter caisson.
30
Horizontal Load, H (N)

4 MOMENT CAPACITY Pushed installation


20
Once the caissons were installed, moment capacity
tests were carried out. These tests are similar to those
Suction installation
reported by Byrne et al. (2003), and consist of rota- 10
tion and translation of the footing at a specified ratio
of M/2RH under a constant vertical load. The tests
were carried out slowly, so that the conditions were
0
fully drained. They are thus relevant to only one of a 0 0.25 0.5 0.75 1
series of possible conditions in the field, where, Horizontal Displacement, u (mm)
depending on caisson size, sand type and loading rate,
conditions may vary from fully drained to almost Figure 8. Two moment capacity tests, load-displacement
undrained. Summary data for the moment tests are response showing yield points Hy.
presented in Table 4, and further data about initial
conditions can be found in Table 3 (installation
method, Vo, initial Rd, etc.) effect on the load-displacement behaviour. The load-
Figures 7 and 8 compare the moment and horizon- displacement curves have been interpreted using
tal load capacities for different installation methods the method described by Byrne et al. (2003), with fit-
for the 200 mm diameter caisson. It is clear from ting linear expressions to the elastic and plastic com-
these figures that the installation method has a strong ponents of the curve. The intersection of the lines

414

Copyright 2005 Taylor & Francis Group plc, London, UK


1.25 60
Horizontal Displacement, u (mm)

by suction FV6_2_2 :V = 40N; D = 200mm

Moment Load, M/2R (N),


suction installed
by pushing FV6_3_2: V = 40N; D = 200mm
50 pushing installed
1

2Rd(theta)
40
0.75
30
0.5
20

0.25 10

0 0
0 0.25 0.5 0.75 1 1.25 1.5 1.75 2 -40 -30 -20 -10 0 10 20 30 40 50 60
Rotational Displacement, 2R (theta) (mm) Vertical Load, V (N) - dw

Figure 9. Plastic displacements increments during the Figure 11. Pushing and suction installation calculations
tests: horizontal displacement with respect to rotational for a 200 mm diameter caisson and V  5N, 40N and 60N.
displacement.

Table 5. Yield surface parameters for L/2R  0.5.


-1
by suction FV6_2_2 :V = 40N; D = 200mm
Vertical Displacement, w (mm)

Eccentricity of yield surface, a


0.75
-0.75 by pushing FV6_3_2: V = 40N; D = 200mm Horizontal dimension of yield surface, ho 0.337
Moment dimension of yield surface, mo 0.122
-0.5 Curvature factor at low V, 1 0.99
Curvature factor at high V, 2 0.99
-0.25

0 in Figure 11. It is possible to fit through these data


points a surface such as expressed by the formula:
0.25
0 0.25 0.5 0.75 1 1.25 1.5 1.75 2
Rotational Displacement, 2R (theta) (mm)

Figure 10. Plastic displacements increments during the tests:


vertical displacement with respect to rotational displacement.

represents a yield point. These points are shown on


(1)
the figure and given in Table 4 for all the tests.
The displacements paths from the tests are shown in
Figures 9 and 10. In general the rotational displacement in which ho, mo, to, a, 1 and 2 define the shape of
causes an initial elastic response that gradually changes the surface an 12  ( 1 2) 1 2/( 1) 1( 2) 2.
to an almost perfectly plastic response, which can be fit- The surface is fitted through the data points using
ted with a straight line. parameter values given in Table 5. These values were
The values of the slopes of these plastic displace- found for a series of rotational tests performed with
ment increments are presented in Table 4 as a ratio the same 293 mm diameter caisson in dry sand.
between horizontal
. and rotational displacement incre- Also shown on Figure 11 are the directions of the
.
ments u /2R and between . vertical
. and rotational dis- displacement increment vectors.
placement increments w/2R. The data for both footing diameters can be pre-
Figure 10 illustrates the change of vertical displace- sented on the same figure by normalising with respect
ment during the rotation of the caisson. The suction to Vo, the maximum applied vertical load. These results
installed caisson experiences a lower magnitude of are shown on Figure 12. Equation (1) has been included
uplift compared with the caisson installed by pushing. in this plot with a value of to  0.064 for the smaller
footing and 0.040 for the larger footing. It is necessary
to use different values of to in this plot because the ten-
4.1 Yield Surface and velocity vectors
sile capacity scales with 2RL2 whilst the Vo value
Using the yield points ((M/2R)y, V ) in Table 4, a plot scales principally with 2RtL. Since the two footings
of the yield surface for low vertical loads is illustrated have the same L/2R value but different t/2R values

415

Copyright 2005 Taylor & Francis Group plc, London, UK


0.15 suction installed ; D = 293 mm (c) The ratio of horizontal and rotational
. plastic dis-
.
Moment Load, (M/2RVo),

pushing installed ; D = 293 mm placement increments, u p/2R p, was independent


suction installed ; D = 200 mm of the installation method.
pushing installed ; D = 200 mm (d) Under rotations more vertical uplift was observed
2Rd(theta)

0.1
for the pushed installed caisson than the suction
installed caisson although this was also depend-
ent on the applied vertical load.
0.05 (e) The yield surface (equation (1)) was applied suc-
cessfully to two different size suction caissons
after normalisation by Vo, but requires differing
0 values of to, to account for different ratios of 2R/t.
-0.1 -0.05 0 0.05 0.1 0.15
Vertical Load, V/Vo, dw

Figure 12. Summary of experimental yield points (nor- REFERENCES


malized by Vo) and incremental plastic displacement
vectors. Bolton, M.D. 1986. The strength and dilancy of sands.
Gotechnique, Vol 36, No. 1, pp 6578
their tensile capacities differ on the normalised plot. Bye, A., Erbrich, C.T., Rognlien, B. and Tjelta, T.I. 1995.
However, the normalisation by Vo merges the two Geotechnical design of bucket foundations. Offshore
curves shown in Figure 11 for any one caisson, thus Technology Conference, Houston, paper 7793
suction or pushing installation has only a minor effect Byrne, B.W. 2000. Investigations of suction caissons in
on the normalised curve. dense sand, DPhil thesis, University of Oxford
In more detail, however, the yield surfaces pre- Byrne, B.W. and Houlsby, G.T. 2003. Foundation for off-
sented in Figure 12 serve as lower bounds for the shore wind turbines, Phil. Trans. of the Royal Society of
moment capacity in the case of a caisson installed by London, Series A 361, 29092300
Byrne, B.W., Houlsby, G.T., Martin, C.M. and Fish, P.M.
pushing. On the other hand, it represents an upper 2002. Suction caisson foundations for offshore wind tur-
bound for a suction installed caisson. The differences bines. Wind Engineering, Vol. 26, No 3
are thought to be due to disturbance in the installation Byrne, B.W., Villalobos, F.A., Houlsby, G.T. and Martin,
process due to suction. C.M. 2003. Laboratory Testing of Shallow Skirted
The incremental plastic displacement vectors were Foundations in Sand, Proc. Int. Conf. on Foundations,
also compared. The vectors for suction installation Dundee, 25 September, Thomas Telford, 161173
tests have a smaller component in the w-direction Gottardi, G., Houlsby, G.T. and Butterfield, R. 1999. The
compared with the vectors for pushed installation tests Plastic Response of Circular Footings on Sand under
(see last column in Table 4 for values). Therefore, General Planar Loading, Gotechnique, 49, 4,
pp 453470
there was less uplift during the rotation of a caisson Houlsby, G.T. and Cassidy, M.J. 2002. A plasticity model for
when a suction installation procedure was used. the behaviour of footings on sand under combined load-
ing, Geotchnique, Vol. 52, No. 2, pp 117129
Houlsby, G.T. and Byrne, B.W. 2005. Calculation pro-
5 CONCLUSIONS cedures for installation of suction caissons in sand. Proc
ICE, Geotechnical Engineering, in press
A series of experiments comparing the moment House, A.R. 2002. Suction Caisson Foundations for
response for suction installed caissons and those Bouyant Offshore Facilities, PhD thesis, the University
installed by pushing have been carried out. The main of Western Australia
results are: Kelly, R.B., Byrne, B.W., Houlsby, G.T. and Martin, C.M.
2004. Tensile loading of Model Caisson Foundations for
(a) The use of suction beneficially reduces the resist- Structures on Sand, Proc. ISOPE. Conf., Toulon
ance to penetration of the caisson. Martin, C.M. 1994. Physical and Numerical Modelling of
(b) The moment resistance of a suction caisson Offshore Foundations under Combined Loads, DPhil
depends on the method of installation. thesis, University of Oxford

416

Copyright 2005 Taylor & Francis Group plc, London, UK


The theoretical modelling of a suction caisson foundation using
hyperplasticity theory

Lam Nguyen-Sy & Guy T. Houlsby


Department of Engineering Science, Oxford University

ABSTRACT: A theoretical model for the analysis of suction caison foundations, based on a thermodynamic
framework (Houlsby & Puzrin 2000) and the macro-element concept is presented. The elastic-plastic response is
first described in terms of a single-yield-surface model, using a non-associated flow rule. To capture hysteresis
phenomena, this model is then extended to a multiple yield surface model. The installation of the caisson using
suction is also analysed as part of the theoretical model. Some preliminary numerical results are given as demon-
strations of the capabilities of the model.

1 INTRODUCTION LRP
VR
In developments of offshore wind turbines, the foun-
dations account for a significant fraction of the overall HR
installed cost, approximately 15% to 40% of the total MR mudline
cost (Houlsby & Byrne 2000). To satisfy the increas-
ing need for renewable energy, there are a number of
offshore wind farms to be constructed off the coast of
the UK within the next few years. The possible use V d
of caisson foundations for these turbines is therefore
an important economical issue.
From previous research, there are elastic-plastic H M
theoretical models available for analysis of shallow Soil surface level
offshore foundations, such as Model B for a jack-up
footing on clay (Martin 1994) and Model C for foot- 2R
ings on sand (Cassidy 1999). These models are based
on the idea of a macro-element, representing the
foundation behaviour. The loading on the footing is Figure 1. Geometry of caisson footing.
represented by force resultants at a chosen reference
point on the footing, and the movement by the corre- of gravity and suction within the caisson. In Figure 1,
sponding displacements of this point. In this paper, a d is the distance between the Load Reference Point
macro-element model for a caisson is presented in (LRP) and an idealised soil surface position just as
outline. The main goal of this work is to establish a installation begins. The position of the LRP is arbi-
theoretical framework to model correctly the cyclic trary, but is conveniently taken at the joint between
behaviour of a caisson foundation, and this necessi- the caisson and the support structure. The conventions
tates extension of previous modelling concepts to use for forces are shown in Figure 2. The forces VR, H2R,
of multiple yield surfaces. H3R, QR, M2R, and M3R are applied at the LRP. In the
analysis, however, it is often convenient to use the
force system, i  (V, H2, H3, Q, M2, M3) at the ideal-
2 CAISSON FOUNDATIONS ised soil surface level. The relationships between
these two systems are: V  VR; H2  H2R; H3  H3R;
A caisson foundation consists essentially of two parts: Q  QR; M2  M2R  dH3R; M3  M3R
dH2R.
a circular top plate and a perimeter skirt, see Figure 1. The displacement vector at soil surface level is i 
The whole foundation is installed by the combination (w, u2, u3, , 2, 3). The corresponding displacements

417

Copyright 2005 Taylor & Francis Group plc, London, UK


in which: K1  2GRk1; K2  GR3k2
8GR2k4d 
2GRd2k3; K3  2GRk3; K4  4GR2k4
2GRk3d;
K5  8GR3k5 and D  K2K3
K42. G is the shear mod-
ulus of the soil, and the factors k1 k5 are dimen-
sionless stiffness coefficients as proposed by Doherty
et al. (2004), who give elastic solutions for a caisson
foundation, in which the elastic stiffness of the cais-
son itself is taken into account.
The g2 term, which is the work of the plastic dis-
placements, specifies the kinematic hardening of
the model. A simple linear hardening relationship is
achieved if g2 is a quadratic function of the plastic
strains, and this form will later be used here as the
basis for the multiple-surface model:

Figure 2. Conventions for forces.

at the LRP are given w  wR; u2  u2R  d3R; (4)


u3  u3R
d2R;   R; 2  2R; 3  3R.

where H 1* H5* are hardening parameters which can


3 RATE-INDEPENDENT SINGLE YIELD conveniently be expressed in terms of the elastic stiff-
SURFACE HYPERPLASTICITY MODEL ness factors K1 K5. Details of these functions are
discussed later. For the time being, however, we sim-
Based on the hyperplasticity framework (Houlsby & ply take g2  0 for the single surface model since the
Puzrin 2000), a mechanical model can be derived yield surface in this case does not undergo kinematic
from two scalar functions: the Gibbs free energy g, hardening.
and either the dissipation function d or yield function y.
The yield function is used here, since it can be identi-
fied directly from test results. The macro-element 3.2 The yield function y
model is expressed in terms of the force vector i and
displacement vector i. It is also necessary to intro- In the hyperplasticity framework (Houlsby & Puzrin
duce the generalized force vector i  (V, H2, H3, 2000), the yield function has been recognised as the
Q, M2, M3) and finally the plastic displacement (or singular Legendre transform of the dissipation func-
internal variable) vector i  (V, H2, H3, Q, M2, tion d in case of rate-independent materials. The yield
M3) . In general, for a model without elastic-plastic surface is therefore expressed as a function of the
coupling, the free energy g and yield function y can generalised forces i. Furthermore, the appearance of
be expressed as: force components, (analogous to the true stresses ij
in continuum plasticity models), in the form of the
yield function leads automatically to non-associated
(1) flow rules (Collins & Houlsby 1997), which are known
as the appropriate to describe soil behaviour. Conse-
(2) quently, association factors which play an interpo-
lation role between true and generalised forces are
proposed in the yield function. Martin (1994) and
3.1 The Gibbs free energy function g Cassidy (1999) have established the yield functions in
In Eq. (1), g1 represents the elastic response of the foun- elastic-plastic models (Model B and Model C for jack-
dation and is independent of plastic displacements. up and circular footings of offshore structures on clay
For linear elasticity it takes the following form: and sand). Based on these results, a yield function for
a caisson footing is proposed with certain modifica-
tions to include aspects such as the non-associated
flow rule:

(3)
(5)

418

Copyright 2005 Taylor & Francis Group plc, London, UK


in which: The parameters aV1, aV2, aM, aH, aQ are the asso-
ciation factors; V, H2, H3, Q, M2 and M3 are the
back stresses which are the difference between true
force and generalised force, and are in turn expressed
as functions of the internal variable i.

and further definitions follow below: 4 RATE-INDEPENDENT CONTINUOUS


HYPERPLASTICITY MODEL
(6)
The main reason for the introduction of continuous
hyperplasticity, which is in effect models an infinite
number of yield surfaces, is to simulate a smooth trans-
(7) ition between elastic and plastic behaviour, and cap-
ture with reasonable precision the hysteretic response
of a foundation under cyclic loading. Such behaviour
can not be described by a conventional single yield
(8) surface model.

4.1 The Gibbs free energy function g


(9) Starting from the form of Gibbs free energy function
for a single yield surface model as in Eq. (1), further
developments for a continuous hyperplasticity model
can be made. The Gibbs free energy function now
(10) becomes a functional as follows:

(11)

(12)

(14)

(13) where  is a dimensionless parameter which varies


from 0 to 1 and expresses the relative sizes of the yield
It is convenient to note that the vertical load at surfaces. When   0, no plastic behaviour occurs.
which the maximum dimension of the yield surface is Once   1, fully plastic behaviour occurs. The hat
achieved is V/V0    ( 1
2t0)/( 1  2), lead- notation is used to denote any function of . The
ing to 1/ 2  (t0  )/(1
). hardening parameters in Eq. (4) now become func-
tions of . These functions determine the shapes of
V0 is the vertical bearing capacity of the foundation
the force-displacement curves. Hyperbolic curves
(the intercept of the yield surface on the positive
may conveniently be used and for this case the hard-
V-axis). Appropriate values of the parameters specify-
ening functions have the form:
ing the yield surface shape for a typical fully embed-
ded caisson are 1  2  0.99; t0  0.1088. The
parameters m0, h0 and q0 are factors which determine (15)
the sizes of the yield surface in the moment, horizon-
tal and torsion directions: typical values are 0.15, where Ai, bi and ni are parameters defining the shape
0.337 and 0.1 respectively. of the curves.

419

Copyright 2005 Taylor & Francis Group plc, London, UK


4.2 The yield function y The above is appropriate provided that the N value cho-
sen is large enough to result in a small di to achieve
For a certain value of , the yield function can be
a reasonable approximation to the integral by use of a
expressed as follows:
summation. Eq. (15) now becomes:
(16)
(20)
where

5.2 The yield function y


(17)
Using the same style of yield function as in Eq. (16),
All the definitions of variables in Eqs. (16) and (17) the jth yield surface can be expressed as:
are as for single-yield model, but these variables are
determined for the yield surface corresponding to .
(21)
5 MULTIPLE-YIELD SURFACE
HYPERPLASTICITY MODEL
where equations exactly similar to (6) to (13) apply,
The concept of an infinite number of yield surfaces can but with each definition applying for the jth surface,
model very well the response of a foundation under thus Eq. (7) becomes for example:
cyclic loading. However, to implement the model in a
numerical analysis, it is necessary to discretise the
continuous plasticity model to a multiple-yield-surface (22)
model. Firstly, the integrals in the Gibbs free energy
become summations. Secondly, the continuously vary-
ing functions of  are replaced by discrete variables.
The factors, S, 12, 1, 2 and t0, have the same val-
ues as in the single yield surface model. The defin-
5.1 The Gibbs free energy function g itions of the generalised forces can be expressed as
The hat notation as in Eq. (14) is now abandoned to follows:
express the fact that the variables are no longer func-
tions, but a series of discrete values. N is the number
of yield surfaces chosen to simulate the continuous (23)
yield surface. We replace  by the factor j/N where j
is the number of the yield surface which is being con-
sidered. In the summation, the increment d in the
integral becomes:

(18)
(24)
The free energy function is therefore:

(25)

(19) (26)

420

Copyright 2005 Taylor & Francis Group plc, London, UK


consistency conditions is not straightforward in
numerical analyses. The use of rate-dependent behav-
iour has been proposed as a means to simplify the
numerical difficulties by Houlsby & Puzrin (2002)
(27) and Puzrin & Houlsby (2003). The dissipation func-
tion d is in this case separated into two functions;
force potential function z and flow potential w.
Houlsby & Puzrin (2002) note that w can take alter-
native forms, depending on the form of the viscosity
assumed. In this paper, linear viscosity is used and the
flow potential functions can be defined as:

(28)
(30)

The coordinates of center of jth yield surface in stress where  is the viscosity; yj is the jth yield function
space can be defined as: which no longer needs to be identically zero. Note
that wj is only zero when the rates of change of plas-
(29) tic displacements are all zero. The incremental changes
of plastic displacements caused by the jth yield sur-
and likewise for the other variables. face can be defined as:
Figure 3 shows the form of yield surfaces after a
purely vertical loading. The size of the smallest yield
surface in the vertical load direction is set as a certain (31)
fraction of the size of the outer yield surface. Between
the inner and outer surfaces a uniform distribution of
sizes of yield surfaces is used. The purpose of using a The total displacement increments are now calcu-
non-zero size of the first yield surface on the V-axis is lated as:
to control the development of vertical plastic dis-
placement on vertical unloading.
(32)
5.3 Incremental response
In the multiple-yield-surface model, using rate-
independent behaviour, the loading point must always
be within or on each yield surface. This condition 6 NUMERICAL ILLUSTRATIONS
requires that the y-values for all active yield surfaces
must be identically zero. The imposition of these Firstly, a result modelling the suction assisted pene-
tration process using the concepts of Houlsby &
Byrne (2005) is shown in Figure 4.
t Secondly, a numerical example is given to illus-
trate test results which are obtained from laboratory
testing of model caissons. In this example, AV  1.0;
AH2  AH3  AQ  AM2  AM3  0.5; bV  bH2 
b H3  b Q  b M2  b M3  1.0; n V  n H2  n H3 
1.0
nQ  nM2  nM3  3.0. Twenty yield surfaces are used.
t0 v = V/V0 The values of yield function parameters are: am 
ah  0.7; aV1  0.297; aV2  1.0; t0  0.1088; m0 
0.15; h0  0.337; the shear modulus of the soil is
G  0.7 MPa, self-weight   15.74 kN/m2, Poisson
ratio   0.2; initial fraction for the first yield func-
Initial fraction = 0.8 tion  0.8. The radius of caisson R  146.5 mm; the
length of the perimeter skirt H  146.5 mm. The cais-
son is installed to the full penetration position and
Figure 3. Multiple yield surfaces. then the horizontal and moment loads are applied.

421

Copyright 2005 Taylor & Francis Group plc, London, UK


300 0.1445
Vertical penetration without Test results
0.1444
250 suction
Vertical loads (kN)

0.1443 Theoretical
200 Vertical penetration with 0.1442
suction assistance

w (m)
0.1441
150
0.144
100 0.1439
50 0.1438
0.1437
0 -0.004 -0.002 0 0.002 0.004
0 0.5 1 1.5 2
Theta3 (rad)
Vertical penetrations (m)
Figure 7. Vertical movements under cyclic loading.
Figure 4. Installation processes with and without suction.

7 DISCUSSION
20
There are four main points that must be addressed in
15
this model: the choice of the hardening functions, the
10
values of the association factors, the effects of suction
5
M3 (Nm)

pressures and the use of the rate dependent solution.


0 Firstly, the hardening functions, Hi*, determine the
-0.004 -0.002 -5 0 0.002 0.004
distributions of plastic displacements which are caused
-10 Test results by each yield surface. Therefore, the solutions can
-15 Theoretical results become stiffer or softer by increasing or decreasing the
-20 factors Ai, bi or ni. It is very important to determine
Theta3 (rad) the appropriate value of the shear modulus G. Since
the hardening functions depend on the elastic stiff-
Figure 5. Rotation under cyclic loading. ness factors, which include the shear modulus, the
value of G strongly affects the solutions.
Secondly, the association factors play the role of
60 determining the direction of the flow vectors of the
Test results plastic strains. To choose suitable values for these fac-
40 tors, it is necessary to consider some special aspects
Theoretical results
20 of the yield functions, such as the positions of the
parallel points where the vertical plastic displace-
H2 (N)

0 ment increments are zero. The directions of the flow


-0.0004 -0.0003 -0.0002 -0.0001 0 0.0001 0.0002 0.0003
-20
vectors in the (V, M) plane, (M, H) plane or (V, H)
plane can be obtained from tests. Furthermore, during
-40 the application of lateral loads, the upward or down-
-60
ward movements of the footing are also depend on the
position of the parallel point and the value of the ver-
u2 (m)
tical load. However, the details of these expressions
Figure 6. Horizontal displacement under cyclic loading.
are beyond the scope of this paper.
Thirdly, as shown in Figure 4, by using suction, the
vertical load which must be applied for installation is
The vertical load V increases to the value of V  945 N rather small compared with that for installation using
during the penetration process and decreases to the purely vertical load. This feature is very useful because
value of V  50 N before the lateral loads are applied. it is impossible to apply a large value of vertical load
During the application of the cyclic loads, the vertical to install the caisson in practice. Consequently, by using
load is kept constant at 50 N. The resulting moment- suction assisted penetration, this obstacle can be
rotation behaviour is shown in Figure 5, and horizontal overcome.
load-displacement in Figure 6. Finally the vertical Lastly, in order to avoid numerical difficulties, the
movements during the cycling are presented in Figure 7. rate-dependent solution has been proposed. The most
In each case the analyses are compared to test results, important aspect of using the rate-dependent solution
and it can be seen that a satisfactory agreement is is the relationship among the viscosity , the time step
achieved. dt and the load step. Suitable values must be chosen to

422

Copyright 2005 Taylor & Francis Group plc, London, UK


maintain accuracy and stability for the numerical solu- Materials. Proc. Royal Society of London, Series A,
tion. There are as yet no precise procedures for select- Vol. 453, pp 19752001.
ing for these parameters. However, by using some Doherty, J.P., Deeks, A.J. and Houlsby, G.T., 2004.
trials, one can determine suitable values for , dt and Evaluation of Foundation Stiffness Using the Scaled
Boundary Method. Proc. 6th World Cong. on Comp.
load step. Mech., Beijing, 510 Sept.
Houlsby, G.T. and Byrne, B.W., 2000. Suction caisson foun-
dations for offshore wind turbines and anemometer
8 CONCLUSIONS masts. Wind Engineering, Vol. 24, No. 4, pp 249255.
Houlsby, G.T. and Byrne, B.W., 2005. Design Procedures for
Installation of Suction Caissons in Sand, Proc. ICE,
This paper presents a multiple yield surface hyper-
Geotechnical Engineering, in press.
plasticity model for caisson foundations. Preliminary Houlsby, G.T. and Cassidy, M.J., 2002. A plasticity model for
choices for the parameters are made. The model the behaviour of footings on sand under combined load-
captures reasonably well the behaviour of a caisson ing. Gotechnique, 52, No 2, pp 117129.
foundation under cyclic loading, and could be incorp- Houlsby, G.T. and Puzrin, A.M., 2000. A Thermomechanical
orated in numerical analyses of caisson/structure Framework for Constitutive Models for Rate-Independent
systems. Dissipative Materials. Int. J. of Plasticity, Vol. 16, No. 9,
pp 10171047.
Houlsby, G.T. and Puzrin, A.M., 2002. Rate-Dependent
Plasticity Models Derived from Potential Functions. J. of
REFERENCES Rheology, Vol. 46, No. 1, Jan./Feb., pp 113126.
Martin, C.M., 1994. Physical and numerical modelling of off-
Cassidy, M.J., 1999. Non-linear analysis of Jack-Up struc- shore foundations under combined loads. DPhil thesis,
tures subjected to random waves. DPhil thesis, University University of Oxford.
of Oxford. Puzrin, A.M. and Houlsby, G.T., 2003. Rate Dependent
Collins, I.F. and Houlsby, G.T., 1997. Application of Ther- Hyperplasticity with Internal Functions, Proc. ASCE, J.
momechanical Principles to the Modelling of Geotechnical Eng. Mech. Div., Vol. 129, No. 3, March, pp 252263.

423

Copyright 2005 Taylor & Francis Group plc, London, UK


Shallow foundations: vertical bearing capacity

Copyright 2005 Taylor & Francis Group plc, London, UK


Bearing capacity of parallel strip footings on non-homogeneous clay

C.M. Martin & E.C.J. Hazell


Department of Engineering Science, University of Oxford, UK

ABSTRACT: On soft seabed soils, subsea equipment installations are often supported by mudmat foundation
systems that can be idealised as parallel strip footings, grillages, or annular (ring-shaped) footings. This paper
presents some theoretical results for the bearing capacity of (a) two parallel strip footings, otherwise isolated;
(b) a long series of parallel strip footings at equal spacings. The soil is idealised as an isotropic Tresca material
possessing a linear increase of undrained strength with depth. The bearing capacity analyses are performed
using the method of characteristics, and the trends of these (possibly exact) results are verified by a companion
series of upper bound calculations based on simple mechanisms. Parameters of interest are the footing spacing,
the relative rate of increase of strength with depth, and the footing roughness. An application of the results to
the design of perforated mudmats is discussed.

1 INTRODUCTION depths greater than a few hundred metres, the


undrained strength at seabed level can be as low as
Shallow foundations are usually designed on the 2 to 10 kPa, increasing with depth at 1 to 2 kPa/m
assumption that they act in isolation. When two foot- (Randolph 2004). A typical offshore mudmat might
ings (or a group of footings) are closely spaced, be 5 m wide, and if supported by a single strip footing
however, there is a beneficial interaction that can be (without perforations) this would imply typical values
quantified in terms of the efficiency, i.e. the ratio of of the ratio kB/su0 in the range 0.5 to 5. When calcu-
the overall (group) bearing capacity to the sum of the lating the bearing capacity of an isolated footing at
individual (isolated) bearing capacities. The literature the upper end of this range, the influence of non-
on this topic has been surveyed by Hazell (2004). For homogeneity on the bearing capacity would certainly
footings on sand, numerous theoretical and experi- be accounted for, either by adopting appropriate plas-
mental studies have shown that the effect of inter- ticity solutions (Davis & Booker 1973, Houlsby &
action becomes highly significant for friction angles Wroth 1983), or by selecting a representative strength
greater than about 30 and spacings less than about su  su0. It is therefore of interest to investigate the
one footing width B. In contrast, the undrained bear- undrained bearing capacity of closely spaced footings
ing capacity of closely spaced footings on clay has for a similar range of kB/su0, and to assess the effect
received very little attention, perhaps because the of using a mudmat with perforations in place of a
early theoretical work of Mandel (1963) showed that continuous foundation (this is sometimes done to
the beneficial effect of interaction was insignificant, save weight, and to make the structure easier to
even for fully rough footings. This was confirmed remove). For a small degree of perforation it might be
experimentally by Hazell (2004), though only a few envisaged that there will be arching over the gap(s),
of his tests were conducted on clay. such that there is no loss of bearing capacity, and even
When considering the relevance of these findings if some soil is squeezed through, there may still be a
to the design of grillages or closely spaced footings beneficial interaction effect. Here we investigate
on soft offshore soils, it is important to note that the these issues using plasticity analyses.
theoretical studies by Mandel (1963) were confined
to homogeneous soil, and although in the experiments
2 BEARING CAPACITY ANALYSES
of Hazell (2004) there was a marked increase of
undrained strength with depth, the dimensionless
2.1 Isolated footings
ratio kB/su0 was no more than 0.2 for the small model
footings tested (su0  mudline strength intercept, The bearing capacity of an isolated strip footing on non-
k  rate of increase of strength with depth). In water homogeneous clay was first studied by Davis & Booker

427

Copyright 2005 Taylor & Francis Group plc, London, UK


(a) (a)
10
B S B
Smooth
9 x
Rough
Nc = Qu/Bsu0

8 su = su0 + kz
u = 0
7 z

6 (b)
S B S B S B S B S
5 etc. etc. x
0 1 2 3 4 5 su = su0 + kz
kB/su0 u = 0
z
(b)
1 Figure 2. Parallel strip footings on non-homogeneous clay:
(a) pair of footings (b) many footings, equally spaced.
Smooth
0.75 Rough
Critical S/B

0.5
but this deteriorates somewhat with increasing non-
homogeneity, especially for the rough footing.
0.25
2.2 Interacting footings methodology
0
0 1 2 3 4 5 In principle, bearing capacity analyses can be per-
formed for an arbitrary number of parallel strip foot-
kB/su0
ings at arbitrary spacings, but the calculations can
become tedious, particularly when using the method
Figure 1. Isolated strip footing on non-homogeneous clay:
of characteristics. Figure 2 shows the two problems
(a) bearing capacity (b) critical edge-to-edge spacing for
interaction between parallel footings. Results from simple considered here, both of which allow a favourable
UB calculations (after Kusakabe et al. 1986) shown dotted. exploitation of symmetry: a pair of parallel strips, and
an infinite number of parallel strips at equal spacings.
Although impossible to realise in practice, the latter
(1973). They showed that for any value of kB/su0 (zero case is relevant to the interior members of a grillage
to infinity) the stress and velocity fields computed containing many bearing elements. Note that in this
using the method of characteristics furnished lower and paper S refers to the edge-to-edge spacing, not the
upper bounds that were coincident. Davis & Bookers centre-to-centre spacing as preferred by some authors.
analyses have since been verified by several authors, Note also that the footings are assumed to be rigidly
and they can also be replicated using the free computer connected, such that they move down together with-
program ABC (Martin 2004). Figure 1a shows the vari- out any horizontal displacement or rotation (for a pair
ation of the bearing capacity factor Nc as a function of of closely spaced footings there is a tendency for sep-
kB/su0. Note that Nc is defined with respect to the mud- aration and tilting to occur).
line strength, i.e. as Qu/Bsu0 where Qu is the ultimate The main bearing capacity analyses for interact-
bearing capacity (per unit run). Figure 1b shows, for the ing footings were performed using the method of
same range of kB/su0, the extent of the zone of plastic characteristics. A modified version of the Matlab
deformation adjacent to each side of the footing. This program InterBC, developed by Hazell (2004) for
distance is important because it also corresponds to the interacting footings on a homogeneous c- - soil,
critical spacing at which parallel strip footings first was used. For a pair of footings, the program con-
begin to interact and give an overall bearing capacity siders the right-hand footing and builds two meshes
that is greater than the sum of the individual capacities. of characteristics, one commencing from the exterior
As kB/su0 increases, the zone of plastic deformation soil surface, and one from the gap between the foot-
becomes smaller, so the footings need to be closer for ing and the axis of symmetry (see Fig. 3). An iterative
any interaction to occur. adjustment process is used to ensure that the two
Also shown in Figure 1 are the results of upper meshes are fully compatible at their common point C
bound calculations performed using the generalised (same coordinates, same stresses). Having calculated
Hill- and Prandtl-type mechanisms of Kusakabe et al. the stress field, the program works back through the
(1986), which were originally developed for circular mesh and constructs the associated velocity field.
footings on non-homogeneous clay. There is close These calculations are more complicated than those
agreement with the exact curves when kB/su0 is small, for an isolated footing, since the inward-flowing soil

428

Copyright 2005 Taylor & Francis Group plc, London, UK


(a) Smooth (a) Smooth

A C B
O A C
O Note: half of one
D footing shown
D

(b) Rough

A B (b) Rough
O

D
C A
O Note: half of one
footing shown
C
D
Figure 3. Pair of parallel strip footings on non-
homogeneous clay (kB/su0  1, S/B  0.15): characteristics
and velocity vectors, with mechanism outlines from simple
UB calculations.
Figure 4. Many parallel strip footings on non-homogeneous
clay (kB/su0  1, S/B  0.15): characteristics and velocity vec-
tors, with mechanism outlines from simple UB calculations.
crosses a velocity discontinuity AD. Note that for
non-homogeneous soil, the velocities outside AOD
are not always parallel to the characteristics. determining the variation of bearing capacity with
Two separate calculations of the bearing capacity spacing, it is always necessary to check this alterna-
are performed: one by integrating the stresses along tive mechanism, for which the bearing capacity can
ACB (deducting the self-weight of the false head if be determined using ABC.
applicable), and the other by equating the internal and Figure 4 shows some typical solutions for the case
external work rates of the collapse mechanism. In all of infinitely many, equally spaced footings on non-
of the analyses for this study it was found that, as the homogeneous clay. Here the double symmetry means
mesh of characteristics was refined, the two calcula- that it is only necessary to analyse half of one footing,
tions of the bearing capacity converged to identical val- so analysis using the method of characteristics is rel-
ues. While this indicates that there are no regions of atively straightforward. A single mesh is constructed,
negative plastic work, it does not necessarily mean that starting from the soil surface and bouncing character-
the calculated bearing capacity is the exact collapse istics off the centerline of the gap as before. The mesh
load, since it has not been demonstrated that the stress is adjusted until point C lies directly beneath the cen-
field can be extended throughout the soil. Construction tre of the footing, with the major principal compres-
of such an extension is straightforward for isolated sion aligned vertically. Calculation of the associated
footings (see e.g. Davis & Booker 1973, Martin 2005), velocity field again incorporating a discontinuity
but for interacting footings it would be necessary to along AD can then be performed. As in the two-
incorporate a non-plastic zone to allow the major prin- footing case, it was always found that the stress- and
cipal compression to flip from horizontal to vertical at velocity-based calculations of the bearing capacity
some point on the z axis. This is not easy, and sug- converged to the same value as the mesh was refined.
gests that the method of characteristics cannot readily This converged value represents a rigorous upper
be used to obtain strict lower bounds for interacting bound, but not a rigorous lower bound since the stress
footings (finite element limit analysis could be used). field is incomplete. If the number of footings is truly
For a pair of smooth footings, the squeezing type infinite then there is no alternative overall failure
failure of Figure 3a is always critical (except on homo- mechanism squeezing failure is the only option, and
geneous soil, where there is no interaction effect, and the bearing capacity must approach infinity as the
an infinite number of one- and two-sided mechanisms spacing tends to zero.
giving Nc  2  can be devised). For a pair of rough As well as the analyses performed using InterBC,
footings that are very closely spaced, overall failure as some simple Hill- and Prandtl-type upper bound
a single footing of width 2B  S may be more critical mechanisms were devised for the problems shown in
than the squeezing failure of Figure 3b. When Figure 2. These were based on the mechanisms of

429

Copyright 2005 Taylor & Francis Group plc, London, UK


Kusakabe et al. (1986), but modified to allow a rigid (a) kB/su0 = 0
wedge of soil to be extruded vertically between the 1.3
interacting footings. In Figures 3 and 4 the outlines of
the optimal upper bound mechanisms are superimposed
Smooth
on the method of characteristics solutions, and there is a 1.2

Efficiency
Rough
fairly close correspondence between the two. Note that
for rough footings, the Prandtl-type mechanism shown
in Figures 3b and 4b is only critical when kB/su0 is 1.1
small; otherwise a Hill-type mechanism (similar to Figs
3a, 4a) governs. Smooth effic. = 1 for all S/B
1
0 0.1 0.2 0.3 0.4 0.5
2.3 Interacting footings results S/B

Concentrating first on the pair of interacting footings, (b) kB/su0 = 1


Figure 5 shows the variation of efficiency (as defined 1.3
at the start of the paper) with the normalised spacing
S/B. For a spacing of zero, the two footings behave as a
Smooth
single footing of width 2B. Although this has no net 1.2

Efficiency
Rough
effect when the soil is homogeneous, there is a benefi-
cial interaction when the strength increases with depth
since the influence of non-homogeneity is enhanced 1.1
(2 kB/su0 is greater than kB/su0, so the operative Nc is
greater in Fig. 1a). For smooth footings the efficiency
1
immediately begins to drop as soon as a gap is intro- 0 0.1 0.2 0.3 0.4 0.5
duced, while for rough footings there is a brief increase S/B
in efficiency prior to the transition between overall and
squeezing failure (see Section 2.2). In all cases there is (c) kB/su0 = 2
a gradual decline towards unit efficiency as the critical 1.3
spacing plotted in Figure 1b is approached.
The results for homogeneous soil (Fig. 5a) agree
Smooth
with those of Mandel (1963): there is no gain in effi- 1.2
Efficiency

Rough
ciency for a pair of smooth footings, and a peak of
just 1.07 (at S/B  0.15) for a pair of rough footings.
When the strength increases with depth, the potential 1.1
gains in efficiency are rather greater, but the spacing
needs to be small ( 0.1B), and the benefit from inter-
1
action is almost all attributable to the effective aug-
0 0.1 0.2 0.3 0.4 0.5
mentation of kB/su0 rather than a genuine arching
S/B
effect. In fact, the rough footing curves in Figure 5
clearly show that the influence of arching diminishes (d) kB/su0 = 5
rapidly as non-homogeneity becomes more signifi- 1.3
cant; when kB/su0  2 there is an almost immediate
transition from overall failure to squeezing failure as
Smooth
the spacing is increased from zero. 1.2
Efficiency

Rough
The predictions from the method of characteristics
are consistent with those from the simple upper
bound analyses, shown as dotted lines in Figure 5. 1.1
The efficiency curve for a pair of rough footings on
homogeneous clay (Fig. 5a) also agrees remarkably
1
well with that obtained by Galloway (2004) using the
0 0.1 0.2 0.3 0.4 0.5
finite element program ABAQUS. This suggests that
S/B
the results obtained here from the method of charac-
teristics may well be exact, though it is not immedi- Figure 5. Pair of parallel strip footings: variation of effi-
ately clear why the squeezing stress field should ciency with edge-to-edge spacing. Efficiency  ratio of
suddenly cease to become extensible at the same overall (group) capacity to sum of individual (isolated)
moment that overall failure becomes critical. It is capacities. Results from simple UB calculations shown
noteworthy and surely no coincidence that dotted.

430

Copyright 2005 Taylor & Francis Group plc, London, UK


(a) Smooth (a)
1.5 W

1.4
Efficiency

kB/su0u0==
1.3 B* S B*
1, 2,
0, 1, 2, 55
1.2 S
Perforation ratio R =
W
1.1
(b)
1
W
0 0.1 0.2 0.3 0.4 0.5
S/B .....
(b) Rough B* S
1.5 S
Perforation ratio R =
kB/su0 = B*+S
1.4 0, 1, 2, 5
Efficiency

1.3 Figure 7. Mudmat with (a) single central perforation (b)


many identical perforations. Perforation ratio R  fraction
1.2 of overall width W that has been removed.

1.1

1 practice such structures would usually have square or


0 0.1 0.2 0.3 0.4 0.5 circular geometry, with the bearing elements taking
S/B the form of a bidirectional grillage, or perhaps cre-
ated by making a series of regularly spaced holes in
Figure 6. Many parallel strip footings: variation of efficiency an initially solid base. Although the actual failure
with edge-to-edge spacing. Efficiency  ratio of overall mechanisms in these cases are complex (requiring 3D
(group) capacity to sum of individual (isolated) capacities. analysis), it is nevertheless instructive to perform
Results from simple UB calculations shown dotted. some simplified calculations for plane strain condi-
tions, to examine the general effect of introducing
Galloways analyses also predict a peak efficiency of perforations. For brevity, only the two extreme
1.07 at a spacing of S/B  0.15, coinciding with an cases shown in Figure 7 are considered. The notation
abrupt transition from overall to squeezing failure. adopted is the same as that used in the paper by White
Corresponding results for an infinite number of et al. elsewhere in these proceedings: W is the overall
equally spaced footings are shown in Figure 6. The width of the mudmat, B* is the width of the individ-
higher the value of kB/su0, the closer the footings need ual bearing elements, and R is the perforation ratio
to be before there is any significant gain in efficiency (i.e. the fraction of W that has been removed). In
(say 10%). For a given spacing, the increase in effi- Figure 7a, the ratios S/B* and kB*/su0 both change
ciency is greatest for the homogeneous soil, and con- constantly as R is increased from zero (the former
siderably higher (by a factor of up to 2) for rough increases and the latter decreases). In Figure 7b, S/B*
footings than for smooth footings. The critical spa- again increases as R is increased, but the ratio kB*/su0
cings at which the curves in Figure 6 reach unit effi- is effectively zero from the outset (since the number
ciency are the same as those in Figure 5. of perforations is very large, the width B* of each
The dotted lines in Figure 6 show that the upper individual bearing element is very small). In both
bound calculations give the same general trend, but cases, as R is increased there eventually comes a point
they overpredict the efficiency factor quite seriously where there is no longer any interaction between adja-
as the spacing becomes small. This is because the cent bearing elements of the mudmat.
simple collapse mechanisms (consisting only of fan Results for the single perforation scenario of Figure
zones and rigid blocks) are unsuitable for modelling 7a are shown in Figure 8. If the overall width W is
the increasingly complex velocity field as S/B 0. taken as given, it is appropriate to characterise the non-
homogeneity by kW/su0, and to define a gross bearing
capacity factor with respect to W, i.e. Nc  Qu/Wsu0.
3 APPLICATION: PERFORATED MUDMATS Using this convention, Nc for a smooth, singly-
perforated mudmat on homogeneous soil decreases
A common situation involving closely spaced foot- linearly from 2  to 0 as R increases from 0 to 1.
ings is the design of a mudmat with perforations. In The other curves in Figure 8a show that breaking W

431

Copyright 2005 Taylor & Francis Group plc, London, UK


(a) Smooth (a) Smooth
10 10
9 9
8 8
Nc = Qu/Wsu0

Nc = Qu/Wsu0
7 7
6 6
5 5
4 4 kW/su0 =
3 kW/su0 = 3 0,0,1,
1, 2,
2, 55
2 0, 1, 2, 5 2 envelope continues
1 1 linearly to (1, 0)
0 0
0 0.1 0.2 0.3 0.4 0.5 0 0.1 0.2 0.3 0.4 0.5
R = S/W R = S/(B*+S)
(b) Rough (b) Rough
10 10
9 9
8 8

Nc = Qu/Wsu0
Nc = Qu/Wsu0

7 7
6 6
5 5
4 4 kW/su0 =
3 kW/su0 = 3 0,0,1,
1, 2,
2, 55
2 0, 1, 2, 5 2 envelope continues
1 1 linearly to (1, 0)
0 0
0 0.1 0.2 0.3 0.4 0.5 0 0.1 0.2 0.3 0.4 0.5
R = S/W R = S/(B*+S)

Figure 8. Mudmat with single central perforation: vari- Figure 9. Mudmat with many identical perforations:
ation of bearing capacity with perforation ratio. Initial variation of bearing capacity with perforation ratio. Initial
plateaus in (b) correspond to overall failure with arching plateaus correspond to overall failure with arching across
across perforation. perforations.

into two smaller widths causes most of the benefit


derived from the increase of strength with depth to be 4 CONCLUSIONS
lost quite quickly, followed by a more gradual decline
once R passes about 0.1. The corresponding curves This paper has presented theoretical solutions for the
for rough-based mudmats with a single central perfo- vertical bearing capacity of rigidly connected, parallel
ration (Fig. 8b) have an initial plateau where overall strip footings on clay exhibiting a linear increase of
failure is more critical than squeezing failure. However undrained strength with depth. Both a pair of footings
the beneficial effect of this arching across the perfora- and a large group of equally spaced footings have
tion is only significant when kW/su0 is small. been considered. Results have been obtained using
Figure 9 presents results for the other scenario of a the method of characteristics, and confirmed by inde-
mudmat with numerous perforations (Fig. 7b). In this pendent upper bound calculations based on simple
case both the smooth and rough curves exhibit mechanisms. The former results are believed to repre-
plateaus corresponding to overall failure, but once sent exact solutions, though they have only been
squeezing failure becomes critical the bearing cap- established as upper bounds at this stage. A practical
acity starts to decline (somewhat more rapidly than in application to the design of perforated mudmats on
Fig. 8). Regardless of the value of kW/su0, the appro- soft offshore soil has been explored.
priate squeezing curve is always that for homogenous
soil, for the reason mentioned above: the ratio kB*/su0
is effectively zero because B* W. (The curved
envelopes in Figs 9a, b are simply alternative presen- REFERENCES
tations of the data for kB/su0  0 in Figs 6a, b.)
Davis, E.H. & Booker, J.R. 1973. The effect of increasing
Perhaps the most interesting prediction in Figure 9 is strength with depth on the bearing capacity of clays.
that for a rough mudmat on homogeneous soil, a per- Gotechnique 23(4): 551563.
foration ratio in excess of 15% can be tolerated with- Galloway, M. 2004. Interaction between adjacent footings in
out any reduction in capacity. Unfortunately this is offshore foundation systems. Final year project report,
clearly not the case when kW/su0  0. Department of Engineering Science, University of Oxford.

432

Copyright 2005 Taylor & Francis Group plc, London, UK


Hazell, E.C.J. 2004. Interaction of closely spaced strip Martin, C.M. 2004. ABC Analysis of Bearing Capacity.
footings. Final year project report, Department of Software and documentation available for download
Engineering Science, University of Oxford. from www-civil.eng.ox.ac.uk/people/cmm/software/abc.
Houlsby, G.T. & Wroth, C.P. 1983. Calculation of stresses Martin, C.M. 2005. Exact bearing capacity calculations
on shallow penetrometers and footings. OUEL Report using the method of characteristics. Issues lecture, Proc.
No. 1503/83, Department of Engineering Science, 11th Int. Conf. of IACMAG, Turin, to appear.
University of Oxford. Randolph, M.F. 2004. Characterisation of soft sediments for
Kusakabe, O., Suzuki, H. & Nakase, A. 1986. An upper- offshore applications. Keynote lecture, Proc. 2nd Int.
bound calculation on bearing capacity of a circular foot- Conf. on Site Investigation, Porto.
ing on a non-homogeneous clay. Soils and Found. 26(3):
143148.
Mandel, J. 1963. Interfrence plastique de foundations
superficielles. Proc. Int. Conf. on Soil Mech., Budapest:
267280.

433

Copyright 2005 Taylor & Francis Group plc, London, UK


Study on bearing behaviors of foundations on multi-layer subsoil

F.F. Yuan1, M.T. Luan1, 2, 3


1
Institute of Rock and Soil Mechanics, The Chinese Academy of Sciences, China;
2
Institute of Geotechnical Engineering, School of Civil and Hydraulic Engineering, Dalian University of
Technology, China;
3
State Key Laboratory of Coastal and Offshore Engineering, Dalian University of Technology, China

S.W. Yan
Institute of Geotechnical Engineering, School of Civil Engineering and Architectures, Tianjin University, China

ABSTRACT: Assuming the punching shear failure occurs in the layered subsoil, Meyerhofs bearing cap-
acity methods for two-layer subsoil was extended to estimating the bearing capacity of multi-layer subsoil. In
order to verify the extended methods, in-situ loading tests in Tianjin New Harbor were carried out, where there
was two-layer natural soil deposit and in turn that was overlain by sand and rubble artificially. And the experi-
mental results are compared with the predictions from the extended methods. It is shown that the calculated
results can well agree with the in-suit test results. Meanwhile, the three-dimensional nonlinear numerical analy-
ses by finite element method are conducted to search for the failure mechanism of layered subsoil. It is found
that under the ultimate loads, the failure model discovered from numerical analyses shall be same as the
assumed punching shear failure model.

1 GENERAL INSTRUCTION The first is the so-called projected area method


(Terzhaghi & Peck 1948, Jacobsen et al. 1977, Baglioni
The bearing capacity of foundations has always been et al. 1982, Kraft & Helfrich 1983). The second is
one of the subjects with major interest in soil mechan- punching shear method (Meyerhof 1974, Meyerhof &
ics and foundation engineering. In conventional engin- Hanna 1980). The third is the averaged strength param-
eering practice, a number of bearing capacity theories eters method (Sreenivasulu 1967, Satyanarayana &
(e.g. Terzaghi bearing capacity theory) is provided to Garg 1980, Hanna 1981). Meanwhile, many experi-
predict the ultimate bearing capacity of foundations. mental investigations have also been preformed to
However most of the theories relate to the foundations study the bearing capacity behavior of two-layer subsoil
on homogeneous soils and are not, in general, valid for (Siva Reddy & Srinivasan 1967, Mitus Okamura &
the case that the subsoil properties vary with depth. Jiro Takemura 1997). However, few experimental and
In practice, the subsoil profile often consists of theoretical investigations have been performed refer-
layered soils, especially in harbor and offshore engin- ring to the bearing behaviors of the subsoil with the
eering (e.g. jacking-up foundations and artificially profile of more than two layers.
filled foundations). Within each layer the soil may be In China, projected area methods and averaged
assumed homogeneous, although the strength param- strength parameters methods are usually recommended
eters of adjacent soils are different from each other for predicting the ultimate bearing capacity of layered
greatly. For this sake the mechanism of failure and subsoil. But the predictions from the two methods
bearing capacity theories for foundations on layered may be unacceptable for the case that the subsoil is
soils should differ from those on homogeneous soils. consisted of more than two layers (Yuan 2001).
However, up to now, the mechanism of failure is still In this paper, Meyerhof & Hannas punching shear
not clear, and to achieve the exact solution for bearing method to estimate the ultimate bearing capacity of
capacity of foundations on inhomogeneous soils is foundations on two-layer soils is introduced in details
difficult. In recent years, many approximate solutions and the method is extended to cover the case that the
have been presented in an attempt to provide acceptable subsoil profile is consisted of more than two layers.
results. Theses methods can be divided into three cat- In-situ loading tests on natural two-layer soils over-
egories according to the assumed failure mechanism. lain by sand and rubble artificially have been carried

435

Copyright 2005 Taylor & Francis Group plc, London, UK


out to verify the extended method. Finally, three- unit weight of soils  (4) the equivalent significant
dimensional nonlinear numerical analyses by finite depth Zmax, for the layered soils. With the strength
element method (using ABAQUS) are conducted to parameters, the ultimate bearing capacity can be
search for the failure mechanism of layered subsoil. determined using the bearing capacity factors Nc, Nq,
N based on Terzhaghi theory for homogeneous soils.
The ultimate bearing capacity can be presented as:
2 BEARING CAPACITY SOLUTIONS
FOR LAYERED SUBSOILS
(2)
2.1 Projected area methods
In China, projected area methods are recommended
by the Codes of civil engineering. The method assumes
that the upper layers served principally to spread the (3)
footing load to a larger area on the lower layer sur-
face, so that to reduce its intensity, and that the foun-
dation fails when bearing capacity failure occurs Where   coefficient of depth; i  value of unit
within the lower layer. The vertical stresses associated weight; i  angle of internal fraction; ci  value of
with the footing load are confined to a zone defined cohesion; Hi  and thickness of each layer.
by lines inclined at angle  to the vertical, as shown in
Figure1. Load from the footing is assumed to be dis-
2.3 Punching shear methods
tributed uniformly over a width B at the interface of
layers, where B  B  2Htan . However, it is hard In this method, the bearing capacity of the footing on
to make clear how the value of  should be selected. layered soils is estimated by considering a simplified
In practice, a value of  of tan
1 0.5 is often adopted mechanism in which the underlying weak soil is
(Houlsby et al. 1989).The ultimate bearing capacity assumed to be in a state of passive failure along a verti-
of the footing can be presented as: cal plane, beneath each edge of the footing, as shown in
Figure 2. Immediately beneath the footing, the vertical
(1) stress acting on the upper layer is assumed to be qb; a
passive earth force Pp inclined at an angle is assumed
Where qb  the ultimate bearing capacity of the to act on the vertical plane beneath the footing edges.
foundation on the lower soft layer; D  surcharge The ultimate bearing capacity can be presented as:
height; B  width of the foundation; H  thickness (4)
of upper layer.
Meyerhof (1974) suggested that the value of Pp
2.2 Averaged strength parameter methods may be obtained from the expression:
In China, averaged strength parameter methods are
widely employed in the design standards of harbor (5)
engineering because of its simplicity. The method
provided empirical equations to determine: (1) The
(6)
average value of cohesion C; (2) the average value of

angle of internal fraction ; (3) the average value of
qu
qu
.D

.D

B Hard soil H
H B
B' Pp Pp
Soft soil
qb qb

Figure 1. Mechanism adopted in projected area methods. Figure 2. Mechanism adopted in punching shear methods.

436

Copyright 2005 Taylor & Francis Group plc, London, UK


Substituting (5) and (6) into (4) gives For square footing, adding shape factor Eq. (9)
becomes:

(7)
(8)

Where ks  coefficient of punching shear; 1 and


2  the unit weight of the upper layer and underlying (10)
layer respectively; c2  cohesion of the underlying
layer; kph  coefficient of passive earth press. The Where ksi  coefficient of punching shear of each
values of ks and  may be obtained from the charts layer; i  unit weight of each layer; i  angle of
presented by Meyerhof & Hanna (1980). internal friction of each layer; ci  value of cohesion;
Hi  thickness of each layer. The value of ksi can also
2.4 Extension of punching shear method to be obtained from the charts presented by Meyerhof &
multi-layer soils Hanna.
The punching shear model proposed by Meyerhof &
Hanna is based on well-established theory and pro-
vides a useful insight into the behavior of subsoil con- 3 EXPERIMENTAL INVESTIGATION
sisting of two-layer soils. However, in the case the
subsoil consisting of soils with more than two layers, 3.1 In-situ loading tests
the bearing capacity equation provided by Meyerhof & In order to investigate the bearing behavior of multi-
Hanna will not be available any more. layer subsoil and to verify the validity of the bearing
On the base of punching shear failure mechanism capacity solutions above, some in-situ loading tests
it is assumed that, at the ultimate load, soils mass of were conducted. The tests were processed in the area
upper layers is vertically pushed into the lower layer, of Tianjin New harbor where the natural subsoil is
shown in Figure 3, the bearing capacity of the foun- consisted of two layers. The upper layer is mainly
dation can be divided into two parts: the shearing consisted of hard clay with the thickness of 0.75 m;
resistance provided by the upper layers and vertical the lower layer is mainly consisted of soft clay. In
resistance provided by the lower layer which is simi- order to enhance the bearing capacity of the natural
lar to the bearing behaviors of piles. Then the ultimate subsoil and reduce the settlement of the foundation,
bearing capacity can be presented as: soils with high bearing capacity such as rubble and
sand were artificially paved on the natural soils,
shown as Figure 4. The strength parameters of each
layer are shown as Table.1. For these different subsoil
profiles, square loading plants with the width of
0.707 m, 1.0 m, 2.0 m and 3.0 m were used in the load-
ing tests respectively.
(9)

qu Loading

Jack
.D
Hard Clay 0.75m Sand 0.5m
Soft Clay Hard Clay 0.75m
Soil-1 Pp1 Pp1
Soft Clay
1 1 Loading tests on nature subsoil
Loading tests on artificial
Soil-2 three-layer subsoil
Pp2
2 2
Pp2 H
Soil-3
3
B 3
Pp3 Pp3 Rubble 0.5m
Sand 0.5m
Soil-4 Hard Clay 0.75m
Soft Clay
qb Loading tests on artificial four-layer subsoil

Figure 3. Mechanism adopted in the extended methods. Figure 4. In-situ loading tests on layered subsoil.

437

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 1. Strength parameters of each layer. Load applied on the footing P/kPa
0 20 40 60 80 100 120 140 160 180 200
Cohesion Friction angle Unit weight 0

Settlement of the footing s/cm


B=0.707m
Soil layers (kPa) () (kPa) 2 B=1.0m
B=2.0m
B=3.0m
4
Hard clay 7.0 19.5 19.0
Soft clay 5.0 10.0 19.0 6
Sand \ 34.0 18.0
Rubble \ 40.0 18.0 8
10
12
3.2 Experimental results
(a) Relationship between applied load and footing
The load-settlement relationships obtained from in- settlement for natural subsoil
situ tests are shown in Figure 5. Since local or punch-
Load applied on the footing P/kPa
ing shear failure may occur in the layered soils, there
0 50 100 150 200 250 300
is no obvious ultimate load in the load-settlement 0

Settlement of the footing s/cm


curve. Obviously, it is difficult to establish the ultim- B=0.707m
1 B=1.0m
ate loads of the foundations. Terzhagi (1943) defined B=2.0m
the ultimate load for these cases as the point where 2 B=3.0m
the load-settlement curve becomes relatively steep 3
and straight. Christiaens (1967) found, by plotting the 4
settlement against the load on a log-log scale, that the 5
diagram consisted of an upper curved part and a lower
6
part with a straight line. The intersection of these two
parts was considered as the rupture point. In this 7
investigation, Christiaens criteria were used to deter- 8
mine the failure points. The experimental ultimate (b) Relationship between applied load and footing
loads qu are given in Table 2. settlement for sand-paved subsoil
From the experimental results, it has been founded
that the ultimate bearing capacity increases with Load applied on the footing P/kPa
increasing thickness and strength of upper layers, 0 200 400 600 800 1000
0
decreases with increasing footing width. The artificial B=0.707m
Settlement of the footing s/cm

paved sand and rubble can largely reduce the settlement B=1.0m
2 B=2.0m
of the footing by approximately 50%80%. B=3.0m

4
3.3 Examination of the bearing capacity methods
For the purpose of examining the validity of each 6
bearing capacity theory introduced above, the predicted
bearing capacities from each method were compared 8
with experimental results, which were shown in
Figure 6.
The comparison shows that, generally, the calcu- 10
lated bearing capacity values using extended methods (c) Relationship between applied load and footing
and projected area method are very close to the experi- settlement for sand & rubble-paved subsoil
mental results, while the calculated bearing capacity
Figure 5. Load-settlement relationship of layered subsoil.
values using averaged strength parameters method
have considerable difference from the experimental
result. Moreover, the projected area method tends to
overestimate the bearing capacity of the foundations
and the punching shear method tends to underesti- error. In these cases, the failure mechanism of the
mate the value slightly. Since the strength parameters subsoil is much closer to that assumed by Meyerhof &
of the layers differ from each other sharply, obviously, Hannas punching shear methods and projected meth-
using the average strength parameters to calculating ods; therefore the predictions from the two methods are
the ultimate bearing capacity must lead to considerable close to the experimental results.

438

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 2. Ultimate bearing capacity of the foundations. 4 NUMERICAL ANALYSIS
Ultimate bearing capacity (kPa) 4.1 Nonlinear elastoplastic constitutive model
Width of Natural Sand-paved Sand & rubble-paved In numerical analysis; modified Drucker-Prager model
footing (m) subsoil subsoil subsoil was adopted to describe the constitutive relationship
of the soils. The yielding surface in the p-t plane and
0.707 150 260 850 -plane is shown in Figure 7.
1.0 120 180 500 The yielding surface of the modified D-P model
2.0 80 100 240
can be written as:
3.0 65 80 135

(11)
ultimate bearing capacity qu/10kPa

24
Developed punching shear method
22 Projected area method
20 Averaged strength parameters method
Experimental result
18 (12)
16
14
12
10
8
6
1.0 1.5 2.0 0.5
2.5 3.0 (13)
Width of loading plane B/m
(a) Comparison for natural subsoil

80
(14)
ultimate bearing capacity qu/kPa

70 Developed punching shear method


Projected area method
60 Averaged strength parameters method
Experimental result
50 Substituting (12) and (14) into (11) gives:
40
30
20 (15)
10
0
1.0 1.5 2.0 0.5
2.5 3.0 Where p  pressure stress; q  Mises equivalent
Width of loading plane B/m stress, r  third invariant; s  stress invariant;  
(b) Comparison for sand-paved subsoil Lode angle;  the inclined angle of yielding sur-
100
face; d  cohesion of the material.
ultimate bearing capacity qu/kPa

90
Obviously, in the modified D-P model, the influ-
80
Developed punching shear method ence of the magnitude of the intermediate principal
70
Projected area method
Averaged strength parameters method
stress and the direction of the invariant  to yielding
60 Experimental method behavior has been taken into consideration.
50
For geotechnical material, the Mises equivalent
40 stress q may lead to plastic volumetric strain for clay
30 and sand. Therefore, the dilation angle  ranged from
20 0 to has been generally used to solve the problem.
10 The plastic potential G  t-ptan. For dilation angle
0   0, the material is no dilatational; and if   ,
0.5 1.0 1.5 2.0 2.5 3.0 the model is fully associated, in this case the volumetric
Width of loading plane B/m deformation of the material is to be often overesti-
(c) Comparison for sand & rubble paved subsoil mated (Zheng Yingren & Sheng Zhujiang 2002). In
this numerical investigation, the flow rule of the
Figure 6. Comparison of experimental results and predic- material is considered to be associated with the dila-
tion from theoretical methods. tion angle   0.5 .

439

Copyright 2005 Taylor & Francis Group plc, London, UK


S3
dpl
D-P model

t

Most-probable rupture surface


d 2 1

S1 S2 p
Modified Figure 8. Contour of equivalent plastic strain.
D-P model

Figure 7. Yielding surface in p-t plane and -plane.

4.2 Numerical model of layered subsoil


The bearing behavior and failure mechanism of the
foundations on natural subsoil, shown in Figure 89,
has been investigated using finite element methods. 2

3
1

Since the footing-subsoil system is centre-symmetric


in the in-situ loading tests, quarter of the system was
analyzed. The width of the numerical model is 8B and Figure 9. Distribution of equivalent plastic strain.
the depth of the model is 5B, shown in Figure 10.
Considering the shear-lock of linear element, the
quadratic-order hexahedral element with 20 nodes Table 3. Soil parameters in numerical analysis.
was used to mesh the model. And the soil parameters
used in the numerical analysis are shown in Table 3. Soil layers Hard clay Soft clay
In the numerical analysis, the loading was specified
by increasing the total node force applied to the rigid Cohesion (kPa) 7.0 19.5
footing. The failure point was defined that at which the Friction angle () 19.5 10.0
finite element program terminate due to not converge United Weight (kPa) 19.0 19.0
(Zheng Yinren & Zhao Shangyi 2003). Consequently, Poissons ratio 0.3 0.35
the load applied to the footing at this point was the Elastic modulus (kPa) 5  103 2  103
ultimate bearing capacity of the foundation.

Table 4. Numerical analysis results.


4.3 Numerical analysis results
Width of the footing (m)
Under the definition given above, the ultimate bear-
ing capacity of the footing on natural subsoil with dif- Ultimate bearing
ferent width was obtained, and the numerical results capacity (kPa) 0.707 1.0 2.0 3.0
were compared with the experimental results, listed in
Table 4. It is shown that predicted bearing capacity Experimental results 150 120 80 65
from the numerical methods agrees well with the exper- Numerical results 140 110 86 67
Ratio of error 6.7% 8.3% 7.5% 3.1%
imental results. Hence, the reliability of the numerical
method is verified.
Moreover, from the numerical results the plastic
zone under the ultimate loads was found, and the most- 5 CONCLUSIONS
probable rupture surface of layered subsoil in 3-D space
could be drawn up as shown in Figure 8. A detailed study has been carried out on the bearing
It can be observed that the failure surface and plas- capacity behaviors of multi-layer subsoil. During the
tic zone in the upper layer is about to be vertical, study, both in-situ experimental investigation and non-
while in the lower layer, the plastic zone expand to a linear numerical investigation have been performed
depth of about 1.5 B. By comparing, it is shown that the carefully. The following conclusions were drawn:
failure mechanism found by numerical analysis agrees 1 The ultimate bearing capacity of layered subsoil
with that of Meyerhof & Hannas punching shear the- increases with increasing thickness and strength of
ory. And the reliability of bearing capacity prediction upper layers, decreases with increasing footing
through punching shear theory is verified again. width. The artificial paved sand and rubble can

440

Copyright 2005 Taylor & Francis Group plc, London, UK


largely reduce the settlement of the footing by Meyerhof G G, Hanna A M. 1980. Design charts for ultim-
approximately 50%80%. ate bearing capacity of foundations on sand overlying
2 The extended methods and projected area methods soft clay. Canadian Geotechnical Journal, 17:300303.
can well predict the ultimate bearing capacity of Michalowski R, Shi L. 1995. Bearing capacity of footings
over two-layer foundation soils. Journal of Geotechnical
multi-layer subsoil. Engineering, ASCE, 16:421428.
3 The 3-D elastoplastic finite methods presented in Okamura M, Takmura J, Kimura J. 1997. Centrifuge model
the paper can well describe the bearing behaviors tests on bearing capacity and deformation of sand layer
of layered subsoil. overlying clay. Soils and Foundations, 37:7388.
4 Verified by both in-situ loading tests and numer- Debeer E E. 1967. Proefndervindelijke Bijdrage Tot de
ical analysis, the punching shear theory can actually Studie Van Het Gransdraagvermogen Van Zand Onder
describe the failure mechanism of layered subsoil. Funderigen op Staal; Bepaling Van derr Vormfactor Sb,
Annales des Travaux Publics de Belgique, Brussels,
Belguim, 6:481506.
Reddy A S, Srinivasen R J. 1967. Bearing capacity of foot-
ings on layered clays. Journal of the Soil Mechanics and
REFERENCES Foundation Division, ASCE, 93:8389.
Sreenivasula, V. 1967. Bearing capacity of footings on two-
Baglioni V P, Chow G S, Endley S N. 1982. Jacking-up layer soils. Proceedings of GEOCN-india IGS
foundation stability in stratified soil profiles. Proc. 14th Conference on Geotechnical Engineering.
Offshore Technical Conference, 4: 363369. Stayanarayana B, Garg R K. 1980. Bearing capacity of foot-
Burd H J, Frydman S. 1997. Bearing capacity of plane-strain ing in two-layered cohesive friction soils. Journal of
footings on layered soils. Canadian Geotechnical Journal, Geotechnical Engineering Division, ASCE, 106(7):
34:241253. 819824.
Hanna A M. 1981. Experimental study on footings in Terzaghi K, Peck R B. 1948. Soil Mechanics in Engineering
layered soil. ASCE, 107(8):11131127. Practice. New York: Wiley.
Houlsby G T, Milligar G W E, Jewell R A, Burd H J. 1989. Terzaghi K. 1943. Theoretical Soil Mechanics, New York:
A new approach to the design of unpaved roads. Ground Wiley.
Engineering, 22(3):2529. Yuan Fanfan, Yan Shuwang, Shun Wanhe. 2001. Methods
Jacobsen M, Christensen K V, Sorensen C S. 1977. for estimating the ultimate bearing capacity of layered
Geennemlokning af tynde sandalg, Vag-och Vatten foundations. Journal of Hydraulic Engineering, 3:4145.
Byggaren, Stockholm, Sweden: 2325. Zheng Yinren, Sheng Zhujiang, Gong Xiaonan. 2002. The
Kraft L M, Helfrich S C. 1983. Bearing capacity of shallow Principles of Geotechnical Plastic Mechanics. China
foundation on sand over clay. Canadian Geotechnical Architecture and Building Press.
Journal, 20(1): 182185. Zheng Yinren, Zhao Shangyi. 2003. Numerical simulation
Meyerhof G G. 1974. Ultimate bearing capacity of footings on failure mechanism of rock slope by strength induction
on sand layer overlying clay. Canadian Geotechnical FEM. Chinese Journal of Rock mechanics and
Journal, 11(2):223229. Engineering, 22(12): 19431952.

441

Copyright 2005 Taylor & Francis Group plc, London, UK


Numerical study of shallow foundations on calcareous sand

N. Yamamoto, M.F. Randolph & I. Einav


The University of Western Australia, Perth, Australia

ABSTRACT: Results from finite element computations of the response of shallow vertically loaded circular
footings resting on calcareous sand under drained conditions are presented. The predictions are based on the use
of the MIT-S1 soil model (Pestana & Whittle 1999). A relatively minor adjustment to the model parameters has
enabled prediction of the quasi-linear bearing response of centrifuge model tests reported by Finnie (1993), and
different failure modes to be distinguished for compressible and incompressible granular materials. Finite elem-
ent analyses of rigid-based circular foundations have shown that the bearing stiffness (ratio of average applied
pressure to imposed settlement) decreased slightly with increasing foundation diameter, whereas in the cen-
trifuge results it was apparently insensitive to the diameter. However, the high compressibility of calcareous
sand confirmed a failure mechanism consisting of a compressional bulb of material beneath the foundation; this
was in contrast to comparative analyses undertaken with a dilatant siliceous sand, where a classical failure
mechanism was observed.

1 INTRODUCTION 2 MIT-S1 MODEL

Shallow foundation systems are used extensively in All numerical simulations in this paper were undertaken
offshore engineering, with dimensions ranging from using the soil model, MIT-S1. This was developed as
10 m or lower for small structures and jack-up rigs, a unified model for predicting the rate independent
to over 100 m for large gravity based structures. In behaviour of different types of soils ranging through
calcareous sediments, found extensively in tropical uncemented sands, silts and clays. Full details of the
regions, shallow foundations may undergo significant model may be found in the PhD thesis (Pestana 1994)
settlements at moderate load levels, due to the high and a subsequent journal paper (Pestana & Whittle
compressibility of the soil matrix. Centrifuge tests 1999). Briefly, the model incorporates additional curve
(Finnie 1993) showed an almost identical displace- fitting features to simulate the different mechanical
ment response for a given applied pressure, for foun- characteristics found in natural soils into a framework of
dation diameters in the range 3 to 10 m, together with critical state soil mechanics. The additional features of
a compressional failure mode, and the aim of the this model include: (1) a lemniscate shaped yield sur-
work presented here was to try and simulate those face with non-associated flow rule; (2) a unique Limit
results numerically. Compression Curve (LCC) to describe the isotropic
The constitutive model chosen is the rate independ- hardening of the yield surface; (3) a rotational harden-
ent elasto-plastic model, known as MIT-S1 (Pestana ing function to describe the evolution of the anisotropic
1994). This choice was motivated primarily because of stressstrain-strength properties; (4) small strain non-
the models capability to simulate the response of highly linearity in shear; (5) non-linear hysteretic response dur-
compressible calcareous sand. This study has compared ing unloadreload cycles; (6) critical state conditions
predicted response of circular foundations with results described by an isotropic failure criterion. These fea-
from physical model tests, and in particular explor- tures contribute to the ability of the model to predict
ing the effect of foundation size on the load-settlement many characteristics of soil behaviour such as the non-
response and mode of failure. For comparison, parallel linear compression curves and critical state lines on
analyses were undertaken for foundations on calcareous e-lnp plots, dilatancy behaviour of sands and the vari-
and incompressible siliceous sands. Note that our aim ation of peak friction angle as a function of stress level
was not to verify whether or not the MIT-S1 model is and density. Here, the model will be used to simulate
successful but to use it simply because it encompasses two extremes of sand, one with high compressibility and
the main soil features we wish to model. the other dilatant.

443

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 1. Index properties of siliceous and calcareous sands.

Siliceous Calcareous
Property sand sand

Mineralogy Quartz, Feldspar, Calcium


magnetite carbonate
Grain shape Subangular Skeletal
grain
Specific Gravity 2.65 2.72
Mean particle size (mm) 0.160.20 0.10.2
Coefficient of uniformity 1.31.7 1015
Max. void ratio 0.98 2.321.97
Min. void ratio 0.580.61 1.410.94

Figure 2. MIT-S1 predictions of undrained stress paths for


siliceous and calcareous sands.

significant mechanical differences between siliceous


and calcareous sands, in particular the strong dila-
tancy of the former and the greater compressibility of
the latter.
The MIT-S1 model requires 13 parameters for
freshly deposited sand, all of which may be derived
from relatively standard laboratory test results. Pestana
et al. (2002) proposed a procedure to select the material
parameters for the MIT-S1 model. Table 2 summarises
the MIT-S1 parameters of Toyoura sand presented by
Pestana et al. (2002) and Goodwyn sand deduced by the
authors. Unfortunately, the proposed procedure did not
Figure 1. MIT-S1 predictions of compression behaviours for allow good predictions of both the triaxial and cen-
siliceous and calcareous sands.
trifuge test data and small modifications were made for
this purpose. We should, however, emphasise that our
3 MATERIALS AND MODEL PARAMETERS main objective was not to validate the model, but to gain
as much understanding as possible on the foundation
This paper compares the foundation response for two response for different input parameters. Therefore, the
extreme sand types, siliceous and calcareous. Toyoura calcareous parameters in Table 2 are those that gave the
sand was chosen for the siliceous sand as it has been best fit to the physical model results, after some adjust-
used extensively by Japanese research groups and its ment of the compression parameters, despite the inher-
properties are well documented; it can be classified as a ent inconsistency in this approach.
poorly-graded uniform fine sand. Goodwyn sand was Significant differences in the parameters pref and cs
selected for the calcareous sand, partly because of the may be seen for the two sands. The smaller pref value for
existence of a laboratory test database (Ismail 2000, calcareous sand reflects the high compressibility (sub-
Sharma 2004), but also because it has direct applic- ject to the parameter selection procedure). High friction
ability to offshore design on the North West Shelf of angles are also characteristic of calcareous sands.
Western Australia (where the Goodwyn A platform is
located). The main properties of both sands are shown in
Table 1. 4 FINITE ELEMENT ANALYSIS
Typical compression responses for Toyoura and
Goodwyn sands are shown in Figure 1, while Figure 2 The MIT-S1 model has been implemented in the
compares undrained stress paths for both sands with ABAQUS (HKS 2002) finite element code. The analy-
similar relative density (63%). These results reveal ses simulate circular foundations of varying diameter

444

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 2. MIT-S1 model parameters for siliceous and calcareous sands.

Test type Symbol Physical meaning Siliceous sand Calcareous sand

Compression test c Compressibility at large stresses (LCC regime) 0.370 0.350


pref Reference stress at unity void ratio for the H-LCC 5500 kPa 2500 kPa
 First loading transition parameter 0.200 0.900
K0 compression test K0NC K0 in the LCC regime 0.490 0.490
0 Poissons ratio 0.233 0.180
 Parameter for non-linear  1.00 2.0
Shear test cs Critical state friction angle 31.0 39.6
mr Peak friction angle at unity void ratio 28.5 60.0
np Constant of peak friction angle 2.45 2.0
m Geometry of bounding surface 0.55 0.35
 Rate of evolution of anisotropy 50.0 50.0
Shear test with small Cb Small strain stiffness parameter 750 450
strain measurement s Small strain non-linearity parameter 2.50 3.0

that are subjected to vertical loading at the surface of


a saturated seabed. The analyses are two-dimensional
(axisymmetric), and are carried out under displacement
control with horizontal constraint, thus simulating per-
fectly rigid and rough-based footings. Fully drained and
homogeneous ground conditions are assumed, with the
initial overburden stress varying with depth. The initial
void ratio is calculated according to the MIT-S1 com-
pression model and thus varies (with the soil becoming
denser) with increasing depth.

5 MODEL PREDICTIONS FOR SHALLOW


FOUNDATIONS

5.1 Comparison with measurements


Figure 3 shows MIT-S1 predictions of the bearing
responses with different foundation sizes (310 m)
on Goodwyn calcareous sand, together with physical
model results, which are conducted on different sized Figure 3. Numerical predictions of physical model results
footings under the same centrifugal acceleration, by with different foundation diameters.
Finnie (1993). The predictions compare the measured
result for void ratio of 1.3 (dense) and effective unit
weight of 7.0 kN/m3. For convenience, the numerical consolidation component dominates the settlement of
predictions were most easily achieved by using a sin- large foundations, thus leading to increasing settle-
gle mesh, but artificially varying the magnitude of the ment with diameter, for a given applied pressure.
soil unit weight (similar to varying the acceleration
level in a centrifuge). The initial bearing stiffness pre-
5.2 Foundation size effect
dicted, and also that throughout the entire response,
decrease with increasing size of foundation, while the Figure 4 shows the bearing capacity factor, N as a
physical model responses were insensitive to the foun- function of diameter together with the results obtained
dation size. Randolph & Erbrich (2000) suggested that from the physical modelling. The bearing capacity fac-
the shear strain component dominates the settlement tor of surface foundations on cohesionless soils can be
of small foundations (D 10 m) and that it increases expressed as N  2q/D, where  is the soil effective
with decreasing foundation size. The absolute settle- unit weight ( 7.0 kN/m3) and D is the diameter. No
ment of small diameter foundations, for a given bear- ultimate capacity was observed in the bearing response
ing pressure, will then stay approximately constant, of the calcareous sand, and hence the values at a nor-
i.e. independent of diameter. They also argued that the malised displacement (/D) of 0.05 and 0.2 are plotted.

445

Copyright 2005 Taylor & Francis Group plc, London, UK


Figure 5. Load-settlement predictions on calcareous and
Figure 4. Foundation scale effect of bearing capacity fac-
siliceous sands.
tor, N.

Numerical results for large foundations (20 to 100 m) that with minor adjustment of parameters derived from
are also shown. triaxial data, the MIT-S1 model is capable of distin-
The reduction of N with increasing diameter, which guishing between the behaviour of compressible and
is a well known scale effect (e.g. De Beer 1965), can incompressible granular materials (see Table 2).
be observed for both calcareous and siliceous sands. Figures 6 and 7 compare the kinematic mechanisms
In fact, the bearing capacity of large foundations, from at collapse (or a displacement of 20% of the diameter)
a practical view-point relevant for offshore design, for foundations on calcareous and siliceous sands, in
decreases with increasing size of foundation. The postu- terms of the incremental displacement and the cumula-
lation of Randolph & Erbrich (2000) is thus verified tive volumetric strain contours. Figure 6 shows failure
numerically. However, Terzaghis bearing capacity equa- patterns typical for these types of sand, with a bulb
tion, which is adequate for design of shallow founda- of compressed material and a punching failure (Vesic
tions on siliceous sand is inappropriate for foundations 1975) for calcareous sand, and a classical rupture pat-
on calcareous sand, as the bearing response is virtually tern accompanied by surface heave for the siliceous
linear from the start, with no obvious yield point sand. In Figure 7, significant volumetric strains are
(Finnie & Randolph 1994). Finnie & Randolph mod- observed beneath the footing on calcareous sand. It is
elled this using a bearing modulus factor, M (ratio of found that appreciable volumetric compression occurs
gradient of the pressuredisplacement response to the over a depth of at least one footing diameter. By con-
unit effective weight of the soil), and noted that the mag- trast, very small volumetric changes can be seen for the
nitude of M was not dissimilar to Nq for relevant friction siliceous sand, with maximum volumetric strain under-
angles. However, for compressible sands the failure neath the footing of only 1.5%. These two observations
mechanism is totally different from that from which Nq imply that the compression properties do not affect sig-
values are derived. nificantly the foundation behaviour on siliceous sand, at
least for foundation diameters of up to 10 m, but they
do dominate the response for calcareous sand (with
5.3 Comparison of calcareous and
increasing effect for larger diameters).
siliceous sands
Analysis of different size foundations on siliceous
Figure 5 shows the bearing response of calcareous and sand led to a scale effect in much the same way as for
siliceous sands, with the latter based on Toyoura sand calcareous sand, as shown on Figure 4. Note that, for
parameters (Table 2). The initial void ratio is 0.8 small foundations (less than 20 m diameter) the bearing
(medium loose) and effective unit weight of soil is capacity factor, N, is similar for displacements of 0.05
8.0 kN/m3. A clear ultimate bearing capacity is observed and 0.2 times the diameter. This is because a clear ultim-
in the siliceous sand analysis (for a foundation diameter ate capacity is reached at small displacements, as
of 10 m), whereas a quasi-linear response with no peak shown in Figure 5. However, at large diameters, the
pressure is observed for the calcareous sand. It is clear curves for /D of 0.05 and 0.2 diverge, as the footing

446

Copyright 2005 Taylor & Francis Group plc, London, UK


Figure 6. Incremental displacements on (a) siliceous sand Figure 7. Volumetric strain contours on (a) siliceous sand
and (b) calcareous sand. and (b) calcareous sand.

response becomes more strain-hardening with no clear


limiting capacity. Ultimately, the foundation response
on siliceous sand becomes similar to that for calcar-
eous sand with the diameter increases. This reflects the
importance of stress level, relative to the point at which
volumetric stiffness decreases, as crushing starts to
become significant. It is worth noting that the values
of N for large foundations on siliceous and calcar-
eous sands are very similar. This is further evidence that
the compression component dominates the foundation
behaviour of relatively large foundations on both sands.
In such cases, the conventional bearing capacity equa-
tion is no longer applicable even for siliceous sand. It
would then be appropriate to adopt other methods, for
example as Finnie & Randolph (1994) proposed.
Another interesting observation of the different fail-
ure mode is the effect of foundation roughness. Figure 8
shows loaddisplacement curves of perfectly smooth-
and rough-based circular footing on siliceous and cal-
careous sands. The roughness is modelled numerically
by constraining, or releasing the horizontal freedom
of the nodes representing the footing-soil interface.
Significant differences of ultimate bearing capacity can Figure 8. Roughness effect on calcareous and siliceous sands.

447

Copyright 2005 Taylor & Francis Group plc, London, UK


be seen with different footing roughness on siliceous calcareous and siliceous sand. A number of effects have
sand. However, the bearing responses for calcareous been demonstrated:
sand are insensitive to the foundation roughness. This
1. Based on a minor modification to the model param-
difference is presumed to arise because of a change in
eters derived from triaxial test data, the MIT-S1
failure mode for siliceous sand (with outward sliding
model was found capable of predicting the quasi-lin-
wedges of soil beneath the smooth-based foundation),
ear response of settlement against bearing pressure
whereas for the calcareous sand the same bulb of com-
found from centrifuge model tests. The prediction
pressed material develops beneath the foundation.
of the model based on this set of parameters also
allowed distinguishing of the foundation responses
6 DISCUSSION on compressible and incompressible granular
materials.
In the finite element analysis, similar settlement for a 2. The effect of foundation size showed that in the finite
given applied pressure, independent of diameter was element analyses, the bearing stiffness (applied pres-
observed by Finnie (1993), at least for relatively small sure divided by settlement) decreases with increasing
foundations; this result was not replicated numerically, diameter of circular foundations, whereas the cen-
however, with the finite element analyses showing trifuge results showed apparent insensitivity to diam-
decreasing bearing stiffness with increasing diameter, eter for small footings. Significant scale effects on
particularly for foundations of a size relevant to large bearing capacity factors were also observed both
offshore structures. from computed and model test results.
The significant difference in the failure mode for 3. The effect of the compression characteristics of cal-
siliceous and calcareous sands was also demonstrated. It careous sand was examined by comparing results
is not fully clear which parameters in the MIT-S1 model for more dilatant siliceous sand. Interestingly, the
dominate the difference in the failure mechanism, kinematic features of strains, velocities and displace-
although the difference is clearly associated with com- ments are consistent with expected patterns for the
pression characteristics. It appears likely that the pref two sand types, ranging from localised compression
parameter, which determines the position of the LCC, beneath the foundation for the calcareous sand, to a
has a strong influence; the smaller the value of pref, the more classical plasticity failure with adjacent heave
more compressible the material becomes, leading to for siliceous sand.
the foundation response being dominated by compres-
It was found that the MIT-S1 model is capable of
sion of material below the foundation. Simplistically,
simulating a wide range of material behaviour through
the consolidation settlement component dominates
appropriate choice of material properties. This feature
the bearing response for any practical foundation size
has proved effective in evaluating shallow foundation
in highly compressible calcareous sands. On the other
behaviour for different types of sand. Moreover, useful
hand, the bearing response of siliceous sand is con-
information has been derived on aspects such as the
trolled by high shear stress ratios for small size founda-
effect of the foundation scale for two extreme mater-
tions, but these changes gradually as the foundation size
ials, calcareous and siliceous sand. Further studies are
increases. Figure 4, showing the similarity of N values
underway for other calcareous soils and intermediate
for large foundations on siliceous and calcareous sands,
materials such as silt.
supports this interpretation.
It must be emphasised that the conventional bearing
capacity equation does not appear to be applicable to
ACKNOWLEDGEMENTS
the design of shallow foundation on calcareous sand.
Terzaghis equation was originally formulated for foun-
The work presented in this paper forms part of the
dations on relatively incompressible (siliceous) sands.
research activities of the Centre for Offshore Foundation
However, for calcareous and other compressible sands a
Systems (COFS), established and supported under
simpler representation of the footing response is that it
the Australian Research Councils Research Centres
is simply yielding continuously, throughout its penetra-
Program. The authors are grateful to Professor Andrew
tion, crushing the underlying material with N virtually
Whittle for supplying the ABAQUS implementation of
zero and a constant bearing modulus factor (similar in
the MIT-S1 model.
magnitude to Nq) as suggested by Finnie and Randolph
(1994).
REFERENCES
7 CONCLUSIONS
De Beer, E. E. 1965. The scale effect on the phenomenon of
This paper has employed the MIT-S1 model for predict- progressive rupture in cohesionless soils. 6th ICSMFE, 2,
ing the bearing response of circular footings resting on Montreal, Balkema, Rotterdam: 1317.

448

Copyright 2005 Taylor & Francis Group plc, London, UK


Finnie, I. M. F. 1993. Performance of Shallow Foundations Pestana, J. M., Whittle, A. J. & Salvati, L. A. 2002. Evaluation
in Calcareous Soil. PhD thesis, University of Western of a constitutive model for clays and sands: Part 1 Sand
Australia, Australia behaviour. Int. J. Num. Anal. Meth. Geomech., 26(11):
Finnie, I. M. F. & Randolph, M. F. 1994. Bearing response 10971121.
of shallow foundations in uncemented calcareous soil. Randolph, M. F. & Erbrich, C. 2000. Design of shallow foun-
Centrifuge 94, 1, Singapore, Balkema, Rotterdam: dations for calcareous sediments. Engineering for cal-
535540. careous sediments, Vol. 2, Bahrain: Balkema, Rotterdam:
HKS 2002. ABAQUS version 6.3-1, Hibbitt, Karlsson and 361378.
Sorensen, Inc.. Sharma, S. S. 2004. Characterisation of Cyclic Behaviour of
Ismail, M. A. 2000. Strength and Deformation Behaviour Calcite Cemented Calcareous Soils. PhD thesis, University
of Calcite-Cemented Calcareous Soils. PhD thesis, of Western Australia. Australia.
University of Western Australia. Australia. Vesic, A. S. 1975. Bearing capacity of shallow foundations.
Pestana, J. M. 1994. A Unified Constitutive Model for Foundation Engineering Handbook. H. F. Winterkorn et al.
Clays and Sands. ScD thesis, Massachusetts Institute of (eds) New York, Van Nostrand Reinhold: 121147.
Technology, USA.
Pestana, J. M. & Whittle, A. J. 1999. Formation of a unified
constitutive model for clays and sands. Int. J. Num. Anal.
Meth. Geomech., 23(12): 12151243.

449

Copyright 2005 Taylor & Francis Group plc, London, UK


Undrained bearing capacity of shallow foundations on structured soils

D.S. Liyanapathirana
University of Wollongong, NSW, Australia

J.P. Carter
University of Sydney, NSW, Australia

ABSTRACT: This paper examines the undrained bearing capacity of shallow circular foundations on struc-
tured soil deposits. Guidelines are given to identify the importance of structural features of the soil when assess-
ing its bearing resistance. Results obtained using a finite element model have been compared with those from
existing bearing capacity formulae based largely on plasticity theory. A new bearing capacity equation has been
proposed incorporating critical state soil parameters and additional parameters that quantify the effects of soil
structure on its mechanical behaviour.

1 INTRODUCTION Geotechnical engineers have long experience in the


use of a factor of safety in design, and generally they
Many structures are founded on shallow footings which have been successful in designing shallow founda-
bear directly on either natural or man-made soils. tions on natural and man made soils using this
Usually these soils in situ have a structure and behave approach. With this method, the effects of soil structure
differently from the same material in a reconstituted on its bearing behaviour are generally ignored.
state (e.g. Burland 1990, Cuccovillo & Coop 1999). However, for special cases, such as offshore structures,
At a fundamental level, there have been useful advances where there is often less experience, incorporating the
in formulating constitutive models incorporating the complex behaviour of structured soils directly in bear-
influence of soil structure, such as those proposed by ing capacity predictions may be very important
Wheeler (1997), Rouainia & Muir Wood (2000), (Leroueil 2002). Therefore, the main objective of this
Kavvadas & Amorosi (2000), Liu & Carter (2002) paper is to examine the significance of soil structure
and Carter & Liu (2005). However, our understanding on the undrained bearing capacity of shallow founda-
of the influence of soil structure on the bearing cap- tions resting on the surface of structured soil deposits.
acity of shallow foundations is still relatively modest. A series of numerical simulations has been carried
One of the important features of the mechanical out to investigate the influence of soil structure on the
behaviour of structured soils is the occurrence of a load-displacement response of shallow foundations.
destructuring phase as these soils are loaded. During These simulations have been carried out by incorp-
this phase, the structure of the soil may be completely orating the Structured Cam Clay (SCC) model (Liu &
or partially lost and only a small change in stress state Carter 2002, Carter & Liu 2005), which is an exten-
may induce very large strains. Consequently, signifi- sion of the widely used Modified Cam Clay model
cant errors in prediction of foundation behaviour can developed originally by Roscoe & Burland (1968), into
arise if the influence of soil structure is not incorp- the finite element program AFENA (Carter & Balaam
orated into these predictions. 1995) developed at the University of Sydney. SCC is
Well-established formulae for determination of the a relatively simple elastoplastic model, which is fully
undrained bearing capacity of shallow circular foun- defined by relatively few parameters, each of which
dations have been proposed by Terzaghi & Peck (1967), has a clear physical meaning and can be conveniently
Salenon & Matar (1982), Kusakabe et al. (1986) and identified by standard soil mechanics tests.
Tani & Craig (1995). These equations take into account Based on these numerical simulations, guidelines
the foundation shape, size, depth of embedment and are provided to identify when the structural features
variation of soil properties with depth, but they do not of the soil would become important in assessing the
directly take into account the influence of soil structure undrained (short-term) bearing capacity of shallow
on the bearing resistance. circular foundations. The factor, Nc, used in the

451

Copyright 2005 Taylor & Francis Group plc, London, UK


classical bearing capacity theory has been improved Table 1. Properties of stiff and soft clay.
by incorporating directly the structural features of the
soil to quantify the undrained bearing capacity of Property Stiff clay Soft clay
shallow foundations on structured soils.
* 0.161 0.22
* 0.033 0.022
2 STRUCTURED CAM CLAY MODEL M* 1.0 1.3
* 0.2 0.4
e*cs 2.75 2.86
The Structured Cam Clay model uses six parameters b 4.0 4.0
to define the soil structure in addition to the usual  0.5 0.5
parameters used to define destructured, reconstituted  1.0 1.0
soil behaviour in the Modified Cam Clay model  (kN/m3) 8.19 7.19
(Roscoe & Burland 1968). The additional parameters
are b, pco, , , a and c. The destructuring index, b, *  gradient of the normal compression line in e
ln(p)
quantifies the rate of destructuring with increasing space,
mean stress and pco defines the size of the initial yield *  gradient of the unloading and reloading line in
e
ln(p) space,
surface. The two model parameters  and  define
M*  gradient of the CSL in q
p space,
respectively the influence of soil structure on the plas- *  Poissons ratio,
tic potential of the soil and the effect of shearing on e*cs  void ratio at p  1 kPa on the CSL in e
ln(p)
destructuring. The latter is directly proportional to the space.
value of . The two parameters a and c are used to
define the additional voids ratio, ei, sustained by the govern the structural features of the soil for typical
soil, which is given by: examples of both stiff and soft clays. The values of
other parameters were kept constant and they are the
same as those given in Table 1.
(1) For reconstituted soils loaded under both drained and
undrained conditions, it was found that for a unique
value of the combined parameter 
B/pco
*, the non-
where p
s is the size of the current yield surface. For
* plotted against
dimensional bearing pressure qav /pco
most natural soils, two of these parameters, pco and
the non-dimensional footing settlement /B was almost
ei, are linked, as explained in Section 4.
the same, irrespective of the individual values of each
The yield surface of the SCC model in p
q
variable. Here, qav is the average applied footing pres-
space is elliptical and passes through the origin,
sure,  is the footing settlement, B is the footing
similar to the Modified Cam Clay model, but non-
diameter and  is the effective unit weight of the soil.
associated plastic flow is assumed. In what follows,
For a particular soil, the size of the yield surface of
all properties of the destructured reconstituted soil
the structured soil, pco, is always greater than the size
are denoted by the superscript *.
*, due
of the yield surface of the reconstituted soil, pco
to the additional voids ratio sustained by the soil
3 FINITE ELEMENT MODEL structure, ei. The variation of pco depends on *, *
and ei, as illustrated in Figure 1, and is given by:
The finite element mesh used for the analysis con-
sisted of fifteen-noded cubic strain triangles, and a (2)
16-point Gaussian quadrature was performed. Accord-
ing to Sloan & Randolph (1982), this element is cap-
able of accurate computations in the fully plastic range Therefore, for a particular soil the degree of structure
during undrained loading in problems which involve with respect to the reconstituted soil can be defined
axial symmetry. In all analyses presented in this
* or ei/(*
*). It was found
by using either pco/pco
paper, it is assumed that the contact between the foot- that the influence of ,  and the destructuring index,
ing and the soil is rough. b, do not have a significant influence on the undrained
bearing capacity. Thus, in the parametric study the
influence of soil structure has been studied by varying
4 EFFECT OF SOIL STRUCTURE ON only ei/(*
*).
BEARING CAPACITY
4.1 Bearing capacity of stiff structured clay
To study the effect of soil structure on the bearing
capacity, a parametric study was carried out by varying Figure 2 illustrates the influence of the degree of soil
the footing diameter and the model parameters which structure on the undrained bearing response of a 2 m

452

Copyright 2005 Taylor & Francis Group plc, London, UK


gradient = 50
p'co1 p'co2 p'co3 40
Dei1

qu/p' *co
30
Dei2
20
e
Dei3 10
gradient =
0
c 0 5 10 15
'B/p' *co
Dei /(*-*) = 0.78 Dei /(*-*) = 1.07
p'*co ln(p') ei /(*-*) = 1.75 Dei /(*-*) = 2.16
Dei /(*-*) = 2.56
Figure 1. Variation of p
co with ei for a structured soil.
Figure 3. Variation of qu/pco* with ei/(*
*) for a 2 m
diameter circular footing on stiff clay.
300
ei /( *-*) = 2.56
250 significant influence on the bearing capacity of the
2.16 foundation. Based on the above observation, an equa-
200
tion for the bearing capacity calculation can be for-
qu (kPa)

150 mulated based on the critical state and structural


1.75
parameters of the soil, as shown below:
100
1.07
50 0.78 (3)
0.0
0
0 1 2 3 4 5 6
/B %
/

Figure 2. Variation of qu for different values of ei/(*


*)
for a 2 m diameter circular footing on stiff clay. The new equation takes into account the influence of
soil structure on the bearing capacity.
diameter circular footing. Clearly, the bearing cap-
4.2 Influence of degree of soil structure on
acity reaches an ultimate value during undrained load-
undrained shear strength
ing, in contrast to the approximate bilinear load-
displacement response observed during drained In the determination of the short term or undrained
loading of similar materials (Liyanapathirana et al. bearing capacity, the undrained shear strength of the
2003a, 2003b). This indicates that for these cases the soil, Su, is an important parameter. However, the
soil beneath the footing should fail in the general undrained shear strength is not an input parameter of
shear failure mode. the Structured Cam Clay model. For the Modified
With an increase in the degree of soil structure, the Cam Clay model, which simulates the constitutive
bearing capacity increases significantly. For example, behaviour of reconstituted soil, the undrained strength
when e i /(*
*) increases from 1.07 to 2.16, there can be derived from the following equation (e.g.
is a four-fold increase in the bearing capacity of the Zdravkovic & Potts 2003):
footing. If the undrained bearing capacity is calculated
based on the reconstituted soil properties, neglecting
the structural properties of the soil, the predicted bear-
ing capacity can be a very low value compared with the (4)
true theoretical load carrying capacity of the footing.
Figure 3 shows the non-dimensional ultimate bear-
ing capacity, qu/pco
*, obtained from the finite element
analysis for different values of degree of soil struc- where KoNC  1
sin , KoOC  KoNC OCRsin
and
ture, ei/(*
*), and B/pco
* for stiff clay. For the
range of B/pco
* considered in the parametric study,
it can be seen that the degree of soil structure has a

453

Copyright 2005 Taylor & Francis Group plc, London, UK


 is the Lode angle, OCR is the overconsolidation 0
ratio, and KoNC and KoOC are the coefficients of earth OCR = 1.5, Dei(**) = 0.78
pressure at rest for a normally consolidated and an OCR = 2.0, Dei(**) = 1.07
overconsolidated soil, respectively. If we assume a 2 OCR = 4.0, Dei(**) = 1.75
circular yield surface in the deviatoric plane, g() can OCR = 6.0, Dei(**) = 2.16
be replaced by M*/3. OCR = 9.0, Dei(**) = 2.56

Depth (m)
For a structured soil ei/(*
*) increases with 4
the degree of soil structure and pco is related to ei
according to Equation (1). The OCR of the soil is also
6
related to soil structure and according to Equation (4),
Su increases with OCR. Therefore, the influence of the
degree of soil structure can be incorporated into the 8
undrained shear strength of the soil via equation (4).

10
4.3 Bearing capacity of stiff clay based on 0 50 100 150 200 250
undrained shear strength Undrained Shear Strength, Su (kPa)
Figure 4 shows the variation of Su with depth in the
structured soil deposit for different values of Figure 4. Undrained shear strength profiles for stiff clay.
ei/(*
*) and the corresponding OCR. In each
case it is assumed that OCR is independent of depth.
The ultimate bearing capacity of a surface founda- 16
tion, qu, is normally expressed in the form: 14
12
(5) 10
qu/Suo

8
where sc is the shape factor, Suo is the undrained shear 6
strength of the soil at the ground surface, and Nc is the 4
bearing capacity factor. 2
Based on plasticity theory, several authors have 0
0 2 4 6 8
recommended values for Nc and sc (or their product)
kB/Suo
as functions of kB/Suo, where k is the gradient of Su
over the depth of the soil deposit, e.g. Davis & Booker OCR = 1.5 OCR = 2
OCR = 4 OCR = 6
(1973), Salenon & Martar (1982), Kusakabe et al. OCR = 9 Kusakabe et al. (1986)
(1986) and Tani & Craig (1995). Salencon and Matar (1982) Tani and Craig (1995)
In Figure 5 the bearing capacity obtained from Terzaghi and Peck (1967)
selected plasticity solutions and the finite element
analysis of the structured soil are compared. According Figure 5. Comparison of bearing capacity obtained from
to Terzaghi & Peck (1967), the value of Nc is 5.14 for the SCC model with solutions based on plasticity theory.
a soil of uniform strength. Three other recommended
values are also plotted in Figure 5. In these solutions,
qu/Suo increases with kB/Suo. However, according to intact undisturbed soil, i.e. the peak strength of the
Figure 5, in addition to kB/Suo, qu/Suo depends on the structured soil, then Equation (6) can be used to obtain
individual values of Suo or in other words on the OCR the ultimate bearing capacity. This may be more con-
of the soil. To sufficient accuracy this effect can be venient than using the alternative, Equation (3), which
incorporated into the bearing capacity calculation as requires direct knowledge of the structural parame-
follows: ters of the soil.

(6) 4.4 Bearing capacity of soft clay


Usually soft clay deposits have a crust above the ground
water table and are close to normally consolidated
below the water table. In this section the influence of
where and pa is atmospheric pres- a surface crust is studied for a shallow, 2 m diameter
circular foundation, and it is assumed that the water
sure (101.32 kPa). If a value of Suo is available for the table is 2 m below the ground surface. The undrained

454

Copyright 2005 Taylor & Francis Group plc, London, UK


0 200
Finite element
150 Terzaghi and Peck (Suo)
2

qu (kPa)
100 Terzaghi and Peck (Suave)

4 OCR = 1.0
Depth (m)

50
OCRo = 3.0
0
6 OCRo = 6.0 0 2 4 6 8 10
OCR
OCRo = 9.0

8 Figure 8. Bearing capacity for different OCR values at the


surface of the soft clay deposit.

10 significantly increases the bearing capacity of the


0 10 20 30 40 footing. When the OCR at the surface of the crust is
Undrained Shear Strength, Su (kPa) 6, the bearing capacity has nearly doubled compared
to the bearing capacity obtained without any crust.
Figure 6. Undrained shear strength profile for soft clay. Unlike the case of stiff clay, for these soft clays, the
bearing capacity does not reach an ultimate value but
continues to increase slowly with increasing footing
100 displacement.
OCRo = 9.0
For soft clay deposits, it is not possible to compute
80 qu using the methods proposed by Salenon & Matar
(1982), Kusakabe et al. (1986) and Tani & Craig (1995),
60 OCRo = 6.0 because in all these methods the bearing capacity is a
qu (kPa)

function of kB/Suo. For these soft clays, the distribu-


OCRo = 3.0 tion of Su is not a simple linear increase with depth, as
40
assumed in the derivation of those methods. Therefore,
20 the finite element results have been compared only
OCR = 1.0, No crust with the predictions of Terzaghi & Peck (1967) method.
Figure 8 shows the bearing capacity obtained from
0
0 1 2 3 4 5 6 the finite element analysis when the footing displace-
ment is 10% of the footing diameter, and from the
/B %
Terzaghi & Peck (1967) method. In the Terzaghi &
Figure 7. Variation of bearing capacity with and without Peck method (Equation 5), the bearing capacity has
surface crust (B  2.0 m). been calculated using the undrained shear strength at
the ground surface and the average undrained shear
strength over the depth interval equivalent to the
shear strength profiles for the four cases considered diameter of the footing. It is clear that where there is
are shown in Figure 6. In the case of OCR  1, there a crust with OCRo greater than about 3, the bearing
is no surface crust. In the other three cases, the OCR capacity equation proposed by Terzaghi & Peck (1967)
values at the ground surface are 3, 6 and 9, respect- significantly over predicts the load carrying capacity
ively, and each has a polynomial variation of strength of the foundation. However, if the OCRo at the surface
above the water table. Below the water table however, it is less than or equal to about 3, the bearing capacity
is assumed OCR  1 for all four cases. Usually in soft calculated using the Terzaghi & Peck equation matches
clays, this surface crust provides the necessary strength reasonably well with the finite element results.
.
to carry surface loading (e.g. Zdravkovic & Potts
2003). In each case, ei was allowed to vary through the
crust and values were calculated using Equation (1), 5 FAILURE MECHANISMS
assuming pco*  10 kPa. Below the surface crust ei
was assumed to have constant value of 0.1. Figures 9 (a) and (b) show, respectively, the incremental
Figure 7 shows the variation of average bearing pres- soil displacements for stiff and soft clay deposits when
sure with footing displacement for the four cases given the cumulative displacement of the footing is 7.5% of
in Figure 6. Clearly, the presence of the surface crust the footing diameter. The footing considered has a

455

Copyright 2005 Taylor & Francis Group plc, London, UK


they do not show a flow pattern in a radial shearing
zone. According to the load displacement response
shown in Figure 7, no visible collapse is observed and
a continuous increase in the vertical load is needed to
maintain the footing movement in the downward direc-
tion. Therefore, it can be concluded that the deform-
ation of the structured soft clay beneath the footing
occurs predominantly as local shear failure under
undrained conditions.

6 CONCLUSIONS

Bearing capacity of circular footings resting on struc-


tured soil has been studied using a finite element
model based on the Structured Cam Clay soil model.
Figure 9 (a). Incremental displacements at /B  7.5% in Both stiff and soft structured clays have been con-
stiff clay (OCR  6).
sidered. The parametric study carried out for stiff clay
shows that the influence of the soil structure signifi-
cantly increases the mobilised bearing resistance of
surface footings. An equation based on the critical state
soil parameters has been proposed to predict the
undrained bearing capacity of shallow foundations on
stiff clay.
Bearing capacity obtained from the Structured
Cam Clay model for the stiff clay has been compared
with published solutions. A modified bearing cap-
acity factor, Nc (or strictly Ncsc), has been introduced,
taking into account the complex behaviour of struc-
tured soil. Finally it is concluded that the failure of
footings on stiff structured clays occurs mostly as a
result of general shear failure but in soft structured
clays, failure occurs as a local shear failure.

Figure 9 (b). Incremental displacements at /B  7.5% in


soft clay (OCRo  6).
REFERENCES
diameter of 2 m . The stiff clay deposit considered has Burland, J.B. 1990. On the compressibility and shear
an OCR of 6 throughout the soil deposit and the soft strength of natural soils. Gotechnique, 40(3), 329378.
clay deposit considered has an OCR of 6 at the Carter, J.P. and Balaam, N.P. 1995. AFENA User Manual,
ground surface, which decays in a polynomial trend Version 6, University of Sydney.
to an OCR of 1 at 2 m below the ground surface. Carter, J.P. and Liu, M.D. 2005. Review of the structured
It can be seen that in stiff and soft clays, the failure cam clay model, Geo-Frontiers, Austin, Texas, Jan 2425,
2005.
mechanisms are not the same. In stiff clay, soil beneath Cuccovillo, T. and Coop, M.R. 1999. On the mechanics of
the centre of the footing moves predominantly in the structured sands. Gotechnique, 49(6), 741760.
vertical direction, but towards the outer edge of the Davis, E.H. and Booker, J.R. 1973. The effect of increasing
footing soil movement is predominantly in the radial strength with depth on the bearing capacity of clays.
direction and the soil heaves around the footing, simi- Gotechnique, 23(4), 551562.
lar to a general shear failure. The load displacement Kavvadas, M. and Amorosi, A. 2000. A constitutive model
response given in Figure 2 also confirms this, as the for structured soils. Gotechnique, 50(3), 263273.
bearing pressure increases with footing penetration Kusakabe, O., Suzuki, H. and Nakase, A. 1986. An upper
and reaches an ultimate value. bound calculation on bearing capacity of a circular foot-
ing on a non-homogeneous clay, Soils and Foundations,
In soft clay, the soil flow pattern shown in Figure 9 (b) 26(3), 143148.
is different to that observed for stiff clay. Soil flow Leroueil, S. 2002. Well known aspects of soil behaviour so
beneath the footing is predominantly in the vertical often neglected. Annual workshop of the Centre for
direction confined largely to a zone beneath the footing. Offshore Foundation Systems, University of Western
Although some vectors are at an angle to the vertical, Australia, July 2002.

456

Copyright 2005 Taylor & Francis Group plc, London, UK


Liu, M.D. and Carter, J.P. 2002. Structured Cam Clay model. Salenon, J. and Matar, M. 1982. Bearing capacity of axially
Canadian Geotechnical Journal, 39(5), 13131332. symmetrical shallow foundations. J. Mechanique Theorique
Liyanapathirana, D.S., Liu, M.D., Airey, D.W. and Carter, J.P. at Appliquee, 1(2), 237267.
2003a. Predicting the behaviour of foundations on Sloan, S.W. and Randolph, M.F. 1982. Numerical prediction
structured soils. XIIIth European Conference on Soil of collapse loads using finite element methods, Inter-
Mechanics and Geotechnical Engineering, Czech national Journal of Numerical and Analytical methods in
Republic, 2, 255260. Geomechanics, 6, 4776.
Liyanapathirana, D.S., Carter, J.P. and Airey, D.W. 2003b. Tani, K. and Craig, W.H. 1995. Bearing capacity of circular
Bearing response of shallow foundations on structured foundations on soft clay of strength increasing with depth,
soils. In Proceedings, International Conference on Soils and Foundations, 35(4), 2135.
Foundations: Innovations, Observations, Design and Terzaghi, K. and Peck, R.B. 1967. Soil Mechanics in Engi-
Practice, Dundee, 521530. neering Practice, John Wiley and Sons, NY.
Roscoe, K.H. and Burland, J.B. 1968. On the generalised Wheeler, S.J. 1997. A rotational hardening elasto-plastic
stressstrain behaviour of wet clay. Engineering Plasticity, model for clays. 14th International Conference on Soil
Cambridge University Press, Cambridge, 535609. Mechanics
. and Foundation Engineering, 1, 431434.
Rouainia, M. and Muir Wood, D. 2000. A kinematic harden- Zdravkovic, L. and Potts, D.M. 2003. Numerical study of the
ing model for natural clays with loss of structure. effect of preloading on undrained bearing capacity,
Geotechnique, 50(2), 153164. International Journal of Geomechanics, ASCE, 3(1), 110.

457

Copyright 2005 Taylor & Francis Group plc, London, UK


An investigation into the vertical bearing capacity of perforated mudmats

D.J. White
Cambridge University Engineering Department, UK

A.J. Maconochie
Technip Offshore UK Ltd, Aberdeen, UK

C.Y. Cheuk & M.D. Bolton


Cambridge University Engineering Department, UK

D. Joray & S.M. Springman


Eidgenssische Technische Hochschule, Zrich, Switzerland

ABSTRACT: The optimal design of a mudmat foundation for a seabed structure maximizes the ratio of ver-
tical compression to pullout capacity, which can be achieved by adding perforations. This paper describes an
investigation led by Technip in which solid and perforated mudmats were installed into soft clay. The installa-
tion load of the solid mudmats matched the theoretical bearing capacity. Perforations reduced the installation
load in proportion to the perforated area. In heterogeneous conditions, a further change in installation load arose
due to the shallower failure mechanism. Peak pullout resistance was governed by separation beneath the mud-
mat, and depended on the width of the grillage elements formed by the perforation.

1 INTRODUCTION

1.1 Mudmat foundations


This paper describes an investigation into the vertical
bearing capacity of mudmat foundations. The term
mudmat describes pre-fabricated shallow founda-
tions made from steel. Mudmats provide temporary
support to jacket structures prior to piling, and are used
to support seabed structures such as pipeline mani-
folds and well completions (Fig. 1).
Mudmat foundations for seabed structures are typ-
ically an integral part of the structure and are fabri-
cated from steel plates reinforced by a grillage. This
type of integral foundation is favoured for deep water
frontiers since installation of the subsea structure can
be completed in a single operation. This investigation
is concerned with the design of mudmat foundations
on soft clay soils as are found in many deep water
regions currently under development.
The load on a seabed structure comprises the sub- Figure 1. Subsea intervention valve protection structure
merged self-weight in addition to small horizontal with mudmat foundation (Fisher & Cathie 2003).
forces due to seabed currents and any loads applied
by the connected equipment, for example due to ther- Removal and reuse of the seabed structure is desir-
mal expansion of pipelines. In shallow waters, loads able. During decommissioning, the structure is removed
from trawlgear are also considered. A typical mudmat using a crane; pullout resistance should be reduced to
is 5 metres square in plan. ease this process.

459

Copyright 2005 Taylor & Francis Group plc, London, UK


Mudmat foundations are therefore required to sup- this paper, su0 is defined as the strength at the founda-
port primarily downward-acting loads and have min- tion level, (depth h below mudline) so su0 increases
imal pullout capacity. Skirts are rarely used, unless with penetration.
unusually high horizontal loads are present. The influence of vertical heterogeneity can be cap-
tured by defining N*c as a modified bearing capacity
factor to be applied to su0 (Equation 2). N*c therefore
1.2 Perforation exceeds the equivalent homogenous value due to the
One method of reducing the uplift capacity of a mud- added strength available at depth.
mat is to introduce perforations. Instead of a solid
cross-section, a perforated mudmat comprises a grill- (2)
age with steel plates covering some of the cells.
Perforations offer two further benefits. Firstly, the The modified N*c is a function of the dimensionless
weight of the structure (and cost of material) is reduced, vertical strength gradient kD/su0 (for circles of diam-
allowing easier handling by a smaller crane. Secondly, eter D) or kB/su0 (for strips of breadth B). Exact plas-
the structure is less influenced by hydrodynamic forces ticity solutions for N*c for strip and circular
during deployment, since waves and currents can pass foundations have been derived using numerical meth-
through the perforations. ods by Davis & Booker (1973) and Salenon & Matar
(1982) and have been replicated for this paper using
freely-available method of characteristics software
1.3 Vertical bearing capacity of mudmats (Martin 2003). The effect of dimensionless strength
1.3.1 Effect of shape heterogeneity on bearing capacity can be illustrated
Design methods for the vertical bearing capacity of a either by the variation in N*c or instead by the change
perforated mudmat on clay are based on fundamental in dimensionless effective depth, zeff /D (or zeff /B) for
solutions derived using classical plasticity. The short which the unmodified bearing capacity factor Nc
term bearing capacity of clay, qmax, (minus the sub- applies (Equation 1). Prior to the availability of mod-
merged weight of overburden at the embedded foun- ern plasticity solutions for N*c, it was more common
dation depth, 
h) is the undrained shear strength, su, to consider this effective depth at which an operative
times a bearing capacity factor Nc: soil strength is chosen, and applied to Equation 1.
Figure 2 shows the increase in N*c with heterogen-
(1) eity, and corresponding decrease in effective depth.
The shallower failure mechanism below a circular
The exact bearing capacity factor, Nc, of an infinite foundation leads to a 50% shallower effective depth.
strip footing on Tresca soil is the Prandtl solution Skempton (1951) suggested that if the strength did
(Nc,strip  2   5.14) whilst for a rough circle the not vary by more than 50% of the average value
exact solution is Nc,circle  6.05 (Cox et al. 1961). No within a depth of 2/3B, the mean value over this range
exact solution exists for a square footing, but finite
element limit analyses bracket the true solution between
5.52 and 6.22 (Salgado et al. 2004). Conventional
finite element analysis indicates a value of Nc,square/
Nc,strip  1.15 (Gourvenec et al. 2005).
In light of these recent solutions, the shape factor
Nc,square/Nc,strip  1.2 (Skempton 1951) applied by
many design codes to the Prandtl solution, leading to
Nc,square  6.17, is slightly optimistic. Instead, these
newer results for square foundations indicate that the
change in geometry from an infinite strip to a square
could increase the bearing capacity by only 15%.

1.3.2 Effect of vertical strength heterogeneity


Most soft seabed clays show increasing undrained
shear strength with depth from a value at the mudline
in the range 210 kPa. The strength usually increases
linearly; a gradient, k, of 12 kPa/m is typical
(Randolph 2004). For a square mudmat foundation of
width W  5 m, these values correspond to a dimen- Figure 2. Effect of strength heterogeneity on bearing
sionless strength gradient, kW/su0, of 0.55. Throughout capacity.

460

Copyright 2005 Taylor & Francis Group plc, London, UK


should be used in Equation 1. For linear vertical het- conducted in a large-scale test tank (plan dimensions
erogeneity, this approach corresponds to zeff /D  0.33 4500 1250 mm) located at the Schofield Centre,
for kD/su0 3. The lower bound solutions in Figure 2 Cambridge University Engineering Department. The
indicate that this assumption is appropriate for strip tank is equipped with bottom drainage and the lower
foundations in slightly heterogeneous conditions, but is 600 mm is filled with a gravel layer. Two layers of
at least two times deeper than is appropriate for circular geotextile were placed above the gravel to prevent
foundations, and presumably also for square mudmats. escape of the test soil.

1.3.3 Effect of embedment


The embedment, h, of a mudmat leads to an increase 2.2 The clay bed
in vertical bearing capacity because of a) the add- A bed of soft natural clay, sampled from the seabed
itional overburden stress acting at the foundation (depth 01 m) in 700 m of water in the Gulf of
level, h, (Equation 1) and b) the additional shearing Guinea, was prepared by vacuum consolidation from
above the foundation level required for penetration, slurry. The properties of the clay, as measured during
which increases the bearing capacity factor Nc. For the field investigation are shown in Table 1.
uniform clay, lower bound solutions for an embedded A partial vacuum was applied to drainage layers
circular footing by Martin (2001) indicate a 30% above and below the clay slurry. This created consoli-
increase in Nc for h/D  0.5 compared to a surface dation in a similar manner to an oedometer with two-
footing. These analyses use the smooth sides of the way drainage. Instead of applying an increase in total
foundation to support the overlying soil. However, stress, a decrease in pore pressure is imposed, whilst
large strain finite element analyses by Hossain et al. the total stress remains atmospheric. The depth of
(2005) indicate a similar increase in Nc for spudcans, clay after consolidation was 300 mm. A freeboard of
which have an open cavity above. water was maintained above the clay throughout the
test programme.
1.3.4 Effect of perforations A site investigation was conducted to establish the
If perforations are introduced to a mudmat founda- strength profile and compare the reconsolidated mois-
tion, the net area over which bearing resistance can be ture content to the field conditions. A laboratory T-bar
mobilized is reduced. Figure 2 illustrates a further (diameter 12 mm, length 100 mm, penetration rate
weakening effect of perforating a mudmat foundation. 2 mm/s) and a motorized vane penetrometer (diam-
The strength gradient range described earlier as typ- eter 33 mm, length 50 mm, rotation speed 13/s) were
ical for mudmat foundations (kW/su0  15) spans a used to measure strength. Moisture content was meas-
doubling of N*c (assuming the circular solutions are ured from tube samples.
permissible for a square). However, if perforations are
introduced to a mudmat, the foundation may respond
as a grillage of independent strip foundations of width 2.3 The experimental programme
B, where B is only a fraction of W. If a 5 m square
Square model mudmats (width, W  125 mm) were
mudmat on the typical soil profile described earlier is
fabricated from aluminium alloy plate of thickness
perforated into 3 strips of width 1 m, the dimension-
4.8 mm and perforated with square or circular holes
less strength gradient, kB/su0 is 0.251. Within this
(Table 2, Fig. 3). The perforated mudmats are charac-
range, N*c (for a strip) reaches only 10% more than the
terised by the width (or diameter) of perforation, P,
homogenous solution for a circle. Therefore, the doub-
the resulting effective strip width, B*, defined as the
ling in capacity of a solid mudmat created by hetero-
width of the resulting grillage, and the perforation
geneity is lost.
ratio, R, which is the perforated fraction of the gross
For design, this detrimental reduction in downward
vertical bearing capacity created by perforation must
be weighed against the beneficial reduction in pullout
resistance. This paper describes an experimental Table 1. Soil properties: soft seabed clay from Gulf of
investigation into the influence of perforation on ver- Guinea.
tical compression and pullout capacity.
Soil property Value

In situ moisture content 150200%


2 METHODOLOGY In situ bulk unit weight 12.5 kN/m3
Peak vane shear strength 13 kPa
2.1 The test tank Liquid limit 175%
Plastic limit 80%
The investigation was commissioned by Technip to Plastic index 95%
aid the design of perforated mudmats, and was

461

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 2. Test programme and mudmat dimensions.

Width Perforation Effective strip Perforation


Test W, mm P, mm B*, mm ratio R

1 125 0
2 125 25 50 0.04
3 125 50 37.5 0.16
4 125 75 25 0.36
5 125 25 25 0.16
6 125 25 12.5 0.36
7 125 12.5 15 0.16
8 125 18.75 10 0.36
9 100 0
10 75 0
11* 125 84.6 66.3 0.36
12 125 0
13 125 0
14 200 0
15* 200 135.4 106.1 0.36 Figure 4. Results of test bed characterisation.

* Circular perforation. B* is mean of mid-side & corner


values. 3.2 T-bar and vane shear tests

Installation and extraction at slower rate: 0.2 mm/s.
Twelve T-bar tests were conducted around the tank
prior to testing. The penetration and extraction resist-
ance has been converted to su using a T-bar factor of
10.5 (Randolph & Houlsby 1984).
The scatter between the twelve T-bar profiles indi-
cates some lateral heterogeneity (Fig. 4). Backanalysis
of the mudmat tests has been carried out using the
mean of the two nearest T-bar profiles.
The top 50 mm shows a strong crust, with the T-bar
strength reaching 0.7 kPa, before returning to 0.5 kPa,
Figure 3. Mudmat shapes.
then increasing with depth. This strength nonlinearity
prevents simple correction of the measured bearing
capacity for vertical heterogeneity; no appropriate
value of kD/suo can be chosen.
area. Additional larger and smaller mudmats were Figure 4 also shows the mean T-bar extraction resist-
used to investigate the effect of size. ance. The ratio of penetration to extraction resistance
Each mudmat was installed vertically at a velocity decreases with depth from 2 to 1. The soil did not
of 2 mm/s (usually to a depth of 0.4 W), held at this posi- close around the back of the T-bar during the first few
tion for one hour, then extracted at 2 mm/s. The applied diameters of penetration. Therefore, the resistance
load and displacement were recorded throughout. ratio at shallow depths is may reflect the clay softening
A pore pressure transducer (PPT) was fitted in the when exposed to free water rather than a drop from
centre of the mudmat in test 13. peak to remoulded strength.
Thirteen vane shear tests were conducted around
the tank at 4 depths. The average back-analysed peak
3 RESULTS: TEST BED and residual vane strengths at each depth are shown
CHARACTERISATION on Figure 4. The peak vane strengths lie close to the
T-bar extraction strength, perhaps reflecting distur-
3.1 Moisture content bance during vane insertion (Randolph 2004).
Samples taken before and after the one month test
period were oven-dried to identify the moisture con-
4 RESULTS: MUDMAT FOUNDATIONS
tent profile (Fig. 4). The moisture contents reflected
the in situ values (Table 1), and were above the liquid
4.1 Installation of solid mudmats
limit. Swelling is evident over the test period, sug-
gesting that negative pore pressure from consolida- Tests 1 and 13 represent the base case of a solid square
tion was not equilibrated when testing started. mudmat and allow the repeatability of the testing to

462

Copyright 2005 Taylor & Francis Group plc, London, UK


Figure 6. Measured and predicted Nc for solid mudmats.

with increasing mudmat width (Fig. 6). This trend


could be attributed to progressive failure as the soil
softens from peak to residual strength along the
longer failure planes of the larger mudmats. However,
Figure 5. Installation and pullout resistance of solid mudmat. the small mudmats reach peak Nc at a higher embed-
ment, for which a higher value of Nc applies.
The experimental values of peak Nc and theoretical
be examined. A PPT mounted in the centre of the values accounting for embedment (Fig. 6) have a
mudmat during test 13 revealed the pore water pres- mean discrepancy of 4%, which shows no clear trend
sure at the interface. The penetration and extraction with mudmat size. The greatest uncertainty in this
resistance, q, is shown in Figure 5. Throughout this comparison is the back-calculated strength profile and
paper, q, is expressed as the applied vertical force, V, selection of an effective depth. The possible discrep-
divided by the gross mudmat area (W2), irrespective ancies due to the use of Nc,circle instead of Nc,square or
of any perforation. under-prediction of the overburden due to surface
The close agreement between the installation resist- heave are less significant.
ance during tests 1 and 13 indicates excellent repeat-
ability. The PPT used during test 13 follows the profile
4.2 Installation of perforated mudmats
of bearing pressure, although recording a lower value.
Since the effective stress is zero at the mudline, the PPT Without arching of the applied load across the perfor-
would equal the bearing pressure if the load was dis- ations, a linear decrease in peak Nc with perforation
tributed uniformly across the mudmat base. However, ratio, R (as defined in Section 2.3), would occur in
these lower values of pore pressure recorded at the uniform clay, where Nc is calculated based on the
centre of the mudmat reflect the theoretical elastic gross mudmat area, W2 (ignoring the minor effect of
distribution of load under a rigid footing, which is shape on Nc). However, the measured peak Nc values
concentrated around the edge. (calculated using su at 0.18 W) exceed this linear
Normalisation of q by su is hampered by the het- trend (Fig. 7a) due to the strength heterogeneity.
erogeneous strength profile. The strength at a depth The strong crust in the test tank led to a higher
of 0.18 W below the mudmat has been used to calculate operative strength during installation of the perfor-
Nc, after correction for the overburden (Equation 1). ated mudmats, since the failure mechanism was shal-
This value was chosen since it is the theoretical effect- lower. For this unusual strength profile, the higher
ive depth for circular footings in gently heteroge- effective su balanced the reduced solid area, leading
neous conditions (Fig. 2). The resulting profiles of Nc to peak Nc values in tests 3, 5 and 7 that are compar-
for tests 1 and 13 vary by only 10% between embed- able to the solid tests (based on su at 0.18 W).
ments of 10 and 50 mm, during which su0 varies from This influence of strength heterogeneity can be
0.38 to 0.65 kPa. Theoretically, Nc should increase by accounted for by approximating the effective depth
20% due to the increasing embedment, whilst con- below a perforated mudmat as 0.18B* instead of
currently the operative strength drops from peak to 0.18 W. The resulting values of peak Nc lie close to
remoulded as the soil softens. the linear reduction with R as predicted if arching and
The peak values of Nc for the five solid mudmat shape effects are ignored (Fig. 7b). The profile of
tests show a slight size effect of decreasing capacity decreasing Nc with depth is anchored at Nc  7 for

463

Copyright 2005 Taylor & Francis Group plc, London, UK


Figure 8. Peak pullout Nc.

Figure 7. Peak Nc vs. perforation ratio, R. mudmats did not reach the steady value evident in
Figure 5 prior to the crack opening and separation
being observed.
R  0, as this is the average for the solid 125 mm The peak Nc during pullout is calculated using an
mudmats. effective depth of 0.18B* (as previously for installa-
The linear variation in peak Nc shown on Figure 7b tion, Fig. 7b), but with the opposite sign used for the
leads to a mean over-prediction of the perforated peak overburden correction in Equation 1. This modification
Nc values by 8%. This discrepancy could be attributed arises because the overburden acts to reduce, rather
to the 1015% difference between Nc,square and Nc,strip, than enhance, the pullout resistance. These peak val-
since a strip failure mechanism is more appropriate ues of Nc are lower than recorded during installation
along a grillage of a highly perforated mudmat. (Fig. 8a). The lower Nc cannot be wholly due to the
However, this transition to a strip mechanism would remoulded (instead of peak) strength acting during
also cause an increase in effective depth (Fig. 2). An extraction; this effect would reduce the resistance by
additional modification to Nc would be difficult to a constant fraction, and Nc would remain proportional
verify due to the heterogeneous test bed. to R. Instead, peak pullout Nc vs. R shows wide scat-
ter and no trend (Fig. 8a).
The peak pullout resistance is limited by separ-
4.3 Extraction of mudmats ation. Therefore, it is appropriate to plot peak pullout
The uplift resistance of each mudmat showed a sudden Nc against the effective strip width, B*, which is
reduction from the maximum value to almost zero, equal to twice the distance that a crack must open in
coincident with separation of the clay from beneath order for full separation to occur (Fig. 8b). This com-
the mudmat (Fig. 5). The matching profiles of bearing parison shows less scatter than Figure 8a. The mud-
pressure and pore pressure during pullout show that the mats with thinner grillage sections (i.e. lower B*)
uplift resistance is sustained by negative pore pressure offer reduced pullout resistance.
at the mudmat-soil interface until a water-filled crack In prototype conditions, consolidation during the
opens beneath the mudmat, eliminating the excess life of the structure could increase the strength of the
pore pressure. The pullout resistance of the perforated underlying soil, raising the uplift resistance.

464

Copyright 2005 Taylor & Francis Group plc, London, UK


5 CONCLUSIONS REFERENCES

Small-scale tests to examine the vertical response of Cox A.D., Eason G. & Hopkins H.G. 1961. Axially symmet-
mudmats on soft seabed clay have been conducted. ric plastic deformation in soils. Proc. R. Soc. London
The installation resistance of solid mudmats agreed 254:145.
with theoretical bearing capacity solutions. Perforation Davis E.H. & Booker J.R. 1973. The effect of increasing
of a mudmat changed the response by: strength with depth on the bearing capacity of clays.
Gotechnique 23(4):551563.
reducing installation resistance in proportion to the Fisher R. & Cathie D. 2003. Optimisation of gravity based
perforated area, design for subsea applications. Proc. ICOF, Dundee.
283296.
reducing the depth of the failure mechanism, lead-
Hossain M.S., Hu Y., Randolph M.F. & White D.J. 2005.
ing to a change in the operative strength in hetero-
Cavity stability and bearing capacity of spudcans on clay.
geneous conditions, and In review.
reducing the peak pullout resistance, which was Gourvenec S.M., Randolph M.F. & Kingsnorth O. 2005.
governed by separation on the underside. This reduc- Undrained bearing capacity of square and rectangular
tion was more dependent on the effective width of footings on clay. In review.
each grillage element, B*, than the proportion of Martin C.M. 2001. Vertical bearing capacity of skirted cir-
perforated area. cular foundations on Tresca soil. Proc. 15th Int. Conf.
Soil Mech. & Geotech. Engng. Istanbul. 1:743746.
The theoretical change in bearing capacity arising Martin C.M. 2003. New software for rigorous bearing
from shape effects is small. Capacity enhancement capacity calculations. Proc. Int. Conf. on Fndns, Dundee.
due to arching over the perforations was minimal. 581592.
An optimal mudmat design maximizes the ratio of Randolph M.F. 2004. Characterisation of soft sediments for
vertical compression to uplift capacity. These results offshore applications. Keynote lecture, Proc. 2nd Int.
suggest that the optimal arrangement of perforation Conf. on Site Investigation, Porto. 209232.
on homogenous clay is a large number of small perfor- Randolph M.F. & Houlsby G.T. 1984. The limiting pressure
on a circular pile loaded laterally in cohesive soil.
ations; this arrangement minimizes B* (and hence Gotechnique 34(4):613623.
pullout resistance) for a given net area (and hence Salenon J. & Matar M. 1982. Bearing capacity of circular
vertical capacity). However, in vertically heterogen- shallow foundations. In Foundation Engineering ed.
ous conditions this may not be optimal since a G. Pilot 159168. Paris: Presses de lENPC.
reduced B* leads to a lower operational strength since Salgado R., Lyamin A.V., Sloan S.W. & Yu H.S. 2004. Two-
the failure mechanism is shallower. and three-dimensional bearing capacity of foundations in
clay. Gotechnique 54(5):297306.
Skempton A.W. 1951. The bearing capacity of clays. Proc.
ACKNOWLEDGEMENTS Building Research Congress 1:180189.

This investigation was funded by Technip. Technical


support was provided by Chris Collison.

465

Copyright 2005 Taylor & Francis Group plc, London, UK


Shallow foundations: combined loading

Copyright 2005 Taylor & Francis Group plc, London, UK


Yield loci for shallow foundations by swipe testing

G. Gottardi & L. Govoni


University of Bologna, Bologna, Italy

R. Butterfield
University of Southampton, Southampton, UK

ABSTRACT: It is now well established that the loaddisplacement response of a shallow foundation can be
modelled as a work-hardening system in a three-dimensional (V,M/B,H) load space. Nevertheless, establishing
the form of a representative set of yield-loci from physical experiments, using load-controlled tests, is a formid-
able task. The paper presents results from, and analytically justifies, a displacement controlled swipe testing
technique that generates a complete yield locus from a single test. Such tests are simple to simulate numerically
and results are shown which closely replicate the experimentally determined yield loci for surface footings.
Preliminary results are also presented from an extension of the numerical swipe test modelling to investigate
buried footings.

1 INTERACTION DIAGRAMS FOR


(1b)
PREDICTING LOAD CAPACITY OF
FOOTINGS
where: h  H/Vmax, m  M/BVmax, v  V/Vmax and
The shortcomings of conventional bearing capacity h0  Hmax/Vmax and m0  Mmax/BVmax.
analyses based on Brinch Hansens well-known equa- In these dimensionless equations Vmax is the verti-
tion (i.e the need for the fundamental centre-line cal, central-load capacity of the footing, which there-
load capacity to be modified by empirical multiplica- fore automatically takes account of its shape.
tion factors for each of its terms to allow for footing The values of (m0, h0), established from these tests
shape, load inclination, eccentricity and so on), led to a and from very many more sophisticated ones on dense
proposal (Butterfield & Ticof 1979; Butterfield 1980) sand (Ticof 1977, Gottardi 1992, Gottardi & Butterfield
that a better solution might be possible using the con- 1995), loose sand (Nova & Montrasio 1991) and a
cept of an interaction diagram a well-established Kaolin clay (Martin & Houlsby 2000), are consistently
procedure in structural engineering. about (0.09, 0.123).
The idea, applied initially to shallow foundations The form of the full three-dimensional surface
on dense sand, was that all load combinations leading to (sketched in Figure 3) was explored by Butterfield &
failure of such a foundation, when plotted in a three- Gottardi (1994). Longitudinal sections of this figure
dimensional load space (V,M/B,H) i.e. vertical, are parabolas and transverse sections geometrically
moment/footing-breadth and horizontal loads will similar, inclined ellipses.
define a failure surface for a specific soil-footing The equation to the complete surface is:
system (Butterfield 1980).
Early experiments, using footings with different (2)
aspect ratios, bearing on dense sand and brass-rod-
model material (Butterfield 1981), established that
both the (V,H) and (V,M/B) sections of this surface in which a 
0.22, corresponding to a 12.7 degree
were remarkably close to simple parabolas of the form, rotation of the ellipse.
It is interesting to note that essentially identical
parabolic and elliptical yield loci have been obtained
(Zdravkovic et al. 2002) from finite element modelling
(1a) of strip footings on sand.

469

Copyright 2005 Taylor & Francis Group plc, London, UK


2 DEVELOPMENT OF A WORK-HARDENING 0.20
MODEL H/V0
0.15
Since the failure-load points for footings on dense M2 A3
sand identified accurately the onset of a work-softening 0.10 D4 B9 B8 G2
D1
process, it was also hypothesised (Butterfield 1980, D3
D2
A5
1981) that all such points might lie on the outermost 0.05 B6
yield-surface of a work-hardening system and, fur-
thermore, that the yield-surfaces of the system might G1
0.00
be a family of geometrically similar figures with the 0.00 0.10 0.20 0.30 0.40 0.50 0.60 0.70 0.80 0.90 1.00
maximum previous vertical centreline-load serving V/V0
as the hardening parameter (i.e. that the yield surfaces
might be a nested set of surfaces described by suitably Figure 1. Experimental load-paths in load-controlled tests
scaled versions of equations (1) and (2)). (from Gottardi & Butterfield 1995).
The key ingredients of such a model are (Nova &
Montrasio 1991): (a) the nested family of yield-
surfaces; (b) the work-hardening process (e.g. by using
H H
the results of a vertical centreline-load test, w versus
V analogously to the role of an e-p curve in Critical
State soil models) and (c) a set of plastic potentials. A
successful model of this kind would enable plastic
displacement increments, vertical, rotational and hori-
zontal (w p, Bp, up), corresponding to load incre- V u
ments (V, M/B, H), to be predicted for any loading
programme applied to the footing. The total displace- V u
ment increments are then the sum of the plastic and
the associated elastic components (e.g. for the verti-
cal displacement, w  we  w p).

w w
3 EXPERIMENTAL DETERMINATION OF
YIELD-LOCI
Figure 2. Output from a horizontal swipe test on a footing.
3.1 Load controlled tests
Identification of yield-loci lying on the (V,M/B,H) volume change) triaxial test on a saturated soil sam-
yield surface is clearly a task of paramount import- ple, consolidated to p0, say, compressed axially to
ance in validating such models. failure and interpreted in a CSM framework and (b) a
One method of doing so, used by Gottardi (1992) horizontal-displacement test to failure of a footing
is to load the footing to V0, unload it to a lower verti- (preloaded to V0) under conditions of zero vertical
cal load followed by reloading to yield along a displacement and zero rotation.
selected load-path. For example, Figure 1 shows a Tan concluded that, if the elastic displacements in
number of his load-paths in the (V/V0H/V0) plane. both tests were essentially zero, the loadpaths gener-
This is clearly a very arduous process in which a ated in the tests would trace the associated yield-
sophisticated load-controlled test and a laboriously loci. He was, in fact, using the previously mentioned
prepared soil bed are needed in order to establish a analogy between specific volume versus p work-
single point on a yield-locus passing through the hardening in CSM and w versus V work-hardening
selected value of V0. for a centrally loaded footing. Figure 2 illustrates a
side-swipe test which generates an approximate
yield locus for a soil-footing system.
3.2 Displacement controlled tests
Martin & Houlsby (2000) extended the analogy to
Fortunately, there is an alternative method, whereby a encompass a rotation-controlled swipe test; one car-
single displacement-controlled test (a swipe test) ried out, from a specific V0 value, under conditions of
can generate a yield locus, embedded in the 3D yield- zero vertical and horizontal displacement. A general
surface corresponding to any value of V0. The idea of swipe test on a footing (Gottardi et al. 1999) then
a side-swipe test was first mooted by Tan (1990) becomes one in which no vertical displacement is
based on an analogy between (a) an undrained (zero allowed (w  w0  constant) but any specified (B, u)

470

Copyright 2005 Taylor & Francis Group plc, London, UK


1.2
H
q

1.0

M/2R
0.8

V
0.6
Swipe paths

Figure 3. Sketch of load paths followed in swipe tests 0.4


(from Gottardi et al. 1999).

0.2
displacement regime may be imposed throughout the
test (Figure 3).
A method of establishing a formally correct yield-
locus from any swipe test is presented in Section 4. 0.0
0.0 0.2 0.4 0.6 0.8 v 1.0

3.3 Test results Figure 4. Dimensionless parabolic sections of yield surface.


The validity of the swipe-testing process was investi-
gated by Gottardi et al. (1999) using a 100 mm diam-
hn
eter, rigid, circular footing on dense sand. 0.15
Best fit
The key results from this work were that, for foot-
GG03
ings preloaded to a vertical load V0: 0.10 GG04
(a) There was indeed a family of nested yield-surfaces GG05
GG06
of the same form as equations (1) and (2), with (V0, 0.05
GG28
H0) replacing (Vmax, Hmax) in (h, m, h0, m0). The
GG29
values of (m0, h0, a), obtained from a statistically 0.00 mn GG07
best-fitted cigar to the swipe-determined yield- -0.15 -0.10 -0.05 0.00 0.05 0.10 0.15 GG08
loci, were (0.09, 0.121,
0.223), values essentially GG10
0.05
identical to those that described the failure enve- GG12
lope. Figure 4 shows a normalised plot of a set of
parabolic yield-loci in the (v, q) plane. 010
.
These cover a range of swipe tests (e.g.
increasing u only; increasing  only and increas- 0.15
ing both u and  in a fixed ratio) for V0  4/5 Vmax.
The normalised parabolas have the equation q
Figure 5. Dimensionless elliptical section of yield surface
4v(1
v)  0, which is derived from equation (2) (from Gottardi et al. 1999).
(Butterfield and Gottardi 1996) by defining q as,

which had been loaded to V0 and partially


(3) unloaded before swiping to yield (see Section 5).
(c) The simple, one-degree-of-freedom, hardening
model could provide reasonable predictions of
(b) Alternatively, by dividing equation (2) through- plastic displacement increments.
out by [4v(1
v)]2 and defining (mn, hn)  (m, h)/ (d) The flow was approximately associated in
[4v(1
v)], all points on the cigar can be reduced (M/B,H) planes (i.e. the plastic potential and the
to a single, central ellipse (Gottardi et al. 1999). yield surface could be assumed identical), but this
Figure 5 shows their data plotted this way, was not so for load-paths along which V did not
together with the best-fit ellipse. Points inside the remain constant. A possible plastic potential for
ellipse belong to elastic reloading paths for footings such cases was later defined by Cassidy (1999).

471

Copyright 2005 Taylor & Francis Group plc, London, UK


4 JUSTIFICATION AND INTERPRETATION OF 1
SWIPE TESTING H locus 2
0.8
H2 locus 1 b2
If, in a classical work-hardening plasticity model of
0.6
the soil-footing system, hardening were controlled 'swipe' test
solely by V0, which in general might be a function of load-path
0.4 H1
the three plastic displacement components (w p, Bp, b1
up), then V0 would change according to: 0.2
V2 V1
0 V01 V02
(4) 0.6 0.7 0.8 V 0.9 1

In the model being considered hardening only occurs


w = w0 a1
as a result of increasing w p. Hence, to avoid harden-
a2
ing, and thereby, ensure no change of yield-locus, any
displacement path imposed on the footing must ensure
that w p  0 (i.e. that the vertical plastic displace- w
ment component does not change during the process).
Since, in a swipe test, the footing is driven along an
arbitrary, extensive horizontal and/or rotational dis- Figure 6. Correction procedure for a horizontal swipe test.
placement path this will necessarily impose plastic
deformations there can be no question of such a
process being one of simple elastic unloading, even 0.20
h=H/Vo Uncorrected
though the vertical load may decrease. Corrected for rig flexibility
The differential form of a yield surface f(V,M,H, Corrected for rig flexibility
0.15
V0) is, and soil elasticity

0.10

(5) 0.05

Since a perfect swipe test ensures V0  0 then


equation (5) requires the two vectors, 0.00
0.00 0.20 0.40 0.60 0.80 1.00
v=V/Vo
(6)
Figure 7. Swipe test output (w  constant) corrected to
provide a w p  constant locus.
to be orthogonal. The first of these is normal to f and,
therefore, the second, the load increment vector, is
locally tangential to f, i.e. the load path in any perfect compressibility of the soil-footing system and the
swipe test necessarily traces a curve embedded in the loading frame are known, for instance from a vertical
three-dimensional yield-surface corresponding to V0 unloadingreloading path.
(Butterfield & Gottardi 2003). The remaining source of error, which cannot be
Unfortunately, from an experimental point of view, corrected directly, is that the hardening parameter V0
it is very much simpler to impose the condition may depend on ( p, up) in addition to w p. Unless
w  0 rather than the formal requirement w p  0. such dependence is also small, as it appears to be
Therefore, due to the decrease in V, which occurs in a from the results of Gottardi et al. (1999), the concept
swipe test, the associated elastic vertical displace- of w p  0 swipe testing is clearly no longer valid.
ment recovery of the soil means that w p  0. A sim- The correction procedure for a w  0 horizontal
ilar error can develop in a practical swipe test due to swipe test is illustrated in Figure 6. At any stage in the
the elastic deformation of the loading frame. In their test, the load point b1 will lie on a yield locus, defined
experiments using dense sand, Gottardi et al. (1999) here by V01. In the hardening plane it will lie on an
found the latter to be much the more significant elastic unloading line at a1, with w  w0, say. A small
source of error. Both of these errors, if small, can be horizontal swipe increment generates a load path
corrected if reasonable values of the vertical elastic increment b1 to b2 and a2 will be the point in the Vw

472

Copyright 2005 Taylor & Francis Group plc, London, UK


plane corresponding to b2 (since w  0 throughout
the test). If the unloading line through a2 intersects w q
u
the virgin Vw curve at V02 then b2 lies on the yield
locus corresponding to V02. If all the yield loci are
geometrically similar then the succession of points
{(V1/V01,H1/V01), (V2/V02,H2/V02), } will trace
out the generic yield locus in a dimensionless (V,H)
plane. Figure 7 shows the output of a horizontal swipe
test corrected this way.

5 NUMERICAL MODELLING
w = applied constant vertical rate
Even with the improved efficiency of yieldcurve u = applied constant horizontal rate
determination provided by swipe testing, it is still = applied constant rotation
a formidable experimental task to verify a three-
dimensional yield surface in an adequate way, espe- Figure 8. Geometry of the finite difference numerical
cially when investigating embedded footings model.
(Bransby & Randolph 1999). One possible approach
is to use numerical modelling for this purpose.
In principle, swipe tests can be performed with
w p  constant and the problems of soil bed replica-
tion and rig stiffness eliminated. Constructing sets of
yield loci this way would be particularly convenient.
A preliminary numerical investigation, using a 2D
finite-difference code (ITASCA 2002), was carried
out to establish whether the geometrical form of the
curves already determined experimentally could be
replicated. The soil bed (18 m wide  8 m deep) was
modelled as a standard elastic, perfectly plastic, Mohr-
Coulomb frictional material (  30;   15) and the
footing (breadth B  2 m) as a rigid structural beam,
initially laying on top of it, with a fully rough inter-
face (Figure 8).
A buried footing was then modelled in two ways: Figure 9. Vertical load (V ) versus displacement (w) curves.
(1) a 1 m deep soil surcharge (i.e. 17.6 kPa) was added
before displacements start and (2) the footing was
physically buried at 1 m depth (D/B  0.5) and the H/Vo
bed depth increased to 9 m. 0.16
Horizontal and rotational swipe tests were per- 0.14
formed as follows. In all tests a vertical downward 0.12
movement sufficient to impose a specific V0, 80% of 0.1
Vmax, was applied; in some cases the footing was
0.08
unloaded to about V0/8 before swipe testing (pre- Surface footing
loaded footings). In all tests swiping was continued 0.06
Surface footing with surcharge
until a steady state, at which all loads remained con- 0.04
Embedded footing
stant, was achieved. 0.02
The vertical load test curves for the three cases 0
mentioned (Figure 9) show that: values of Vmax are 0 0.2 0.4 0.6 0.8 1 V/Vo
very close to the Terzaghi predicted values; there is
no work softening (since the M/C model does not Figure 10. Horizontal swipe test load paths.
allow it); the difference between the first loading and
unloadreload slopes is much smaller than those The load paths for the three horizontal swipe tests
measured in experiments. Since the M/C model has are shown in Figure 10 (each curve has been normal-
not generated w p  constant loci the following com- ized by the V0 value specific to each test). Results for
parisons relate only to the form of the swipe curves both normally loaded and preloaded footing tests
themselves. converge satisfactorily.

473

Copyright 2005 Taylor & Francis Group plc, London, UK


footing caused not only an approximate doubling of
its vertical load capacity (Figure 9), but also a small
expansion of its (V/V0, H/V0) yield envelope (from
about V0/8 to V0/7) and, in the case of the physically
buried footing, a small lateral shift towards the origin
of its peak load point. The (V/V0, M/BV0) envelope
also expanded, but its peak load point shifted slightly
away from the load origin.

6 FINAL REMARKS

There is now a considerable body of evidence that the


Figure 11. Rotational swipe test load paths. response of shallow, rigid foundations to either changes
(V, M/B, H) in the applied load or changes in the
associated displacements (w, B, u) can be mod-
250 elled, with reasonable accuracy, as a work-hardening
H (N) plastic system with a single hardening parameter V0
200 GG03 dependent only on the vertical plastic displacement
150 component w p.
The determination of yield-loci (defining in aggregate
100 a three-dimensional yield-surface for the foundation-
GG07 soil system) from experiments is a crucial part of the
50 validation of such models. A technique for doing this
efficiently, a displacement-controlled swipe test, has
0 been described and, when corrected for soil compress-
0 200 400 600 800 1000 1200 1400 1600
ibility, formally shown to generate the correct yield
V (N) locus, if soil hardening is governed solely by the plas-
tic component of the Vw curve for the system. The
200
M/B (N) results obtained replicated very closely those obtained
GG04
150 from vastly more laborious load-controlled tests.
GG08 A preliminary numerical investigation, using a
100 well established constitutive model, has been able to
reproduce the key features of experimental results for
50 surface footings with surprising accuracy. Encouraged
by these results, the analysis was extended to include
0 embedded footings (D/B  0.5). The dimensionless
0 200 400 600 800 1000 1200 1400 1600 load paths generated were again of parabolic form
V (N) with, as expected, peak values increasing to around
0.14 for both horizontal and rotational swipe tests with
Figure 12. Experimental data, horizontal and rotational w  constant. The analysis is currently being refined
swipe tests on surface footing (from Gottardi et al. 1999).
to perform perfect swipe tests throughout with
w p  constant, in order to provide yield loci directly.
In the rotational swipe tests a tensile bond strength Scale effects, if any, between model and prototype
of 10 kPa was included at the footing-soil interface. foundations have yet to be fully determined, although,
The bonded and unbonded swipe curves (not shown) since they will be reflected in the centreline-load
were almost identical. capacity of the system, the dimensionless equations
Figure 11 shows the output from these tests and it to the yield loci from model tests may still be applic-
is interesting to note that whereas the preloaded foot- able. Large-scale or centrifuge swipe testing may well
ing paths from rotational swiping proceed only in a be a practical proposition.
clockwise direction, those from horizontal swiping
move in either direction.
Figure 12 shows the Gottardi et al. (1999) horizontal REFERENCES
and rotational swipe data from their experiments, which
is clearly very similar to the computer model output. Bransby M.F., Randolph, M.F. 1999. The effect of embedment
The buried footing swipe tests also generated depth on the undrained respond skirted foundations to
essentially parabolic load paths. Embedment of the combined loading. Soils and foundations 39(4), 1933.

474

Copyright 2005 Taylor & Francis Group plc, London, UK


Butterfield R. 1980. A simple analysis of the load capacity condizioni di carico. Ph.D. Thesis, University of Padova,
of rigid footings on granular materials. Journe de Italy.
Gotechnique, ENTPE, Lyon, France, 128137. Gottardi G., Butterfield R. 1995. The displacement of a
Butterfield R. 1981. Another look at gravity platform foun- model rigid surface footing on dense sand under general
dations. SMFE in Offshore Technology, CISM, Udine, Italy. planar loading. Soils and Foundations, 35(3), 7182.
Butterfield R., Gottardi G. 1994. A complete three- Gottardi G., Houlsby G.T., Butterfield R. 1999. The plastic
dimensional failure envelope for shallow footings on response of circular footings on sand under general
sand. Gotechnique, 44(1), 181184. planar loading. Gotechnique, 49(4), 453469.
Butterfield R., Gottardi G. 1996. Simplified failure ITASCA Consulting Group (2002) FLAC: Fast Lagrangian
envelopes for shallow foundations on dense sand. Int. Analysis of Continua. Minneapolis, USA.
Jour. Offshore and Polar Engineering, 6(1), 5864. Martin C.M., Houlsby G.T. 2000. Combined loading of
Butterfield R., Gottardi G. 2003. Determination of yield spudacan foundations on clay: laboratory test. Gotech-
curves for shallow foundations by swipe testing. Proc. nique 50(4), 325338.
1st International symposium on shallow foundation, Nova R., Montrasio L. 1991. Settlements of shallow founda-
FONDSUP, Paris, France, 1, 111118. tions on sand. Gotechnique 41(2), 243256.
Butterfield R., Ticof J. 1979. The use of physical models in Tan K. 1990. Centrifuge and theoretical modelling of foot-
design (discussion). Proc.7th ECSMFE, Brighton, UK, 4, ings on sand. Ph.D. Thesis, University of Cambridge, UK.
259261. Ticof J. 1977. Surface footings on sand under general planar
Cassidy M.J. 1999. Non-linear analysis of jack-up structures loads. Ph.D. Thesis, University of Southampton, UK.
subjected to random waves. DPhil Thesis, University of Zdravkovic L., Ng P.M., Potts D.M. 2002. Bearing capacity
Oxford, UK. of surface footings on sand subjected to combined load-
Gottardi G. 1992. Modellazione del comportamento di ing. Proc. Int, Conf. on Num. Methods in Geotech. Eng.,
fondazioni superficiali su sabbia soggette a diverse NUMGE V, 323330.

475

Copyright 2005 Taylor & Francis Group plc, London, UK


Investigating 6 degree-of-freedom loading on shallow foundations

B.W. Byrne & G.T. Houlsby


Department of Engineering Science, University of Oxford, United Kingdom

ABSTRACT: Previous laboratory studies of the response of shallow foundations have only considered planar
loading. This paper describes the development of a loading device capable of applying general loading on model
shallow foundations. Loading involving all six degrees of freedom {vertical (V ), horizontal (H2, H3), torsion (Q)
and overturning moment (M2, M3)}, can be applied experimentally to the model foundations. Aspects of the design,
including the loading rig configuration, development of a six degree-of-freedom load cell, numerical control algo-
rithms and an accurate displacement measuring system are described. Finally results from initial experiments are
presented that provide evidence for the generalisation of existing work-hardening plasticity models from planar
loading to the general loading condition.

1 INTRODUCTION
2R
Reference position
1.1 Motivation
The response of shallow foundations subjected to gen-
eral loading is an important area of civil engineering,
particularly in the offshore industry, where foundations Current position
w M
must be designed for loadings due to harsh environmen- H
tal conditions. These conditions may lead to large ver-
tical (V ), horizontal (H) and moment (M ) loads on the u
foundations. Whilst earlier studies considered overall
stability, more recent studies have attempted to model V
the displacements, using model tests to calibrate work
hardening plasticity theories (Houlsby et al. 1999,
Martin & Houlsby 2000, 2001, Byrne & Houlsby 2001, Figure 1. Sign conventions for 3 degree-of-freedom load-
ing (Butterfield et al. 1997).
Cassidy et al. 2002, Houlsby & Cassidy 2002).
Recently, this work has focussed on suction cais-
son foundations (Byrne et al. 2002, Byrne & Houlsby
2003). With geometry rather like an upturned bucket,
1.2 Background theory
the caisson is simply installed by sucking the water out,
and thus forcing the skirts into the seabed. This type of Figure 1 shows a shallow foundation under three
foundation has potential applications in the developing degree-of-freedom loading as defined by Butterfield
offshore wind energy industry. In this application the et al. (1997). This problem has received much attention
loading consists of very high moment and horizontal over the past twenty years, and the load displacement
loads, but low vertical loads. This is a very different pat- behaviour of the foundation can be captured well by
tern of loading from that experienced by heavier struc- work-hardening plasticity theories (as shown by the
tures in the oil and gas sector. In addition, the wind and papers cited above). A key component of the plasticity
wave directions may not coincide, so the base shear and theories is the definition of a suitable yield surface.
moment are not in the same direction. Considerable Figure 2 shows the shape of a yield surface that has
uncertainty surrounds how these foundations may per- been defined experimentally, for shallow foundations
form under these loading conditions (Byrne & Houlsby under three degree-of-freedom loading. This shape can
2003). be expressed mathematically as equation 1.

477

Copyright 2005 Taylor & Francis Group plc, London, UK


2
H H2
2r
M2
M3

H3 3
M/2R
Q
V

1
V
Yield surface
Figure 3. 6dof loading on a circular foundation.
Figure 2. Yield surface for shallow foundations.

2 DESIGN OF A 6 D-O-F LOADING RIG

2.1 The loading system


(1) Previous experimental work at Oxford has used a three
degree-of-freedom (3dof) loading device designed by
where v  V/Vo, m  M/2RVo, h  H/Vo, ho is the Martin (1994). This planar loading device achieves ver-
normalised horizontal load capacity, mo is the nor- tical and horizontal motion by using a system of sliding
malised moment capacity, a is the eccentricity of the plates, and rotational movement by rotating the loading
arm relative to these plates. All motions are independ-
ellipse in the h:m plane, ent of each other, and are each driven by a stepper
motor these features are useful for implementing
and 1 and 2 are shaping parameters for the section in load and displacement control systems. However, this
the vertical load plane. Numerous studies have identi- type of system would become too cumbersome for six
fied the parameter values for the yield surface for a var- degree-of-freedom (6dof) motions, and so an alternative
iety of footing types and for different soils for example approach is required. Typically, in robotics applications,
see Houlsby et al. (1999) for shallow circular founda- the Stewart Platform (Stewart 1965) is considered to be
tions on sand, or Martin & Houlsby (2000) for spudcans the most elegant approach to achieving 6dof movement
on clay. A natural extension of the these theories is to of a platform. There are numerous applications of this
six degrees-of-freedom and Martin (1994) proposed an system in robotics, but the authors do not believe the
expression for this case: system has been used for the testing of civil engineering
structures, and in particular testing of foundations. The
arrangement described in this paper is a variant of the
Stewart platform, and similar arrangements are used,
for instance, in the automobile industry for dynamic
testing of vehicles.
The system uses six actuators which, at one end, are
connected to the loading platform, and at the other are
(2) connected to a stiff reaction frame. Provided that six
properly arranged actuators are used, and are pinned at
where h2  H2/Vo, h3  H3/Vo, m2  M2/2RVo, m3  both ends, then it is possible to achieve 6dof motion of
M3/2RVo and q  Q/2RVo. Figure 3 shows the definitions the loading platform by changing the lengths of the
of the loads from Butterfield et al. (1997). The displace- actuators in a co-ordinated fashion. By careful selection
ments work-conjugate to the loads (V, H2, H3, Q, M2, of the actuator geometry, it is possible to ensure that the
M3) are (w, u2, u3, , 2, 3). There has been no system- control problem is well-conditioned, so that calcula-
atic study of footing response to full six degree-of- tions proceed in a straightforward fashion.
freedom loading to verify the extension of the planar The disadvantage with the Stewart Platform is that
loading theories to the general case. In the following the the simple motions are not linearly or independently
development of a loading device capable of applying the related to the motion of any individual actuator, unlike
general loading is discussed, and some initial experi- the 3dof system of Martin (1994). Therefore, quite com-
mental results are presented that can be used to verify plex control routines are required to ensure that all actu-
equation 2. ators move in concert to achieve the desired motion.

478

Copyright 2005 Taylor & Francis Group plc, London, UK


Actuator Forward Platform
Lengths Kinematics Pose
A1 w
A2 u2
A3 u3
A4
A5 Inverse 2
A6 Kinematics 3

Figure 5. Calculation procedures used in computer program.


Figure 4. Photos of the 6dof loading rig including a close-
up of the small LVDT measurement system.
extending/retracting each actuator to its required length.
This calculation procedure is known as the inverse kin-
Figure 4 shows the loading rig as constructed, show- ematics problem and is a simple analytical calculation.
ing three actuators approximately vertical and three The opposite calculation, called the forward kinemat-
actuators approximately horizontal. This arrangement ics problem, is not so straightforward, and requires a
ensures that the problem is well conditioned, as the main numerical solution. If the lengths of each actuator are
motions can be directly related to the motions of a known, then it is possible to calculate the new pose of
sub-set of the actuators. For example, to achieve vertical the platform. Within the actuators are linear potentiom-
movement the three vertical actuators must move the eters that allow the user to determine the current length
same distance, while only a slight adjustment of the hori- of the actuator, and therefore the pose of the platform.
zontal actuators is required. Both inverse and forward kinematics procedures are
The actuators, supplied by Ultra Motion, are linear used within the software as shown in Figure 5.
actuators each powered by an Animatics Smart Motor. A typical test proceeds by determining the initial
This brushless DC servo-motor incorporates an inte- platform pose using the forward procedure. The user
grated control system featuring a motion controller, then specifies a sequence of moves in terms of platform
encoder and amplifier. The actuators have a maximum pose. These moves are broken into a series of small
extension of 200 mm and can move at rates of up to moves so that the non-linearity of motion of each actu-
5 mm/s. Commands to the actuators can specify relative ator can be captured. The inverse procedure is used to
motions, position, velocity or acceleration. The actu- calculate for each of the moves the required length
ators are daisy-chained together and commands can be of each actuator. A file of actuator lengths with time
sent to individual actuators and then executed simultan- (position-time data) is recorded. The relevant data from
eously with a global command. More importantly, a this file are sent to each actuator, and each movement
number of moves can be downloaded to on-board mem- is executed simultaneously. An on-board buffering
ory on the motors, and then executed according to a system allows moves to be downloaded to each actu-
synchronised clock system common to all actuators. ator. The actuators themselves use sophisticated control
This makes it possible to execute complicated platform processes to determine the velocity and accelerations
motions provided one can determine, in advance, a time required, so that the actuator reaches each position at
history of the individual actuator motions required. the time required, thereby ensuring a smooth motion.
While the moves are being performed the control
2.2 The control program program logs the data. In particular the actuator lengths
are recorded and the platform pose is calculated and
A program has been written in Visual Basic to control displayed.
the loading system. The program allows input of a
sequence of moves in terms of the motions (w, u2, u3,
2.3 The load cell
, 2, 3) of the platform, known as the pose. These
motions can be described in terms of a rotation and The load cell was constructed using a thin walled cylin-
translation matrix (i.e. a transformation matrix). This der of radius r  27.5 mm, wall thickness t  0.475 mm
matrix can be applied to the co-ordinates of the pinned and length 70 mm. It was fabricated from Aluminium
connections of the actuators with the loading platform alloy with a Youngs Modulus of 72 GPa and a shear
to produce a new set of co-ordinates for the platform in modulus of 27.1 GPa. The thin walled section was
its new position. If the co-ordinates of the other (fixed) machined from a larger block, leaving heavy end
ends of the actuators are known, then it is possible flanges. The transition from thin-walled section to
to determine the required lengths of each actuator. To flange was smoothed at an appropriate radius to min-
move the platform to the new position simply requires imise stress concentrations. A total of 32 strain gauges

479

Copyright 2005 Taylor & Francis Group plc, London, UK


2.4 Small LVDT system
One determination of the platform pose is by using the
linear potentiometers within the actuators. This, how-
ever, provides only a coarse measurement of the plat-
form pose. In particular there are issues of electrical
noise and rig stiffness which have a significant impact
on both the resolution and accuracy of this measure-
ment. To achieve a more accurate determination of the
foundation movement a system of small LVDTs (20 mm
range) are used. These are placed in a similar configur-
ation to the actuators, but supported on a separate frame
as shown in Figure 4. The program carries out the for-
ward kinematics calculation to determine the pose of
the platform, given the measured lengths of the LVDTs.
Figure 6. The 6 degree-of-freedom load cell.
This allows very fine resolution of the foundation move-
ment to the order of a few microns (Williams 2005).

1
C1
C2
0.8 C3
C4 3 EXPERIMENTAL RESULTS
Circuit Output (V)

C5
0.6 C6

Some preliminary experimental results on a 150 mm


0.4
diameter flat circular footing using only displacement
0.2 control are presented here. At the time of writing load
control routines were being developed and are antici-
0 pated to be implemented in the near future. The experi-
-0.2 ments were carried out on Leighton Buzzard 14/25
0 50 100 150 200 250 300 350 400 450 500 silica sand. This is a uniform sand with particle sizes
Vertical Load (N)
ranging from 0.6 mm to 1.18 m. The maximum and
Figure 7. Calibration curves for the loadcell under vertical minimum void ratios are 0.79 and 0.49 respectively. The
loading. sand was prepared in a loose state with a relative density
estimated as 20%. Fuller details of the experimental
work are reported by ap Gwilym (2004), Stiles (2004)
are fixed to the outer surface of the cylinder to measure and Williams (2005). The experiments were designed
the appropriate strains. Figure 6 shows the completed to determine the shape of the yield surface in the six
cell. The strain gauges were arranged in six Wheatstone dimensions. A number of swipe tests were performed
bridge circuits, each corresponding to the measurement with various combinations of translations and rotations
of a particular load component. Each circuit was fully at a constant vertical displacement. The swipe test
compensated for temperature. Eight gauges were used has been used extensively to determine the shape of
for the vertical and torque circuits, and four gauges for yield surfaces, see Martin (1994), Gottardi et al. (1999),
the moment and horizontal load circuits. The cell was Martin & Houlsby (2000), Byrne & Houlsby (2001).
calibrated by applying known loads and measuring the
output from all six circuits. By varying the loads one
at a time, it is possible to determine components of
the matrix X relating loads to voltages in the equation 3.1 Vertical loading
C  XF where C is the circuit output vector and F is Prior to carrying out the swipe tests it was necessary to
the load vector. Figure 7 shows the results from the six perform vertical loading tests, as these give information
circuits for changes in the vertical load. The slopes for for the hardening law. Five experiments are shown in
these six curves represent the components of the part of Figure 8, showing good repeatability of the results. Note
the matrix relating to vertical load (i.e. the first column that the measurement of the displacement is coarse, as in
of the matrix X). Inverting X produces a six by six cali- these experiments the small LVDT system was not used.
bration matrix that can be incorporated into the control
program, so that loads are calculated during the experi-
ment. Note that the design of the circuits is such that the
3.2 Swipe tests
off-diagonal terms are small. This is indicated in Figure
7 where only one circuit is responsive to the change in A number of swipe tests were performed to investigate
applied load. the suitability of equation 2. A typical experimental

480

Copyright 2005 Taylor & Francis Group plc, London, UK


600 80

Horizontal Loads, H2, H3 (N)


60
500
40
Vertical Load (N)

400
20
300 0
-20
200
-40
100
-60
0 -80
0 5 10 15 20 25 30 0 100 200 300 400 500 600
Vertical Displacement (mm) Vertical Load, V (N)

Figure 8. Typical vertical loading results. Figure 10. Horizontal swipe results.

80 0.14
H2
70 H3
Loads, H, M/2R, Q/2R (N)

M2/2R 0.12 Horizontal Swipes


60 M3/2R

H/Vo, M/2RVo, Q/Vo


Q/2R
50 0.1
40
0.08
30 Rotational Swipes
20 0.06
10
0.04
0
-10 0.02
Twisting Swipes
-20 0
0 100 200 300 400 500 600 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Vertical Load, V (N) V/Vo

Figure 9. A horizontal swipe result. Figure 11. Results normalised by Vo.

result for a swipe test is shown in Figure 9. In this test Table 1. Parameter values for work-hardening model.
the footing was displaced vertically to a pre-specified
This Gottardi Byrne &
distance at which point the vertical load reached approxi- Parameter study et al. 1999 Houlsby 2001
mately 530 N. At this load the footing was translated
horizontally. The figure shows that as the footing trans- ho 0.122 0.122 0.154
lates horizontally the relevant horizontal load traces a mo 0.077 0.090 0.094
path around a yield surface. In this particular test the qo 0.033 N/A N/A
translation was u2 so the only horizontal load developed 1 0.688 1.0 0.82
was H2. It is instructive to observe that the other load 2 0.709 1.0 0.82
components are all relatively unaffected by the transla- a
0.212
0.223
0.25
tion, as was expected. It is also possible to carry out tests
involving translations u2, u3,
u2 and
u3. The results
of these translations are shown in Figure 10 where the vertical load axis. It is clear that the results depend on
load paths for H2 and H3 are plotted. Note that each of the mode (i.e. translation/twisting/rotation) of the swipe
the tests starts at a different vertical load. However, it is test but not on the direction. Equation 2 can be fitted to
clear that the magnitudes and the shapes of the load the above results to give the parameter values in Table 1,
paths are similar for the different translations. This con- which are compared to data for footings on sand under
firms the expectation that similar load paths will be planar loading.
traced out regardless of the translation direction. Similar In determining these parameters it was also neces-
experiments were carried out for rotations and twists sary to use results for combined swipes, that is swipes
with the same results (i.e. the results were independent involving simultaneous rotation and translation and
of direction). other combinations of movements. For instance Figure
The data, such as shown in Figure 10, can be easily 12 shows the results from a test where a translation of u3
compared by normalising all the loads by Vo as sug- and rotation of
3 were applied simultaneously to the
gested in equation 2. Results are plotted in Figure 11 foundation. A number of these tests (twenty included in
for all possible pure horizontal, rotational and twisting the above analysis) were performed as they are neces-
swipes with the negative swipes reflected about the sary in determining the fit, and in particular determining

481

Copyright 2005 Taylor & Francis Group plc, London, UK


0.1 ACKNOWLEDGEMENTS
Normalised Loads, H/Vo, M/2RVo

0.08 H3
M3/2R
0.06 The authors acknowledge the funding from the
0.04
Lubbock Trustees (pilot project grant), the Royal
Society (equipment grant), the Department of
0.02
Engineering Science at Oxford University and EPSRC.
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 We acknowledge the work of Clive Baker in construct-
-0.02 ing the 6dof load cell and Chris Waddup who made
-0.04 the actuator frame and LVDT support frame. We also
-0.06 acknowledge the work carried out by final year under-
-0.08 graduate project students: Llywelyn ap Gwilym, Ed
V/Vo Stiles and Rachel Williams. The experimental work
described here was conducted by these students under
Figure 12. Non co-planar loading applied to the foundation. the direction of BWB.

0.08
0.07 M2/2R REFERENCES
M3/2R
0.06
0.05 ap Gwilym, T.Ll. ab E. 2004. Control of a six degree of free-
0.04 dom loading rig. Fourth year project report, Department
M/2RVo

0.03 of Engineering Science, University of Oxford.


0.02 Butterfield, R., Houlsby, G.T. and Gottardi, G. 1997.
0.01 Standardised sign conventions and notation for generally
0 loaded foundations. Gotechnique 47, N 5, pp 10511054;
-0.01 corrigendum Gotechnique 48, N 1, p 157.
-0.02 Byrne, B.W. and Houlsby, G.T. 2001. Observations of footing
-0.03 behaviour on loose carbonate sands. Gotechnique 51,
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
V/Vo N 5, pp 463466.
Byrne, B.W. and Houlsby, G.T. 2003. Foundations for off-
Figure 13. A swipe test where consecutive rotations are shore wind turbines. Phil. Trans. Roy. Soc. A 361, Dec.,
performed. pp 29092930.
Byrne, B.W., Houlsby, G.T., Martin, C.M. and Fish, P.M. 2002.
Suction caisson foundations for offshore wind turbines.
the parameter a which gives the rotation of the ellipse in Wind Engineering 26, No 3.
the h:m plane. The test shown in Figure 12 could not Cassidy, M.J., Byrne, B.W. and Houlsby, G.T. 2002. Modelling
have been performed using the previous 3dof loading the behaviour of a circular footing under combined load-
ing on loose carbonate sand. Gotechnique 52, N 10,
rig as it involves non co-planar loads. Equally Figure 13 pp 705712.
shows a test unique to the 6dof device in that during the Gottardi, G., Houlsby, G.T. and Butterfield, R. 1999. The
swipe test the footing was first rotated by 2 and then plastic response of circular footings on sand under gen-
rotated by 3. (i.e. orthogonal and consecutive rota- eral planar loading. Gotechnique 49, N 4, pp 453470.
tions). Initially under the rotation 2 the load path for M2 Houlsby, G.T. and Cassidy, M.J. 2002. A plasticity model for
tracks around a yield surface. When 2 stops and 3 the behaviour of footings on sand under combined load-
starts the response for M2 drops off and the response for ing. Gotechnique 52, N 2, pp 117129.
M3 picks up and eventually tracks around the same yield Martin, C.M. 1994. Physical and numerical modelling of off-
surface that M2 tracked. shore foundations under combined loads. DPhil Thesis,
University of Oxford.
Martin, C.M. and Houlsby, G.T. 2000. Combined loading
of spudcan foundations on clay: laboratory tests.
4 CONCLUSIONS Gotechnique 50, N 4, pp 325338.
Martin, C.M. and Houlsby, G.T. 2001. Combined loading
In this paper the description of a unique loading device of spudcan foundations on clay: numerical modelling.
capable of applying six degree-of-freedom motion to Gotechnique 51, N 8, pp 687700.
a model foundation is presented. The resulting loads Stewart, D. 1965. A platform with six degrees of freedom.
on the foundation are measured using a six degree-of- The Institution of Mechanical Engineers 180, N 15,
freedom load cell. A number of experiments, mainly pp 371384.
Stiles, E. 2004. Experiments using a six degree of freedom
displacement controlled swipe tests, are presented and loading rig. Fourth year project report, Department of
interpreted to provide verification of the extension of a Engineering Science, University of Oxford.
three degree-of-freedom plasticity model to six degrees- Williams, R. 2005. Six degree of freedom loading tests on
of-freedom. Further experimental work is required to sand and clay. Fourth year project report, Department of
verify the model fully. Engineering Science, University of Oxford, in preparation.

482

Copyright 2005 Taylor & Francis Group plc, London, UK


Single surface hardening model a system law to describe the
foundation-soil interaction

A. Kisse & K. Lesny


Institute of Soil Mechanics and Foundation Engineering, University of Duisburg-Essen, Germany

ABSTRACT: A new model is presented which describes the behaviour of shallow foundations over the whole
loading range up to the ultimate limit state. Hence, the separate analysis of different limit states is no longer
necessary. The model is based on the isotropic hardening concept of the plasticity theory and consists of two
components, a failure condition which describes the ultimate bearing capacity of the foundation and a dis-
placement rule which describes the load displacement behaviour from beginning of loading up to the failure of
the system. The theoretical concept of the model is presented and its application is shown using the example of
a small scale model test and a vertical breakwater.

1 INTRODUCTION uplift

A thorough understanding of the structure-soil-


F2 or
M3/b2 F2 or M3/b2
interaction is the basis for a safe and economical design.
In todays codes of practice, e.g. Eurocode 7 (2004) base failure
S
different ultimate limit states and serviceability limit or
= arctan (adm. e/b2) sliding or overturning
states have to be distinguished. For each limit state
it has to be ensured that the foundation of a structure,
e.g. a shallow foundation, does not fail to function in
accordance with defined partical safety factors or the
displacements and rotations are allowable. However, F1
the different limit states are inconsistent and strongly
correlated. Thus the structure-soil-interaction and Figure 1. Resultant ultimate bearing capacity of a founda-
therefore the safety of the structure can only be tion in the loading space (Lesny & Richwien 2002).
approximated.
The concept of the Single Surface Hardening Model,
which is described here, includes two components.
sliding, uplift and limitation of eccentricity) are inte-
The first component is a failure condition which
grated into a unique limit state equation. So for the
describes the ultimate limit state (ULS) of a shallow
design of foundations it needs to be checked only if
foundation without distinguishing different failure
the loadpath is located inside the failure surface or not
modes. The second component is a displacement rule
(Fig. 1).
which reflects the complete load-displacement rela-
Here, the failure condition for the basic case of
tion before the system reaches its ultimate limit state,
shallow foundations on non-cohesive soil without
thus integrating the serviceability limit state (SLS).
embedment is presented. Further cases are described
in Lesny (2001) and Lesny & Richwien (2002).
Generally, a single footing is loaded by a vertical
2 FAILURE CONDITION
load F1, horizontal load components F2 and F3, a tor-
sional moment M1 and bending moment components
In analogy to the concept of constitutive laws of plas-
M2 and M3 (Fig. 2). The load components are sum-
ticity the failure condition describes the ultimate
marized in the load vector:
bearing capacity of the foundation similar to a yield
condition. Hence, all the former isolated ultimate
limit states (for shallow foundations: base failure, (1)

483

Copyright 2005 Taylor & Francis Group plc, London, UK


Figure 2. Geometry and loading. Figure 3. Interaction F1/F10-F2/F10-M3/F10 in the F2/F10-
M3/F10-plane from small scale model tests (Lesny 2001).

In the basic case (c  0, d  0) the geometry of the


footing described by the side ratio b2/b3, weight ,
shear strength  of the soil and a quantity S describ-
ing the roughness of the footing base have to be
considered (Fig. 2).
With these input parameters the failure condition
of the general form

(2)

has been defined by the following expression:

Figure 4. Failure condition for general loading vs. failure


loads from small scale model tests (Lesny & Richwien
2002).
(3)
The parameter  controls the position of the max-
imum of the failure surface. Many experimental data
This formulation is an extension of the more basic show that the peak of horizontal or moment loading
form from Nova & Montrasio (1991). Their failure occurs at vertical loads F1/F10 0.5. So another
condition considers only the load combination of a value as the one postulated by Nova & Montrasio
horizontal load and a bending moment. (  0.95) is used here. The limit state uplift is
The quantity F10 represents the resistance of a foot- already included in Equation 3, because only positive
ing under pure vertical loading which can be calcu- vertical loads are admissible.
lated using traditional bearing resistance formula The failure condition (Eq. 3) defines a failure sur-
(e.g. pREN 1997-1 2004). face which is parabolic in the F1-Fi,i1 and F1-Mi-
The parameters ai govern the inclination of this planes and has, in opposite to other authors (e.g.
failure surface for small vertical loading where the Gottardi & Butterfield 1993), none rotated elliptical
limit states sliding and overturning have been relevant cross sections in the M2-F3 or M3-F2-planes. Since the
(see Fig. 1). These limit states are integrated by the experimental data from Perau (1995) shows only a
following formulations of the parameters ai and : small rotation of the yield surfaces for ratios of
F1/F10 0.14, a rotation of the yield surface is not
taken into account (Fig. 3).
The parameters have been derived from an analysis
of numerous small scale model tests. As an example
(4) Figure 4 shows the failure condition for general loading
compared to failure loads of some small scale model

484

Copyright 2005 Taylor & Francis Group plc, London, UK


tests. The curvature of the failure surface was deter- and a strain-hardening function H:
mined with failure loads from various tests with sim-
pler loadings, which ranges over the whole loading
area (Lesny 2001). (9)

3 DISPLACEMENT RULE The yield surface expands due to isotropic harden-


ing until the failure surface defined by the failure
The failure condition spreads out a failure surface condition is reached. Thus, the parameters ci and in
which represents the outer border of the loading (Fig. Equation 8 have to be determined as functions of ai
1). So the displacements ui and rotations  i of the and , respectively. The hardening law a is given by
foundation are caused by arbitrary loading inside the the following expression:
failure surface. They are described by the displace-
ment rule and summarized in a displacement vector:
(10)
(5)
In Equation 10 the parameter A describes the initial
Due to the complex interaction of load components, stiffness of the load displacement relationship.
displacements and rotations the displacement rule has Thus, the expansion of the yield surface is deter-
been formulated using the well-known strain harden- mined depending on the vertical displacement u1. The
ing plasticity theory with isotropic hardening (e.g. dependence on the horizontal displacements and the
Zienkiewicz 1988). Hence, displacements and rota- rotations as considered in the strain hardening expres-
tions are calculated according to Equation 6 where sion by Nova & Montrasio (1991), has been neglected.
elastic components are neglected. This additional complexity may not be necessary,
especially if a practicable formulation of the model
should be maintained. This opinion is represented also
(6) by Cassidy et al. (2000).
Many hardening laws (e.g. Houlsby & Cassidy
2002) require small scale model tests under centric
The components of the displacement rule are a yield vertical loading to determine the factor A. Since this
surface described by the yield condition F: is not convenient for practical applications, a pro-
cedure has to be developed to determine A from soil-
mechanical standard tests. A first approximation is
presented in Equation 11.

(11)

Here in max and min is the bulk density for the min-
imum and maximum density. The void ratio emin
(7) refers to the minimum density of the soil. The density
D relates the actual density of the soil in proportion to
the minimum and maximum density:
with the parameters ai and  of the failure condition,
a plastic potential G:
(12)

The exponent m is derived from an oedometer test


in which the initial density of the soil represents the
conditions in the field. It is assumed that the settle-
ment *  s/h0 depends linearly on the normal stress
 if depicted in doublelogarithmic load settlement
(8) curves. This condition has been proven to be valid by
Moussa (1961) on the basis of extensive experimental

485

Copyright 2005 Taylor & Francis Group plc, London, UK


1.4 represented best by a tensorial function (Lesny et al.
F1 [kN] with measured F10 2002):
1.2

1
(14)

0.8 Here, the stiffness matrix can be derived from the for-
mulation in Equation 6:
with theoretical F10
0.6

0.4
(15)
- experimental
0.2 + theoretical
u1 [mm]
So the matrix K defines the relation between the load-
0
0 1 2 3 4 5 6 7 8 ing and the corresponding deformations and rotations.
This matrix is an ordering matrix for the Single
Figure 5. Comparison of experimental results and theoret- Surface Hardening Model. The components kij of the
ical prediction of a small scale model test under centric ver- matrix show which displacements or rotation comp-
tical loading with dense Essen sand (D  0.78). onents are generated by the applied load. Results of
small scale model tests with eccentric vertical loading
F1 have shown that only vertical displacements u1 and
investigations on sand samples. The value m is the rotations 2 are generated (see e.g. Nova & Montrasio
gradient of the double logarithmic load-settlement 1991 or Bay-Gress 2001). In this case, Equation 14
curve: can be simplified to:

(13)
(16)
Therefore it does not depend on the pressure level.
However, it is influenced by the initial void ratio e0
and the kind of sand. The settlement resulting from the vertical load is
Figure 5 shows the experimental result for a small determined by K11 and the settlement resulting from
scale model test under centric vertical loading from the moment M2 is determined by K15.
Perau (1995) and a simulation of the same test by the Thus, it is possible to indicate the influence of the
hardening law according to Equation 10. The tests individual load component on the respective displace-
were carried out on Essen sand (min  14.7 kN/m3, ment or rotation.
max  17.7 kN/m3, emin  0.49 and emax  0.79) with The simulation of a small scale model test with such
quadratic steel plates without embedment. The theor- an eccentric vertical loading using the Single Surface
etical value of the initial stiffness A of the load dis- Hardening Model is shown in Figures 6 and 7. The
placement behaviour has been determined with the test was carried out on dense Essen sand (D  0.82)
procedure described above. with an eccentricity of the vertical load of e2  15 mm
The failure load F10 has been determined accord- and footing dimensions of b2  b3  90 mm.
ing to the traditional bearing resistance formula (e.g. In this case the initial stiffness parameter A has
pREN 1997-1 2004). been determined to A  0.61 according to Equation
Additionally, it has been adjusted to the failure 11. The other parameters are c3  0.8 and  0.65.
loads measured in the tests. These failure loads are As shown before, the foundation behaviour strongly
obviously larger than the theoretical values resulting depends on the value F10.
in a steeper load-displacement curve in comparison to The tests carried out by Perau (1995) show a dif-
theoretical results. ferent load displacement behaviour compared to
Altogether a good agreement of the theoretical other tests on sand (e.g. Nova & Montrasio 1991 or
curve with the experimental curves is shown. Bay-Gress 2001). The curves do not show a curva-
ture. Hence, in the range of the failure load the theor-
etical values deviate from measured values
4 MODELLING OF SMALL SCALE significantly (Fig. 6).
MODEL TESTS However, it can be shown that with the formulation
described here it is possible to simulate the load dis-
The complex load-displacement behaviour of the placement behaviour from the beginning of loading
foundations under various loading situations can be up to failure of the system.

486

Copyright 2005 Taylor & Francis Group plc, London, UK


0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 BC
0
F1 [kN] uplift
-1 wave crest
rotation failure

-2
wave trough Fh (t)
eC SWL
with measured F10 h* hC
-3
with theoretical F10 rubble mound FG
sliding along the base hs
-4
hr
- experimental
-5
+ theoretical subsoil
bearing resistance failure in
Fu (t) rubble mound
-6 bearing resistance failure in subsoil
u1 [mm]
-7 Figure 8. Vertical Breakwater, wave loading and failure
modes according traditional design.
Figure 6. Simulation of a small scale model test with
eccentric loading on dense Essen sand, F1
u1 curves.
Table 1. Important input parameters and their variation
range.
0.14
M3/B [mm] Parameter Variation range
0.12
with measured F10
Significant inshore wave 0.0 Hsi/hs  0.6
0.1 height Hsi/hs
Width of caisson BC/hs Design parameter
0.08 height of caisson hC/hs 0.8 h*  hC/hs  1.2 h*
with theoretical F10 Weight of caisson CC/w 1.4  CC/w  2.3
0.06 Eccentricity of caisson
0.2  eC/BC  0.2
dead load eC/BC
0.04
- experimental Friction angle of rubble mound r 30  r  45
+ theoretical Friction angle of subsoil s 25  s  45
0.02 Remarks
w3 * B [mm] hs: height of still water level, w: weight of water,
0 C: cross-section/(BChC).
0 2 4 6 8 10 12

Figure 7. Simulation of a small scale model test with


(Lesny et al. 2000). On the basis of an extensive
eccentric loading on dense Essen sand, M3/B
3B curves.
parameter study the interactions between the different
load cases and failure modes and the corresponding
5 EXAMPLE: VERTICAL CAISSON input parameters were analysed. The study was
BREAKWATER restricted to the ULS.
All input parameters like water and wave param-
In the following the application of the new system law eters, geometric parameters, unit weights and strength
is shown using the example of a vertical breakwater. parameters of the soil have been varied within ranges
The vertical breakwater considered here is placed that cover typical design situations. The variation
on a thin rubble mound on sandy subsoil. For the ranges of the most important design parameters are
design the loading under still-water level (SWL), summarized in Table 1.
wave crest and wave trough are considered (Fig. 8). In the design process the width of the caisson BC
According to todays design the isolated limit states of was determined so that none of the limit states in
uplift, rotation failure, sliding and bearing resistance Figure 1 was reached. The caisson dimensions were
failure in the rubble mound or in the subsoil have to limited to 0.5  BC,crit/hC  2.0. Table 2 shows the
be checked. Rotation failure, however, is often substi- number of cases in which a certain combination of
tuted by limiting the eccentricity of the resultant ver- load case and failure mode determine the required
tical loading to 1/3 of the foundation width. caisson width. Here, only load cases and failure
An investigation of the influence of these individ- modes which were generally relevant are specified.
ual failure modes on the design was accomplished in Other design situations were not critical at all (e.g.
the context of the EU-MAST III PROVERBS project loading at SWL, failure due to uplift) or restricted to

487

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 2. Results of the parameter study. cannot be neglected. Their importance depends on
certain combinations of the input parameters of Table 1.
Failure modes Although several important interactions are evi-
dent the results finally show that none of the failure
Bearing
modes clearly dominates over the whole range of the
resistance Sliding
failure in Limitation of along of the input parameters. The reason is that the failure modes
Load cases subsoil eccentricity caisson are inconsistent and strongly correlated by their input
parameters. Hence, splitting up the limit state of a
Wave crest 45.0% 15.0% 20.8% shallow foundation into single failure modes is artificial
(5,903,478) (2,224,422) (2,711,880) and cannot reflect the physical behaviour very well.
Wave trough 12.8% 4.02% 0.006% As a consequence the evaluation of the significant
(1,664,388) (525,384) (720) failure mode for a shallow foundation is difficult.
In contrast to the traditional limit state design con-
cept the Single Surface Hardening Model offers an
alternative formulation to the modelling of the behav-
iour of shallow foundations with an unique limit state
equation.
Figure 9 show the results of the proposed model
applied to the example breakwater of Figure 8. On
Figure 9(a) the failure condition and the loading in
the F1-F2-plane and in the F1-M2/BC-plane are shown.
Obviously, the stability of the breakwater is governed
by the high horizontal loading. Only an increase in the
vertical loading, i.e. of the caisson weight, would lead to
a sufficient safety. Figure 9(b) shows the vertical and
horizontal displacements of the breakwater depend-
ing on the corresponding load components F1 and F2.

6 CONCLUSION

A design model has been presented which describes


the complex behaviour of a shallow foundation under
loading up to failure. The two components of this
model, failure condition and corresponding displace-
ment rule, consider both, ULS and SLS. Hence, the
separate analysis of different limit states is no longer
necessary. The theoretical concept of this design
model has been provided. Its advantages have been
pointed out by simulating a small scale model test and
designing a vertical breakwater.

ACKNOWLEDGEMENT

The authors wish to thank the German National


Figure 9 Failure condition (a) and load-displacement- Research Council (DFG) for supporting this research
curves (b) for a vertical breakwater under various loading. work.

very special design situations (e.g. bearing capacity


failure in the rubble mound). The results show that REFERENCES
bearing resistance failure in the subsoil is the most Bay-Gress, C. 2000. Etude de linteraction sol structure,
important failure mode, it determines the caisson comportement non lineare sol fondation superficielle.
dimensions in over 50% of all investigated cases. These Universit Louis Pasteur, Stra burg (in french)
However, the limitation of the eccentricity substitut- Cassidy, M.J., Byrne, B.W. & Houlsby G.T. 2002. Modelling
ing rotation failure and sliding along the caisson base the behaviour of circular footings under combined loading

488

Copyright 2005 Taylor & Francis Group plc, London, UK


on loose carbonate sand. Geotechnique 52 (10): Lesny, K & Richwien, W. 2002. A Consistent Failure Model
705712 for Single Footings embedded in Sand. Proc. of the Int.
Gottardi, G. & Butterfield, R. 1993. On the Bearing Workshop on Foundation Design Codes and Soil
Capacity of Surface Footings on Sand Under General Investigation in View of International Harmonization and
Planar Loads. Soils and Foundations 33 (2): 6879 Performance Based Design, Kamakura, Tokyo, Japan:
Houlsby, G.T. & Cassidy, M.J. 2002. A plasticity model for 159165
the behaviour of footings on sand under combined load- Moussa, A. 1961. Die Zusammendrckbarkeit von Sand.
ing. Geotechnique 52 (2): 117129 Mitteilungen des Instituts fr Verkehrswasserbau,
Lesny, K. 2001. Entwickling eines konsistenten Versagens- Grundbau und Bodenmechanik der Rheinisch-
modells zum Nachweis der Standsicherheit flachgegrn- Westflischen Technischen Hochschule Aachen, Heft 23
deter Fundamente. Mitteilungen aus dem Fachgebiet (in German)
Grundbau und Bodenmechanik der Universitt Essen, Nova, R. & Montrasio, L. 1991. Settlements of shallow
Heft 27 (in German) foundations on sand. Geotechnique 41: 243256
Lesny, K., Kisse, A. & Richwien, W. 2002. Proof of Perau, E. 1995. Ein systematischer Ansatz zur Berechnung
Foundation Stabililty Using a Consistent Failure Model. des Grundbruchwiderstands von Fundamenten.
Proc. of the Int. Conf. on Probabilistics in Geotechnics Mitteilungen aus dem Fachgebiet Grundbau und
Technical and Economic Risk Estimation, Graz, Austria: Bodenmechanik der Universitt-GH Essen, Heft 19 (in
95103 German)
Lesny, K., Perau, E., Richwien W. & Wang, Z. 2000. Some pREN 1997-1. 2004 Geotechnical Design Part 1: General
Aspects on Subsoil Failure of Vertical Breakwaters. Rules. European Committee for Standardization, Brussels
Forschungsbericht aus dem Fachbereich Bauwesen, Heft Zienkiewicz, O.C. 1988. The Finite Element Method.
83, Universitaet Essen (in German) London: Mc Graw Hill

489

Copyright 2005 Taylor & Francis Group plc, London, UK


Bearing capacity of strip footings subjected to complex loading

L. Thorel
Laboratoire Central des Ponts et Chausses, Route de Bouaye, Bouguenais Cedex, France

A.-H. Soubra
GM, Universit de Nantes, Bd. de lUniversit, Saint-Nazaire cedex, France

J. Garnier & R. Assaf


Laboratoire Central des Ponts et Chausses, Route de Bouaye, Bouguenais Cedex, France

ABSTRACT: For shallow water systems installed on a sandy soil, the load applied to the footing may be com-
plex (i.e. inclined and eccentric) and may include cyclic effects. This paper is limited to the study of the static
behavior of a soil-footing system under inclined, eccentric or complex loading. Centrifuge tests on a small scale
model of a strip footing placed on a horizontal sandy soil have been performed in the LCPC laboratory (Nantes)
in France. Quasi-static force-controlled load, using hydraulic servo-jack, is applied to the footing. Experimental
results of the bearing capacity are compared to those given by the kinematical approach of the limit analysis the-
ory. Also, the experimental findings concerning the footing kinematics are presented and discussed. Finally, a
comparison between the reduction coefficients obtained by centrifuge model tests and those given by the limit
analysis theory is presented for different loading configurations.

1 INTRODUCTION Therefore, there is an urgent need to properly evalu-


ate these reduction coefficients based on rigorous
The calculation of the ultimate bearing capacity of experimental model tests and to perform some the-
shallow foundations is an old problem of soil mechan- oretical models in agreement with the experimental
ics. Numerous methods of computation of the ultimate findings.
bearing capacity are available in practice in the case of This paper is devoted to the determination of the
a central vertical load and a horizontal ground surface. ultimate bearing capacity of a strip footing resting on
These methods are based on either empirical tech- a horizontal ground surface and subjected to a com-
niques (e.g. Eurocodes 1996, DTU 1988, Fascicule62 plex loading based on a centrifuge experimental
1993, Vesic 1975) or simplified approaches (e.g. model and a theoretical limit analysis model. It
Caquot & Krisel 1953). Some of them make use of should be emphasized that complex loading is a fre-
laboratory shear strength characteristics of the soil (c, quent load for an offshore structure founded on the
!), others refer to in situ test results (e.g. cone pene- sea-bed. It results from both vertical loading (weight
tration or pressuremeter tests). of superstructures, overloading, etc.) and horizontal
For a complex (inclined and eccentric) loading, the loading (e.g. wave, wind, earthquake).
determination of the ultimate bearing capacity is First, a brief description of the experimental small
more difficult. For instance, the methods based on scale model is given. Then, a presentation of some
in-situ tests are convenient only for a strip footing rest- useful experimental results follows. Some insights
ing on a horizontal ground surface and subjected to a into the footing kinematics due to several complex
central vertical load. Also, the computation of the loading are given. Finally, the kinematically admis-
ultimate bearing capacity of a strip footing subjected sible failure mechanisms used for the computation of
to an inclined and/or eccentric load based on labora- the ultimate bearing capacity under several loading
tory test results makes use of approximate empirical configurations are presented. The paper concludes by
reduction coefficients to take into account the effect a comparison between the experimental findings of
of load inclination and eccentricity. These empirical the ultimate bearing pressures and the theoretical
coefficients may significantly overestimate the true solutions given by the limit analysis theory for differ-
ultimate bearing pressure (Soubra et al. 2003b). ent complex load configurations.

491

Copyright 2005 Taylor & Francis Group plc, London, UK


It should be mentioned that similar comparisons tests on a horizontal ground surface (  0) and those
between centrifuge experimental results and upper- where the footing edge is placed at a distance d greater
bound solutions from limit analysis theory have been than 6B from the crest of the slope have been pre-
made by Soubra et al. (2003a) in the case of an eccen- sented. No embedment of the foundation (i.e. D  0)
tric loading and Soubra et al. (2003b) in the case of an is considered here. Several sets of parameters have
inclined loading. Also, the effect of the proximity of a been tested : e/B  0, 1/8;   0, 15,
15 and 20.
slope on the bearing capacity of a strip footing was
investigated by Soubra et al. (2004) in the case of a 2.2 Experimental setup
vertical load and Thorel et al. (2004) in the case of a
complex loading. A rectangular strong box (120 cm  80 cm) as shown
in Figure 2 has been filled with white Fontainebleau
sand using dry raining technique (Garnier et al. 1999).
2 CENTRIFUGE MODELLING The main characteristics of the white Fontainebleau
sand are presented in Tables 2 and 3. Following the
Physical modelling with centrifuge, a widespread results from Ovesen (1979), the scale effect can be neg-
technique in the geotechnical field (Cort 1988, Ko lected if BM/D50  30. Here, as BM  40 mm, the ratio
et al. 1991, Leung et al. 1994, Kimura et al. 1998 & BM/D50 is about 180.
Phillips et al. 2002), makes it possible to reproduce in A typical strong box includes 6 foundations of
situ stresses in a small scale model. The full-scale width B  40 mm and length L  280 mm. It is
geotechnical model, named prototype (P) and the divided into 2 lanes delimited by 3 vertical thick glass
small scale model (M) are linked together through plates (Figs 24), which were used in order to simulate
scaling laws, deduced from equilibrium equations. plane strain conditions. When the strong box is filled
The main scaling factors x*  XM/XP used in this up, the surface of the sand mass is levelled.
study are listed on Table 1 in which N is the centrifuge
acceleration or g-level.

2.1 Experimental program


Data presented here have been extracted from several
experiments performed in the LCPC on shallow foun-
dations (Bakir 1993, Bakir et al. 1994, Marchal et al.
1998, Marchal 1999, Garnier et al. 2000 and Assaf
2004). The tests concern the cases of inclined and/or
eccentric load, and the foundation may be placed in
the proximity of a slope, at a distance d from the crest
of the slope. Figure 1 shows the different geometrical
parameters used in the analysis. In this paper, only

Table 1. Scaling factors. Figure 2. Raining of dry sand using an automatic mobile
hopper.
Parameter Scaling factor
Table 2. Fontainebleau sand characteristics.
Length, Displacement *  1/N
Density *  1 s ( d)min ( d)max c !
Acceleration g*  N (kg/m3) (kg/m3) (kg/m3) (kPa) (degrees)
Stress *  1
Force F*  1/N2 2650 1422 1739 06* 3940*
Angle of rotation *  1
*d  16 kN/m3 and normal stress ranging from 50 to
300 kPa.
<0 e<0 <0 e>0
Table 3. Particle size analysis of Fontainebleau sand.
D
d Particle
e e size (mm) 0.80 0.50 0.315 0.2 0.125 0.08 0.05
B
Finer (%) 100 99.9 96.1 31.7 4.1 2.6 2.3
Figure 1. Notations used in complex loading of footings.

492

Copyright 2005 Taylor & Francis Group plc, London, UK


Four calibrated boxes (Fig. 4), laying on the floor of 280 mm long (i.e. L/B  7 for M footing type and
the strong box are used to check the density of the sand L/B  9.33 for B footing type).
model. As a rain gauge gives a sample of the mass of The upper faces of the footings are equipped with
water rained, the density of sand is obtained from the semi-cylindrical joints (Fig. 5) which can be placed
average mass of sand rained in those boxes during plu- on the footing centreline or eccentrically (e/B  1/8).
viation. At the end of a centrifuge test, the strong box is This allows the foundation to rotate freely around the
emptied and calibrated boxes are levelled and weighed. load application point.
The loading device consists of a hydraulic servo
2.3 Devices and loading process controlled jack with a force transmission rod
equipped with a ball joint in its middle part (Fig. 6).
The geotechnical centrifuge at LCPC laboratory, This system allows the foundation to move freely in
Nantes (Cort & Garnier 1986) has a radius (distance the horizontal direction without any spurious moment
axis/ basket platform) of 5.50 m. in the transmission rod.
Two different accelerations have been used during The load applied to the model is measured using a
the tests (Fig. 5) : the 50 g (M type foundations 5000 daN transducer placed on top of the jack and
with BP  2 m and BM  40 mm) and the 30g (B linked to the transmission rod. The displacement of
type foundations with BP  0.9 m and KM  30 mm). the rod is also measured during loading (Fig. 6). The
Also, two types of model foundations have been used displacement of the foundation is measured using four
to simulate two different widths (2 m and 0.9 m) of LVDT transducers placed on its corners (Fig. 7). Three
the prototypes. of these transducers measure the vertical displacements,
The foundations are made of aluminium. Sand par- and the last one gives the horizontal displacement.
ticles are glued at the footing base to simulate rough
base conditions. However, the footing sides are kept
smooth. Finally, friction at the glass-footing interface displacement transducer
(Figs. 34) has been reduced with Teflon strips fixed rod upper fixing
on the footing. Both foundations are 4 mm thick and force transducer actuator ram

oil inlet
(outlet)
280 mm 280 mm
model foundations ball joint
support plate with
glass plates a 15 inclined side

1200 mm
force
800 mm transmission
rod
360 mm container wall

force transmission
rudder bar
Figure 3. Scheme of a strongbox equipped with glass plates.
Figure 6. Side view of the loading device and detail of the
force transmission rudder bar.

Figure 4. Layout of the density boxes: 2 boxes per lane.

40mm 30 mm

15mm 15mm

("M" type) ("B" type)


Figure 7. Location of the displacements transducers on the
Figure 5. Model foundations used in the tests. foundation.

493

Copyright 2005 Taylor & Francis Group plc, London, UK


0

Failure
5
s/B [%]

10

15

A-1-1-cvc1
20
0 500 1000 1500 2000
q [kPa]

Figure 8. Example of loading curve (test A-1-1 : vertical


central load).
Figure 9. Strip footing in the initial and current position.

Two other transducers measure the rotation of the 32


footing. 361
Each strong box follows the same experimental 728
33 Initial positions :
setup, as described below: Y [mm] 1078
Preparation of the soil sample and installation in C : X = 0, Y = 32 mm
34
the centrifuge with the foundations, the trans- J : X = 20, Y = 32mm
ducers and the loading device. 1446
35 J' Length in model scale.
Macrogravity self-weight loading of the soil sam- C'
ple during five 1 g 50 g cycles. 1813 kPa A-3-3
Step by step loading of the foundation until failure 36
-5 0 5 10 15 20 25
(Fig. 8). X [mm]
Control of the sample density during the emptying
of the strong box (embedded calibrated boxes). Figure 10. Movement of the foundation base for vertical
The displacement of the centre of the foundation is central loading (e  0,   0).
then calculated by combining these measurements. 31 32

33 33
Y [mm]

3 RESULTS
Y [mm]

3.1 Footing kinematics 35


J'
34 J'

The application of a complex load on a footing may C'


A- 2-2 C' A- 1-3
37 35
induce different types of movement depending on the 0 5 10 15 20 25 0 5 10 15 20 25
direction and intensity of the load. When both e and  X [mm] X [mm]
have the same sign (Fig. 1), the moments reported to
the footing centre due respectively to horizontal and Figure 11. Movement of the foundation base for vertical
vertical components have the same sign (i.e. cumula- eccentric loading (left) and inclined central loading (right).
tive effect). However, opposite signs of e and  leads 32
32
to an antagonistic effect of the moments due to the
horizontal and vertical forces. 33
During each test, the movement of the foundation 33
Y [mm]

Y [mm]

may be calculated assuming that it is a rigid body with 34 C'


J'
a 2D movement. The displacement of the foundation is 34
determined with one rotation and two displacements 35 J'
(Fig. 9). The kinematics of the strip footing is useful in C'
A-1-4 A-2-5
order to validate the movement predicted by the fail- 36 35
0 5 10 15 20 25 25 20 15 10 5 0 -5
ure mechanisms used in the limit analysis approach. X [mm] X [mm]
Measurements of displacements and rotations
make it possible to follow the movement of the base of Figure 12. Movement of the foundation base for antag-
the foundation (Figs 1012). This data confirms the onist loading (left) and cumulative combined loading (right).

494

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 4. Experimental and limit analysis results.

Test d e/B  (qr) (qr)m (qr)m/(qr)m(0)

EXP LA EXP LA EXP LA


Test (kN/m3) (
) () (kPa) (kPa) (kPa) (kPa) (
) (
)

A-1-1 16.20 0 0 1916 2140


A-2-1 16.28 0 0 1673 2148
A-3-3 16.29 0 0 1813 2150 1808 2138 1 1
M-1-1 16.04 0 0 1734 2125
M-1-4 16.04 0 0 1908 2125
M-10-1 16.02 0 0 *825 1156
M-10-3 16.02 0 0 *888 1156 856 1156
A-1-3 16.20 0
15 871 752
A-3-6 16.29 0
15 707 755
M-2-3 15.98 0
15 844 745
M-3-4 16.09 0
15 856 749 800 749 0.44 0.35
M-7-5 16.07 0
15 848 748
M-11-2 16.07 0
15 730 748
M-11-3 16.07 0
15 742 748
M-4-2 16.04 0
20 569 490
M-4-6 16.04 0
20 571 490 553 490 0.31 0.23
M-5-2 16.09 0
20 519 491
A-1-2 16.20 1/8 0 1455 1570
A-2-2 16.28 1/8 0 1439 1576
M-2-4 15.98 1/8 0 1473 1553 1436 1562 0.79 0.73
M-2-7 15.98 1/8 0 1455 1553
M-3-2 16.09 1/8 0 1460 1561
M-4-4 16.04 1/8 0 1333 1557
A-1-4 16.20 1/8
15 884 1057
A-1-5 16.20 1/8
15 764 1057 802 1058 0.44 0.49
A-2-6 16.28 1/8
15 759 1061
A-1-6 16.20 1/8 15 791 592
A-2-5 16.28 -1/8
15 643 594 717 593 0.40 0.28
A-2-3 16.28 1/8 20 482 393
A-2-4 16.28 1/8 20 512 393 497 393 0.27 0.18

*tests performed with a B type foundation.


qrm(0) is the mean value for vertical central load (e    0).

assumptions made of using a non-symmetrical mech- P


anism even for a footing subjected to a vertical load at
the centre. P.tan
1 1 n
3.2 Ultimate bearing pressures i Vi,i+1
n
For each loading configuration (i.e. e and ), Table 4 Vi+1
gives the mean value (qr)m of the bearing pressure as i Vi
given by the experimental (EXP) and theoretical (LA)
models. Also, the final two columns of the same table
give the reduction coefficients (qr)m/(qr)m(0) correspond- Figure 13. Translational multiblock failure mechanism
ing to different loading configurations using the mean (Soubra 1999).
values of the experimental and theoretical results.
The theoretical results are obtained by the kin- translational multiblock non-symmetrical mechanism
ematical approach of the limit analysis (LA) theory. presented by Soubra (1999) and Soubra et al. (2003b)
These values are presented in column qr LA of is used (Fig. 13). However, when the eccentricity is
Table 4. For vertical or inclined central load, the present, two different rotational failure mechanisms

495

Copyright 2005 Taylor & Francis Group plc, London, UK


<0 2,5
=0
e<0 e=0

2
A B E
=0
e = B/8
Rh

Calculated [MPa]
R0 1,5
r0 R
r C
r1 =0
' D = -15
e = 0 (B-type)

1 e = B/8
= -15
O e=0
= -15
e = B/8 = 20
Figure 14. Double-spirals rotational failure mechanism 0,5
e=0
(cumulative loading, Soubra et al. 2003a). = 20
e = B/8
O 0
0 0,5 1 1,5 2 2,5
1 0 Experimental [MPa]
<0
e<0 r1 Figure 16. Calculated failure stress versus experimental one.
r()
r0
A C loading process of the footing, particularly for inclined
loading.
V() It should be emphasized that the comparison
between the experimental and the theoretical results
have shown that the LA results are, as expected,
B
greater than the experimental results for all loading
Figure 15. Log-spiral rotational failure mechanism (antag- configurations except those that include loading
onistic loading, Soubra et al. 2003a). inclination. This remark may indicate that the present
theoretical models do not properly simulate the
are employed. When the complex loading leads to a experimental model which may include a velocity
cumulative effect of the load components (cf.  3.1), discontinuity along the soil-footing interface. This
the soil failure is simulated by a double spiral failure discontinuity will induce an energy dissipation along
mechanism with three different shear zones (Fig. 14). this interface and increases the bearing capacity of
An antagonistic effect of load components is simu- the footing. Further work should be undertaken to
lated by a rigid block failure mechanism bounded by study the possible effect of this discontinuity on the
a log-spiral slip surface (Fig. 15). For both rotational bearing capacity of the foundation.
mechanisms, the moments due to horizontal and ver-
tical components of the footing load have the same
sign when reported to the rotation centre of the failure 4 CONCLUSIONS
mechanism. The laboratory shear strength parameters
used in the theoretical models are c  4.5 kPa, Strip footings resting on a horizontal sandy soil sub-
!  39.4. These values were obtained by direct shear jected to inclined and/or eccentric loading have been
box tests. investigated from an experimental and a theoretical
The comparison between centrifuge results and point of view using respectively centrifuge modelling
theoretical solutions given by the kinematical approach and limit analysis theory. Reasonably good agreement
of limit analysis for the different load combinations is is obtained between experimental and theoretical
presented in Figure 16 and Table 4. Reasonably good results in terms of the ultimate bearing capacity and
agreement between experimental and theoretical the reduction coefficients for several load combin-
results is observed in terms of both the ultimate bear- ations. The experimental findings have shown that even
ing pressures and the reduction coefficients. It can be for vertically loaded footings without any load eccen-
seen that the experimental results do not differ from tricity, the failure mechanism is non-symmetrical if
the LA predictions by more than 34% (with a mean the foundation is free to rotate. Further work should
value of 15%) for a given loading configuration. The investigate more elaborate mechanisms that may
discrepancy may be attributed mainly to the ageing include a velocity discontinuity along the soil-footing
effect of the re-used sand (Thorel et al. 2003) and the interface when the footing load is inclined. Additional

496

Copyright 2005 Taylor & Francis Group plc, London, UK


tests may be undertaken in the future to study the case Ovesen N.K. 1979. The scaling law relationships.
of a and frictional cohesive soil material. Design parameters in geotechnical engineering, 7th
E.C.S.M.F.E., Brighton, Vol. 4, 319323.
Phillips R., Guo P.J., Popescu R. 2002. Physical modelling in
Geotechnics. ICPMG 02. Balkema. 1025.
ACKNOWLEDGEMENTS Soubra A.H., Garnier G., Thorel L. 2003a. Effet de lexcen-
tre-ment de la charge sur la portance des fondations
The authors would like to thank C. Favraud and superficielles: Etude thorique et exprimentale. Symp.
N. Thtiot for their help to perform centrifuge tests. Int. FONDSUP 57 nov., Paris Magnan & Droniuc (ed.)
455462.
Soubra A.H., Thorel L., Garnier J. 2003b. Effet de linclinai-
son de la charge sur la portance des fondations superfi-
REFERENCES cielles: Etude thorique et exprimentale. Symp. Int.
FONDSUP 57 nov., Paris Magnan & Droniuc (ed.)
Assaf R. 2004 Fondations superficielles tablies sur sol hori- 463470.
zontal et soumises des chargements inclins et excen- Soubra A.H., Garnier J., Thorel L., Chambon P., El Hachem
trs. Etude exprimentale sur modle rduit centrifug de E., Cortas J., Oueidat H., Chehade R. 2004. Portance des
la capacit portante et du mouvement de la fondation. fondations superficielles tablies proximit dune pente :
Rapport LCPC RMS/MSC n2004-4-04-1/1-a. 249. tude thorique et exprimentale. Coll. int. Gotechnique.
Bakir N.-E. 1993. Etude sur modles centrifugs de la Beyrouth. 1922 mai. 6.
capacit portante de fondations superficielles. Thse Soubra. A.-H. 1999. Upper-bound solutions for bearing
Ecole Centrale de Nantes. Juin. 276. capacity of foundations. J. of Geotech. & Geoenv. Engrg.,
Bakir N-E., Garnier J., Canepa Y. 1994. Etude sur modles ASCE. 125(1). 5968.
centrifuges de la capacit portante de fondations Thorel L., Gaudin C., Rault G., Garnier J. 2003.
superficielles. Etudes et recherches des LPC. GT 59. Vieillissement des sols reconstitus utiliss sur les mod-
Octobre. 188. les physiques en centrifugeuse. IS LYON 03. 3rd Int.
Caquot A. & Krisel J. 1953. Sur le terme de surface dans le Symp. On Deformation Characteristics of Geomaterials.
calcul des fondations en milieu pulvrulent. Proc. 3rd Lyon 2224 sept 03. Di Benedetto et al. (ed.). Balkema.
Int. Conf. on Soil Mech. And Found. Engrg., ICOSOMES, 8996.
Zurich, Vol. I, 336337. Thorel L., Soubra A.H., Garnier J., Chambon P., El Hachem
Cort J.F. (ed) 1988. Centrifuge 88. Proc. Int. Conf. on geo- E., Cortas J., Oueidat H., Chehade R. 2004. Effet dun
technical centrifuge model. Paris. 2527 Paris. 610. chargement complexe sur la portance des fondations
Cort J.F. & Garnier J. 1986. Une centrifugeuse pour la superficielles tablies proximit dune pente : tude
recherche en gotechnique. Bulletin de liaison des labo- thorique et exprimentale. Coll. int. Gotechnique
ratoires des Ponts et Chausses. 146. 528. Beyrouth. 1922 mai. 6.
DTU 13.12. 1988. Rgles pour le calcul des fondations Vesic A.S. 1975. Bearing capacity of shallow foundations.
superficielles. Cahier CSTB, Paris, AFNOR DTU P11- In : Foundation engineering handbook, Winterkorn &
711, 2830. Fang (ed.). Van Nostrand, 121147.
Eurocode 7. 1996. Calcul gotechnique. XP ENV 1997-1.
Fascicule 62 Titre V. 1993 Technical rules of design and cal-
culation of the foundations of the works of Civil NOTATION
Engineering. M.E.L.T. Ministre de lquipement, du
Logement et des Transports. 182. d  unit weight of dry soil
Garnier J., Amar S., Mezazigh S., Marchal O. 2000. New  angle of the slope with the horizontal
results for slope-foundation interactions. GEOENG   load inclination
2000. Melbourne. !  effective angle of internal friction
Garnier J., Derkx F., Cottineau L.-M., Rault G. 1999. Etudes
gotechniques sur modles centrifugs. Evolution des
( d)max  maximum density of dry soil
matriels et des techniques exprimentales. Bulletin de ( d)min  minimum density of dry soil
liaison des laboratoires des Ponts et Chausses. 223. s  density of solid particles
2750. B  breadth of foundation
Kimura T., Kusakabe O, Takemura J. (ed) 1998. Int. Conf. c  effective cohesion intercept
Centrifuge 98. Tokyo, Balkema. 919. D  embedment depth
Ko H.Y. & McLean F. (ed). 1991. Int. Conf. Centrifuge 91. D50  diameter corresponding to 50% finer
Boulder. 1314 June. Balkema. 616. e  eccentricity
Leung C.F., Lee F.H., Tan E.T.S. (ed). 1994. Int. Conf. g  centrifuge acceleration
Centrifuge 94. Singapore. 31 aug2sep. Balkema. 836.
Marchal O., Garnier J., Canepa Y., Morbois A. 1998. Bearing
L  length of foundation
capacity of shallow foundations near slopes under com- P  vertical component of the applied force
plex loads. Centrifuge 98. Tokyo. vol.1. 459464. q  applied stress
Marchal. O. 1999. Portance de fondations superficielles qr  failure stress
tablies proximit de talus et soumises des charges (qr)m  mean failure stress of a loading configuration
inclines et excentres. Thse Doct. Univ. de Nantes. 357. s  settlement

497

Copyright 2005 Taylor & Francis Group plc, London, UK


Numerical analysis of bearing capacity of foundation under
combined loading

Zhao Shao-fei1,2, Luan Mao-tian1,3,4 & Lu Ai-zhong2


1
Institute of Geotechnical Engineering, School of Civil and Hydraulic Engineering, Dalian University of
Technology, Dalian, China;
2
School of Civil Engineering, Shandong University of Science and Technology, Qingdao, China;
3
State Key Laboratory of Coastal and Offshore Engineering, Dalian University of Technology, Dalian, China;
4
Institute of Rock and Soil Mechanics, The Chinese Academy of Sciences, Wuhan, China

ABSTRACT: It is important to define stability/failure envelopes and failure mechanism of foundation soil
under combined loads (vertical, horizontal and moment) in design and construction of engineering structures.
Swipe test method is effective to probe the failure envelopes in (V, M, H) space. Here a modified swipe test pro-
cedure is presented by inserting a transition step in the conventional two-step swipe test procedure. As an example,
a strip footing subjected to combined loads resting on a homogeneous clay deposit is analyzed by the two-
dimensional finite element method based on the proposed swipe test procedure. Numerical results show that the
failure envelopes computed by the proposed swipe tests are more accurate than those computed by the conven-
tional swipe tests. According to the combined load characteristic in engineering, the modified swipe test proced-
ures are used to investigate influences of vertical component on the failure envelopes in the M-H load space and
soil foundation failure mechanisms. Some remarkable soil failure mechanisms under different constant vertical
loading conditions are produced.

1 INTRODUCTION Martin & Houlsby (2000, 2001) and for conical foot-
ings by Tan (1990).
Offshore structures and their foundations are usually The Swipe test method plays an important role in
subjected to large horizontal load and overturning experimental and theoretical studies for searching
moment due to wave and wind action in addition to failure envelopes. Results by this method are con-
massive vertical gravity load. Under this pattern of sidered to approach the true envelopes in the load space
combined loading, the ultimate bearing capacity of (V, M, H). This paper aims to improve envelopes
offshore foundation is of significant importance. How- obtained by numerical analyses. Here numerical ana-
ever, few appropriate analysis and design methods are lyses are conducted using the commercially available
available at present. In the design and construction of finite element program ABAQUS (HSK 2002).
offshore structures and facilities, an important issue The horizontal load and moment are usually caused
is to define the ultimate loads and failure mechanisms by wave or wind loading and vertical load is mainly due
of foundations rationally for different combinations to the gravity of structures and facilities. Consequently,
of horizontal (H ), vertical (V ) and moment (M ) com- the vertical load generally varies in a relative narrow
ponents of the applied load. Under combined load, it range or almost keeps constant compared with other
is reasonable to analyze foundation stability by the two components. Therefore, it is more realistic to study
failure envelopes in a three-dimensional load space the soil failure envelopes and failure mechanisms under
(V, M, H). This subject has received wide attention the combined load pattern in which vertical load
in the geotechnical communities. A number of efforts component V is a constant.
have been made on failure envelopes in the load space
(V, M, H) for skirted foundations by Bransby &
Randolph (1998, 1999), for caisson foundations by 2 FINITE ELEMENT MODEL
Cassidy (2004), for foundations under flat footings by
Gottardi et al. (1999), Gourvenec et al. (2003), Luan The strip footing of width B used for the analysis is
et al. (1988, 2002), Taiebat & Carter (2000), Ukritchon assumed to be rigid, resting on the surface of the
et al. (1998), for spudcan footings by Martin (1994), foundation soil. The contact between the footing and

499

Copyright 2005 Taylor & Francis Group plc, London, UK


Load application point

v M
H
h
V

Figure 2. Sign convention for the loads and the


displacements.

3.2 Method of Probing failure envelopes in the


load space (V, M, H)
In the finite element analysis, load is usually applied
by the load controlled method or the displacement
controlled method. As different from the load con-
Figure 1. Finite element model used in numerical analysis. trolled method, the displacement controlled method
enables to simulate the post-failure response. From
the load-displacement response of the footing it may
the foundation soil is assumed to be rough. An elastic be concluded that the soil foundation has attained its
perfectly plastic constitutive model based on Mohr- limit equilibrium state. In this state, the slope of the
Coulomb failure criterion is employed for purely load-displacement curve tends to zero that means the
cohesive soil with a cohesion of c. Full adhesion is strip footing displacement continually increases with-
assumed between the footing and the soil. out changing the applied load. The applied load cor-
Plane-strain analyses are carried out for surface responding to the limit equilibrium state is the ultimate
strip footings. The finite element mesh should be bearing capacity. For combined loading pattern, the
large enough to eliminate the influence of artificial displacement controlled method is found to be more
boundary conditions. In the model shown in Figure 1, suitable for searching the failure envelopes than the
a width of 20B and depth 10B has been chosen. The load controlled method as demonstrated by Bransby
finite element model consists of 10440 8-node quadri- and Randolph (1997).
lateral isoparametric elements and a reduced integra- Usually, soil failure mechanisms are illustrated by
tion technique using a (2  2) Gaussian quadrature is displacement vectors of the finite element mesh nodes.
used to compute the element stiffness. However, this method is not convenient in this analysis
when the mesh is very fine. The distribution of equiva
lent plastic strain and the deformed
3 NUMERICAL ANALYSIS
soil domain corresponding to the limit equilibrium
state are used instead to display the soil failure mech-
3.1 Sign convention for loads and displacements
anisms. The finite element model and the suggested
The application point of the combined loads compon- analysis methods have been verified in a simple clas-
ents V, M and H is assumed at the midpoint of the strip sical problem, the vertical ultimate bearing capacity
footing as shown in Figure 2. Under combined loads, of a strip footing on purely cohesive clay deposit by
displacements of the strip footing include a vertical Zhao et al. (2004).
displacement v, a horizontal displacement h and a rota-
tion angle . Sign convention for the three loads and 3.2.1 Swipe test
their corresponding displacements follows the recom- Analysis procedure of a swipe test was first intro-
mendations of Butterfield et al. (1997) as shown in duced by Tan (1990) in centrifuge tests and referred
Figure 2. Because only the positive vertical component as a side-swipe test. Two steps are included in a wipe
V is considered in this paper and the foundation test procedure. For a failure envelope in ij load plane
responses are symmetric for the two combined loading as an example, in its first step a displacement ui in
cases (V, M, H) and (V, -M, -H), it is satisfactory to ana- i-direction is imposed from a zero load state to the final
lyze foundation failure responses under the two loading state at which the ultimate load is reached, and in its
patterns in which the moment component M is applied second step a displacement uj is imposed in j-direction
in its positive direction and the horizontal component H keeping the constant displacement in i-direction until
either in its positive direction or in its negative direction. the footing loading does not vary with the increased

500

Copyright 2005 Taylor & Francis Group plc, London, UK


0.8
ui
utj
uti
(a) Step 1 (b) Step 2
0.7

uj

(c) Step 3 (d) Final position


0.6
0.8 0.9 1.0
Figure 3. Procedure of a modified swipe test.
0.9

displacement in j-direction. The resulting load path in 0.6


the second step will approach the failure envelope in
the ij load plane.

M/(B c)
2
0.3
3.2.2 Modified swipe test
Envelopes by swipe test method are within the true
envelopes in load space (V, M, H) as discussed by
Bransby and Randolph (1997). In order to improve its 0.0
result, a modified swipe test procedure is developed -1.2 -1.0 -0.8 -0.6 -0.4 -0.2 0.0 0.2 0.4 0.6 0.8 1.0 1.2
in which a transition step is put in between the two H/(Bc)
steps. In the transition step, displacements in i-direction
Swipe tests (HM) dh/(Bd) = 1.0
and j-direction are imposed simultaneously and the A swipe test (MH) du/(Bd) = 4.0
ratio of the displacement increments in two directions Modified swipe tests (HM) Combined load and
is kept constant until the footing loading does not dh/(Bd ) = 0.2 displacement controlled tests
vary with the increased displacements in i-direction
and in j-direction. The composite load path formed in Figure 4. The failure envelope in the M-H load plane
both transition step and the former second step con- (V  0.0).
stitutes the final failure envelope in ij load plane. Full
procedure of a modified swipe probe test consists of
three steps as shown in Figure 3. The last step may be
equal to zero and 0.5 Vult respectively. Vult is the ana-
cancelled if it is unnecessary.
lytical solution (2  )Bc given by Prandtl (1921).
For the case of V  0, failure envelopes by different
3.2.3 Load-displacement controlled probe test
probe methods are shown in Figure 4. Most results by
A load-displacement controlled test is an additional
the swipe tests (
HM, MH and HM) match well with
method to probe the failure envelopes. This method
the points by load-displacement controlled probe tests.
consists of two steps. In its first step, a load is directly
However, when the moment M component is approach-
applied in i-direction by load controlled method which
ing its maximum value Mmax, the two-step swipe tests
is less than the ultimate bearing capacity in this direc-
can not present satisfactory results comparing with the
tion and a point within failure envelope is gained. In its
results by load-displacement controlled probe tests.
second step, keeping loading in i-direction constant, a
Three modified swipe tests (HM) are carried out to
load in j-direction is applied by displacement con-
probe the envelope portion including the point of
trolled method until applied load does not vary with
moment Mmax. Ratios of dh to (Bd) in the second
the increased displacement in j-direction. A point, at
steps of modified swipe tests (HM) are equal to 0.2,
which load value in i-direction can be predefined, is
1.0 and 4.0 respectively. Compared with the results
probed on the envelope in the ij load plane.
from the load-displacement controlled tests, it is evi-
dent that the results by the modified swipe tests (HM)
4 NUMERICAL RESULTS are better than those given by the two-step swipe test
(HM). In the three modified swipe tests, both cases of
4.1 Failure envelopes in the M-H plane dh:(Bd)  1.0 and h:(Bd)  4.0 present more sat-
isfactory results than the case of h:(Bd)  0.2.
Failure envelopes in the M-H load plane are studied Results by the load-displacement controlled tests
for two cases in which vertical load component V is show that the moment, Mmax, on the envelope is equal

501

Copyright 2005 Taylor & Francis Group plc, London, UK


0.8 becomes deeper as H increases. When the horizontal
0.7 load H is larger than 0.95Bc and smaller than
0.6 1.0Bc, the plastic strain p comes forth in the soil,
0.5
which is adjacent to the two sides and the base of the
strip footing, besides that on the scoop surface as
M/(B2c)

0.4
shown in Figures 6(f)(h). According to the distribu-
0.3 tion of the plastic strain p and the deformed soil
0.2 shape accompanying the scoop failure mechanism,
0.1 another soil failure mechanism occurs closely to the
0.0 strip footing, which is the same with the double-
-0.1 wedge mechanism named by Bransby & Randolph
-1.0 -0.5 0.0 0.5 1.0 (1998). If the horizontal load is H  1.0Bc, the dou-
H/(Bc) ble-wedge mechanism occurs only in the soil, as
Swipe tests(-HM)
shown in Figure 6(i). If the horizontal load H is
Modified swipe tests(-HM)
Modified swipe tests( HM) Load-displacement applied in the horizontal displacement controlled
Swipe tests(MH) controlled probe tests method until its ultimate value Hult, failure occurs
under the soil-footing interface as shown in the
Figure 6(j).
Figure 5. The failure envelope in the M-H load plane For the case of V  0.5 Vult, soil failure mecha-
(V  0.5 Vult). nisms are shown in Figure 7 for different values of
horizontal load H. When horizontal load H is equal
to
H0 (V  0.5 Vult and M  0), its negative ulti-
to 0.80B2c corresponding to H  0.95Bc and the ultim- mate horizontal load, failure also occurs under the
ate moment M0 is 0.71B2c corresponding to the zero footing as shown in Figure 7(a) like the case of V  0.
horizontal load H. If horizontal load H is larger than
H0 and smaller
When the vertical load V is equal to 0.5 Vult and than
0.4Bc, failure mechanisms are scoop mechan-
kept constant, the failure envelope in the (V, M, H) isms which become deeper as the H value increases.
load space is shown in Figure 5. In the second step At the same time, plastic failure zone develops in the
of the modified swipe tests, the ratio of dh to (Bd) soil at the right side of the strip footing in addition to
is equal to 1.0. It is shown clearly that the modified the scoop soil failure mechanism under the footing, as
swipe probe tests can provide better prediction than shown in Figures 7(b) and 7(c). When horizontal load
those given by the conventional swipe tests and their H varies from
0.4Bc to 0.6Bc, mechanisms are
results coincide with the points probed by the load- illustrated in the Figures 7(d)-(i). Each mechanism
displacement controlled probe tests. consists of three parts: scoop part below the strip
The maximum moment Mmax on the envelope is footing, shear transitional zone and a triangle part at
equal to 0.72B2c corresponding to the horizontal the right side of the strip footing. The scoop part is not
load H  0.4Bc and the moment M0 is 0.69B2c symmetrical as that in the V  0 case although it
corresponding to H  0. The M-H envelope shape in becomes deeper as horizontal load H increases. It
the case of V  0.5 Vult is different from that in the seems that such a failure mechanism has not been
case of V  0 though both of them are not symmetri- reported in previous studies.
cal about the line of H  0. For the two cases, the If H varies from 0.6Bc to 0.8Bc, plastic strain p
maximum horizontal loads Hmax are almost same and in the shear transitional zone at the right side of the
equal to 1.038Bc (V  0) and 1.032Bc (V  0.5 Vult) strip footing extends to the soil under the footing
respectively. towards left side and traverses the asymmetrical
scoop region and the Hansens mechanism (Hansen
1970) is formed with a triangular plastic zone as shown
4.2 Soil failure mechanisms
in Figures 7(j) and 7(k). Although the Hansens mech-
Soil failure mechanism under combined loading anism and the asymmetrical scoop mechanism exist
may be illustrated, as in the benchmark problem, in the soil simultaneously, the Hansens mechanism
by the plastic strain p and the deformed soil seems to play a more dominant role compared
shape computed by the load-displacement controlled with the asymmetric scoop mechanism when H is
tests. closer to 0.8Bc. When H varies from 0.8Bc to H0,
For the case of V  0, soil failure mechanisms only the Hansens mechanism is valid for the soil as
are shown in Figure 6 with different horizontal load shown in Figure 7(l). The failure mechanism in the
H. When H is less than 0.95Bc, scoop failure mechan- case of H  H0 is the same as that in the case with
isms named by Bransby & Randolph (1998) occur H 
H0 except for the sliding direction of strip
and are shown in Figures 6(a)-(e). Scoop mechanism footing.

502

Copyright 2005 Taylor & Francis Group plc, London, UK


(a) H =-0.8Bc (f) H =0.95Bc

(b) H =-0.4Bc (g) H =0.98Bc

(c) H =0.0Bc (h) H =0.99Bc

(d) H =-0.4Bc (i) H =1.00Bc

(e) H =0.8Bc (j) H =Hult

Figure 6. Distribution of plastic strain p in the limit equilibrium state under combined loading (V  0.0).

5 CONCLUSIONS results than those given by the conventional swipe tests.


In order to search for the accurate location of a point
The foundation response under combined loads includ- on the failure envelopes corresponding to a value of
ing vertical load V, horizontal load H and moment M load which is defined beforehand along a given direc-
has been studied for a strip footing using the finite tion, it is necessary to jointly employ load-displacement
element method. In experimental and numerical stud- controlled method of loading. Compared with the
ies, swipe test, a kind of displacement controlled probe results by load-displacement controlled methods, modi-
test, is often adopted to probe the failure envelopes in fied swipe tests can offer a better approximation when
the (V, M, H ) load space. In order to improve results the ratio of two displacement increments is equal to
from the numerical swipe test, a modified swipe test unity in the transition step. In offshore environment,
procedure is presented in which a transition step is the vertical component of load V is usually a constant.
inserted between two steps of the conventional swipe To reproduce the failure mechanism of soils under
probe test. The modified swipe tests provide better combined loading with this characteristics, the case

503

Copyright 2005 Taylor & Francis Group plc, London, UK


(a) H =-H0 (g) H = 0.2Bc

(b) H =-0.8Bc (h) H =0.4Bc

(c) H =-0.6Bc (i) H = 0.6Bc

(d) H =-0.4Bc (j) H = 0.7Bc

(e) H =-0.2Bc (k) H =0.75Bc

(f) H =-0.0Bc (l) H =0.8Bc

Figure 7. Distribution of plastic strain p in the limit equilibrium state under combined loading (V  0.5 Vult).

504

Copyright 2005 Taylor & Francis Group plc, London, UK


of V  0.5Vult is investigated by the finite element Gourvenec, S. & Randolph, M. 2003. Effect of strength non-
method and a comparative study with the case of homogeneity on the shape of failure envelopes combined
loading of strip and circular foundations on clay.
V  0 is made. Numerical results show that soil failure
Gotechnique, Vol.53, No.6, pp. 575586.
mechanisms are quite different for these two cases. In Hansen, J. B. 1970. A revised and extended formula for bear-
the case with V  0, soil failure pattern varies from ing capacity. Danish Geotechnical Institute Bulletin, No.
the symmetrical scoop mechanism to the scoop- 28, pp. 511.
double-wedge mechanism and then to the double- Hibbit, Karlsson and Sorensen, Inc. 2002. ABAQUS users
wedge mechanism as H increases. However, in the Manual: Version 6.3-1. Hibbit, Karlsson and Sorensen, Inc.
case with V  0.5Vult, soil failure pattern varies from Luan, M.T., Jin, C. P. & Lin G. 1988. Ultimate bearing
the asymmetrical scoop mechanism to the scoop-fan- capacity of shallow footings on non-homogeneous soil
triangle mechanism, to the scoop-Hansen mechanism foundations. Chinese Journal of Geotechnical Engineering,
Vol.10, Mo.4, pp. 1427.
and then to the Hansen mechanism as H increases.
Luan, M. T. Wang, D. & Guo Y. et al., 2002. The state-of-the-art
on seabed dynamics and offshore foundation design. Soil
Dynamics and Geotechnical Earthquake Engineering
ACKNOWLEDGEMENTS (Edited by Liu H.L.), China Architecture & Building
Press: Beijing, pp. 2848.
The authors wish to express their gratitude to Professor Martin, C. M. 1994. Physical and numerical modeling of
Dahong Qiu of Dalian University of Technology for his offshore foundations under combined loads, PhD thesis,
continuing support and invaluable advice. The finan- University of Oxford, UK.
cial support for this study provided through a grant Martin, C. M. & Houlsby, G. T. 2000. Combined loading
of spudcan foundations on clay: laboratory tests.
(50179006) from National Natural Science Foundation
Gotechnique, Vol.50, No.4, pp. 325328.
of China is grateful acknowledged. Martin, C. M. & Houlsby, G. T. 2001. Combined loading of
spudcan foundations on clay: numerical modeling.
Gotechnique, Vol.51, No.8, pp. 687699.
REFERENCES Prandtl, L. 1921. Uber die Eindringungsfestigkeit (Harte)
plastischer Baustoffe und die Festigkeit von Schneiden.
Bransby, M. F. & Randolph, M. F. 1997. Shallow founda- Journal of Applied Mathematics and Mechanics (ZAMM)
tions subject to combined loadings. Computer Methods 1(1): 1520.
and Advances in Geomechanics (Edited by Yuan, J.X.), Taiebat, H. & Carter, J. P. 2000. Numerical studies of the
Rotterdam: Balkema, pp. 19471952. bearing capacity of shallow footings on cohesive soil
Bransby, M. F. & Randolph, M. F. 1998. Combined loading subjected to combined loading. Gotechnique, Vol.50,
of skirted foundations. Gotechnique, Vol.48, No.5, No.4, pp. 409418.
pp. 637655. Tan, F. S. 1990. Centrifuge and theoretical modelling of con-
Bransby, M. F. & Randolph, M. F. 1999. The effect of ical footings on sand. PhD thesis, Cambridge University,
embedment depth on the undrained response of skirted UK.
foundations to combined loading. Soils and Foundations, Ukritchon, B., Whittle, A. J. & Sloan, S. W. 1998. Undrained
Vol.39, No.4, pp. 1933. limit analyses for combined loading of strip footings on
Butterfield, R., Houlsby, G. T. & Gottardi, G. 1997. Standard clay. Journal of Geotechnical and Geoenvironmental
sign conventions and notation for generally loaded foun- Engineering, Vol.124, No,3, pp. 265276.
dations. Gotechnique, Vol.47, No.5, pp. 10511054. Zhao, S. F. Luan M. T. & Lu, A. Z. 2004. FEM-based non-
Cassidy, M. J., Byrne, B. W. & Randolph, M. F. 2004. linear numerical analyses for limit-equilibrium problems in
A comparison of the combined load behaviour of spud- geotechnics considering non-associated flow rule. Rock
can and caisson foundations on soft normally consoli- and Soil Mechanics, Vol.125, No.Supp.2, pp. 121125.
dated clay. Gotechnique, Vol.54, No.2, pp. 91106.
Gottardi, G., Houlsby, G. T. & Butterfield, R. 1999. Plastic
response of circular footings on sand under general pla-
nar loading. Gotechnique, Vol.49, No.4, pp. 453469.

505

Copyright 2005 Taylor & Francis Group plc, London, UK


Centrifuge tests on improving offshore foundation systems

Henderikus G.B. Allersma


University of Delft, Deflt, The Netherlands

ABSTRACT: The centrifuge testing technique is a suitable method for investigating new ideas for construc-
tion methods in the geotechnical field. In particular the small geotechnical centrifuge of the University of Delft
has proven to be very convenient for this purpose. The small scale models can be modified quickly and easily.
The tests can be reproduced accurately, so that the effect of small changes in the design can be made visible.
Several test programs have been carried out in the centrifuge of the University of Delft to examine methods of
improving the loading capacity of offshore foundation elements. The paper focuses on improving the loading
capacity of circular footings and suction caissons. It was found that parameters, such as roughness, have a sig-
nificant influence. Some new ideas for suction caissons were tested. An unexpected observation was that the
pullout capacity could be improved by removing a part of the caisson.

1 INTRODUCTION out to simulate new design concepts. The small cen-


trifuge appeared to be very suitable for this task. The
Several types of foundation systems are in use in the small models lend themselves to easy modification, and
offshore industry. Some of them are used to fix floating the accurate reproducibility of the tests ensures that the
structures, while others are used to support vertical effects of slight differences in design can be made
loads. A typical example of the latter are spudcans, visible.
which are large circular conical footings used for bear- Several interesting, and in some cases unexpected,
ing mobile platforms (jack-up units). Foundations of results were obtained. Some of the ideas tested would
this kind can exhibit problems in sliding behavior when seem to have potential application in engineering
they are founded on sand. Several test programs have practice.
been carried out in the centrifuge to analyze the sliding
behavior, and the experimental results were compared
with the governing foundation criterion for a site-
2 TEST FACILITY
specific integrity assessment (Allersma 1997). How-
ever, an interesting question that presents itself is how
2.1 Centrifuge facility
far the sliding resistance can be improved. The small
centrifuge was used to test some new design concepts. The tests were performed in the geotechnical centrifuge
A more generally applicable foundation element is of the University of Delft. The centrifuge (Allersma
what is known as a suction caisson. The main attraction 1994a) is a relatively small device with a diameter of
of this system is the convenient installation method. 2.5 m and a maximum sample weight of approxi-
A caisson with a diameter of 9 m and a height of 10 m mately 300 N. The maximum space available for the
can be installed in a few hours using only a pump. The model and the actuator is 400  400  400 mm3. The
caissons are subjected predominantly to horizontal small size of the equipment and samples has proven
loading when they are used as an anchoring system to be very convenient in operation.
for floating structures. However, other applications A two-dimensional loading system (Allersma
exist in which the caissons are subjected to vertical 1994b) was used in the test program (Fig. 1). An accu-
loading. An obvious example would be the support of rate and almost frictionless translation in two perpen-
a gravity load, but the caissons may also be subjected dicular directions was provided by tempered steel shafts
to a vertical pullout load, for example in tension leg and linear ball bearings. The horizontal load (max.
platforms, where the load has a cyclic character. 2 kN), vertical load (max. 5 kN), horizontal dis-
In order to improve the loading capacity of suction placement (150 mm), and vertical displacement
caissons, several centrifuge tests have been carried (30 mm) can be adjusted independently by means of

507

Copyright 2005 Taylor & Francis Group plc, London, UK


Figure 2. Different modifications to influence the sliding
Figure 1. Photograph of the 2 dimensional loading system. capacity of spudcans.

small DC motors, load sensors, and displacement depth in clay bottoms is 0.5 to 1 times the diameter of
transducers. The loading device is controlled by a PC, the footing, in which case sufficient horizontal load
which is located in the spinning part of the centrifuge can be mobilized to eliminate any possibility of sliding.
and is accessible via slip rings. The fact that the meas- However, when the footings are used on sand, only
uring signals do not have to travel over long lines or limited penetration is possible. A common governing
slip rings means that the measuring and control system foundation criterion for a site-specific integrity
is relatively insensitive to noise. assessment is the sliding of the windward leg. Wave
The loading path during a test can be defined in action results in a partial unloading of the windward leg
advance and entered into a computer program. The and some bending of the long legs (80 m), whereby
program is written in a common computer language, the footings are subjected to horizontal loading. In a
so that it can be modified easily by the user. previous test program (Allersma et al. 1997), the sliding
behavior was examined and compared with the usual
design criteria. It was found that some of the criteria
2.2 Model preparation
are fairly critical. This has stimulated the perform-
An important aspect of the centrifuge modeling tech- ance of tests to examine methods of improving the
nique is the preparation of the sand bed. In order to sliding capacity.
visualize the effect of slight design variations, it is The tests were performed at 150 g. At this g-level,
important for the sand beds to be reproduced accurately, a diameter of 14.4 m is simulated, taking into account
for which purpose a computer-controlled sand pouring that the effective stress in dry sand is 1.6 times greater
machine has been developed (Allersma 1994b). The than in saturated sand. The dimensions of the model
density of the sand sample can be controlled by keeping spudcan and several variants are shown in Figure 2.
the falling height constant during raining. The medium- Type (a) is considered to be the standard spudcan.
dense sand beds showed a high degree of repro- Because tipped spudcans are not uncommon, type
ducibility, with a standard deviation of mean porosity of (b) was included to investigate what influence the tip
0.2%. The height of the sand layer that was used was has. Type (c) is a flat footing and (d) represents an
100 mm, with a ground area of 250  250 mm2. In unconventional shape. The effect of skirts in combi-
general, dry sand was used with a D50  0.2 mm. Most nation with a tip is demonstrated with footing types
of the tests were performed with medium dense sand (e) and (f). Type (g) is equipped with three wings on
with the following parameters: friction angle !  33; the conical surface and type (h) has three pins in
E  15000 kPa;   17 kN/m3; porosity n  37%. combination with a tip. Furthermore, some types
Because drained conditions were being simulated, dry were tested with both a smooth steel surface and with
sand was used. a rough surface (covered with sand paper).
A typical diagram of a sliding test is shown in
Figure 3. Initially, the footing was subjected to a vertical
3 SPUDCAN FOUNDATIONS pre-load of 150 N, after which the load was decreased
to 50 N. The horizontal load was increased while
Spudcans are large circular (e.g. d  20 m) conical holding the vertical load constant. Failure is defined as
footings used to support jack-up units. The footings the occurrence of a vertical displacement. Each test
are actually designed for clay soils. The penetration was performed twice and plotted on the same figure.

508

Copyright 2005 Taylor & Francis Group plc, London, UK


4 SUCTION CAISSONS

Several test programs have been carried out to inves-


tigate the behavior of suction caissons subjected to
different loading conditions. In the first instance, com-
monly applied caisson types were used. The parameters
that were varied were the loading angle, attachment
point and the height to diameter ratio (Allersma et al.
2000). It was found that the pullout load increases
when the attachment point is lowered and when the
loading is closer to the horizontal. As may be expected,
larger caissons yield a larger bearing load, both hori-
zontally and vertically. A question that arose was
Figure 3. Typical diagram of a sliding test on a spudcan whether smarter solutions exist for increasing the
foundation. bearing load than simply making the caissons bigger.

Table 1. Comparison of the sliding capacity of different


4.1 Horizontal loading
types of footings.
In this application, a long cable or chain is attached to
Sliding capacity the caisson in order to anchor floating structures. The
caisson is loaded approximately horizontally, at an
Factor angle of 15 to the horizontal axis. The attachment point
Footing Smooth Rough
type [N] [N] Smooth Rough
is h  2H/5, where h is the distance between attach-
ment point and tip and H is the total height. Having
cone 193 247 1 1.28 found that larger diameters increase the anchor capacity,
cone  tip 243 285 1.26 1.48 a cheaper variant in which the increasing diameter is
flat 106 0.55 simulated by vertical wings (Fig. 4) was investigated.
piramidal 180 0.93 Two different wing sizes were tested with a caisson
skirt 5 mm 303 1.57 with a diameter of D  30 mm. In Figure 5 it can be
skirt 8 mm 363 1.88 seen that the effect of wings is relatively insignificant.
rib 250 1.29 Wings that increase the caisson surface area by 20%
pin 283 305 1.47 1.58
had no effect whatsoever. An increase in the bearing
capacity of only 10% was measured when the pile
The close fit of the two diagrams demonstrates the surface was increased by 50% using wings. Figure 6
reproducibility of the test procedure. shows some of the results (tests end FEM calcula-
The sliding capacity of the different types of footings tions) of increasing the caisson diameter by 50%. It
are summarized in Table 1, where the factor for the appeared in this case that the horizontal bearing cap-
smooth spudcan without tip was given a value of 1 as a acity increases by approximately 30%. The reason that
reference. It is interesting to note the significant influ- strips are less effective can be explained by the dia-
ence of roughness. gram of the failure surface, as shown in Figure 7.
Making the surface of the spudcan as rough as pos- Strips allow the soil to flow fairly easily around them
sible may well be the cheapest way of increasing the to fill the space formed during caisson displacement.
sliding capacity. The tip also appears to have a signifi- This means that the volume of the mobilized soil plug
cant influence, and a tip with a rough surface increases differs little from that of a caisson without strips.
the sliding capacity by almost 50%. While it is no sur- Increasing the diameter, however, causes a signifi-
prise that skirts are effective, they are probably a rela- cantly larger mobilized soil volume.
tively expensive solution. The ribs seem to have the Because the soil around the top of a caisson has a low
same effect as a tip. The effect of three pins (to the same stress level, it is interesting to investigate how far the top
depth as the tip) was surprising, where, especially in section contributes to the bearing capacity. Especially
combination with a rough surface, they appear to have when suction caissons are used in soft soils, it is
the same effect as a skirt. believed that the upper part makes no significant con-
It is striking that the flat footing yields a much tribution to the horizontal bearing capacity. In order to
lower sliding capacity than the conical types. The examine this phenomenon, a test series was carried out
main purpose of testing the pyramidal footing was to with piles that were penetrated to an equal depth, and
examine whether this shape yields a larger vertical where different lengths of the top section were removed.
load than a flat footing with a rough surface. The cen- The original height of the pile was H  50 mm. In all
trifuge tests showed no significant difference. tests, the attachment h  2H/5, and the loading angle

509

Copyright 2005 Taylor & Francis Group plc, London, UK


Figure 4. Using wings instead of increasing the diameter.
Figure 8. Horizontal bearing capacity of a caisson in sand
after removing the top section of the caisson.

was 15 to the horizontal. The test results in sand are


shown in Figure 8. Each height was tested twice in
order to demonstrate the reproducibility of the test
procedure. It is striking that the bearing capacity
increases with decreasing caisson height. There
appeared to be an optimum when the height was
decreased by 30%. In this case, the horizontal bearing
capacity increased by almost 30%. After the optimum,
the effect decreases with decreasing height. However,
even when half the height was left, the bearing cap-
Figure 5. Effect of wings on the horizontal bearing capacity, acity was still 16% higher. An explanation for this
duple tests.
phenomenon is that, in the case of a full-size caisson,
part of the friction resistance is between caisson and
soil. However, when the top is removed, more friction
has to occur in the soil itself, which yields a larger
bearing capacity. The experimental results were com-
pared with 3-dimensional FEM calculations. It was
found that the calculations show a similar tendency.
Some tests were performed to investigate whether a
similar behavior could be observed in clay. It appeared
that the effect in clay was less pronounced, but, in this
material too, shorter caissons have a larger bearing
capacity. Some orientational tests have shown that the
vertical pullout load also increases with increasing
Figure 6. Effect of diameter on the horizontal bearing caisson height. These findings should stimulate the
capacity. development of techniques to install suction caissons
below the sea floor.

4.2 Vertical loading


In the interests of further increasing the vertical pullout
capacity of suction caissons, an idea based on the
umbrella anchor was tested. This technique uses a
full-size caisson with a dome to install two segments
that are joined by a hinge at a given depth. After instal-
lation, the dome section is removed and the hinged
section is pulled up until the structure opens fully
(Fig. 9). Small-scale tests at 1 g have shown that a
length approximately equal to or greater than the cais-
son height was required to open the structure com-
pletely. The aim of the centrifuge tests was to
Figure 7. Diagram to demonstrate the difference in failure investigate the difference in pullout capacity between a
mode with strips of a larger diameter. closed and an opened caisson. Therefore, no attempt

510

Copyright 2005 Taylor & Francis Group plc, London, UK


Figure 9. Diagram showing the operation of the umbrella
mechanism.

Figure 11. Diagram of the test setup to simulate active suc-


tion in the centrifuge.

occur when stabilizing a vessel with cranes mounted


on deck for the purpose of transporting large units.
The advantage of a vessel compared with a crane
barge is its speed, but a disadvantage is its instability
in waves, which severely hampers crane load position
control. One way of improving the vessels stability
would be to use pre-tensioned cables anchored to the
sea floor, where suction caissons are envisaged as a
temporary anchoring system. The suction caissons
Figure 10. Comparison of the pullout capacity of the open
are installed in the usual way by means of a pump.
and closed caisson, with load given in prototype values. After installation, the pump remains attached and is
used to maintain a suction pressure during the few
days that the crane is in operation. A problem in this
was made to test the actual opening mechanism. application is that the height of the usual suction cais-
Instead, two models were made, simulating the two sons is too great, so that they hang in the water in the
conditions of interest, and they were installed during the lifted position, causing an unacceptable drag on the
preparation process of the sand box. The results of vessel. Furthermore, numerous applications are in
loading tests performed in the centrifuge at 150 g are relatively shallow water areas (1020 m). Consequently,
shown in Figure 10. If the pullout load is compared for it is worthwhile investigating whether a limited caisson
structures at the same depth it is clear that the open cais- height can be compensated for by a larger area, and to
sons yields a higher pullout load. However, a good com- what extent the shallow water can be compensated for
parison requires account to be taken of the necessity of by active suction. It would also be worthwhile investi-
raising the caisson by at least 1 caisson height. This gating various anchoring system configurations, while
effect is demonstrated by the line h  H in Figure 10. In keeping within the limits of constructional feasibility.
this light it can be seen that the umbrella mechanism In the first instance, centrifuge tests were carried
has no advantage at low depths. At depths greater than out to investigate the effect of active suction (Allersma
8 m, however, it was found that the opened caisson et al. 2003). A diagram of the test setup is shown in
starts to yield a larger pullout load. There is probably a Figure 11. A small model of a suction caisson was
transition to another failure mechanism at that depth. installed in a sand (D50  0.1 mm) layer at 1 g condi-
tions. The caisson was attached to a loading system to
measure the vertical pullout resistance. A pre-defined
4.3 Maintaining suction
suction pressure was maintained by means of a small
Suction caissons are generally used for long-term computer-controlled gear wheel pump, using the signal
foundation applications, for which it is impractical to from a pressure transducer.
increase the pullout resistance by active suction. Figure 12 shows the relationship between the
However, it would be possible to maintain the suction active suction pressure and the vertical pullout cap-
if the anchor load only persisted for a relatively short acity, which is expressed as a load per unit area. It can
period. Short anchoring system loads of this kind be seen that the additional pullout capacity increases

511

Copyright 2005 Taylor & Francis Group plc, London, UK


Figure 12. Vertical pullout resistance of caisson during
active suction. Figure 14. Photograph of different caisson types used to
examine the effect of the shape.

Figure 13. Conversion of the centrifuge results to signifi-


cant prototype values by extrapolation.
Figure 15. Comparison of a single caisson with three
in-line caissons and a sardine-can-shaped device.
linearly with increasing suction. Tests at 1 g and 30 g
were carried out. The 30 g test was performed at 1/3
of prototype scale. In the 1 g tests, the contribution of the three in-line caissons. Because some arching may
the friction was too small to be measured. Since the be expected in the narrow space between the caissons, it
tangent of the relationship seems to be independent of would seem reasonable to expect a larger uplift capac-
the g level the idea is supported that the prototype ity. It may be also be expected that the additional soil
bearing capacity can be deduced by extrapolation, as body underneath the caisson would have a greater
is demonstrated in Figure 13. It was found that 4 volume when more caissons are in line. With a single
units, each consisting of three caissons with a height caisson, the soil body has a roughly conical shape. If the
of 4.5 m and a diameter of 9 m are needed to yield suf- caissons are small enough and close enough together, a
ficient bearing capacity (ca. 90 MN) at a depth of bridge may be formed between the soil bodies, resulting
15 m to stabilize the vessel. Constructional prefer- in a larger additional soil volume per caisson. In the
ences and issues surrounding the optimum shape gave case of the sardine-can-shape, the additional soil
rise to a test program in which different configurations body would be expected to be wedge shaped, rounded
are compared with each other. The different types at the ends. This results in a larger volume than with
used are shown in Figure 14. All systems have almost three widely-spaced single caissons.
the same surface area and height, and the in-line cais-
sons are in contact with each other. Comparison of the
uplift capacity of two in-line caissons with a single 5 CONCLUSIONS
caisson has shown that the profiles of the ratio of nor-
malized uplift capacity to suction pressure are very Several test series have been carried out in the geotech-
similar. Figure 15 compares three in-line caissons with nical centrifuge of the University of Delft to examine
a single caisson and a sardine-can-shaped caisson. some ideas for improving the capacity of foundation
This test series revealed some differences between the systems. The small size makes the centrifuge very
different caisson shapes. The single caisson shows a suitable for performing trial and error tests. The models
lower normalized uplift capacity than the unit with of the foundation elements can be modified with simple

512

Copyright 2005 Taylor & Francis Group plc, London, UK


tools and techniques. The small soil containers allow a Allersma, H.G.B. 1994b. Development of miniature equip-
quick and accurate preparation of the soil layers. The ment for a small geotechnical centrifuge. Innovation in
reproducibility of the sample preparation allows the instrumentation and data acquisition systems.
effect of small changes in design to be made visible. Transportation Research Record 1432, 99105.
Allersma, H.G.B., B. Hospers, J.G. den Braber, 1997.
In spite of the small size, complicated tests can be Centrifuge tests on the sliding behaviour of spudcans.
performed. An example is the tests on caissons with Canadian Geotechnical Journal, 34(5), 658663.
active suction, in which both the load and the suction Allersma, H.G.B., A.A. Kirstein, R.B.J. Brinkgreve, 2000.
pressure have to be controlled in flight. Centrifuge and numerical modelling of horizontally
Several phenomena and tendencies have been loaded suction caissons. Int. J. of Offshore and Polar
observed. Some of them are also predicted by FEM Eng.,Vol.20 No.3, 222228.
calculation. It seems likely that some of the findings Allersma, H.G.B., J.A. Jacobse, R.L. Krabbendam, 2003.
will be of interest in engineering practice. Centrifuge tests on uplift capacity of suction caissons
with active suction. Int. Offshore and Polar Eng. Conf.,
Honolulu, 734739.
REFERENCES

Allersma, H.G.B. 1994a. The University of Delft geotechnical


centrifuge. Proc. Centrifuge94, Singapore, A.A. Balkema,
4752.

513

Copyright 2005 Taylor & Francis Group plc, London, UK


Shallow foundations: mobile jackup units

Copyright 2005 Taylor & Francis Group plc, London, UK


Extraction of jackup spudcan foundations

O.A. Purwana, C.F. Leung & Y.K. Chow


Centre for Offshore Research and Engineering, Department of Civil Engineering, National University of Singapore

K.S. Foo
KeppelFELS, Singapore

ABSTRACT: Centrifuge model tests were carried out to simulate the installation, operation, and extraction of
a jackup spudcan foundation in normally consolidated clay. The uplift resistance during spudcan extraction was
examined with particular attention paid to the development of suction at the spudcan base. The test results
revealed that the suction developed at the interface between the spudcan base and the underlying soil can be the
major resistance during extraction of spudcans with a relatively long operation period.

1 INTRODUCTION 2 EXPERIMENTAL SETUP AND PROCEDURE

The use of mobile jack-up rigs has significantly 2.1 Centrifuge model setup
increased recently. A mobile jackup rig is typically
Figure 1 shows a schematic illustration of the centrifuge
supported by three to four independent legs where
model setup. All the centrifuge tests were conducted
each leg is equipped with a footing known as spud-
at 100 g. The 125 mm diameter model spudcan, which
can. The spudcan is generally circular or polygonal
is scaled down 100 times proportionally from a typ-
in plan, with a shallow conical underside and a sharp or
ical prototype spudcan used in the field, corresponds
truncated conical tip. These days, the spudcan diam-
to a 12.5 m diameter prototype spudcan. To facilitate
eter can be in excess of 20 m. This type of foundation
the monitoring of total vertical and pore pressures
is not custom-designed for a specific site condition
around the spudcan, the model spudcan has been
(Poulos 1988). As a result, the spudcans may experience
instrumented with a total of 5 pore pressure trans-
very deep penetration during installation, particularly
ducers and 4 total pressure transducers placed at the
in soft seabeds. Under static preload, penetrations of
top and bottom of the spudcan, as shown in Figure 2.
2 times spudcan diameter are commonly encountered
Pore pressure transducers were also installed in the
in the field (Craig & Chua 1991). The deep penetra-
soil just beneath the final penetration depth of the
tion of spudcan in impermeable sediments often
spudcan. The installation and extraction of the spud-
results in great difficulties in extracting the legs dur-
can was carried out by a loading actuators controlled
ing its removal. The extraction can be time consum-
by a closed-loop hydraulic servo-valve control sys-
ing which has economic consequence to the offshore
tem. In addition, potentiometers were employed to
industry.
monitor the penetration elevation of the spudcan, the
To date, relatively few studies have been carried out
ground surface settlements and water level. Through-
to investigate the phenomenon of suction developed
out the test, the water level was kept above the ground
at the spudcan base. Craig & Chua (1990) reported
surface. Figure 3 shows the photograph of the experi-
that for the extraction of shallow-embedded spudcans
mental setup.
in uniform soft clay, good adherence and sustainable
base suctions could develop. In view of this, a centrifuge
model study has been carried out at the National
2.2 Soil sample
University of Singapore to simulate the installation,
operation and extraction of a jackup spudcan in nor- The soil sample is normally consolidated clay consti-
mally consolidated clay. Particular attention is paid to tuted from Malaysian kaolin clay with a liquid limit
the development of total vertical and pore pressures of 80%, plastic limit of 35% and a specific gravity of
around the spudcan. The experimental procedure and 2.60. The dry clay powder was mixed with water
results are presented in this paper. to produce clay slurry at a water content of 120% or

517

Copyright 2005 Taylor & Francis Group plc, London, UK


Figure 2. Arrangement of sensors for the model spudcan
(all dimensions in mm).

Figure 1. Centrifuge model setup (all dimensions in mm).

1.5 times the liquid limit of the clay. After the clay
mixing process inside a vacuum container for 3 hours,
the clay slurry was then poured on top of a 30 mm thick
sand layer in the model container. The clay was then
subjected to a consolidation pressure of 20 kPa applied
in stages for 1 week. Pore pressure transducers were
installed in the clay after the preconsolidation process.
The model container was then placed on the cen-
trifuge to subject the soil to self-weight consolidation at Figure 3. Photograph of the experimental setup.
100 g resulting in a final clay sample of 370 mm depth.
After self-weight soil consolidation has been completed,
the centrifuge model was brought to rest and the test the cone penetration test with a cone factor of 11.8 for
assembly including the loading actuator attached with a typical test is shown in Figure 4.
the load cell and model spudcan, potentiometers and The 5 kPa strength recorded at the ground surface
cone penetration test actuator was assembly on the is due to the effect of 1 g consolidation under 20 kPa.
model container, as shown in Figure 1. The rate of increase in undrained shear strength of
During the test proper and prior to spudcan extrac- 1.43 kPa per m depth is fairly close to the value of
tion, cone penetration test was carried out in-flight at 1.48 kPa per m depth obtained by Liao (2001) for the
a distance of 16 cm away from the centre of the spud- same clay and consolidation procedure determined by
can. The undrained shear strength profile obtained from in-flight T-bar tests (Stewart & Randolph 1991).

518

Copyright 2005 Taylor & Francis Group plc, London, UK


Undrained shear strength prior to extraction (kPa) refers to the net force excluding the effective weight
0 5 10 15 20 25 30 35 40 of the spudcan. Figure 5 shows the variations of force,
0 penetration, total vertical and pore pressure at the spud-
2
can top and base as well as pore pressures in the soil
beneath the final penetration depth of the spudcan
4 with time. It should be noted that prototype time is
determined assuming soil consolidation is the domin-
6
ant event such that the prototype time is model time
8 multiplied by N2 where N is the gravitational field. Thus
Undrained shear
the prototype time is only applicable for the spudcan
Depth (m)

10 strength rate = 1.43 kPa/m


operation period. It does not apply to the spudcan instal-
12 lation and extraction processes and the time scale is
only used as a reference for comparison during spud-
14 can installation and extraction.
16
18 3.1 Installation

20
The spudcan in Test A was installed into the clay
using load control mode until it reached its final pene-
22 tration depth of 189 mm or 18.9 m in prototype scale.
This penetration depth is equivalent to about 1.5 times
Figure 4. Undrained shear strength profile. the spudcan diameter. In the installation, intermediate
stops were taken to help ensuring that the spudcan
2.3 Experimental procedure finally rest close to the target depth. The prototype
penetration resistance of 30.9 MN was maintained
The simulation mainly consisted of three stages i.e. until there was no further significant settlement of the
installation, operation, and extraction. The spudcan spudcan (Figs 5a, b). The total vertical and pore pres-
installation was carried out using load-controlled sures at the top of spudcan increased almost linearly
mode at a rate of 1 kPa/s until the spudcan reached the with penetration depth, as shown in Figures 5c, d. At
target depth of about 1.5 times the spudcan diameter. the base of the spudcan, the magnitudes of total vertical
Subsequently, the installation load was reduced to a and pore pressures were considerably larger than those
certain working load simulating the operation of the at the top (Figs 5e, f).
spudcan. At the end of the operational period, the In general, the maximum increase in pore pressure
extraction was performed using displacement- occurs at the center and decreases radially outwards
controlled mode at a rate of 1 mm/s. This rate ensures from the centre of the spudcan. Figure 5g reveals that
an undrained condition according to the dimension- positive excess pore pressures had generated in the
less velocity group defined by Finnie (1993). soil beneath the spudcan during installation but the
magnitude decreases with increasing depth beneath
(1) the spudcan base and radial distance from the spud-
can centre. A comparison of the total and pore pressure
magnitudes above and beneath the spudcan confirms
where v  velocity of the footing, B  diameter of the that an undrained loading condition prevails during
spudcan, and cv  the coefficient of consolidation. the installation process.
According to equation 1, substituting the diameter of
the model spudcan B  125 mm, consolidation coef- 3.2 Operation
ficient cv  40 m2/year and displacement rate v 
1 mm/s results in V  99 which is well above the limit The operation of the spudcan was initiated when the
for undrained condition. installation load was reduced to a certain working load.
In the present study, a ratio of 0.75 between the work-
ing load and installation load was adopted. In Test A,
3 EXPERIMENTAL RESULTS the operation load was maintained for a model time of
about 350 seconds which corresponded to a 53-day
The performance of the model spudcan during instal- operation period in the field (Fig. 5a). At the spudcan
lation, operation and extraction of Test A is described top, the installation-induced excess pore pressure grad-
in this section. The results will be presented in proto- ually dissipated (Fig. 5d) revealing the reconsolidation
type scale hereinafter, unless otherwise stated. It should of the heavily remolded soil. On the other hand,
be noted that the spudcan force presented hereinafter excess pore pressure at the spudcan base dissipated at

519

Copyright 2005 Taylor & Francis Group plc, London, UK


Figure 5. Time records of various responses (Test A).

520

Copyright 2005 Taylor & Francis Group plc, London, UK


a considerably higher rate (Fig. 5f). The total vertical This phenomenon would be further confirmed by
pressure transducers at the two sides of the spudcan conducting a half-spudcan test in the next stage of the
(Figs 5c,e) registered fairly constant values. This sug- experiment.
gests that effective stress has progressively developed
during the operation period.
4 UPLIFT RESISTANCE
3.3 Extraction
To further interpret the test results, the net uplift resist-
Extraction was carried out by releasing the compres- ance of the spudcan R is decomposed into two compon-
sive load on the spudcan and subsequently applying a ents i.e. (i) soil resistance above spudcan top, denoted
continuous upward movement to the spudcan (Fig. 5a). by R1, and (ii) resistance beneath the spudcan, denoted
In this case, a change in control mode from load con- by R2. It is noted that R is simply the load cell reading
trol to displacement control was executed beforehand. minus the effective self-weight of the spudcan. R1 can
The extraction created sudden increases in both total be obtained from the readings of total vertical pres-
vertical and pore pressures at the top of spudcan and sure transducers placed at the top of the spudcan
they reach the peak values after a 0.5 m upward dis- minus the hydrostatic pressure at the corresponding
placement (Figs 5c,d). On the other hand, instantan- elevation. It is postulated that the resistance beneath
eous reduction in both pressures takes place at the base the spudcan is mainly attributed to the suction at the
of spudcan immediately after the extraction (Fig. 5f). interface of spudcan base and the surrounding soil. R2
The amount of extraction-induced change in pressures is thus obtained from the negative excess pore pres-
at the base of spudcan is considerably higher than that sure readings at the spudcan base, i.e. the total pore
at the top. pressures at the base less the hydrostatic pressure. Since
The pressure change in term of pore pressure and there are at least two measurement points for both
total vertical pressure are of similar magnitude. This total vertical and pore pressures, the average values
suggests that the applied uplift force is translated mainly are calculated in proportion to the respective tributary
to the change in pore pressure rather than the change circular area of individual transducer over the total
in effective stress. The uplift resistance, registered by plan area of the spudcan.
the load cell, continues to increase as the spudcan moves Test B was conducted with identical procedures as
further up and reaches its maximum after a consider- Test A except that the prototype operation period
ably large displacement i.e. around 1.8 m or 15% of the increases from 53 days to 843 days. Figure 6 presents
spudcan diameter (Figs 5a, b). the net uplift resistance and its two components of
This maximum uplift resistance is termed as the both tests. It is evident that the breakout force of Test
breakout force. In this case, the magnitude of proto- A is significantly smaller than that of Test B. This is
type breakout force of 19.4 MN is about 63% of the mainly contributed to the significant increase in the
installation load of 30.9 MN. At nearly the same ele- base suction R2 for Test B. Thus the increase in net uplift
vation, the maximum drop in both total vertical and resistance R is mainly attributed to the increase in R2.
pore pressures also take place.
In terms of pore pressure, the maximum reduction
takes place at the centre and decreases radially out- Force (MN) Force (MN)
wards with a difference of about 100 kPa between the 0 5 10 15 20 0 5 10 15 20 25 30 35 40
centre and the edge. During extraction, the spudcan 0 0
base in fact experiences an artificial downward pres- 2 2 R1
R2
sure that develops due to the decrease in pore pressure 4 4 R
below the hydrostatic pressure. That is negative excess
Spudcan Depth (m)

6 6
pore pressure or suction has occurred at the spudcan
base. After reaching the breakout force, the uplift resist- 8 8
ance subsequently drops markedly whereas the total 10 10
vertical and pore pressures increase toward the hydro- 12 12
static pressure value. After complete extraction, a mass
14 14
of relatively stiff clay was observed stuck at the top of
the spudcan. This is also indicated by a residual proto- 16 16
type load of around 4 MN registered by the load cell. 18 18
This observed behavior seems to tally with Craig & 20 20
Chua (1990) which postulated that the drop in uplift Test A Test B
resistance takes place due to water entering the inter-
face between the soil and the spudcan with the adhe- Figure 6. Uplift resistance and its components at different
sive bond or suction at the interface being overcome. operation periods.

521

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 1. Summary of ultimate force. spudcan base reveal that significant negative pore
water pressure or suction had developed and the mag-
R1 R2 R nitude of suction is considerably larger for the spud-
Operation period can with a longer operation period. This finding
Test (days) (MN)
suggests that suction can be a major component of
A 53 9.5 8.5 19.4 breakout force during spudcan extraction. Further
B 843 12.2 21.8 36.0 centrifuge model tests are currently being conducted
to evaluate the contribution of suction in spudcan
extraction in greater detail.
Table 1 summarizes the ultimate forces generated
during extraction for Tests A and B. It is apparent ACKNOWLEDGEMENTS
that the resistance of the overlying soil against uplift
increases only slightly with operation period. On the The support of equipment from the Keppel Professor-
other hand, a fairly long operation period will result in ship fund is gratefully acknowledged. The first author
a significantly larger breakout force and suction at the wishes to thank the National University of Singapore
spudcan base. For fairly long operation period, the base for the award of a research scholarship for his doctoral
suction can contribute up to about 60% of the break- studies. The assistance of the laboratory officers in the
out force. This finding establishes that base suction is Geotechnical Centrifuge Laboratory of the National
a major contributor of the breakout force for spudcans University of Singapore in conducting centrifuge model
with long operation periods as the base suction is tests is gratefully appreciated.
found to significantly increase with operation period.

5 CONCLUSION REFERENCES

A centrifuge model study has been carried out to Craig, W.H. & Chua, K. 1990. Extraction forces for offshore
foundations under undrained loading. ASCE Journal of
examine the performance of an extensively instru- Geotechnical Engineering 116(5): 868884.
mented jackup spudcan during installation, operation Craig, W.H. & Chua, K. 1991. Large displacement perform-
and extraction in normally consolidated clay. The ance of jack-up spudcans. Proc. Centrifuge 91: 139144.
results of 2 tests with different spudcan operation Rotterdam: Balkema.
periods are reported in this paper. It is found that the Finnie, I.W.S. 1993. Performance of shallow foundations in
spudcan installation is essentially an undrained process. calcareous soils. Ph.D Thesis, University of Western
A much larger breakout force is required to extract the Australia, Perth.
spudcan with a longer operation period. The net break- Liao, B. 2001. Behaviour of piles subject to negative skin
out force can be decomposed into soil resistance above friction and vertical load. Master Thesis, National Uni-
versity of Singapore.
spudcan and resistance beneath the spudcan. It is estab- Poulos, H.G. 1988. Marine Geotechnics. London: Unwin
lished that the increase in soil resistance above the Hyman.
spudcan with operation period is much smaller than Stewart, D.P. & Randolph, M.F. 1991. A new site investigation
the corresponding increase in the resistance beneath tool for the centrifuge. Proc. Centrifuge 91: 461466.
the base. The pore pressure transducers placed at the Rotterdam: Balkema.

522

Copyright 2005 Taylor & Francis Group plc, London, UK


Numerical simulation of the breakout process of an
object at the ocean bottom

X.X. Zhou, Y.K. Chow & C.F. Leung


Centre for Offshore Research and Engineering, Department of Civil Engineering,
National University of Singapore, Singapore

ABSTRACT: This paper presents a numerical model to simulate the breakout process of an object at the
ocean bottom. In the present model, a tiny gap (its initial width can be zero) is assumed during the pullout
process. The fluid motion in the gap is assumed to be creeping flow. The seabed is assumed to be a linear elas-
tic porous medium in which Darcys law is valid. The results from the present model are compared to those from
Mei et al. (1985) and they are found to agree well.

1 INTRODUCTION 2 DEVELOPMENT OF MODEL FOR LONG


PLATE PROBLEM
The breakout process of objects lying at the ocean
bottom is a subject of great interest in offshore engin- It is assumed that there is a tiny gap between the base
eering. Sawicki & Mierczynski (2003) suggested that of the object and the seabed surface, and water fills
the whole breakout process can be separated into two this gap. The width of the gap is infinitesimal at the
stages. In the first stage, the object is lying on the initial stage. During the pullout process, the gap will
seabed surface and no gap exists between them. In the enlarge, and meanwhile water outside of the gap will
second stage, there is a tiny gap between the base of flow in. As a result, the negative excess water pres-
object and the seabed surface. To date, relative few sure (suction) will be generated in the gap.
numerical methods are available to simulate the In this article, we will examine a long plate prob-
breakout process. In Sawicki & Mierczynski (2003), lem in order to compare the present model with Mei
a one-dimensional model is proposed for the first et al. (1985). The same method can also be applied to
stage of the breakout process. Foda (1982) proposed objects with other geometric shapes, such as circular
an analytical model to describe the second stage of plate or conic footing. The definition sketch of the
the breakout phenomenon. Despite the many approxi- long plate problem is shown in Figure 1.
mations adopted to simplify the problem, Fodas the-
ory is still rather complex. Mei et al. (1985) proposed 2.1 Governing equation for fluid motion
an alternative analytical theory to simulate the second in tiny gap
stage of the breakout process. In Mei et al.s model,
The Reynolds number R for the fluid motion in the
the deformation of soil skeleton was not taken into
gap between the base of the long plate and the seabed
account. Since the above models are based on very
surface can be expressed as
ideal situations (simple soil properties of seabed, sim-
ple geometrical shape of object, etc.), it is meaningful
to develop a numerical model to simulate more com- (1)
plex cases.
The object of this paper is to develop a numerical
model to simulate the second stage of the breakout where f denotes the fluid density; u denotes the char-
process. The proposed model in this paper can be acteristic fluid speed (here we use the horizontal vel-
used in more complex situations than those of Foda ocity of the fluid in the gap);  denotes the fluid
(1982) and Mei et al. (1985). viscosity and h(t) denotes the width of the gap at time t.

523

Copyright 2005 Taylor & Francis Group plc, London, UK


L: width of long plate
y
h(t): width of gap (not to scale)
F: pullout force
F G': submerged weight of long plate
p(x,t): suction in gap
L2 L2
water Long plate water

Seabed surface p(x,t) h(t)


x
G'
Water Saturated
Linear Elastic Porous Seabed

Figure 1. Definition sketch of breakout process of a long plate.

Mei et al. (1985) introduced a non-dimensional In addition, the continuity equation of fluid motion
variable (w/kL)1/3h(t) to denote the width of the gap
at time t, where L is the width of the long plate; k is (5)
the permeability of the seabed, w is the unit weight
of fluid, and other notations are the same as those in
the above paragraph. The results given in Mei et al. is also used, where u and v are horizontal and vertical
(1985) indicated that for common problems, h(t) is velocities of fluid, respectively.
very small comparing to L before the breakout phe- From Eqs. (3), (4) and (5), we can obtain
nomenon occurs (in our computations, we find usu-
ally h(t) 10
2 m). In the above condition, it can be
easily estimated that the Reynolds number of the fluid
motion in the gap is much less than 1.0. According to
classic theorems of fluid dynamics, the inertia of the
moving fluid in the tiny gap is insignificant under this
circumstance. The fluid motion in the tiny gap may be
assumed to be creeping flow and described by
(6)
(2)
In the above two equations, by deleting A(x, t), and
replacing u, v and p by u, v and p, respectively,
where u denotes the velocity vector of the fluid; p yields
denotes the water pressure in the tiny gap; and 
denotes the fluid viscosity.
In addition, the distribution of suction in the gap (7)
may be assumed to be uniform in the vertical direc-
tion as a result of the fact that the width of the gap is
so small comparing to the width of the long plate. where u, v denote respectively the horizontal and
vertical velocities of fluid motion just above the
2.2 Computational equation seabed surface, and p denotes the suction just above
the seabed surface. Meanwhile we use u
, v
to
The seabed is assumed to be a linear elastic porous denote respectively the horizontal and vertical veloci-
medium, in which fluid motion obeys Darcys law. ties of fluid motion just below the seabed surface, and
Solving Eq. (2), yields p
denote the excess pore water pressure just below
the seabed surface. In the seabed, Darcys law
(3)
(8)
where A(x,t) and B(x,t) are arbitrary functions of x and t.
At the base of the long plate, the no-slip condition is assumed to be valid, where k is the permeability of
for fluid is adopted: the seabed, w is the unit weight of fluid, vsh and vsv
are the horizontal and vertical velocities of soil skel-
(4) eton at the seabed surface, respectively.

524

Copyright 2005 Taylor & Francis Group plc, London, UK


At the seabed surface, the following relationships In Eq. (15), it is observed that if h(t) is zero, dh/dt
are assumed: should also be zero. This means that the object can
not be pulled out if the width of the gap is equal to
(9) zero under the condition that the seabed is imperme-
able and rigid. Of course, here the failure of the
By using Eqs. (7), (8) and (9), we can obtain seabed has not been taken into account. This conclu-
sion agrees with lubrication theory (Batchelor 1967).

2.3 Numerical procedures


By using finite difference method, Eq. (12) can be
written as

(10)

where vf is the velocity of fluid motion from the


seabed into the gap; (t) is the average vertical dis-
placement of the seabed surface. (t) is obtained from

(11) (16)

On the right hand of Eq. (10), the first term is


much less than the other terms. Thus it can be neg-
lected and Eq. (10) can be simplified as Equilibrium equation of the long plate

(17)

is needed, where p(x, t) is suction in the gap, F(t) is


(12) pullout force and G is the submerged weight per unit
length of the long plate.
In addition to Eqs. (16) and (17), another relation
In Eq. (12), the average value of the displacements of
is also utilized to solve the current problem: given
seabed surface (t) is used to simplify the computational
p(x, t), vf (x, t) and (t) can be determined according to
process. In the following, let us consider a limit situation
Biots theory. In this study, the finite element method
of Eq. (12). If the seabed is assumed to be impermeable
is used to solve Biots equation.
and rigid, we have k 0, (t) 0, vsv 0 and vf 0.
The iterative method may be adopted to solve the
As a result, Eq. (12) can be reduced to
proposed model. Firstly, the initial values of vf (x, t)
and vsh(t) at time t are assumed. Then through Eqs.
(13) (16) and (17), p(x, t) can be obtained. Using this p(x, t),
the new values of vf (x, t) and vsh(t) can be obtained
using Biots equation. After some iterations, the accur-
acy of the solution can be ensured.
Applying the boundary conditions at the outside
and middle of the long plate
2.4 Comparison with Mei et al. (1985)
(14) Mei et al. (1985) proposed an analytical model for the
breakout process of the long plate problem. In their
model, the seabed is assumed to be rigid but porous,
to Eq. (13), we can obtain the distribution of suction and a boundary condition proposed by Saffman
in the gap as (1971) was adopted at the seabed surface to consider
the effect of the seabed to the breakout process. This
boundary condition uses an empirical constant ,
(15) which depends on the structure of the porous material,
but is largely independent of the viscosity of the fluid.

525

Copyright 2005 Taylor & Francis Group plc, London, UK


In this study, it is assumed that the seabed is an
1012
elastic body and the soil does not fail in the pullout
process. This phenomenon only occurs when the top F
G' mk
=10
3 =10
12
soil of the seabed is strong enough and the pullout 1011 gw L2 gw L2
force is not very large. Therefore, here we only con-
F
G'

gw Lt
sider quite stiff soil. 1010 =10
2

m
The current problem can be expressed as the fol- gw L2
lowing functional:
109 F
G'
=10
1
gw L2
(18)
108
103 104 105
E
where t is the time needed for the pullout process, F gw L
and G are respectively the applied pullout force and 1015
the submerged weight per meter of the long plate, L is mk
=10
16
the width of the long plate, H is the depth of the per- 1014 F
G'
=10
3 gw L2
meable soil from the seabed surface, k is the permea- gw L2
bility of the seabed, E and v are respectively the 1013

gw Lt
F
G'
Youngs modulus and the Poissons ratio of the seabed, =10
2

m
gw L2
and w and  are the unit weight and the viscosity of 1012
the fluid. F
G'
By using the well-known Buckingham Pi theory, 1011 =10
1
gw L2
Eq. (18) can be rewritten as
1010
103 104 105
E
gw L
(19)
Figure 2. Effect of the Youngs modulus of the seabed on
the pullout time of the long plate.
It is observed that in Eq. (19), the number of the vari-
ables governing this problem has been reduced from ten
to seven. Simplicity of the problem is thus achieved.
According to the above analysis of the effect of the
Eq. (19) can be changed as
normalized E on the pullout time, we let E/wL be
equal to 105 in the present model to compare with
Mei et al. (1985) model. This value can ensure the
(20) seabed condition here we used is the same as that in
Mei et al. (1985) model. In Mei et al. (1985) model,
the parameter  of the seabed may be estimated in a
The effect of the Youngs modulus of the seabed on range according to the permeability of the seabed
the pullout time is shown in Figure 2, where v is equal because obviously other factors will affect , but not
to 0.3, and H/L is equal to 10, which can ensure the only the soil permeability. Mei et al. (1985) estimated
assumption that the permeable seabed is deep enough in that for very fine sand (k  10
5 m/s),  is about
this problem. It is found that the effect of the normal- 0.001 and for coarse sand (k  10
2 m/s),  is about
ized E on the pullout time depends on the normalized 0.1. For clay, the permeability which is about 10
9 m/s,
net pullout force (F
G)/wL2. When (F
G)/wL2 is the  value should be much smaller than that of sand.
greater than 10
2, this effect can be very prominent. Therefore, here for soils with k  10
4 m/s, 10
6 m/s
However, when (F
G)/wL2 is smaller than 10
3, this and 10
9 m/s, we may estimate that  are in the ranges
effect can be neglected. The reason is that the large nor- of [0.001, 0.01], [0, 0.005] and [0, 0.002], respectively.
malized net pullout force can induce larger deformation The comparison results are shown in Figure 3. In
of the seabed surface and this deformation may affect Figure 3, the horizontal axis represents the normal-
the pullout process. From Figure 2, we can also find that ized net pullout force and the vertical axis represents
when the normalized E is larger enough, the pullout the normalized pullout time. And the value of the nor-
time will not increase when the normalized E increases malized permeability k/wL2 will affect the relation-
further. Otherwise, from our computations we find that ship between the pullout force and the pullout time.
the Poissons ratio v can also affect the pullout time From Figure 3, it is evident that the present model
when the normalized E is small. But when the normal- agrees well with Mei et al. (1985) model for the present
ized E is large enough, this effect can be neglected. long plate problem.

526

Copyright 2005 Taylor & Francis Group plc, London, UK


1012 model agree well with those from Mei et al. (1985).
mk The present numerical model can also be applied to
=10
12 more complex situations (for example, an object is of
gw L2
1011 more complex geometric shape, the properties of the
seabed are not homogeneous, and the pullout force
gw Lt

1010 varies with time, etc.), but it should be verified by


m

well-controlled experiments.
109 Present scheme
-4
Mei et al. (1985), k=10 ms, =0.001 and 0.01
-6
Mei et al. (1985), k=10 ms, =0 and 0.005

108 REFERENCES
10-4 10-3 10-2 10-1
F
G' Batchelor, G.K. 1967. Introduction to Fluid Dynamics.
gw L2 Cambridge: Cambridge University Press.
Foda, M.A. 1982. On the extrication of large objects from
1015 the ocean bottom (the breakout phenomenon). Journal of
Fluid Mechanics 117: 211231.
mk Lamb, H. 1945. Hydrodynamics. New York: Dover publi-
=10
16
1014 gw L2 cations.
Mei, C.C., Yeung, R.W. & Liu, K.F. 1985. Lifting of a large
object from a porous seabed. Journal of Fluid Mechanics
1013 152: 203215.
gw Lt
m

Saffman, P.G. 1971. On the boundary condition at the sur-


face of a porous medium. Studies in Applied Mathematics
1012 Present scheme 50: 93101.
-4
Mei et al. (1985), k=10 , =0.001 and 0.01
-9
Mei et al. (1985), k=10 , =0 and 0.002
Sawicki, A. & Mierczynski, J. 2003. Mechancis of the
breakout phenomenon. Computers and Geotechnics 30:
1011 -4
10 10-3 10-2 10-1 231243.
F
G' Sherman, S.F. 1990. Viscous Flow, International Edition.
gw L2 McGraw-Hill series in mechanical engineering,
McGraw-Hill Publishing Company.
Figure 3. Comparison results between present scheme and Vesic, A.S. 1971. Breakout resistance of object embedded in
Mei et al. (1985) for the long plate problem. ocean bottom. ASCE Journal of Soil Mechanics and
Foundations Division 97: 11831205.

3 CONCLUSIONS

A numerical model for simulating the breakout process


of an object at the ocean bottom has been developed.
For the long plate problem, the results from the present

527

Copyright 2005 Taylor & Francis Group plc, London, UK


Spudcan penetration in sand overlying clay

K.L. Teh, C.F. Leung & Y.K. Chow


Centre for Offshore Research and Engineering, Department of Civil Engineering,
National University of Singapore, Singapore

ABSTRACT: Punch through failures of spudcan foundations of mobile jack-up rigs during installation have
been reported in practice. In the present study, centrifuge model tests were conducted to investigate the phe-
nomenon of punch-through of a spudcan installed in a dense sand overlying normally consolidated clay. The
results of tests conducted on 10-m prototype diameter spudcans revealed a distinct transition in soil failure
mechanism from the overlying stiff soil to the soft clay below. The validity of conventional bearing capacity
theories is evaluated against the test results.

1 INTRODUCTION showed that for the case with 2.6 m sand thickness, no
distinctive punch-through failure was observed whereas
Sutheast Asia is an active region for offshore oil and for the cases of 7 m and 9.5 m sand thickness, punch-
gas explorations. The soil conditions in Southeast Asia through failures occurred within the first 2 m of pen-
can be quite different from those in other areas. As an etration. A comparison of the test results with theory
example, Castleberry & Prebaharan (1985) reported revealed that the Hanna & Meyerhof (1980) method
that the Sunda Shelf consists of a layer of stiff soil under-estimated the observed failure loads for all three
crust of 5 to 15 m thick found at about 6 m below the cases. Furthermore, a sand plug with a depth virtually
seabed. Lying beneath the stiff soil is soft clay. This equal to the initial thickness was observed beneath
rather peculiar subsurface profile may lead to punch- the footing. Craig & Chua (1990) hence suggested an
through failure of mobile jack-up rig spudcan founda- additional component of resistance around the plug
tion during installation. In fact, Osbourne & Paisley and footing periphery to be incorporated in the bear-
(2002) revealed that on average there is 1 incident of ing capacity analysis when the spudcan penetration
unexpected spudcan punch-through failure occurring goes beyond the interface.
in Southeast Asia region annually. This statistic Kenny & Andrawes (1997) conducted a series of 1-g
inevitably highlights the need to understand the tests with a 0.12-m diameter (B) model footing. The
mechanism of spudcan punch-through failure in layered upper sand layer had a of 48.5 and a variable thick-
soils in order to provide a better prediction of the spud- ness (H) up to 3 times the footing diameter while the
can behaviour installed in such geological formations. underlying clay had a cu of 10.6 kPa. The test results
showed no obvious punch-through failure even
though the footing settlement was up to 60% of the
2 LITERATURE REVIEW footing diameter. Comparisons of test results with
theories revealed that the load-dispersion angle  (see
In the offshore industry, the recommendations for the Figure 1) increased with H/B.
design of spudcans installed in layered soils are given Michalowski & Shi (1995) presented design charts
in SNAME (2002). The design method generally for strip footing on sand overlying clay using limit
follows the bearing capacity calculations of shall- analysis based on kinematics approach. The solution
ow foundation in layered soils proposed by Jacobsen was based on two defined failure mechanisms, namely
et al. (1977) and Hanna & Meyerhof (1980). rigid block collapse mechanism of two-layer founda-
Craig & Chua (1990) presented the results of a series tion soil and failure mechanism with continual defor-
of centrifuge tests on model spudcans (14 m in proto- mation field in clay. The limit analysis showed that
type diameter) installed in dense sand (with friction the bearing capacity increased with increase in the
angle of 38) overlying a stiff clay layer (with uni- underlying clay strength. However, when the bearing
form undrained shear strength cu of 41 to 45 kPa). The capacity of the overlying sand was achieved, further
results of three tests with different sand thicknesses increase in clay strength did not improve the bearing

529

Copyright 2005 Taylor & Francis Group plc, London, UK


q Hydraulic
cylinder

LVDT
Cone
penetrometer

SAND

CLAY, qclay

Figure 1. Schematic of the load spreading mechanism.


Load
cell

capacity except for the case of very stiff clay with a


relatively thin upper sand layer as compared to the
footing width.
Burd & Frydman (1997) carried out a finite ele- Sand
ment analysis of a strip footing on sand overlying
clay. The results of their analysis revealed a similar PPT
trend in terms of the development of bearing capacity NC Clay
and limiting pressures as reported by Michalowski &
Shi (1995). Furthermore, the analysis showed that the TSC
load-dispersion angle  increased with but reduced
with cu. Interestingly,  was found to be insensitive to (a)
H/B, which is in contrast to the observations by Kenny &
Andrawes (1997) and later Okamura et al. (1998).
A series of centrifuge data was reported by
Okamura et al. (1997) using a relatively small footing
diameter of 2 m to 6 m (in prototype scale). For the
10

case of sand overlying clay with uniform undrained 80


shear strength, it appeared that as H/B increased, the 100 mm
punch-through failure occurred at a higher penetra-
tion ratio (D/B where D is the spudcan penetration). (b)
In addition,  was observed to reduce with increasing
Figure 2a. Centrifuge model set-up; b. Enlarged view of
strength of the clay layer and increased with H/B. model spudcan.
Although many studies had been carried out to
investigate the bearing capacity in layered soil, the
mechanism of spudcan punch through is still not well In all cases, the spudcan was installed using a servo-
understood and many contrasting observations were valve controlled hydraulic cylinder and its penetration
reported by various researches. In view of this, a cen- is measured by a long-travel potentiometer attached to
trifuge model study has been carried out in the pres- the hydraulic cylinder. The bearing load is measured
ent study to evaluate the mechanism of punch through through a load cell attached between the hydraulic
of spudcan installed in dense sand overlying clay. cylinder and the spudcan. Three short-travel poten-
tiometers were employed to measure the ground surface
displacement during spudcan penetration. Several pore
3 CENTRIFUGE MODEL TEST pressure transducers (PPT) and total stress cell (TSC)
were installed in the soil (Figure 2) to monitor the stress
3.1 Experimental set up changes in the soil during the tests. A set of miniature
cone penetration test equipment was also mounted on
A schematic of the experimental set up is shown in the loading frame to measure the soil strength in-flight.
Figure 2. All tests were performed at 100 g. A strong-
box model container of 500 mm in diameter and
3.2 Soil preparation
400 mm in height is used. The thickness of the overly-
ing sand layer ranges from 5 m to 10.5 m in prototype Malaysian Kaolin clay and Japanese Toyoura sand
scale. The spudcan has a prototype diameter of 10 m. were used in the study. The properties of the soils are

530

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 1. Properties of Malaysian Kaolin clay (after Goh 3.3 Test procedure
2003).
Once the saturation process of the sand layer was com-
Specific gravity (Gs) 2.60 pleted, several PPTs were installed in the soil through
Liquid limit (wL) 80 openings at the side wall of the strong box with the
Plastic limit (wP) 35 help of semi-circular guides (Konig et al. 1994). The
Coefficient of consolidation (at 100 kPa) 40 m2/yr potentiometers were then placed, loading frame was
Coefficient of permeability (at 100 kPa) 2.0 10
8 m/s mounted, and the model spudcan and miniature cone
Angle of internal friction,  23 penetrometer were attached to the model setup. The
Modified Cam-clay parameters:
model container was then again subjected to 100 g until
M 0.9
" 0.244 the clay was fully consolidated under the new sand
K 0.053 surcharge. Once the consolidation was completed, the
N 3.35 valve controlling the bottom drainage was closed before
the cone penetration test was carried out at a rate of
2.5 mm/s prior to spudcan penetration. The spudcan
Table 2. Properties of Toyoura sand. was installed using a load control mode at the rate of
0.15 kN/s. By adopting this loading rate, the slowest
Specific gravity 2.65 resultant penetration rate in the linear region (before
Average particle size (mm) 0.2 punch-through failure) is 0.3 mm/s. This rate of pene-
Uniformity coefficient 1.3 tration ensures an undrained installation process based
Dmax, Dmin (mm) 0.3, 0.115
D50, D10 (mm) 0.2, 0.163
on the velocity group parameter proposed by Finnie
Range of density (kg/m3) 13351645 (1993). At the same time, the bearing load based on
the soil response was captured continuously.

shown in Tables 1 and 2. The clay slurry at 1.5 times liq-


uid limit was first mixed under vacuum for 5 hours. The 4 RESULTS AND DISCUSSIONS
clay was then consolidated under 100 g for 8 hours with
two-way drainage. After spinning down, the clay sur- All the test results are presented at prototype scale.
face was leveled and the sand was rained in using spot Four tests with sand layer thickness of 5 m, 7 m, 7.7 m
type air pluviation method with a constant drop height and 10.5 m were performed. The bearing stress is taken
of 600 mm. The travel path and speed were controlled to as the bearing load divided by the cross-sectional area
achieve the desired relative density. The effects of the of the spudcan. The bearing stress-normalised pene-
travel path and speed on the relative density of Toyoura tration depth responses for the four tests are presented
sand are reported by Eio (2003). The strength of the in Figure 3. It is evident that the bearing stress increases
sand layer can be obtained through a correlation chart with thickness of the overlying sand layer. For all the
based on the relationship between relative density and four tests, the bearing stress increases with normalised
friction angle (Ueno 2000). The relative density of penetration depth before the occurrence of spudcan
Toyoura sand recorded in the tests ranges from 88% to punch through, as indicated in Figure 3. It is interesting
95%. The friction angle of this dense sand ranges from to note that despite the large differences in the thick-
41 to 42.5. The sand was then saturated by applying ness of the overlying sand layer, the depth at which the
vacuum on the top and water was allowed to flow into spudcan punches through takes place within a narrow
the strong box under a control manner through a tube range of 10% to 12% of spudcan diameter beneath the
installed at an opening of the strong box wall at the ele- sand surface. However, a much larger loading pres-
vation of sand-clay interface. sure is required for the spudcan to punch through for
The undrained shear strength, cu, of the clay the test with the greatest thickness of overlying sand
obtained from cone penetration tests conducted in layer. The above observations can be attributed to a
flight prior to spudcan installation reveals that the combination of several factors. Firstly, a larger load-
clay is normally consolidated with cu increasing lin- ing pressure is required to penetrate a spudcan into a
early with depth. By adopting an effective unit weight thicker overlying sand layer as the stress bulb lies
of sand and clay as 10 kN/m3 and 6 kN/m3, respect- mostly within the sand layer. Secondly, the increase in
ively, the clay strength profile can be denoted by loading pressure is somewhat compensated by the
reduction in bearing stress at the sand-clay interface
(1) due to a larger area of load spread as the clay is lying
deeper down. Thirdly, the strength of the underlying
where clay becomes stronger under thicker overlying sand.
OCR  over-consolidation ratio of clay, and For the test configuration and soil properties in the
v  vertical effective overburden pressure. present study, the depth at which the spudcan punches

531

Copyright 2005 Taylor & Francis Group plc, London, UK


q (kPa)
800

700

600

500

400

300
B = 10 m
200 case 1 (H = 5 m; = 41)
case 2 (H = 7 m; = 42)
Punch-through case 3 (H = 7.7 m; = 42)
100
failure case 4 (H = 10.5 m; = 42)
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 1.1 1.2 1.3 1.4
D/B

Figure 3. Bearing pressure (q) versus normalized depth (D/B) curves.

through coincidentally falls with a narrow band due 2.2


to the combination of the above factors. Further stud-
2
ies are required to examine whether this phenomenon
applies to other test configurations and soil profiles 1.8
and properties. 1.6
The post punch-through spudcan behaviour is differ- Caquot & Kerisel
1.4
ent for the four tests. For tests with the smallest overly- (1953)
ing sand thickness, the spudcan only punches through a 1.2
H/B

distance of about 40% spudcan diameter after which Hansen


1
a larger loading pressure is required for the spudcan (1970)
0.8
to penetrate further. On the other hand, the spudcan
punches through a massive distance of over 1.3 times 0.6
spudcan diameter for the test with the thickest overlying 0.4
sand layer. This can be attributed to the larger loading
0.2
pressure required for the spudcan to punch through and tan ' mobilised /tan '
the underlying soft clay does not have sufficient bearing 0
capacity to resist the penetrating spudcan. 0 0.2 0.4 0.6 0.8 1
In the present study, a local shear failure is antici-
pated even at the beginning of the spudcan penetra- Figure 4. Relationship between normalized sand layer
thickness and mobilized strength of sand.
tion. Since the strength of the lower normally
consolidated clay layer is partly governed by the over-
burden pressure, the utilized bearing capacity is then where the bearing capacity factor Nq is given by
a function of H. As H increases, cu also increases rep-
resenting a higher bearing capacity. The bearing
capacity factor N of a circular footing under general (4)
shear failure is given by
Alternatively, Hansen (1970) gave
(2)
(5)
where q0 is the surface loading pressure, B is the foot-
ing diameter and  is the effective unit weight of the Using the above equations, the mobilized friction
soil. Caquot & Kerisel (1953) gave angle mobilized can be back calculated and shown in
(3) Figure 4. The results show that as H/B increases,
mobilized approaches . By extrapolating the curve

532

Copyright 2005 Taylor & Francis Group plc, London, UK


40 qu predicted/qu measured Dcr predicted/Dcr measured
H/B qsand/qclay 1.4 1.4
35 0.50 30.18 (1.0) (1.0)
0.70 26.30 (0.9) (0.9)
30 1.05 19.43 1.2 1.2

25 (1.0) 1 1
(0.9)

20
(0.5)
(0.69) 0.8 0.8
15

10 0.6 0.6
(q/qu = 0.4)
5
q/qclay 0.4 0.4
0
1 1.5 2 2.5 3 3.5 4 4.5
0.2 Failure stress of punch-through 0.2
Figure 5. Relationship of load spread angle and measured Critical penetration depth of punch-through
loading pressure. 0 0
0 0.2 0.4 0.6 0.8 1 1.2
H/B
connecting experimental data, it is apparent that gen-
eral shear failure is only fully mobilized when H/B of
Figure 6. Comparison of failure bearing stress and depth
a layered soil system has a magnitude exceeding 2. between experimental data and Hanna & Meyerhof method.
Figure 1 illustrates the schematic of a load spread-
ing mechanism during the initial spudcan penetration
to the state of punch-through failure. The load spread failure mechanism to deep foundation failure mecha-
angle  can be evaluated as follow: nism. At this stage, the intrusion of the sand plug into
the clay layer generates a localized pressure bulb. The
influence zone of the localized pressure bulb usually
(6) extends to some depth behind the footing tip and
hence the resistance behind the footing tip affects the
generated bearing stress (Randolph & Houlsby 1984).
in which qclay is calculated using the bearing capacity During the early intrusion, due to the distance between
formula proposed by Davis & Booker (1973) for clay footing tip and the base of sand layer is the closest, a
with strength increasing linearly with depth. huge force is required to overcome the resistance. As
A back analysis of the test data is carried out to cor- the spudcan penetrates deeper into the underlying soft
relate  with the ratio of bearing pressure to ultimate soil, the soil resistance reduces. The soil behind the
punch-through pressure, q/qu and the bearing ratio, footing tip starts to fail and flow easily. This process
qsand/qclay. The correlations shown in Figure 5 reveal continues until the effect of the sand layer is completely
that as the loading pressure approaches the ultimate diminished.
punch through stress, the load spread angle increases As H/B increases, the spudcan punch through load
denoting a reduction in bearing stress at the sand-clay increases, as reported earlier. During the onset of
interface. This allows the layered system to support spudcan punch-through, there is a sudden loss of
more loads and helps to explain the increase of bearing resistance and the imposed stresses cannot be
observed bearing stresses prior to failure point. In supported. In view of the fact that the bearing capacity
addition, as overlying sand thickness increases, qsand/ of the lower weak soil layer is not capable of resisting
qclay reduces due to the higher cu of the clay layer. the load, this results in a rapid large penetration until
From Figure 5,  is found lower as H/B increases. a level where the soil at greater depth with higher
This leads to higher load concentration beneath the strength is able to sustain the load. In other words, the
spudcan and subsequently higher mobilization of the higher failure load leads to more catastrophic conse-
strength of the sand layer, as indicated in Figure 4. As quence of punch-through failure to the jack-up rig.
a result, a higher ultimate punch through resistance is Figure 6 summarizes the comparisons in terms of
observed. punch-through failure stress (qu) and the correspon-
The authors postulated that the occurrence of the ding penetration depth (Dcr) between the present
second peak after further penetration below the punch- experimental data and that predicted using Hanna &
through elevation as observed in Figure 3 except that Meyerhof method. In general, Hanna and Meyerhof
for case 1 is due to the transition of shallow foundation method under-predicted qu for all cases except for the

533

Copyright 2005 Taylor & Francis Group plc, London, UK


case of H/B  1.05. This exception requires further Caquot, A. & Kerisel, J. 1953. Sue le terme de surface dans
investigation. le calcul des foundations en milieu pulvelent, Proc. III
Int. Conf. Mech. Found. Eng., Vol. 1:336337.
Castleberry, J.P. & Prebaharan, N. 1985. Clay crust of the
Sunda Shelf A hazard to jack-up operations. Proc. Eight
5 CONCLUSION Southeast Asian Geotechnical Conference, Vol. 1:4048.
Craig, W.H. & Chua, K. 1990. Deep penetration of spudcan
The performance of a spudcan installed in dense sand foundations on sand and clay. Gotechnique 40, No.
overlying normally consolidated clay has been exam- 4:541556.
ined in the present centrifuge model study. Four tests Davis, E.H. & Booker, J.R. 1973. The effect of increasing
with different thickness of upper sand layer are stud- strength with depth on the bearing capacity of clays.
ied and the results are compared with the existing Gotechnique 23, No. 4:551563.
Eio, T.L. 2003. Sand preparation for geotechnical model
theories. The following findings are established:
test. BEng thesis, National University of Singapore,
1. For the thickness of dense sand ranging from half Singapore.
to 1 spudcan diameter overlying soft clay in the Finnie, I.M.S. 1993. Performance of shallow foundation in
present study, punch through of spudcan has been calcareous soil. Ph.D. thesis, University of Western
Australia, Australia.
observed in all 4 tests during spudcan installation.
Goh, T.L. 2003. Stabilisation of an excavation by an embed-
The ultimate punch through load increases with ded improved soil layer. Ph.D. thesis, National University
thickness of overlying sand but the elevation of of Singapore, Singapore.
punch through falls within a relatively narrow Hanna, A.M & Meyerhof, G.G. 1980. Design chart for ulti-
band of between 10% and 12% of spudcan diame- mate bearing capacity of foundation on sand overlying
ter below the sand surface. soft clay. Can. Geotech. J . 17:300303.
2. The punch-through failure is more drastic for tests Hansen, J.B. 1970. A revised and extended formula for bear-
with thicker overlying sand as the spudcan is noted ing capacity. Danish Geotechnical Institute Bulletin
to punch through a great depth of over 1.3 times No. 28.
Jacobsen, M., Christensen, K.V. & Sorensen, C.S. 1977.
spudcan diameter. Appropriate action must be taken
Gennemlokning af tynde sandlag. Vag og Vattenbyggaren
when installing spudcan under such scenario. 8/9:2325.
3. Back analysis of the present centrifuge test data Kenny, M.J. & Andrawes, K.Z. 1997. The bearing capacity
reveals that the load spread angle is not a constant, of footings on a sand layer overlying soft clay.
as commonly assumed in conventional bearing Gotechnique 47, No. 2:339345.
capacity theories in layered soils, but increases with Konig, D., Jessberger, H.L., Bolton, G., Phillips, R., Bagge, G.,
qsand/qclay and decreases with H/B. Renzi, R. & Garnier, J. 1994. Pore pressure measurement
4. It is apparent that conventional bearing capacity during centrifuge model test: Experience of five labora-
theories cannot predict the spudcan penetration tories. Centrifuge 94: 101108.
Michalowski, R.L. & Shi, L. 1995. Bearing capacity of foot-
process in sand over clay. Further studies are cur-
ings over two-layer foundations soils. Journal of
rently in progress to examine the mechanism of Geotechnical Engineering, ASCE, 121(5):421428.
punch through of spudcan with an aim to propose Okamura, M., Takemura, J. & Kimura, T. 1997. Centrifuge
a rational design approach of the above problem. model test on bearing capacity and deformation of sand
layer overlying clay. Soils and Foundations, Vol. 37, No.
1:7388.
ACKNOWLEDGEMENT Okamura, M., Takemura, J. & Kimura, T. 1998. Bearing
capacity predictions of sand layer overlying clay based on
The authors wish to acknowledge the assistance of limit equilibrium methods. Soils and Foundations,
laboratory officers of the Geotechnical Centrifuge Vol. 38, No. 1:181194.
Laboratory at the National University of Singapore. Osbourne, J.J. & Paisley, J.M. 2002. SE Asia jack-up punch-
The present study is funded by the National throughs: The way forward?. Proc. Int. Conf. Offshore
University of Singapore research grant RP 264-000- Site Investigation and Geotechnics: 301306.
Randolph, M.F. & Houlsby, G.T. 1984. The limiting pressure
167-112 Punch through of mobile jack-up. on a circular pile loaded laterally in cohesive soil.
Gotechnique 34, No. 4:613623.
SNAME. 2002. Guidelines for site specific assessment of
REFERENCES mobile jackup units. Society of Naval Architects and
Marine Engineers, Technical and Research Bulletin
Burd, H.J. & Frydman, S. 1997. Bearing capacity of plane- 5-5A, New Jersey.
strain footings on layered soils. Can. Geotech. J. 34: Ueno, K. 2000. Methods for preparation of sand samples.
241253. Centrifuge 98, Vol. 2:10471055.

534

Copyright 2005 Taylor & Francis Group plc, London, UK


Punch-through of spudcan foundations in two-layer clay

M.S. Hossain & Y. Hu


Curtin University of Technology, Australia

M.F. Randolph
The University of Western Australia, Australia

D.J. White
The University of Cambridge, UK

ABSTRACT: Spudcan punch-through failure continues to be a major cause of foundation failure of offshore
jack-up rigs. The resulting damage to the rig can range from minor structural damage of its leg and jacking
mechanism to complete loss of the rig. Most punch-through failures happen during the jacking up and preload-
ing and in stratified soil profiles with a relatively thin layer of sand or stiff clay overlying a weaker layer. Punch-
through causes a reduction in bearing resistance with depth, causing instability during the load-controlled
jacking-up process. To reveal the failure mechanism during punch-through, model spudcan foundation tests
have been conducted on a two-layer clay sample (strong over weak) in a drum centrifuge. Half-spudcan model
tests were carried against a transparent window to visualise the soil flow mechanisms around the spudcan dur-
ing penetration. Particle Image Velocimetry (PIV) analysis was used to track the soil particle movements and
hence obtain precise details of the failure mechanisms. Full-spudcan tests were also conducted to measure
the vertical load-penetration responses. The study shows that punch-through failure and associated softening
response are directly linked to the relative thickness of the top clay layer to the foundation diameter, and also
the strength ratio between the two soil layers. A punching failure through the upper layer was observed as shear
zones developed from the spudcan rim down to the layer interface. A soil plug was carried down beneath
the spudcan. The thickness of the soil plug was 80% of the initial top layer thickness, and the cavity formed
above the spudcan remained open until the spudcan fully penetrated into the soft layer. Softening penetration
resistance profiles were observed.

1 INTRODUCTION rate of accidents than most engineering structures,


with a quoted accident rate of 2.6% annually, of which
Spudcan foundations are commonly used to support a quarter were due to punch-through (LeBlanc 1981,
jack-up rigs in offshore exploratory drilling, tempor- McClelland et al. 1981, National Research Council
ary production and maintenance work. Before the 1981, Rapoport & Young 1987, Sharples et al. 1989).
commencement of the jack-up operation, the spud- Most punch-through failures happen during pre-
cans are proof loaded by preloading to obtain an loading in stratified soil profiles with a relatively thin
acceptable margin of safety against the anticipated layer of sand or strong clay overlying a weaker layer
extreme storm loading. Typically preloading is accom- (Baglioni et al. 1982, McClelland et al. 1981, Young
plished by pumping seawater into the hull, called bal- et al. 1984, Craig & Higham 1985, Rapoport & Young
last preloading. The ballast is then discharged and 1987, Senner 1993). A typical load-penetration response
the hull is raised further to provide an adequate air- for these conditions is illustrated in Figure 1. When
gap for subsequent operation. During this initial the preload reaches the initial peak (A in Fig. 1), the
preloading, the footings of a jack-up are essentially foundation needs to penetrate to point A to equili-
subjected to vertical loading. brate the current load at point A. During uncontrolled
Jack-up rigs are now being considered for use in punch-through, rotation of the rig will throw addi-
deeper waters and more hostile environments, and for tional load onto the penetrating leg, and hence the
long term functions in hydrocarbon development pro- penetration will actually exceed the distance AA
ject. However, mobile jack-up rigs suffer a much higher shown in Figure 1.

535

Copyright 2005 Taylor & Francis Group plc, London, UK


Bearing Capacity with the space above the footing assumed to be occu-
partial intended pied by a smooth rigid shaft. Since soil could not
floating preload preload flow-back on top of the footing, the lower resistance
elevated
A B
at depth permitted by a backflow mechanism was not
strong layer
captured.
The offshore industry design guidelines (SNAME
Footing Penetration

1997) suggest approximate solutions from Brown &


weak layer
Meyerhof (1969) and Meyerhof & Hanna (1978) for
"punch-through"
strong over weak clay. These solutions have been com-
monly accepted by practitioners, although it has not
been possible to confirm the assumed failure mechan-
A' isms in the soil from field measurements.
B'
In this study, a series of centrifuge model tests on
strong clay overlying soft clay were carried out to
investigate the conditions leading to catastrophic
punch-through failure. Half-spudcan penetration tests
were performed against a transparent window to cap-
Figure 1. Punch-through failure during preload (replicated
from McClelland et al. 1981).
ture soil flow images. Subsequently, Particle Image
Velocimetry (PIV) analysis allowed detailed quantifi-
cation of the soil flow mechanisms. Separate tests
At most punch-through sites, rapid penetration is were conducted on a full-spudcan to measure the
experienced by only one leg, resulting in structural associated load-penetration responses.
damage to the leg or, worse, overturning of the rig.
Punch-through can also occur in normally consoli-
dated or lightly overconsolidated clays and silts due to 2 EXPERIMENTAL METHODOLOGY
partial consolidation occurring during any delays in
preloading, and the development of a localised strong The tests described here were performed in the drum
crust of soil just beneath the spudcan (McClelland centrifuge at the University of Western Australia
et al. 1981, Young et al. 1984). (Stewart et al. 1998). The outer channel of the drum
Stratified soil deposits with potential for punch- centrifuge is 300 mm high and 200 mm deep (radi-
through failure can be categorised into two groups: ally). In this study pre-consolidated soil specimens
(a) strong clay over weaker clay and (b) sand over were placed into a strongbox fitted within the chan-
clay. This paper considers the former stratification; nel. A strongbox (258  80  160 mm) with a plexi-
geological processes that may lead to this situation glass window was built to allow observation of soil
were described by McClelland et al. (1981). flow. Images were captured in-flight by a high reso-
The literature dealing with experimental and analyt- lution (2592  1944 pixels) digital still camera. The
ical investigations into foundation response in two-layer experimental arrangement is shown in Figure 2.
clay soils is quite extensive. However, model tests have The half-spudcan was 60 mm diameter; however,
mostly been conducted on the laboratory floor (at 1 g), the width of the box was insufficient for a full spud-
and thus do not maintain similarity of the strength ratio can of the same diameter, so the full-spudcan diam-
su/vo, where vo is the effective overburden stress. eter was reduced to 30 mm. Both spudcans were made
The tests will therefore not capture any tendency for from duraluminium, with a 13 shallow conical under-
soil to flow around the spudcan. Also most of these side profile (included angle of 154) and a 76 protrud-
studies (experimental or analytical) were limited to ing spigot. The half-spudcan permitted penetration
either footings resting on the surface of the soil, with tightly against the window of the strongbox so that the
the assumption that the displacement of the footing soil flow could be observed. All tests were conducted
prior to ultimate load was very small or for strip foot- at 100 g, corresponding to a prototype diameter for
ings. Therefore, application to spudcan foundations the half-spudcan of 6 m and for the full-spudcan of
undergoing large penetration may not be appropriate. 3 m. The spudcan model sizes were small enough to
Wang & Carter (2002) presented bearing capacity avoid boundary effects.
factors for a circular footing on strong clay over soft
clay, in terms of the thickness ratio, H/D, of the top
2.1 Preparation of clay specimen
layer (where D is the footing diameter) and the shear
strength ratio, sub/sut (with subscripts b and t repre- The footing tests were performed on heavily over-
senting the bottom and top layer respectively). Large consolidated specimens of Speswhite kaolin (LL 
deformation finite element analyses were undertaken, 61%; PI  27%; Gs  2.6; cv  2 m2/year). The soil
so that the changing geometry was correctly modelled, was consolidated in two-stages. In the first stage, a

536

Copyright 2005 Taylor & Francis Group plc, London, UK


Cradle White painted region soil ingress between the window and the spudcan. Tests
Mini video camera Viewing window were performed at a penetration rate of 0.05 mm/s to
Camera
Ring channel Lamp Plane strain permit more images to be taken, allowing precise
chamber analysis of the deformation. However, undrained con-
ditions were still maintained, according to the criter-
ion defined by Finnie (1993).
Coloured flock powder was used to mark each layer
in the plane of the window. The flock powder added
texture to the white kaolin specimen in order to allow
PIV analysis of the captured images. PIV operates by
tracking the texture (i.e. the spatial variation of bright-
ness) of a mesh of patches through a series of images.
Control markers allowed the variation in image scale
due to fisheye, refraction and other effects to be
accounted for using close range photogrammetry
Inside of window (White et al. 2003).
Radial line: direction of g aligned radially Typical deformation mechanisms at three stages of
punch-through are shown in Figures 35, for the layer
Figure 2. View of drum-channel set-up. combination of H/Dhalf  0.65 (Test T3). In Figure 3,
the spudcan is embedded by d/Dhalf  0.2, so is half a
Table 1. Summary of tested soil profiles. diameter above the initial position of the layer bound-
ary. Figure 3a shows the distorted layers, and Figure
H/D 3b shows the instantaneous velocity field calculated
Test Clay deposit Soil strength (Dfull, Dhalf) using PIV. Initially, penetration was accompanied by a
shallow heave mechanism, but by d/Dhalf  0.2, the
T1 0.60, 0.30
T2 sut  32 kPa 1.00, 0.50
soil surface is stationary and the spudcan penetrated
T3 stiff/soft sub  14 kPa 1.30, 0.65 by punching into the lower soft layer (Fig. 3b). A soil
T4 sub/sut  0.44 1.70, 0.85 plug with the shape of a truncated cone (marked
T5 1.83, 0.92 AABB in Fig. 3a) formed in the upper layer and
T6 2.20, 1.10 moved down with the spudcan. Meanwhile, a transi-
T7 uniform, soft 14 kPa 0 tional zone of shear deformation (ABC or ABC)
T8 uniform, stiff 32 kPa # moves downwards and sideways.
The deformation mechanism at this embedment is
clarified by the contours of normalised horizontal and
homogeneous and de-aired slurry at a moisture content vertical velocity, V/Vspudcan. Closely spaced velocity
of 120% was consolidated one-dimensionally under a contours indicate regions of high strain rate. Below
maximum pressure of 400 kPa. the spudcan rim, the horizontal velocity was higher
A further slurry layer was added on top of the con- within the soft layer than immediately below the
solidated first layer. This second layer was consoli- spudcan in the stiffer soil (Fig. 3c). The soil plug was
dated under a maximum pressure of 100 kPa. The soil undergoing vertical compression, since the vertical
models were then prepared by inverting the sample velocity decreases down the centreline. The shear
(so that the stronger layer was on top) and cutting the deformation in the transition zone is shown by the
pre-consolidated clay to fit the strongbox. The test reduction in vertical velocity from 0.7 Vspudcan to
details and soil strengths are summarized in Table 1. 0.1 Vspudcan between AB and AC (Fig. 3d).
One half-spudcan, one full-spudcan and two T-bar Figure 4 shows the soil flow at a penetration of
tests were conducted in each test specimen. The T-bar d/Dhalf  0.4. The mechanism had changed by this
tests were to measure the soil shear strengths (using a deeper embedment. Comparison of Figures 4b and 3c
T-bar factor of 10.5), and were undertaken at a rate of shows the increased horizontal deformation in the
1 mm/s to ensure undrained behaviour. During each soft layer at the deeper embedment. At d/Dhalf  0.4,
test, a free water depth of about 30 mm was maintained the soil plug was virtually rigid, with the vertical
above the soil surface. V/Vspudcan  0.9 within 0.7 Dhalf of the spudcan
base (Fig. 4c) and negligible horizontal movement
(Fig. 4b). Also, more distinct shear planes had formed
3 OBSERVED DEFORMATION MECHANISMS along AB and AB, indicated by the contours of
vertical V/Vspudcan  0.4 and 0.9 which are parallel
The half-spudcan was penetrated into the soil with the and separated by only 4 mm along AB and
central flat side pushed against the window to prevent AB. Also, distributed shear continued within ABC

537

Copyright 2005 Taylor & Francis Group plc, London, UK


(and ABC) as indicated by the fan of contours of
V/Vspudcan 0.4 (Fig. 4c).
Figure 5 shows the spudcan at an embedment of
d/Dhalf  1.4, with the soil plug fully penetrated into
the soft layer. Comparing Figures 3 and 5, there is
negligible change in the thickness of the soil plug
after it has formed in the stiff layer, and thus created
the shear plane AB (and AB). Initially, shear plane
(a) Annotated image from centrifuge test
AB is in the stiff soil, mobilising sut. In Figure 5,
0 shear plane AB involves shearing on the surface of
the soil plug, mobilising sub. Therefore, as the soil
10 plug penetrates into the soft layer, the mean resistance
A
created by shearing along AB decreases from sut to sub.
20

30

40
C
50 B

0 10 20 30 40 50 60 70 80
(b) Velocity field from PIV analysis (axes in mm)

(a) Annotated image from centrifuge test


0
0
20
20
40
40
V/Vspudcan = 0.1
60
V/Vspudcan = 0.2
60
80
V/Vspudcan = 0.1
V/Vspudcan = 0.1 80
100
100
120
-80 -60 -40 -20 0 20 40 60 80 120
(c) Contours of normalised horizontal velocity, -80 -60 -40 -20 0 20 40 60 80
V/Vspudcan (axes in mm) (b) Contours of normalised horizontal velocity,
V/Vspudcan (axes in mm)

0 0

20 20

40 40

60 60

80 80

100 V/Vspudcan 0.8 0.1 100


V/Vspudcan 0.9 0.1

120 120
-80 -60 -40 -20 0 20 40 60 80 -80 -60 -40 -20 0 20 40 60 80
(d) Contours of normalised vertical velocity, (c) Contours of normalised vertical velocity,
V/Vspudcan (axes in mm) V/Vspudcan (axes in mm)

Figure 3. Penetration mechanism for Test T3 (H/Dhalf  Figure 4. Penetration mechanism for Test T3 (H/Dhalf 
0.65) embedded at d/Dhalf  0.2. 0.65) embedded at d/Dhalf  0.4.

538

Copyright 2005 Taylor & Francis Group plc, London, UK


The transitional shear zone is eventually eliminated forcing a flow-round mechanism through the stronger
with further penetration. The soil plug thickness was top layer (which would fill the cavity, thus limiting
80% of the initial top layer thickness. Hcavity). More tests are needed to verify and quantify
For all the other cases, with H/Dhalf  0.3 to 1.1, this effect.
the soil plugs showed similar shapes, with the transi-
tional zone gradually disappearing as the spudcan
penetrated further through the top layer. 4 BEARING CAPACITY PROFILES
The cavity formed above the spudcan remains open
until the spudcan has passed fully through the top layer Figure 6 shows the bearing resistance during Tests T1,
(Fig. 5). This is because of the high strength ratio of T5 and T6 for which H/Dfull varied from 0.6 to 2.2. A
the upper soil layer. The cavity depth in Figure 5 sharp peak in resistance followed by softening is
exceeds the top layer thickness, reaching H  0.4D. observed in each case, indicating that steady penetra-
This can be compared with maximum stable cavity tion is not reached in the upper layer. This behaviour
depths found for uniform soil (Hossain et al. 2004), can be linked to the observed change in failure mech-
expressed using Equation 1, where Hcavity is the stable anism. In Test T3 (H/Dhalf  0.65, Figs 35), punch-
cavity depth, su is the soil undrained shear strength through began at a very shallow penetration depth
(32 kPa for the upper soil layer), and  is the (effective) (d/Dhalf 0.2) when shear zone ABC formed (Fig.
soil unit weight (7 kN/m3 for kaolin clay under water). 3b), and penetration was permitted by horizontal dis-
placement of the lower layer (Fig. 3c).
This observation can be related to the penetration
(1) resistance profile from Test T1 (H/Dfull  0.6) shown
in Figure 6, where the peak bearing capacity was
reached immediately after the spudcan shoulder
For these tests, Hcavity/Dhalf  0.86 using Equation 1, touched the soil surface at d/Dfull  0.12. Beyond
which is lower than the value of Hcavity/Dhalf  1.1 this depth, the bearing resistance of the spudcan
observed in Test T3 (Fig. 5). This cavity extends decreased with further embedment as the penetration
through the top layer into the lower layer. mechanism moved into the softer layer. The steady
Two cases were tested for which cavity collapse deep bearing resistance was reached at an embedment
was expected within the top layer, having H/Dhalf  of H/Dfull (Fig. 6), which corresponds to full
0.92 (T5) and 1.1 (T6) (0.86, see Table 1). However, embedment of the soil plug in the soft layer (Fig. 5).
an open cavity through the top layer was observed for For Tests T5 and T6, the peak bearing resistance,
these cases. qmax, was reached when the spudcan was 1.5Dfull
This observation suggests that the cavity depth in above the original layer boundary (Fig. 6). Beyond
conditions with a strong layer overlying a weaker this depth, punch-through reduced the bearing resist-
layer may be greater than for homogeneous soil con- ance. The values of qmax are compared to the upper
ditions. This is logical, since in the double layer soil, layer thickness ratio, H/Dfull in Figure 7. Values are
the spudcan can displace the lower soft soil layer lat- also shown for the limiting cases of H/Dfull  0 and
erally (as shown in Figs 3c and 4b) more easily than H/Dfull  $ (associated with uniform soft and uni-
form strong soil respectively).

q (kPa)
0 50 100 150 200 250 300 350
0.0

0.5 stiff clay


H/Dfull = 0.6
soft clay
1.0

1.5
d/Dfull

stiff clay
H/Dfull = 1.83
soft clay
2.0 stiff clay
H/Dfull = 2.2
soft clay
2.5

3.0

3.5
Figure 5. Soil deformation for Test T3 (H/Dhalf  0.65)
embedded at d/Dhalf  1.4. Figure 6. Bearing resistance profiles from full-spudcan tests.

539

Copyright 2005 Taylor & Francis Group plc, London, UK


400 Uniform (H = ) soil strength ratio was sub/sut  0.44 and the top layer
thickness (H/Dhalf) varied from 0.3 to 1.1 for the half
350
spudcan, with H/Dfull values of 0.6 to 2.2 for the full
This study
300 spudcan.
q = sut (2.93H/D + 7sub/sut) Punch-through failure occurs when a peak in the
250
qmax (kPa)

(Meyerhof & Hanna 1978) penetration resistance is reached. This is triggered by


q = sut (3H/D + 6.05sub/sut) the transition to a failure mechanism with shear zones
200 (Brown & Meyerhof 1969)
extending from the spudcan shoulder to the base of
150 the strong layer. A soil plug with the shape of a trun-
Uniform (H = 0)
100 Test cated cone forms in the upper layer below the spud-
sub / sut = 0.44 can and moves down as the spudcan penetrates further.
50 A transitional shear zone surrounding the soil plug
disappears during further penetration, and is not evi-
0
0.0 0.5 1.0 1.5 2.0 2.5 dent once the spudcan is fully embedded in the soft
H/Dfull layer. The soil plug has a depth of 80% of the initial
thickness of the top layer.
Figure 7. Peak bearing resistance, qmax, for sub/sut  0.44.
Surface heave occurs until the shear zone reaches
the bottom of the upper soil layer. Once the shear plane
is formed, the spudcan penetration displaces the lower
Peak resistance increases linearly with H/D, soft soil more easily than the upper strong soil, which
although there is a discontinuity between H/D  0.5 stops further surface heave.
and 0. The layered qmax results intersect the value The peak bearing resistance and the depth at which
for uniform stiff soil at H/D  2.25 indicating that this occurs increases with increasing thickness of the
for sub/sut  0.44, the peak bearing resistance in top soil layer. The depth of cavity formed above the
the upper layer is reduced by punch-through if spudcan exceeds that predicted for a single uniform
H/D 2.25. soil layer, because the spudcan can displace the bottom
The predicted values following Brown and soft soil more easily than force flow-back through the
Meyerhof (1969) and Meyerhof and Hanna (1978) for upper strong soil.
sub/sut  0.44 are also shown. Good agreement is It should be noted that only one strength ratio has
found, even though these prediction methods were been studied so far. Further tests are planned with
developed for circular surface footings. The peak different strength ratios in order to evaluate the effect
resistances found in this study occur at penetration of the strength ratio between the upper and lower soil
depths with an open cavity above the spudcan. These layers.
cases are equivalent to a circular footing. However, if
back flow occurs in the top soil layer, a different
response would occur since a shorter soil plug would ACKNOWLEDGEMENTS
be pushed into the lower layer.
A steady bearing resistance that exceeds the This research was supported by the Australian
homogenous value of 10sub is reached at the end of Research Council through an ARC Discovery Grant
each test (Fig. 6). The additional resistance arises (A00105806). This support is gratefully acknow-
from shear on the sides of the soil plug, and increases ledged. The experiments could not have been per-
with H/D, which governs the thickness of the plug. formed without the support of the Centre for Offshore
The gentle decrease in bearing resistance from qmax to Foundation Systems (COFS) and especially the drum
the steady deep value in Tests T5 and T6 reflects the centrifuge Technician Mr. Bart Thompson.
reduction in mobilised strength on the soil plug from
sut to sub.
REFERENCES

5 CONCLUSIONS Baglioni, V.P., Chow, G.S. & Endley, S.N. 1982. Jack-up
foundation stability in stratified soil profiles. Proc. 14th
Punch-through failure of spudcans penetrating through Offshore Technology Conference, Houston, Texas,
363369, OTC 4409.
strong clay overlying softer clay has been investigated Brown, J.D. & Meyerhof, G.G. 1969. Experimental study of
by centrifuge modelling. Half-spudcan models were bearing capacity in layered clays. Proc. 7th Int. Conf. on
used to examine the deformation mechanisms using Soil Mech. and Found. Eng, 2: 4551.
PIV image analysis. Full spudcan tests were used to Craig, W.H. & Higham, M.D. 1985. The application of cen-
obtain profiles of spudcan penetration resistance. The trifugal modelling to the design of jack-up rig foundations.

540

Copyright 2005 Taylor & Francis Group plc, London, UK


Offshore Site Investigation, 293305. London: Graham & Senner, D.W.F. 1993. Analysis of long term jack-up rig
Trotman. foundation performance. Offshore Site Investigation and
Finnie, I.M.S. 1993. Performance of shallow foundations in Foundation Behaviour, Netherlands, 28: 691716.
calcareous soils. PhD Thesis, Univ. of Western Australia. Sharples, B.P.M., Bennett, W.T. & Trickey, J.C. 1989. Risk
Hossain, M.S., Hu, Y., Randolph, M.F. & White, D.J. 2004. analysis of jack-up rigs. 2nd Int. Conf. on the Jack-up
Limiting cavity depth for spudcan foundations penetrat- Drilling Platform: Design, Construction and Operation,
ing clay. Submitted to Gotechnique for review. City University, London, 101123.
LeBlanc, L. 1981. Tracing the causes of rig mishaps. Offshore, SNAME 1997. Guidelines for site specific assessment of
March, 5162. mobile jack-up units. Soc. Naval Architects and Marine
McClelland, B., Young, A.G. & Remmes, B.D. 1981. Avoiding Engineers, Technical & Research Bulletin 5-5A, New
jack-up rig foundation failures. Proc. Symp. Geotech. Jersey.
Aspects of Coastal & Offshore Structures, Bangkok, Stewart, D.P., Boyle, R.S. & Randolph, M.F. 1998.
137157. Experience with a new drum centrifuge. Proc. Int. Conf.
Meyerhof, G.G. & Hanna, A.M. 1978. Ultimate bearing Centrifuge 98, Tokyo, 1: 3540.
capacity of foundations of layered soils under inclined Wang, C.X. & Carter, J.P. 2002. Deep penetration of strip
loads. Can. Geotech. J. 15: 565572. and circular footings into layered clays. Int. J. Geomech.
National Research Council, Marine Board Committee on 2(2): 205232.
Assessment of Safety of OCS Activities 1981. Safety and White, D.J., Take, W.A. & Bolton, M.D. 2003. Soil deforma-
Offshore Oil, Washington, D.C., National Academy Press. tion measurement using particle image velocimetry (PIV)
Rapoport, V. & Young, A.G. 1987. Foundation performance and photogrammetry. Gotechnique 53(7): 619631.
of jack-up drilling units, analysis of case histories. Int. Young, A.G., Remmes, B.D. & Meyer, B.J. 1984. Foundation
Conf. on Mobile Offshore Structures, City University, performance of offshore jack-up drilling rigs. J. Geotech.
London, 461476. Eng. Div., ASCE 110(7): 841859.

541

Copyright 2005 Taylor & Francis Group plc, London, UK


Influence of jack-up operation adjacent to a piled structure

D.P. Stewart
Golder Associates, Perth, Australia

ABSTRACT: When jack-up rigs are operated in relatively close proximity to a fixed offshore structure, the
jack-up footings may be positioned close to the piles that support the fixed structure. This paper describes a cen-
trifuge-based experimental study into the effect of penetration and extraction of jack-up footings adjacent to a pile.
The experiments involved penetration and extraction of a model jack-up footing adjacent to a single instrumented
pile followed by lateral loading of the pile at the head. The results show that at relatively close footing-pile spacing:
(i) pile head stiffness reduced to about 60 to 75% of the undisturbed value, and (ii) lateral soil resistance and stiffness
reduced to a minimum of about 40% of the undisturbed values. It was found that the most significant degradation in
pile response occurred at a footing-pile clear spacing of less than about 0.75 times the jack-up footing diameter.

1 INTRODUCTION

Jack-up rigs are often operated in relatively close weaker pile loading
proximity to a fixed offshore structure, typically dur- jack-up leg & footing response compared
to fully intact soil
ing a rig work-over. During these operations, the foot-
ings for the jack-up may be positioned relatively close
to the piles that support the fixed structure. This oper-
ation can give rise to several effects, such as:
remoulded soil
1 Inducing bending moments in the piles of the fixed
structure due to soil movement caused by penetra-
tion and extraction of the jack-up footings. This
has been studied by Siciliano et al. (1990) and is
not considered further here. intact soil pile
2 A longer-term reduction in pile lateral load carry-
ing characteristics due to the presence of a zone of
remoulded soil and possibly also a footprint or Figure 1. Effect of zone of remoulded soil and crater resulting
crater next to the pile, Figure 1. This was the main from jack-up footing on lateral load response of adjacent pile.
focus of the study described in this paper.
The work reported here was an experimental study The drum centrifuge features a channel for sample
carried out on the drum centrifuge at The University containment that is 300 mm in width (measured verti-
of Western Australia. This paper describes some of cally) and 200 mm in depth (measured radially), with
the experimental equipment and methods and pre- a diameter of 1.2 m. Its maximum rotational speed is
sents a summary of the main results. 850 rpm, giving a peak effective acceleration of 400 g
for a 150 mm deep sample. The channel rotates about
a vertical axis so that the radial centrifugal accelera-
2 EXPERIMENTAL EQUIPMENT tion field acts horizontally.
The key feature of the centrifuge is a central tool
2.1 Drum centrifuge table that may be stopped and started independently
from the sample containment channel, achieved with
Experiments were undertaken on the geotechnical two concentric drive shafts. This provides the ability
drum centrifuge at The University of Western Australia. to stop the central tool table, modify or change the
A complete description of the facility is given by model, and then carry out further experiments, without
Stewart et al. (1998), and a brief outline follows. stopping the channel. The tool table has actuators to

543

Copyright 2005 Taylor & Francis Group plc, London, UK


enable movement in three directions: vertical, radial allowed to set and gain strength with the centrifuge
and circumferential. spinning.
The experiments were undertaken at a scale of The sample preparation procedure resulted in clay
1:190, so that linear dimensions were reduced by a samples that were overconsolidated over their full depth.
factor of 190 and a centrifugal acceleration of 190
times the earths gravity was applied.
3.2 Pile & footing testing
2.2 Key equipment The model piles were installed on diametrically oppo-
Two identical model piles were fabricated to represent site sides of the channel, connected to the data acqui-
prototype steel piles of approximately 1.5 m diameter. sition system, and the channel re-accelerated to 190 g.
The model piles were fabricated from drawn brass After re-consolidation of the sample, the first pile
tubing of 7.14 mm outside diameter and 0.37 mm wall was loaded laterally at 26 mm (4.94 m or 3.2d, where
thickness and were strain gauged to measure bending d is the pile diameter) above the clay surface, incor-
moment at 11 positions along their length. After porating several unload-reload loops. This test repre-
strain gauging, the piles were coated with a water- sented a base case, with no disturbance from an adjacent
proof sealant and covered with heatshink plastic tub- footing. The model jack-up footing was then pene-
ing. The base of each pile was sealed with a flat shoe. trated to a depth of 45 mm (8.55 m or 0.75D, where D
Fine wiring from the strain gauges was passed through is the footing diameter) and extracted, at a clear spac-
the centre of the hollow section and gathered into a ing of 15 mm (0.25D) from the second pile. During
cable leading from the top of the pile and into the data the footing penetration and extraction, the pile was
acquisition system. The piles were calibrated in bend- not restrained at the head and the pile movement was
ing using dead weights. not recorded. Following extraction of the footing, the
The calculated bending stiffness (EI) of the model pile was loaded laterally directly towards the jack-up
piles, when scaled to prototype units, is about footprint, in the same manner as the base case test.
5890 MNm2, which may be compared to a bending After testing the two piles, a series of T-bar pen-
stiffness of 6300 MNm2 for a 1.5 m diameter steel etrometer tests (Stewart & Randolph 1991) using a
tubular pile with 25 mm wall thickness. The outside 20 mm long and 5 mm diameter bar were performed
diameter of the model piles, including the waterproof to measure the undrained shear strength of the sam-
coatings, is equivalent to 1558 mm. The scaled plastic ple. The channel was then stopped, the model pile
moment of the model pile is approximately 58 MNm. moved to a new location and the process repeated.
A model jack-up footing was fabricated to repre- The jack-up footing was generally installed in-line
sent a nominal 12 m diameter prototype footing. The with the direction of pile loading, although this was
60 mm (11.4 m) diameter footing had a flat base and varied in some tests. The testing programme is sum-
top, with a 6 mm (1.1 m) high conical base projection marised in Table 1.
and was 9 mm (1.7 m) thick. The footing was attached Each time after the channel was stopped, the sam-
to a load cell so that the load required to penetrate the ple was allowed to re-consolidate overnight before
soil could be measured. testing the piles. T-bar penetrometer tests were per-
formed after each series of pile tests.

3 EXPERIMENTAL PROCEDURE 4 RESULTS

3.1 Soil sample preparation 4.1 Soil strength


Two series of experiments were conducted, with Typical results from T-bar penetration tests are pre-
slightly different strength profiles. The clay samples sented in Figures 2 and 3. The results show that the
were initially prepared as a slurry at about twice the samples were relatively uniform in strength, espe-
liquid limit, placed into the drum centrifuge channel cially over the upper half. The undrained shear
and then normally consolidated under its own self- strength profiles can be represented approximately
weight at an acceleration of 300 g. over the upper part of the samples by:
For the first series of tests, 30 mm thickness of
8  1.8 kPa/m for Series 1
sand was then sprayed onto the surface of the sample
8  1.5 kPa/m for Series 2.
and consolidation allowed before removing the sand.
For the second series of tests, it was attempted to A stronger band of material was evident in the
form a stronger surface layer over the clay by placing middle of the sample, presumably caused by drying
a 10 mm thick slurry comprising 98% kaolin and 2% during the preparation stages. However, the stronger
high early strength Portland cement. The slurry was band was present uniformly around the sample. The

544

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 1. Summary of testing programme. Undrained shear strength (kPa)
-60 -40 -20 0 20 40 60
Clear spacing Remarks and 0
Test between pile pile loading
series Test no. and footing direction
5
1 P1 Base case, no extraction penetration
footprint

Depth (m)
P2 0.25D Toward footprint 10
P3 0.75D Toward footprint
P4 0.5D Toward footprint
P5 0.25D At 90 to footprint 15
P6 1D Toward footprint
P8 0.25D Toward shallower
footprint 20

2 Q1 0.25D Toward footprint


Q2 Base case, no 25
footprint
Q3 1D Toward footprint Figure 3. T-bar penetrometer test results for Series 2 with
Q4 0.5D Toward footprint a stronger crust.
Q5 0.5D At 45 to footprint
Q6 0.75D Toward footprint
Q8 0.5D At 90 footprint
1.0
D  Footing diameter.
Normalised strength

0.8

Undrained shear strength (kPa)


0.6
-60 -40 -20 0 20 40 60
0
0.4
This study, without crust
5 0.2 This study, with crust
extraction penetration
Siciliano et al. (1990)
Depth (m)

10 0.0
0 1 2 3
Radial distance from footing centre / footing
15 diameter

Figure 4. Strength measurements within and near to a


20
footprint.

25
4.2 Remoulding due to footing penetration
Figure 2. T-bar penetrometer test results for Series 1 To gain an appreciation for the amount of strength
without a crust. reduction due to extreme remoulding, a cyclic T-bar
test was undertaken at one location, where the T-bar
was moved up and down over a test interval until a
results show strengths during extraction of the T-bar near constant resistance was achieved. This test indi-
that are about 65 to 75% of the penetration strengths, cated a fully remoulded strength that was about 50%
due to remoulding. of the peak strength.
For the sample prepared with a stronger crust, the At one location after penetration and extraction of
effect of the crust on the T-bar penetration resistance the jack-up footing, a series of T-bar penetrometer
is relatively minor. Excavation of the sample at the tests were carried out at various radial distances across
end of the test revealed that the cemented layer was and extending away from the footprint. The results of
less than about 5 mm thick and was overlain by about these tests are summarized in Figure 4 as the strength
5 mm of weak uncemented soil. The T-bar is 5 mm in normalized by that measured at a location that was not
diameter and would not be expected to reveal the true influenced by a footprint plotted against the radial
strength of the crust. distance from the centre of the footing. Data is shown

545

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 2. Approximate radial extent of remoulded zone around 3.0
a footing inferred from calculated collapse mechanisms.
base case
Approximate 2.5
inferred 1.0D
extent of

Lateral load (MN)


2.0
Footing remoulded
Reference type zone Comments
1.5 0.25D
Hu et al. Skirted 0.85D Finite element
(1999)
Hu and Shallow 0.9D Finite element 1.0
Randolph
(1998a)
0.5
Hu and Jack-up 0.75D Large strain
Randolph finite element
(1998b) 0.0
Martin and Shallow 0.95D LB Uniform strength 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6
Randolph 1.3D UB Lateral deflection of pile head (m)
(2001) Shallow 0.8D LB Strength increasing
1.05D UB with depth
Plate 0.75D LB Uniform strength Figure 5. Pile head load-deflection response for Series 1
0.9D UB tests. Clear spacing between pile and jack-up footing is shown.

LB  Lower bound plasticity calculation, UB  Upper bound


plasticity calculation. effect of the footprint, with an increase in deflection
of only 10 to 20% at a spacing of 1.0D.
The test data also showed an accompanying increase
for depths of 0.25D and 0.5D. Data presented by in maximum bending moment as the clear spacing
Siciliano et al. (1990) is also plotted for comparison. between the pile and the footing reduced. For the same
The results show that for the samples used in these load applied to the pile, the maximum bending
experiments, the strength at the centre of the footprint moment increased by a factor of up to about 20% at a
reduced to as low as 25% of the intact strength near to clear spacing of 0.25D. This response was observed
the surface for the sample with a crust. However, in because the weaker remoulded soil provided less resist-
general the strength reduced to between about 40% ance to pile movement near to the surface, thus forcing
and 80% of the intact strength at the centre of the the maximum bending moment to occur at greater
footprint. There is little reduction in strength shown at depth.
a distance of about 1.5 to 2D from the centre of the At a given clear spacing, piles loaded directly
footprint, although the main influence is confined to towards the footprint give a slightly softer response
within a distance of about 0.75D. This distance to the than piles loaded at right angles.
edge of the remoulded zone compares relatively well
with that expected from a review of the results of 4.4 Lateral soil resistance
large-strain finite element and upper and lower bound
calculations, as shown in Table 2. The test data were processed using the software
The summary results in Table 2 are the extent of the Mathcad to derive the lateral soil resistance distribu-
collapse mechanism calculated by the various meth- tion along the shaft of the pile. This was performed by
ods shown. This might be expected to correspond to curve fitting of the bending moment data followed by
the zone of most intensely remoulded soil around the double integration to derive deflection and double
footing. A zone of disturbed but less significantly differentiation to derive lateral soil resistance. After
weakened soil would be expected to extend around the trial and error, a 6th order polynomial curve was used
intensely remoulded zone, as shown in Figure 4. to fit to the bending moment data. Typical results
from the curve fitting are shown in Figure 6.
Figure 6 indicates the good fit that was generally
achieved to the bending moment data and the good
4.3 Pile response
agreement between the measured pile head deflection
The pile head load-deflection response measured in and the deflection predicted from double integration
several typical tests is presented in Figure 5. The of the bending moment curve. The distribution of lat-
results show that the pile test at a clear spacing of eral pressure over the upper part of the pile appears to
0.25D underwent deflections that were about 40 to be reasonable. However, the magnitude of pressure
60% higher than for the base case. Tests undertaken at over the lower part of the pile, below a depth of about
a greater clear spacing showed progressively less 90 mm, may be excessive and was not used further.

546

Copyright 2005 Taylor & Francis Group plc, London, UK


Pressure (kPa) Moment (Nmm) Deflection (mm) Limit pressure / su
-50 0 50 0 2000 4000 -5 0 5 10
0 2 4 6 8 10 12
-40
0
-20

0 1 this study

Depth / pile diameter


20 Broms (1964)
Depth (mm)

2
40

60 3
Murff & Hamilton (1993)
80
4 actual strength profile
100 uniform strength

120 5

Figure 6. Typical curve fitting result with experimental data Figure 8. Pile lateral limit pressure versus depth for the
shown as points, fitted and derived values shown as lines. Series 1 base case.

of limit pressure proposed by Broms (1964) is also


200 shown. A similar, although different, relationship is
base case
used for generating soft clay p-y curves by the
American Petroleum Institute (1993) method.
150
Also shown in Figure 8 are curves of limiting pres-
sure derived using the empirical method given by
1.0D
Pressure (kPa)

Murff & Hamilton (1993), developed from an upper


100
bound plasticity calculation. The curves are for a
0.25D rough pile in close contact with the soil. It is worth
50
noting that limiting pressures at depth of as high as
12su were obtained analytically by Randolph &
Houlsby (1984) and Broms (1964), while Reese et al.
0 (1975) obtained a value of 11su.
In summary, the distribution of lateral limit pressure
with depth appears reasonable, although slightly higher
-50 than expected. However, there may be scatter and dif-
0 0.05 0.1 0.15 0.2 0.25 0.3 ferences introduced through differences in the methods
Displacement / diameter of deriving undrained shear strength, the relative size
of the experimental equipment, measurement errors
Figure 7. Typical p-y curves at a depth of 2 pile diameters associated with the depth below the model soil surface,
for Series 1 tests. Clear spacing between the pile and jack-up the possibility that soil adhered fully around the model
footing is shown. pile rather than forming a gap behind the trailing face,
differences in installation and loading procedures and
This curve fitting process was carried out on each partial drainage in the model tests.
line of data recorded during the pile test, so that for The lateral limit pressure normalized by that meas-
any given point along the pile, a plot of derived lateral ured in the base case test is plotted versus depth for a
soil resistance (p) versus lateral deflection (y) could number of pile-footing offsets in Figure 9. This figure
be generated. Typical p-y curves calculated by this shows a general trend for the drop in limit pressure to
curve fitting technique are shown in Figure 7 at a reduce from a maximum at the soil surface. The foot-
depth of 2 pile diameters for several typical cases. ing was penetrated adjacent to the piles to a depth cor-
The results show a clearly defined limit pressure that responding to 5.5d. Inspection of the figure indicates
reduces due to the influence of penetration of the that for an offset of greater than about 0.75D, there is
footing. A reduction in the stiffness of the ground little or no reduction in limit pressure below the low-
reaction is also evident. est position of the footing. However, at closer offsets
A plot of the limit pressure versus depth is shown a reduction in limit pressure is likely to extend below
for the base case pile in Figure 8. This shows an the lowest position of the footing.
expected increase in limit pressure with depth before The data from Figure 9 are re-plotted as limit pres-
reaching a near constant value of about 11.5 times the sure versus offset in Figure 10. The trends exhibited
undrained shear strength at that depth. The distribution by the results are relatively similar to those indicated

547

Copyright 2005 Taylor & Francis Group plc, London, UK


Normalised limit pressure distance of about 2D appears unlikely. For the kaolin
0 0.2 0.4 0.6 0.8 1 1.2 clay used in these experiments, a sensitivity of about
0 2 (ratio of intact to remoulded strength) was meas-
0.75D 1.0D base case ured using the T-bar. For soils of greater sensitivity it
1
is possible that the zone of weakened soil will be
more closely confined around the footing.
Depth / pile diameter

The footprint and zone of weakened soil remaining


2 after extraction of the jack-up footing was found to
have an effect on the lateral load response of the pile.
3 For a pile located at a clear spacing of 0.25D from the
edge of the footing, the pile head stiffness dropped to
about 60 to 75% of the stiffness in undisturbed
4 ground. For a pile located at a clear spacing of 1D, the
0.25D
0.5D pile head stiffness dropped to about 80 to 90% of the
5 stiffness in undisturbed ground. At a given clear spac-
ing, piles loaded directly towards the footprint give a
Figure 9. Lateral limit pressure normalized by the base case slightly softer response than piles loaded at right angles.
limit pressure versus depth for Series 1 tests. Clear spacing The maximum bending moment in the piles for a
between the pile and jack-up footing is shown. given applied lateral load was slightly greater due to
the effect of the jack-up footing. It is possible that
1.2 where a stronger surface crust, greater strength at the
mudline occurs or where the soil has greater sensitiv-
1 ity, that a more significant increase in bending
Normalised limit pressure

moment could occur.


0.8 The lateral soil resistance was assessed by deriving
p-y curves from the test data. These curves indicated
0.6 both a reduction in soil reaction stiffness and a drop in
the lateral limit pressure. Lateral limit pressures
0.4 reduced to as low as about 40% of the undisturbed
values near the surface. This is a similar reduction to
0.2 that measured in a cyclic T-bar penetrometer test, car-
1d 2d 3d 4d ried out to assess the remoulded strength.
0
0 0.5 1 1.5 2
ACKNOWLEDGEMENTS
Clear offset / footing diameter

Figure 10. Lateral limit pressure normalized by the base The work described in this paper was jointly funded
case limit pressure versus footing offset for Series 1 tests. by Somehsa Geosciences and the Centre for Offshore
Depth in terms of pile diameters is shown. Foundation Systems and was carried out while the
author was a Lecturer at The University of Western
Australia. The experimental work was carried out
in Figure 4 for strength measurements within and under the direction of the author by Ee Sun Lim with
adjacent to the footprint. Note that a radial distance of assistance from Bart Thompson, Clem Ryan, Tuarn
0.5D shown in Figure 4 is equal to a clear offset dis- Brown and Wayne Galbraith.
tance of zero in Figure 10.

5 SUMMARY AND CONCLUSIONS REFERENCES

An investigation into the effect of operation of a jack- American Petroleum Institute 1993. Recommended practice
up rig on the lateral performance of piles supporting for planning, designing and constructing fixed offshore plat-
an adjacent piled structure has been described. forms Working stress design, API-RP-2A, 20th edition.
Broms, B.B. 1964. Lateral resistance of piles in cohesive
The results show a zone of relatively intense
soils. Journal of the Soil Mechanics and Foundations
remoulding around the footing and extending to a Division, Proc. ASCE, 90 (SM2), March: 2763.
radial distance from the centre of the footing of about Hu, Y. and Randolph, M.F. 1998. A practical numerical
0.75D. A zone of less intensely remoulded but still approach for large deformation problems in soil. Inter-
weakened soil then extends further, to a radial distance national Journal of Numerical and Analytical Methods in
of about 1.5D to 2D. Significant remoulding beyond a Geomechanics, 22(5): 327350.

548

Copyright 2005 Taylor & Francis Group plc, London, UK


Martin, C.M. and Randolph, M.F. 2001. Applications of the Siciliano, R.J., Hamilton, J.M., Murff, J.D. and Phillips, R.
lower and upper bound theorems of plasticity to collapse of 1990. Effect of jackup spud cans on piles. Proc. Offshore
circular foundations. Computer Methods and Advances in Technology Conference, Houston, OTC 6467.
Geomechanics, Tucson, Arizona, Balkema, 2: 14171428. Stewart, D.P., Boyle, R.S. and Randolph, M.F. 1998.
Murff, J.D. and Hamilton, J.M. 1993. P-Ultimate for Experience with a new drum centrifuge. Proc. International
Undrained Analysis of Laterally Loaded Piles. Journal of Conference Centrifuge 98, Tokyo, Kimura, Kusakabe
Geotechnical Engineering, 119(1): 91107. and Takemura (eds), Balkema, 1: 3550.
Randolph, M.F. and Houlsby, G.T. 1984. The limiting pres- Stewart, D.P. and Randolph, M.F. 1991. A new site investi-
sure on a circular pile loaded laterally in cohesive soil. gation tool for the centrifuge. Proc. International
Geotechnique, 34(4): 613623. Conference Centrifuge 91, Boulder, Balkema: 531538.
Reese, L.C., Cox, W.R. and Coop, F.D. 1975. Field testing
and analysis of laterally loaded piles in stiff clay. Proc.
Offshore Technology Conference, Houston.

549

Copyright 2005 Taylor & Francis Group plc, London, UK


FE modelling of spudcan pipeline interaction

L. Kellezi
GEO Danish Geotechnical Institute, Lyngby, Denmark

G. Kudsk
Maersk Contractors, A-P-Mller, Copenhagen, Denmark

P.B. Hansen
GEO Danish Geotechnical Institute, Lyngby, Denmark

ABSTRACT: Spudcanpipeline interaction problems were expected during a jack-up rig installation at the
proximity of an oil and gas pipeline in the North Sea. Three positions of the rig were considered assessing the
impact of the installation on the pipeline structure. From the available geotechnical data lower/upper bound soil
profiles, consisting of sand-clay-sand soils from the seabed, are first derived. Parallel to the conventional analy-
ses, 2D FE modelling of the spudcan penetration is carried out evaluating the soil deformation around the spud-
can and at the nearby pipeline. For a more realistic evaluation of the soil conditions, the history of the previous
installations was taken into account. The strength of the clay layer is re-evaluated carrying out back and con-
solidation analyses. The updated soil profiles are used for 2D and 3D large deformations FE analyses of spudcan-
pipeline interaction. The impact of the spudcan penetration on the pipeline structure stability is discussed. The
rig was installed and the results of the analyses were compared with the field observations.

1 INTRODUCTION the finite element (FE) method. The two-dimensional


(2D) and 3D modelling of the spudcan penetration-
For jack-up rigs located in close proximity of other pipeline interaction is carried out employing virgin
offshore structures, embedded or lying on the seabed, and more realistic soil profiles and assuming three
soil displacement caused by the spudcan penetration different rig positions.
may induce deformations into the nearby structures, The stability of the pipeline is investigated and dis-
which can be pipelines, platform pile foundations, quay cussed. The rig was installed at one of the considered
constructions etc. positions and the field observations are compared
Spudcan penetration is generally calculated based on with the FE results.
the conventional analysis, Hansen (1970), Florkiewicz
(1989) etc. However, for complex soil conditions
numerical methods are developed and presented in 2 SOIL CONDITIONS FOR VIRGIN
Ghoah & Kikuchi (1991), Griffiths (1982), Hu & SEABED
Randolph (1998), Sloan & Randolph (1982) etc. Using
these methods the soil pattern at an area around the Geotechnical soil investigation and laboratory testing
spudcan, where other structures might exist, can be are carried out in the area. Based on the available data,
investigated as in this case. lower and upper bound soil profiles are derived applic-
The impact of a jack-up rig installation on an exist- able to spudcan penetration analysis assuming virgin
ing oil and gas pipeline in the North Sea is investi- seabed conditions. The impact of the spudcan pene-
gated assuming the following. Virgin seabed conditions tration on the pipeline stability is evaluated by first
are initially considered, utilizing the geotechnical data considering those profiles.
derived from the site investigation and laboratory test- The soil conditions, lower/upper bound, consist of an
ing carried out before any structure installation at the upper layer of loose to dense sand with thickness 2.0 m
location. For a more realistic evaluation the impact of the and 1.5 m for which friction angle   30 is evalu-
previous installation is taken into account. The inves- ated. Below this layer soft to very soft clay is found to
tigation is generally based on the analysis employing depths 7 and 5 m, respectively. The undrained shear

551

Copyright 2005 Taylor & Francis Group plc, London, UK


strength cu  15 kPa and, cu  20 kPa are assessed
for lower and upper bound profiles, respectively.
Spudcan
Below the clay layer very dense sand is encoun-
tered to approximately 30 m depth. The friction angle
  35 is assessed for this layer.
The effective unit weight is also derived for the dif-
ferent layers according to the available data.

3 SPUDCAN AND PIPELINE DATA

The jack-up spudcans considered for installation have


a contact area of about 249 m2. Distance from spudcan a)
base (full contact) to spudcan tip is 1.6 m. The con-
sidered lightweight (plus variable loads) is 84 MN/leg Spudcan
and the maximum preload is 153 MN/leg.
The rig with the skirted spudcans previously installed
at the location have a full contact area of 252 m2.
Distance from spudcan base to spudcan tip is 1.8 m,
to the bottom of the permanent outer skirt is 4.7 m and
to the chords tip 5.2 m. The lightweight load is
71 MN/leg and the maximum preload is 123 MN/leg.
The pipeline is constructed from steel API5LX52
material. The concrete coating, the corrosion coating
and the steel covers existing along the length near the
considered spudcan make the pipeline a composite b)
steel-concrete structure. The pipeline is 30 to 50%
buried at the investigated location. Figure 1. a) Deformed and adaptive mesh, lower bound soil
profile. b) Deformed model, material regions, lower bound soil
profile.
4 SPUDCAN PIPELINE INTERACTION,
2D FE ANALYSIS, VIRGIN SEABED
CONDITIONS
was used to solve the contact interaction forces. A
Conventional spudcan penetration analyses based on friction coefficient of 0.6 is chosen for steel-sand
new developments of Hansen (1970), which satisfy contact conditions.
DNV (1992) and SNAME (2002), are initially carried Geostatic stresses and gravity loading are applied
out for lower/upper bound soil profiles. From the analy- to simulate the initial conditions.
ses, no punch through risk is found at the location and Instead of a vertical load, an axial deformation or a
spudcan penetrations varying from 6.0 to 7.5 m are rigid body movement is applied to the spudcan axis of
predicted. symmetry and the reaction forces at the spudcan-soil
To analyze the soil push up at the pipeline location interface are recorded.
2D axisymmetric FE modelling of the spudcan pene- The results of the FE analysis are given in Figure 1
tration is carried out using the Elfen FE program for the lower bound soil profile and Figure 2 for the
(Elfen V.3.0.4. (2001)). Implicit large strain elasto- upper bound one, where the geometry of the soil pro-
plastic analysis was chosen. The Mohr-Coulomb soil files and the soil strength parameters can be seen. The
constitutive model is applied considering the avail- deformation parameters are evaluated based on authors
able soil parameters. experience with the soil in the North Sea.
Adaptive re-meshing, which is a powerful tool of The results consist of deformation pattern and adap-
Elfen, is incorporated to avoid excessive element dis- tive mesh design given in Figure 1a, 2a and deformed
tortion. An error indicator based on stress norm pro- model and material regions Figure 1b, 2b.
jection, (L2 Zienkiewicz-Zhu projection type, Elfen V. As mentioned above, no punch through risk is
3.0.4 (2001)), is applied. Additional re-meshing regions expected during rig preloading. However, the soil
in critical areas (outer & inner corner at the bottom of profiles with the soft layer of clay under the sand and
the spudcan and its outside edge) are created. the large spudcan penetration makes possible severe
A Coulomb friction contact between the spudcan squeezing of the soft clay and large soil heave or
and the soil is simulated. An updated penalty method uplift around the spudcan.

552

Copyright 2005 Taylor & Francis Group plc, London, UK


As expected, for maximum preload the spudcan
Spudcan should penetrate deeper until it is in contact with the
bottom sand layer, giving larger soil push-up evalu-
ated to be over 2.0 m. This means that the pipeline
structure, depending on its rigidity is uplifted pos-
sibly at that height, approximately along 10 m length,
at the location close or tangent to the spudcan.

5 MODIFIED SOIL CONDITIONS DUE TO


PREVIOUS RIG INSTALLATION
a) Considering the results derived from section 4 and the
calculated large amount of the soil push-up at the pos-
Spudcan
sible pipeline location, a more realistic assessment of
the current soil conditions was found necessary. In
this framework the history of previous rigs installed at
the location is considered.
Using the data from the field, only one rig with
skirted spudcans was previously installed in the area,
almost at the current position, operating for a period
of 1.5 years. The geometrical profile for the skirted
spudcan of this rig is given in section 2.
Conventional skirted spudcan penetration analysis
was carried out for this rig from the authors, where
b) the same lower/upper bound soil profiles as discussed
in section 3 were applied. The predicted spudcan pene-
Figure 2. a) Deformed and adaptive mesh, upper bound soil tration was 4.6 m as upper bound value and 7.0 m as
profile. b) Deformed model, material regions, upper bound soil lower bound value. From the field observations the
profile. penetration of the spudcan at the position near the
pipeline was 4.6 m, which corresponds to the upper
bound soil profile.
Considering the observations for skirted spudcan
The amount of the soil heave in relation to spudcan
penetration and the time of the rig operation a
penetration can be seen from Figure 1 and 2. For the
re-evaluation of the soil conditions is carried out at
amount of the spudcan penetrations as given in the
the spudcan, near the pipeline.
above figures, only 39% and 71% of the maximum
Firstly, the observed penetration of skirted spudcan
preloads are applied for lower and upper bound soil
shows that the lower bound soil profile is very con-
profiles, respectively.
servative. The upper bound profile seems more realis-
Because of the large deformations and especially
tic although the borehole and the CPT data near the
the geometry of the soil profiles, at a certain spudcan
area support more the lower bound data. So, only the
penetration the sand layer is cut off and the clay layer is
upper bound soil profile will be considered further.
exposed through the sand and the calculation in Elfen
Secondly, the soft clay layer, which lies under the sur-
can not further continue for larger preload. This shows
ficial layer of sand has been consolidated during a
a limitation of the program found also in other previ-
period of 1.5 years due to previous rig installation and
ous analyses carried out by the authors using Abaqus,
operation. This fact is taken into account below.
and Elfen FE program, Kellezi & Stromann (2003).
There can be different failure mechanisms for the
pipeline during spudcan penetration. From the above
5.1 Soil consolidation analysis
2D axisymmetric FE modelling it can be investigated
that the soil heave, up-lift or push-up is maximum at The (2D) axisymmetric FE analyses are carried out
the P1 pipelinespudcan location. For these spudcan for the skirted spudcan employing Plaxis FE pro-
penetration levels, the soil push up at the three con- gram, Plaxis V.8.2 (2002). The effect of the previous
sidered pipelinespudcan relative positions P1 (2.3 m installation, loading and unloading, in the increase of
from the spudcan periphery), P2 (5.0 m), P3 (9.5 m) are the soil strength due to the soil consolidation, is
1.8 m, 1.25 m, 0.2 m and 1.7 m, 0.8 m, 0.1 m, for the investigated. The FE model and the results are given
lower and the upper bound soil profiles, respectively. in Figure 3.

553

Copyright 2005 Taylor & Francis Group plc, London, UK


Skirted Spudcan

Figure 3. Consolidation analyses mean effective stresses.

A 2D axisymmetric model satisfies the 3D condi-


tions for vertical loading. The initial geostatic stresses
are first calculated employing a coefficient of lateral
pressure K0  0.5 based on the available data.
The spudcan and the skirt are modelled as rigid
Figure 4. Spudcan penetration curves, more realistic soil
weightless bodies, with Mohr-Coulomb elasto-plastic
profiles.
constitutive soil model. Drained condition is applied for
the sand and undrained conditions for the clay layers.
From the observed spudcan penetration no conclu- 6 2D FE SPUDCAN PIPELINE
sion can be derived related to the strength of the clay INTERACTION, MODIFIED SOIL
layer. Because of the skirt structure the spudcan can PROFILES
be preloaded to maximum preload if the strength of
the clay is cu  20 kPa or larger, giving the same pene- New conventional and 2D FE calculations are carried
tration, as long as the skirt touches the sand. out employing the more realistic soil profiles and the
The reason is that after full base contact the skirted current rig spudcan to investigate the soil deform-
spudcan behaves as an embedded footing. This is an ation at the P1, P2 and P3 pipeline spudcan relative
observation previously carried out by the authors using positions during spudcan penetration.
FE modelling of the skirted spudcans resting on diff- The penetration curves calculated for the profile
erent soil conditions. This is also verified by other with clay strength cu  50 kPa and cu  65 kPa are
authors such as Hu et al. (1999). given in Figure 4 where results from the conventional
The governing equations of consolidation analysis and FE analyses are presented.
in Plaxis follow Biots theory. Darcys law for fluid The spudcan penetration for cu  50 kPa, at max-
flow and elasto-plastic behaviour of the soil skeleton imum preload, is expected to be 3.4 m and the soil
are assumed. The formulation is based on the small push-up 1.5 m (P1), 0.7 m (P2) and negligible for P3
strain theory. position.
The skirted spudcan is loaded to maximum preload The spudcan penetration for cu  65 kPa, at max-
and unloaded to lightweight load. FE consolidation imum preload, is evaluated about 3.0 m and the soil
analysis for lightweight load is carried out until min- push-up 1.2 m (P1), 0.5 m (P2) and negligible for P3
imum excess pore pressure is reached. position.
For normally consolidated clays the ratio of the The results from the 2D FE analysis carried out
undrained shear strength to the vertical effective stress with Elfen in the form of deformed mesh etc. are not
under which it was consolidated in the field show a listed as they are similar to Figure 1 and 2 with the
close correlation with the Plasticity Index, Bishop & difference that less spudcan penetration is necessary
Henkel (1964). A ratio varying between 0.2 to 0.25 is even for bearing the maximum preload.
considered based on the authors experience. From those results, for the P3 position, distance of
This judgment shows that the strength of the soft about 9.5 m of the pipeline from the spudcan periph-
clay is increased from 20 kPa to approximately 50 to ery, the pipeline stability should not be effected dur-
65 kPa during rig operation. So the upper bound pro- ing spudcan penetration.
file is modified increasing the strength of the clay to However, for the P3 position the spudcan is more
cu  50 kPa (lower bound value) and 65 kPa (upper close to the CPT, which is used as a basis for deriving
bound value) designing more realistic soil profiles. the lower bound soil profile. So the spudcan could be

554

Copyright 2005 Taylor & Francis Group plc, London, UK


placed over non-uniform soil conditions, partly over the
consolidated soil from the previous rig installation and
partly over the lower bound soil profile. This could
cause non-homogeneous penetration for the spudcan
to be taken into account during rig installation.
From the seismic survey and the bathymetry data,
shallow depressions or footprint are found in the seabed
created due to the previous rig installation and removal.
They are not considered critical and are not taken into
account in the current investigation.
What happened with the pipeline structure during
the partial uplifting along its length is investigated
from the 3D modelling of the spudcan pipeline
interaction model presented in the following.

Figure 5. 3D FE model, more realistic soil profiles.


7 3D FE SPUDCANPIPELINE INTERACTION
MODIFIED SOIL PROFILES

3D FE modelling of the spudcanpipeline interaction


is carried out employing the more realistic soil pro-
files as elaborated in section 6. One of the discussed
pipelinespudcan relative positions is considered for
demonstration. The 5 m distance or position P2 is found
more interesting as this is between the P1 position
(most critical case) and P3 (safe case). Discussions
and comments concerning the other two positions are
carried out.
3D FE modelling is carried out with Elfen V. 3.3.0
2001. Explicit dynamic large strain elasto-plastic
analysis is chosen different from the 2D axisymmetric
modelling where implicit static calculations were car-
ried out.
Figure 6. 3D FE model at maximum preload. Spudcan
In the 3D FE model only a quarter of the spudcan penetrationpipeline interaction.
is considered using symmetry conditions. For simpli-
fication, the tip of the spudcan of height 0.8 m is
removed. So the calculated penetration is corrected conditions are applied with this respect. According to
with the tip height when the results are interpreted. this simplification a symmetrical pipeline is assumed
The pipeline and the spudcan structures are both to exist on the other side of the spudcan parallel to the
modelled. The pipeline is considered 50% buried in considered pipeline and symmetric to the spudcan
the seabed and is simplified by modelling it as a solid centre. Although this is not the real situation this will
elastic cylinder. The real dimensions of the pipeline, not affect the evaluation of the pipelinesoilspudcan
of the coatings and of the covers are taken into interaction.
account for deriving equivalent elastic properties. In A Coulomb friction contact between the spudcan
deriving those properties the fact that the covers are and soil is modelled. The two contact surfaces are the
separated elements in contact, is taken into account. seabed and the spudcan area. An updated penalty
In the dynamic analysis spudcan penetration is method was used to solve the contact interaction forces.
applied as a displacement pulse. The mass of the sys- The Coulomb friction contact between the pipeline
tem is scaled, so the analysis is reduced to a pseudo- and soil could not be modelled as several problems
static one. The reason why explicit solution is chosen raised in the 3D model concerning contact modelling.
is that les requirements are regarding computer size The results of the 3D calculations for the more
and les time is needed for computation in comparison realistic soil profile with clay strength cu  50 kPa and
to implicit solution for a 3D modelling. cu  65 kPa are analyzed. The deformed model with
As spudcan penetration or the preloading is a sym- material regions specified is given in Figure 6, only
metrical action regarding soil deformation around the for the case cu  65 kPa.
spudcan, only the part where the pipeline is located is For the considered pipeline position P2 the pipeline
considered as shown in Figure 5. Appropriate boundary structure stability is investigated by the principal stress

555

Copyright 2005 Taylor & Francis Group plc, London, UK


deformation, a new 3D FE calculation is carried out
for the more realistic soil profiles, artificially increas-
ing the pipeline rigidity ten times.
The results showed that the pipeline deformation is
much smaller when a rigid pipe is assumed. As the
pipe cannot deform very much, the soil moves more
relative to the pipeline. This kind of soil deformation
and pipeline deformation mechanism cannot absolutely
be ruled out as not enough details are known concerning
pipeline structure rigidity, however the authors evalu-
ate that this could have been an optimistic evaluation.

8 CONCLUSIONS
Figure 7. Pipeline vertical displacements at different sec-
tions along its length as function of spudcan penetration. 2D and 3D FE modelling of the spudcanpipeline
interaction during a jack-up rig installation in the
North Sea is carried out. Three possible rig positions
1-1, which is maximum at the pipeline length close are investigated to assess the impact of the spudcan
and tangent to the spudcan. This means that part of penetration on the pipeline. The history of the previ-
the pipeline is working in tension, however not neces- ous installation in deriving realistic soil conditions, is
sarily critical. accounted.
The spudcan penetration at maximum preload for Parallel to conventional, large deformation analy-
soil profile with clay strength cu  50 kPa is expected ses with mesh adaptivity are carried out with Elfen
to be 3.4 m referring to spudcan tip. Correcting for FE program.
the change in the spudcan geometry, this corresponds Although the results of the 2D and 3D model are
to 2.6 m penetration in the 3D model. similar, the 3D modelling is considered the best way
The spudcan penetration at maximum preload for to analyze this kind of problem. More research needs
soil profile with clay strength cu  65 kPa is expected generally to be carried out in this area, particularly
to be 3.0 m referring to spudcan tip. Correcting for the regarding pipeline-soil contact interaction during spud-
change in the spudcan geometry, in the 3D model this can penetration.
corresponds to 2.2 m penetration as seen in Figure 6.
For illustration, the vertical deformations of the
different pipeline sections, P2 position, along the ACKNOWLEDGEMENTS
y-direction, are given in Figure 7 for both cases of the
clay strength values. The differential vertical deforma- The current work is carried out at GEO-
tions at y  0 m, and y  12.5 m for spudcan penetra- Danish Geotechnical Institute. The authors thank
tion 2.6 m (the real penetration 3.4 m) is about 0.64 m ConocoPhillips Norway for supplying with the neces-
and for spudcan penetration 2.2 m (the real penetra- sary data and Rockfield Software Ltd for the tech-
tion 3.0 m) is about 0.52 m. These are considered nical support in FE modelling with ELFEN.
unacceptable. The pipeline deforms also horizontally
during spudcan penetration but at a smaller rate.
From the results of the 3D FE analyses, for the
pipeline-spudcan distance of about 9.5 m, P3 pos- REFERENCES
ition, the soil deformation is negligible. This confirms
the results derived from the 2D modelling. Bishop, W. A. & Henkel, J. D. 1964. The triaxial test.
The location of the rig, which corresponds to the London Edward Arnold, Ltd.
P3 spudcanpipeline relative position, was finally DNV (Det Norske Veritas) 1992. Foundations Classification
chosen in the field and spudcan penetrations of 2.9 m Notes No. 30.4. February.
were recorded. No problems with the pipeline structure Elfen V. 3.0.4 2001. User Manual, Rockfield Software.
Elfen V. 3.3.0 2001. User Manual Explicit Rockfield Software.
stability were observed as expected.
Florkiewicz, A. 1989. Upper bound bearing capacity of lay-
ered soils. Can. Geotech. Journal. Vol. 26, 730736.
7.1 Rigid pipeline assumption Ghosh, S. & Kikuchi, N. 1991. An arbitrary Lagrangian-
Eulerian finite element method for large deformation
For demonstration purposes and in order to investi- analysis of elastic-visco-plastic solid. Comp. Meth.
gate the effect of the pipeline rigidity in the pipeline Appll. Mech. Eng., Vol. 86, 127188.

556

Copyright 2005 Taylor & Francis Group plc, London, UK


Griffiths, D. V. 1982. Computation of bearing capacity on Kellezi, L. & Stromann, H. 2003. FEM Analysis of Jack-up
layered soils. Proc. 4th Int. Conf. On Num. Meth. In Spudcan Penetration for Multi-Layered Critical Soil
Geomechanics, Vol. 1, 163170. Conditions. Proc. of BGA Intern. Conf. on Foundations,
Hansen, J. B. 1970. A revised and extended formula for bear- ICOF2003, Dundee, Scotland, 410420.
ing capacity. Bulletin No. 28. The Danish Geotechnical Plaxis, V. 8.2 2002. User Manual 2D, Delft University
Institute. Technology and Plaxis BV.
Hu, Y. & Randolph, M. F. 1998. Deep penetration of shallow SNAME 2002. T&R Bulletin 5-5A. Site Specific Assessment
foundations on non-homogeneous soil. Soil and Found. of Mobile Jack-Up Units. The Society of Naval Architects
Vol. 38, No. 1, p. 241246. and Marine Engineers.
Hu, Y., Randolph, M. F. & Watson, P. G. 1999. Bearing Sloan, S. W. & Randolph, M. F. 1982. Numerical prediction
response of skirted foundation on nonhomogeneous soil. of collapse loads using finite element method. Int. J.
Journal of Geotech. and Geoenviron. Eng. Vol. 125, No. Num. Analy. Meth. Geomech. Vol. 6, 4776.
11, 924935.

557

Copyright 2005 Taylor & Francis Group plc, London, UK


Copyright 2005 Taylor & Francis Group plc, London, UK
Jack-up footing penetration and fixity analyses

L. Kellezi
GEO Danish Geotechnical Institute, Lyngby, Denmark

H.W.L. Hofstede
Marine Structure Consultants bv, Schiedam, The Netherlands

P.B. Hansen
GEO Danish Geotechnical Institute, Lyngby, Denmark

ABSTRACT: For a jack-up rig structure to be located in the North Sea a practical methodology for footing-
soil interaction analysis considering the installation and the operation phases is described. The soil conditions
at the location consist of sand underlain by soft clay followed by sand and stiff clay. Penetration analyses are car-
ried out for the skirted footing without/with application of suction. Most Probable (MP) and Highest Expected
(HE) skirt resistances are calculated. For the operation phase footing stability and ultimate capacities under
storm load conditions are assessed. 2D finite element (FE) modelling is applied as an alternative to the 3D one.
The pore pressure increase for dynamic/cyclic (DC) loading is implemented from the critical suction value. For
lower/upper bound soil parameters, the footing capacities and stiffness are checked for the given factored load
paths. The yield envelopes for V-M (H-constant) and V-H (M  0) are calculated. The horizontal, vertical and
the rotational stiffness are derived applicable to structure analysis.

1 INTRODUCTION 2 FOOTING GEOMETRY

Footing-soil interaction analysis is carried out as The jack-up footing is constructed from a spudcan body
part of the design of a three-leg jack up rig structure and grout boxes installed around it. To improve footing
to be installed at the proximity of an oil platform in stability and to serve as a mean for scour protection, an
the North Sea. The purpose of the analysis is to deter- outer steel skirt is mounted around the perimeter of the
mine strength and stiffness capacities of the jack-up grout boxes. The dimension outer skirt/spudcan tip to
footings during installation (penetration) and opera- the largest contact footing area is about 3.0 m. The foot-
tion (fixity), respectively. The analysis is based on ing is not equipped with dowels. Brackets are installed
SNAME (2002). to reinforce the outer skirt. Inner skirts (diaphragms or
From the preliminary design the footing has diam- ribs), divide in compartments the volume between the
eter D  16 m. From the jack-up structure design pre- spudcan, grout boxes and the outer skirt.
load capabilities and the combined quasi-static (QS) The footing concept design is shown in Figure 1.
(wind and current) and dynamic/cyclic (DC) (due to
wave and jack-up dynamic behaviour) loads at three
headings, are calculated. 3 SOIL CONDITIONS AND WATER DEPTH
Using the available soil data design soil profiles are
established. For the installation phase conventional The geology at the location is characterized by layers
footing penetration, including application of a suction of Pleistocene and Holocene age. The site investigation
system, is calculated. For the elevated conditions consists of four boreholes/CPT (cone penetration tests)
maximum load capacities at yield, and foundation to 20 m depth, one composite CPT and sampling bore-
stiffness are derived. The stability of the foundation is hole to 80 m depth and offshore, onshore laboratory
checked for the applied loads. Similar research is car- testing. The water depth is 44 m.
ried out from Bransby & Randolph (1999), Gourvenec The data are used to derive the designed soil profile
(2003). at the location, consisting of sand layers underlain by

559

Copyright 2005 Taylor & Francis Group plc, London, UK


soft clay, followed by sand and stiff glacio-marine 4.1 Penetration analysis, no suction applied
clay to the depth applicable to foundation analyses.
The footing bearing capacity is calculated mainly based
Lower/upper bound design soil profiles and parameters
on Hansen (1970) and authors experience with spud-
are given in Table 1. In such assessments the charac-
can penetration prediction in the North Sea, Kellezi &
teristic parameters have been selected as a cautious
Stromann (2003). The spudcan body and the grout
estimate of the value effecting the occurrence of the
boxes are simplified to an equivalent circular footing
relevant limit state, Eurocode 7 (1997).
with a flat bottom.
Drained loading conditions are considered for the
The derived lower/upper bound design soil profiles
sand layers and undrained for the clay layers. The design
are applied in the conventional penetration analysis. As
friction angles have been evaluated from relative dens-
sand overlies soft clay punching failure is investigated
ities interpreted from CPT data and drained triaxial test
using a load spread assumption of 1:3. Firstly, the pene-
results. The design undrained shear strength values are
tration of the spudcan tip in the sand layer occurs. Next,
mainly evaluated from the unconsolidated undrained
squeezing of the soft clay is initiated and developed.
(UU) and consolidated undrained (CIU) triaxial and
The penetration resistance of the skirt elements is the
field laboratory (torvane and penetrometer) tests.
sum of skin resistance and the end resistance. Two cal-
culations, respectively Most Probable (MP) and Highest
Expected (HE) are carried out based on DNV (1992).
4 INSTALLATION PHASE, CONVENTIONAL
The latter governs the requirements to the penetration
PENETRATION ANALYSIS
force. The calculation of the skirt resistance is based on
the CPT data, which give a continuous record of the
To define footing penetration depth for the maximum
cone resistance with depth. For the chosen CPTs an
preload assessment of the static bearing capacity of the
average cone resistance at 0.2 m interval is calculated.
spudcan in combination with the grout boxes and the
Skirt resistance is calculated as a function of average
skirt, at various depths, is carried out.
cone resistance qc and empirical coefficients kp and kf,
which relates qc to skirt end resistance and skin fric-
tion. In the current analysis, kp  0.3 and kf  0.001 are
Spudcan chosen for the sand layers when calculating the MP
Grout Box resistance and kp  0.6 and kf  0.003 for HE resist-
ance. About 25% smaller values of these coefficients
Grout Box are applied in the upper first metre due to local piping or
Bracket lateral structure movement.
Several uncertainties remain regarding the conver-
sion from MP to HE resistance considering the effect
of penetration rate and generation of the pore pressure
during CPT testing. Therefore, a consistent set of cor-
relation factors does not exist. Soil resistance versus
Inner Skirt
penetration is given in Figure 2.
Outer Skirt
The results show that there is a slight risk of punch
through for vertical load larger than about 35 MN/leg
Figure 1. Footing cross section view. and lower bound soil parameters. This can be associated

Table 1. Lower/upper bound design soil profile.

Depth of Angle of Internal Undrained Shear


Layer (m) Friction ! () Strength cu (kN/m2)

Unit
Weight  Lower Upper Lower Upper
Soil type Top Bottom (kN/m3) Bound Bound Bound Bound

SAND, silty fine, very loose to medium dense 0 0.9 8 25 30


SAND, silty fine, dense to very dense 0.9 3.5 10 35 40
CLAY, very soft to soft 3.5 5.0 6 15 25
SAND, silty fine, medium dense to dense 5.0 6.0 9 30 35
CLAY, firm to very stiff 6.0 20 10 60  4  d 60  8  d

d: the depth below the soft clay layer.

560

Copyright 2005 Taylor & Francis Group plc, London, UK


Figure 3. Footing penetrations versus suction.
Figure 2. Footing penetration curves.

with a sudden penetration of 0.5 m for MP skirt resist- steady state flow solutions with H/D 0.5. Regarding
ance and footing penetration of 2.75 m. For HE skirt the inside skin friction, if critical suction occurs, there
resistance and footing penetration of 2.5 m, the sudden will be soil plug uplift. For installation purposes it is
penetration can be 0.7 m. However, this load level is not required to have some free space between the footing
reached in a storm condition. or spudcan lid and the soil plug. The heave of the soil
plug is observed in the calculations to be 4 to 5% of the
4.2 Penetration analysis applying suction skirt penetration depth.
The outside skin friction is assumed to be increased
For maximum preload of 30 MN/leg the skirt is not
by 1015% for maximum suction, based on values
expected to fully penetrate into the dense to very dense
reported from previous suction buckets installations in
sand layer, resulting in a penetration of 1.75 m for lower/
the North Sea.
upper bound soil data and HE skirt resistance.
A macro program is created to facilitate the current
The possibility of increasing the skirt penetration by
analyses. The reduction suction factors applied are
applying suction is investigated before the installation
rt  0.1 for the skirt/spudcan tip resistance, ro  0.15
phase. Initially, footing will penetrate during application
for the outer skirt skin resistance and ri  0.9 for the
of the maximum preload of 30 MN/leg. After this, sub-
inner skirt skin resistance. The required suction to reach
sequent suction might be applied to reach the desired
a desired depth is calculated. The penetration depth as a
depth.
function of differential skirt water pressure (suction) is
The driving force consisting of the preload and the
given in Figure 3.
suction, which is applied in the available volume sur-
The results are interpreted in such a way that, if the
rounded by spudcan/grout boxes, the inner skirt com-
footing will not penetrate more than 1.75 m for max-
partments and the sand soil filling the skirts, should
imum preload the proposed required suction to reach
be larger than the resistance from the soil.
for example 2.5 m penetration is 37 kPa, if LB soil
The resistance force can be subdivided into skin
data reveals. If UB soil data applies at the site, max-
friction acting on the inner, respectively the outer
imum 2.3 m footing penetration should be expected
skirt, combined with the skirt tip resistance and spud-
applying suction less than the critical value of 49 kPa.
can/grout boxes bearing capacity. The resistance force
is a function of the suction factors. The skin fric-
tion changes due to suction. The inside skin friction
reduces, whereas the outside skin friction increases, 5 OPERATION PHASE, FE FIXITY
Clausen & Tjelta (1996). In addition, the tip resistance ANALYSIS
is reduced. The maximum reduction is obtained when
critical suction value is reached. The analyses applicable to the operation phase con-
In the current penetrability analysis the critical suc- sider the assessment of footing foundation strength
tion is calculated based upon approximate numerical and stiffness under storm load conditions. Regarding

561

Copyright 2005 Taylor & Francis Group plc, London, UK


Spudcan
Inner Skirt

Waterfilled

Grout Box Vq = V/(B*L)

Outer Skirt

H = 2.75 m

2e = 2*M/V
Hq = V/B
Uq = *H* Hq = M/(2*L)

Figure 4. 2D FE footingsoil interaction model.

foundation stability three-dimensional (3D) finite A  B  L where B  3  R and L  R/3. B is


element (FE) modelling of the footing and the sur- the width of the rectangular in the moment loading
rounding soil employing coupled pore pressure/ plane and L is the length in the out-of-plane. The loads
deformation elements under dynamic loads, would be are applied for unit length L.
the target. The outer and inner skirts are modelled employing
Taking into account the soil conditions at the loca- beam or plate elements in 2D. The stiffness of the outer
tion, the complexity of the problem, the use of lower/ skirt is calculated from the FE analysis, which demon-
upper bound soil profiles and the fact that sufficient strates outer skirt deflection during impact to seabed. An
safeties are found in the V-H plane, two-dimensional equivalent rigidity in 2D, EI  2.5E04 kN/m2/m, is
(2D) FE deformation analyses for static load paths are assessed.
carried out as an engineering alternative. Modelling of the inner skirts is a difficult issue in the
Footing penetration of 2.75 m during the jack-up current 2D footing analysis. The inner skirts comprise
installation is probably most critical (closer to the soft six compartments where suction will develop during
clay layer) for the stability analysis. Therefore, the fixity DC loading. Several ways are tried to approximate the
analyses for lower/upper bound soil parameters is car- inner skirt in 2D. Modelling a deformable out-of plane
ried out for footing at this penetration depth. The ana- plate at the depth of inner skirts tip, simulating the fact
lyses are performed with FE program Plaxis V8.2 that the inner skirts will increase the outer skirt rigidity
(2002) applying updated mesh and Mohr Coulomb non- was found more realistic. After trial and error analyses
linear soil constitutive model for sand and clay layers. it was observed that there is no large differences in the
The soil strength parameters, as applied in the con- footing capacity for different ways of simplifying the
ventional penetration analysis are also implemented in inner skirts.
the fixity FE model. In addition, in absence of labora-
tory data, dilatation angle   !
30 is derived.
5.2 Load paths for storm conditions and
Other relationships such as   ! and   !/2 are also
equivalent 2D loads
investigated.
The deformation parameters are evaluated based on The loads acting on the footings are derived from the
the geotechnical data, but mostly on the authors experi- jack-up structural model where the footing and the
ences with FE modelling in the North Sea. E  200  surrounding soil are modelled as uncoupled springs,
cu is calculated for the clay layers and E  4  qc for reflecting torsion, rotation and displacements. Later
the sand layers. in the design, the springs are coupled, reflecting the
The 2D FE model can be found in Figure 4. The 3D interaction between the different responses.
footing geometry is adapted in the 2D analysis. Factored and unfactored load paths for storm condi-
tions are derived for 3 headings. Only the factored load
combinations (LC1, LC2, LC3), shown in Table 2,
5.1 2D FE footing modelling
are applied in the current analyses at footing axis of
The original 3D footing configuration is considered symmetry, 1.0 m below seabed, SNAME (2002). Weight
circular with radius R. To model it in 2D the circular from spudcan and skirts are included, soil plug is
shape is transformed into a rectangular one with excluded.
the same area and moment of inertia, DNV (1992). Based on the previous experiences with 2D and 3D
Combining these two criteria footing area is modelling of the skirted spudcans using Abaqus FE

562

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 2. Load paths.

Factored Loads Static1) Quasi-Static2) (QS) Dynamic/Cyclic3) (DC)

Load Comb
(LC) Angle () Leg V (MN) V (MN) H (MN) M4) (MNm) V (MN) H (MN) M4) (MNm)

LC1 0 3 22.21 24.112 0.378 12.159 30.020 1.793 25.409


LC2 180 3 22.21 22.078 0.389 12.366 21.906 2.268 39.767
LC3 210 2 22.21 20.125 0.393 12.205 15.273 1.982 42.493
1)
Static (still water no environmental loads).
2)
Quasi-static (wind & current1).
3)
Dynamic/cyclic (wave & inertia (due to jack-up dynamic behaviour)  12).
4)
M is the footing moment reduced by 1/1.1 for the 2D to 3D effect.

program the rotation moment M/leg (calculated from introduce a realistic weakness in the soil taking into
structural analysis) is reduced by 10% taking into account soil-steel interaction for the skirts and the
account the 2D analysis and the 3D effect. spudcan.
The application of the 3D combined V, H, M loads in The FE model is built asymmetric as asymmetric
the 2D model is shown in Figure 4. The fact that, for loads are applied. Standard boundary conditions are
footings with skirt, base level equals skirt tip level, is incorporated at the far field. The size of the model in
taken into account. The V load is applied as a distrib- the horizontal and the vertical direction is chosen so
uted pressure over footing area. The H load is applied as that the boundary conditions will not affect foundation
a line load at the considered reaction point. The moment capacity. The FE mesh is refined around the spudcan
M is applied as a set of two horizontal line loads acting and is coarser far from it.
respectively 1.0 m above and 1.0 m below the reaction
point.
5.4 Suction during DC loads, critical suction
Applying the loads in such a way gives the possibility
of constructing different stages starting from the maxi- In the current 2D fixity analysis, where pseudo-static
mum preloading, unloading to static loads, continuing loads are applied, the suction effect is simulated from
by QS and DC load paths. the critical suction concept based on calculated suction
levels, giving exit gradients equal to one inside the skirt.
This is compared directly to the measured differential
5.3 2D FE updated mesh analysis
pressure inside the skirted footing placed on frictional
Updated mesh analysis is applied to simulate possible material in order to evaluate the state (safe/critical) of
large deformations when defining footing ultimate suction levels.
load capacities. With this analysis the influence of For large diameter footings such as the current one, a
the geometry change on the equilibrium conditions simple approach is to look upon the skirt tip as a sheet
is taken into account by large deformation theory. pile wall, Hansen (1978). A point of interest is the exit
Additional terms are firstly included in the structure gradient, where the seeping water emerges behind the
stiffness matrix to model the effect of large soil-struc- wall. The normalized critical suction is usually given
tural distortions on the FE equations. A procedure is as ucr   H   where H is the outer skirt length.
secondly included to model correctly the stress change, For H/D 0.5, ucr  H  /(1.0
0.68/(1.46 
which occurs when finite material rotation happens. H/D  1)) is given by Clausen &Tjelta (1996) as used in
Update in the FE mesh as the calculation proceeds, is section 4.2.
done automatically and is based on the updated Critical suction is only valid for frictional materials.
Lagrangian formulation. This applies to the soil profile at the considered loca-
The initial geostatic conditions are calculated. tion. When the critical suction is reached, a piping chan-
Triangular 6-noded FEs are used to discrete the soil and nel may start to form but this takes time, and short time
the spudcan. The spudcan is modelled as a rigid weight- suction higher than ucr is therefore possible, even for
less body. fully drained conditions.
Footingsoil interaction is considered by incorporat- Based on the above, the suction effect is approxi-
ing interface elements, which have a reduced strength mated in the 2D fixity analysis for DC LC by applying
derived by applying the coefficient Rint  0,67 at the a distributed load equal to ucr  3  2.75  9.1
contact area with the sand layers. These elements  75 kPa at the level of outer skirt tip on the tension
enhance the flexibility of the FE mesh. They also side of the footing as shown in Figure 4. The effective

563

Copyright 2005 Taylor & Francis Group plc, London, UK


unit weight of the sand filling the skirt is taken as an Table 3. V-H-M and V-H ultimate capacities, lower bound.
average of the top sand layers. The width of the applied
suction load is a function of the eccentricity value Lower bound V-H-M Capacity V-H Capacity
e  M/V and is 2e  1.69 m for DC LC1, 2e  3.63 m
Load Angle V H M V H
for LC2 and 2e  5.56 m for LC3 considering the given Comb () Leg (MN) (MN) (MNm) (MN) (MN)
load paths in Table 2 and the concept of the effective
footing area. LC1 0 3 35.708 1.798 38.182 41.490 4.427
LC2 180 3 21.836 2.268 50.913 21.256 7.117
5.5 Footing stability and ultimate capacity LC3 210 2 15.746 1.982 39.584 6.694 4.439

Two scenarios, lower and upper bound soil parameters


respectively, are considered in the current 2D FE ana-
lyses. For each scenario, the footingsoil system stabil-
ity and stiffness are assessed for the given load paths
and the ultimate limit state calculating the capacity at
yield. The calculations is carried out as follows:
Stage 1: Load the footing with the maximum pre-
load of 30 MN/leg (Plaxis staged construction). Stage 2:
Unload to Static load. Stage 3: Apply the V-H-M for the
QS LC1, LC2 and LC3. Stage 4: Apply the H compon-
ent from the DC load for the three LCs. Stage 5: Apply
the V-M from the DC LCs. Stage 6: Apply the V-H from
the DC LCs after Stage 2 checking the sliding capacity.
Stage 7: Apply additional V-H loads increased by ratio
after Stage 2, defining the ultimate V-H capacities.
Stage 8: Apply additional V-M loads increased by ratio
after Stage 4 to define the ultimate V-H-M capacities.
Except for the two first stages the rest are repeated
for each load path. So in all, 20 stages are calculated
for each scenario. The ultimate footing capacities (the
load paths at yield) for lower bound soil parameters
are given as 3D loads in Table 3.
They can be compared with the 3D load paths given
in Table 2, showing that footing capacity for LC1 and
LC2 is well satisfied (load vectors fall inside the yield
surface) but LC3 is not completely satisfied for LB
(load vectors fall in or slightly out of the yield surface).
Safety in the V-H plane is checked. For UB conditions
the ultimate capacity is larger than LC2. So LC2 is
within the yield surface for this case. The results of the
calculations are given in Figure 5 for the sliding and in
Figure 6 for V-H-M ultimate capacities.

5.5.1 Investigated failure modes


Various failure modes are investigated for different LCs.
Typical failure modes encountered are: Sliding at base
with local failure around skirt tips, Figure 5a; Sliding
along the soft clay layer below skirt tip, Figure 5a,b;
Sliding along base of skirt tip, Figure 5c; Deep-seated
failures governed by moment equilibrium with centre
located anywhere, Figure 6a; Moment equilibrium Figure 5. Sliding at yield. Total displacements. a) LC1;
b) LC2; c) LC3.
centre below the foundation base, Figure 6b, c;

5.5.2 Footingsoil nonlinear stiffness left and the right points in the footing diameter. The fail-
The horizontal, vertical and the rotational stiffness for ure loads are defined from the asymptote of the yield
the footing are derived from the 2D fixity analyses. The curves. This is taken into account when deriving the
rotation is derived based on the vertical stiffness for the foundation stiffness for the structure analysis.

564

Copyright 2005 Taylor & Francis Group plc, London, UK


resistances. Fixity analyses consider the assessment of
footing stability and ultimate capacity under storm
load conditions. Suction effect during DC loading is
implemented based on the critical suction concept.
The horizontal, vertical and the rotational foundation
capacities and stiffness are derived from the 2D fixity
analyses found applicable and an effective engineering
approximation.

REFERENCES

Bransby M.F & Randolph M.F., 1999, The effect of skirted


foundation shape on response to combined V-M-H load-
ings. Int. Journal of Offshore and Polar Engineering,
(IJOPE) 9(3), page 214218.
Clausen C.J.F. & Tjelta T.L. 1996, Offshore Platforms
Supported by Bucket Foundation. Proc. 15th IABSE,
Copenhagen, page 819829.
DNV (Det Norske Veritas) 1992. Foundations Classification
Notes No. 30.4. February.
Eurocode 7, 1997, Geotechnical Design, Part 1, General Rules,
CEN/TC 250, 20042009.
Gourvenec S. 2003, Alternative design approach for skirted
footings under general combined loading. Proceed. of BGA
International Conference on Foundations (ICOF), Dundee,
25 Sept. 2003, page 341349.
Hansen, J.B. 1970. A revised and extended formula for bear-
ing capacity. Bulletin No. 28. The Danish Geotechnical
Institute.
Hansen B. 1978, Geoteknik og Fundering del II. Laboratoriet
for fundering. Danmarks Tekniske Hjskole (kursus 5821-
geoteknik 2 (In Danish).
Figure 6. DC load at yield. Rupture figures. a) LC1; Kellezi, L. & Stromann H. 2003, FEM Analysis of Jack-
b) LC2; c) LC3. up Spudcan Penetration for Multi-Layered Critical Soil
Conditions. Proc. of BGA Intern. Conf. on Foundations,
ICOF2003, Dundee, Scotland, 410420.
6 CONCLUSIONS Plaxis V. 8.2 2002. User Manual 2D, Delft University
Technology and Plaxis BV.
A methodology for a jack-up footingsoil interaction SNAME 2002. T&R Bulletin 55A. Site Specific Assessment
analyses is described. The installation and the operation of Mobile Jack-Up Units. The Society of Naval Architects
and Marine Engineers.
phases are considered.
Penetrations analyses without/with application of
suction are carried out calculating MP and HE skirt

565

Copyright 2005 Taylor & Francis Group plc, London, UK


Breakwater caissons and liquifaction

Copyright 2005 Taylor & Francis Group plc, London, UK


Performance of caisson breakwater subjected to breaking wave loads

X.Y. Zhang, C.F. Leung & F.H. Lee


National University of Singapore, Singapore

ABSTRACT: Centrifuge model tests were conducted to investigate the performance of caisson breakwaters
subjected to breaking wave loads. A servo-controlled electric actuator was developed to apply wave loads on the
caisson during centrifuge flight. The vertical and horizontal movements and tilt of the caisson breakwater were
found to increase progressively with the number of load cycles and much of the movements took place within
the first 2000 wave cycles. The results reveal that a minimum caisson width of 16 m is required to maintain the
stability against sliding and tilting.

1 INTRODUCTION 2 EXPERIMENTAL SETUP AND PROCEDURE

A gravity caisson breakwater is a hollow concrete box 2.1 Centrifuge model setup
which is in filled with sand after it has been placed at
The centrifuge tests were conducted at 100 g. Figure 1
the desired location. It is primarily built to provide
shows the centrifuge model setup. The model gravity
protection against wave attacks and at the same time
caisson used in the experiments is 250 mm (25 m in
reduce the amount of dredging required around a har-
prototype scale) high and 200 mm (20 m) long span-
bour entrance. The recent development of harbours and
ning the whole width of the model container. In this
related activities in deep waters lead to an increasing
paper, the behaviour of caisson breakwaters with widths
demand for vertical breakwaters. Oumeraci (1994)
of 140 mm (14 m) and 160 mm (16 m) were studied.
reviewed a number of breakwater failures and estab-
The 20 mm (2 m) base protrusions at both ends of the
lished that wave breaking and breaking clapotis repre-
caisson base were designed to facilitate a wider lateral
sent the most common source of disasters experienced
by vertical breakwaters. Caisson breakwaters are con-
ventionally designed by considering the wave loading
as a quasi-static load and then checked for their sta-
bility against sliding, overturning and bearing capacity
failures. However, the breaking wave loads are dynamic
hydraulic loads with very short durations and very
high peak loads that may exceed the quasi-static wave
load by over 10 folds. The failures of caisson break-
waters in Japan reported by Hitachi (1994) and
Takahashi (1994a) may be attributed to the severe
impacts created during wave breaks. Hence, conven-
tional design methods may not be appropriate for
caisson breakwaters.
This paper presents the results of centrifuge model
study on the horizontal, vertical and tilt displace-
ments of caisson breakwater subjected to breaking
wave loads. The dimension of caisson width will con-
siderably affect its construction cost and stability. Too
wide a caisson will result in a high construction cost
while too narrow a caisson may not provide sufficient
stability against wave loading. Hence, the effect of
caisson width on the caisson performance will also be Figure 1. Experimental setup and instrumentation (all
investigated. dimensions in mm).

569

Copyright 2005 Taylor & Francis Group plc, London, UK


load spread hence enhancing the bearing capacity of the 2.3 Experimental procedures
caisson. In order to reduce the side friction effects in the
Two loading stages, namely caisson infilling stage and
plane strain model, the sides were lined with grease-
wave loading stage, were simulated in the centrifuge
treated polyethylene sheets (Khoo et al. 1994). The sand
tests. Once a centrifuge acceleration of 100 g has been
bed is 150 mm thick (15 m) made of uniform medium
achieved, ZnCl2 solution was released into the caisson
sand with an effective mean diameter of 0.34 mm, uni-
to simulate the ballasting process. Then the wave simu-
formity coefficient of 0.32, minimum dry density of
lator was activated to apply breaking wave loads on
1280 kg/m3 and maximum dry density of 1540 kg/m3.
the caisson breakwater for approximately 20,000 wave
To overcome the scaling conflicts between time
cycles. The sample rate of measurement was 20 Hz per
scales for consolidation and dynamic events, silicone
instrument during the infilling stage and 300 Hz during
oil with viscosity of 100 cSt and a specific gravity of
the wave loading stage. Unless otherwise stated, the
0.96 at 25C was used as the pore fluid in the model.
test results are presented in prototype scale. It should
When preparing the sand bed, about 22 kg of dry sand
be noted that the downward caisson settlement and sea-
was thoroughly mixed with about 45 kg of silicone oil
ward horizontal caisson movement are both taken as
and the mixture was de-aired under vacuum for about
positive. It is assumed that the caisson tilts along some
1 hour (Lee 1985). Then the sand and silicone oil mix-
point on the caisson base and hence the horizontal cais-
ture were transferred to the container in small portion
son movement can be divided into two components,
of about 500 g each using a stainless steel beaker. The
one due to sliding and another due to tilting.
beaker was then fully immersed in the oil in the con-
tainer and tipped over, thus allowing the sand to plu-
viate through silicone oil without bubble formation.
3 RESULTS AND DISCUSSIONS
During this process, the beaker was moved slowly over
the area of construction. This achieved an approxi-
3.1 Infilling stage
mately equal rise of sand surface at all points. The
process was repeated until the sand bed was built up Figure 2 shows that during the infilling stage, the cais-
to the prescribed level. The relative density of the son settlement increases almost linearly with loading
sand bed was between 55% and 60%. pressure. Similar phenomenon has been observed by
Two displacement transducers V1 and V2, as shown Ng (1998) for the load-displacement response of a
in Figure 1, are employed to monitor the settlement at spudcan footing. Practically no horizontal caisson dis-
the seaward and landward sides of the caisson, while placement is noted revealing that the sand bed below
transducer H1 was employed to measure the horizontal the caisson has been prepared uniformly. As expected,
caisson movement. The pore pressures in the soil were the 14-m wide caisson settled more than the 16-m
measured by miniature pore pressure transducers at wide caisson. This is because under the same vertical
six locations. Pore pressure transducer P7 is located pressure, the reduction in the caisson width and hence
inside the model caisson to measure the pressure of base area will decrease the bearing capacity of the
the zinc chloride solution (ZnCl2) in the caisson. footing. Ng (1998) considered that the sand bed pre-
pared by pluviation would incur substantial amount
of plastic strains, which suggests that the state of sand
2.2 Design of wave simulator system
bed underneath the caisson lies on or close to the
Assuming a design prototype wave height of 3 m and yield surface. This would account for the immediate
wave period of 6 s, the magnitude of the appropriate settlement observed when ballasting the caisson with
breaking wave load on a vertical breakwater can be ZnCl2 solution.
estimated using the Godas (1985) formula and the
extended Goda formula developed by Takahashi et al.
(1994b). It is rather complex and difficult to develop
a wave peddle stroke system to simulate the prototype 60
14-m-width
Caisson settlement

breaking waves due to space constraint in the centrifuge 50


model. Hence, a relatively simple wave simulator was 40
(mm)

fabricated for the present study such that the simula- 30


tor would impart an appropriate transitory load on the 20 16-m-width
model caisson. The servo-valve load-control wave
10
simulator system consists of a torque arm, gear box
and electric motor, as shown in Figure 1. The load cell 0
attached on the head of the torque arm was used to 0 50 100 150 200 250
record the resultant wave loads during centrifuge tests. Caisson pressure (kPa)
Details of the wave simulator have been described by
Leung et al. (2004) and will not be repeated here. Figure 2. Caisson settlements during infilling stage.

570

Copyright 2005 Taylor & Francis Group plc, London, UK


3.2 Wave loading stage the wave loading. Figure 4 shows the corresponding
Fourier spectra for the wave loads and caisson move-
3.2.1 Caisson movement responses
ments from 2000 to 2080 wave cycles. The resonance
Figure 3 shows typical instantaneous caisson move-
frequency is found to be 0.24 Hz and the highest
ments and pore pressure responses during the first 12
amplitudes occur at the resonance frequency for all
wave cycles for the 14-m wide caisson. The simpli-
cases. Since up to 18,000 wave load cycles have been
fied triangular breaking wave load is simulated in the
applied, the recorded instantaneous responses may
centrifuge and normalized by the total weight of the
not clearly reveal the whole trend. Therefore, the data
caisson including the weight of infill material. It is
plot was smoothened by averaging adjacent data
found that there is an immediate build-up of vertical,
points over a window of 50 wave cycles.
horizontal and tilt movements of the caisson breakwater
The comparisons of the smoothened movements of
during the first several wave cycles. The instant-
caissons with different widths are shown in Figure 5.
aneous pore pressures are well coupled throughout
The caisson movements increase steadily with the
number of wave cycles and approximately 90% of the
15
final movements have taken place within the first
wave load (%)
Normalized

10 2000 wave cycles. After that, the caisson breakwater


5
enters the stabilization stage where the movement
magnitude starts to tail off towards an asymptotic
0 value. The rate of increase in settlement in this stage
is very low. Towards the end of the stabilization stage,
15 virtually no progressive movement is detected indi-
cating the soil has a fairly constant void ratio.
(kPa)

It is noted that the magnitude of the caisson settle-


P1

0
ment does not vary much with respect to caisson
-15

12.5
load amplitude (%)

2
Tilt angle amplitude Normalized wave
(kPa)

0.0
P4

1
-12.5
0
-25.0
0.006
0.06
Tilt angle
(degree)

(degree)

0.04 0.004

0.02 0.002
0.00 0.000
amplitude (mm)

0.3
Settlement

9
Settlement

6 0.2
(mm)

3
0.1
0
-3 0.0
Sliding amplitude
Horizontal sliding

10
1.2
(mm)

5
(mm)

0
0.6
-5
-10
0.0
0 2 4 6 8 10 12 14
Number of wave cycles 0.0 0.4 0.8 1.2 1.6
Frequency (Hz)
Figure 3. Typical instantaneous movements and pore pres-
sure response of caisson with 14 m width during wave loading Figure 4. Fourier amplitude spectra of wave loads and
stage. movements of caisson with 14 m width.

571

Copyright 2005 Taylor & Francis Group plc, London, UK


100 14m-width 3.2.2 Instantaneous pore pressure responses
Figure 3 reveals that the pore pressure time history is
Settlement (mm) 80 highly unsymmetrical during the initial wave cycles.
60 The landward readings (P1) consists of a short nega-
tive phase followed by a long positive phase, while
40 16m-width the seaward (P4) consists of a short positive phase
20 followed by a long negative phase. There is a rapid
0 change from negative to positive excess pore pres-
0 5000 10000 15000 sures within the first half of loading cycles under the
Number of wave cycles landward side of the caisson. On the other hand, the
excess pore pressures change from positive to nega-
(a) Caisson settlement of gravity centre tive values under the seaward side of the caisson. This
is attributed to the fact that under maximum forward
0 wave load, the seaward caisson base is unloaded and
the sand beneath the caisson dilates resulting in nega-
Horizontal sliding

-10 tive pore pressure. At the same time, the landward


-20 16m width caisson base is loaded and the sand beneath the cais-
(mm)

son heel continues to be compressed, causing an


-30
increase in instantaneous pore pressure. Similar pore
-40 14m width pressure behaviour at the two sides of a shallow foun-
-50 dation subjected to earthquake loading was noted by
Zeng & Steedman (1998).
0 5000 10000 15000 Figures 6(a) and (b) show the Fourier spectra of
Number of wave cycles landward and seaward pore pressure responses. The
(b) Horizontal sliding of caisson breakwater
responses of the sandbed to the fundamental wave
frequency remain unchanged, suggesting that the
increase in shear stiffness arising from negative residual
0.3 14m width pore pressure generation is insignificant to affect the
Tilt angle (degree)

fundamental wave frequency. However, the sand bed


0.2 undergoes reduction in amplitude of pore pressure
from the surface to bottom at landward side during
0.1 wave loading. Oumeraci (1994) considered that the
16m width
impact wave loads are transferred to the seabed through
0.0 rocking and swaying motions of the structure and the
0 5000 10000 15000 only resistance opposite to these loads is provided by
Number of wave cycles
friction, inertia and damping forces. Hence, the wave
impacts are felt strongest by the landward surface soil
(c) Tilt angle of caisson breakwater just beneath the base of caisson and then transferred
to a greater depth through rearrangement of the soil
Figure 5. Comparisons of movements of caissons with dif- skeleton.
ferent widths.
3.2.3 Residual pore pressure responses
Cyclic loading of sand may cause a change of the
width during the infilling stage, as indicated in Figure effective shear strength and stiffness. This has to do
2. However, during the wave loading stage, the mag- with the development of residual pore pressures.
nitudes of settlement, horizontal sliding and tilt angle Figure 7 shows the residual pore pressure response of
are much larger for the 14-m wide caisson as com- the 16-m wide caisson. The build-up of the residual
pared to the 16-m wide caisson. This is reasonable as pore pressure is quite low although the instantaneous
the factor of safety against sliding and overturning is pore pressure can be up to 15 kPa. It reaches the peak
significantly lower for a narrower caisson with a value at about 400 wave cycles and then gradually
reduced total weight. Figure 5(c) shows that the max- dissipates towards the hydrostatic pressure. The contour
imum tilt angle for the 14-m wide caisson is 0.25 maps of excess pore pressure at different wave cycles
implying that this width may not be adequate to pro- are indicated in Figure 8. It is noted that the build-up
vide the necessary stability against overturning. At of excess pore pressure is localized at the two edges
the same time, the horizontal movement of the 14-m of the caisson base. The landward soil under the cais-
wide caisson is up to 51 mm which also implies that it son base dilates after several wave cycles. On the
may not have sufficient stability against sliding. other hand, bulbs of positive pore pressures develop

572

Copyright 2005 Taylor & Francis Group plc, London, UK


0.3
Amplitude of P1

P4 (kPa)
3 0.0
2
(kPa)

-0.3
1 -0.6
0 1.2

P5 (kPa)
0.9
6 0.6
Amplitude of P3

0.3
4
(kPa)

0.0
2

P1 (kPa)
-0.4
0 -0.8
-0.2 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 -1.2
Frequency (Hz)

P2 (kPa)
(a) Landward pore pressure response -0.3
-0.6
-0.9
Amplitude of P4

3
(kPa)

2
P3 (kPa)
0.3
0.2
1
0.1
0 0.0
0 5000 10000 15000 20000
4 Number of wave cycles
Amplitude of P6

Figure 7. Residual pore pressure response of caisson with


2 16 m width.
(kPa)

0
0 0
-0.2 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6
-5 -5
Frequency (Hz)
-10 -10
(b) Seaward pore pressure response
-15
-15
-15 -10 -5 0 5 10 15 -15 -10 -5 0 5 10 15
Figure 6. Fourier amplitude spectra of seaward and land-
ward pore pressure responses of caisson with 14 m width. (a) 5th cycle (b) 100th cycle

0 0

at the seaward side. With increasing wave cycles, -5 -5


the negative pore pressures gradually pervade in -10 -10
the seaward soil from the bottom to the soil depth
-15 -15
of 10 m. -15 -10 -5 0 5 10 15 -15 -10 -5 0 5 10 15

The residual pore pressure responses of the 14-m (c) 5000th cycle (d) 18000th cycle
wide caisson are shown in Figure 9. The soil under-
neath the caisson base dilates at the commencement Figure 8. Contour maps of residual pore pressure response
of the applied wave loading. The decrease in pore of caisson with 16 m width under different wave cycles.
pressure in the foundation soil increases the effective
stresses of the sandbed. However, this is only local-
ized beneath the caisson base. The swelling sand may the resistance of the deeper soil strata and no slip sur-
possibly be pushed sideward and this may help to face was visible extending to the ground surface.
explain why the shear strength increased during the
wave loading while the caisson keeps on settling and
sliding. After the centrifuge test, the sandbed was 4 FUTURE WORKS
carefully checked. It was observed that soil bulges
had developed at both sides of caisson breakwater as The impact and cyclic wave loads are considered by
soil was displaced from beneath the caisson. There was Oumeraci (1994) to be the two main reasons that lead
no catastrophic collapse or tilt of the caisson break- saturated soils to experience large unacceptable per-
water which remains deeply embedded, mobilizing manent deformations. The difference between impact

573

Copyright 2005 Taylor & Francis Group plc, London, UK


-5 5 CONCLUSIONS
P1 (kPa)

-10 A series of centrifuge experiments has been con-


-15 ducted to examine the influence of caisson widths on
-5 the behaviour of caisson breakwaters during infilling
P3 (kPa)

and wave loading stages. The magnitude of impulsive


-10
wave force is calculated using the original and extended
-15 Goda formulae. In-flight simulation of breaking wave
loads is actuated by means of wave simulator system
-10
P4 (kPa)

at 100 g. It is established that:


-15
-20
1 The caisson movements are found to increase
steadily with increasing number of wave cycles
-10 within the first 2000 wave cycles. After that, the
P5 (kPa)

-15 caisson breakwater enters the stabilization stage


where the movement magnitude starts to tail off
-20 towards an asymptotic value.
0 5000 10000 15000 20000 25000
Number of wave cycles 2 Although the soil gains strength progressively due
to the build-up of negative pore pressures, per-
Figure 9. Residual pore pressure response of caisson with manent caisson deformations still occur, possibly
14 m width. because the sand moves sideward.
3 The results also suggest that the caisson must be at
least 16 m wide such that the caisson tilt angle and
horizontal sliding movement fall within an accept-
able range during wave loading.

ACKNOWLEDGEMENTS

The study is sponsored by the National University of


Singapore Research Grant number R-264-000-119-
112. The assistance of Centrifuge Laboratory staff in
the conduct of centrifuge model tests is also grate-
fully appreciated.

REFERENCES

Goda, Y. 1985. Random seas and design of maritime forces,


Figure 10. Failure of caisson breakwater after partial University of Tokyo Press, Tokyo, Japan.
liquefaction. Hitachi, S. 1994. Case study of breakwater damages Mutsu-
Ogawara port, Proc. Int. Workshop on Wave Barriers
in Deepwaters, Port and Harbour Research Institute,
and cyclic loading phenomena is the applied drag Yokosuka, Japan: 308329.
force on the caisson sides. It is believed that the alter- Khoo, E., Okumura, T. & Lee, F.H. 1994. Side friction effects
nate loading and unloading of the two edges of the in plane strain models. Proc. Centrifuge 94:115120.
caisson may lead to caisson failure mechanism different Lee, F.H. 1985. Centrifuge Modeling of Earthquake
to that resulting from impact loading. Therefore, a Effects on Sand Embankments and Islands. Ph. D Thesis,
two-directional cyclic wave with strong magnitude Cambridge University, England.
from
10% to 10% is currently being studied in the Leung C.F., Zhang X.Y. & Lee F.H. 2004. Wave Impact on
centrifuge. Figure 10 shows the failure of caisson Caisson Breakwater, Proc. 9th Australia New Zealand
breakwater under partial liquifaction when subjected Conference on Geomechanics, Auckland, New Zealand:
874880.
to the strong cyclic wave loads. The liquifaction depth Ng, T.G. 1998. Cyclic Behavior of Spudcan Footing on Sand,
is up to 5 m and the caisson breakwater experiences Ph. D thesis, National University of Singapore, Singapore.
catastrophic failure with significantly tilt, settling and Oumeraci, H. 1994. Review and analysis of vertical break-
sliding. Details of this mechanism are currently under water failures lessons learned, Coastal Engineering,
investigation. Vol. 22: 339.

574

Copyright 2005 Taylor & Francis Group plc, London, UK


Takahashi, S., Tanimoto, K. & Shimosako, K. 1994a. Dynamic composite breakwaters, Proc. International Conference on
response and sliding of breakwater caissons against Hydro-technical Eng. For Port and Harbor Construction,
impulsive breaking wave forces, Proc Int. Workshop on Port and Harbour Research Institute: 489504.
wave Barriers in Deepwaters, Port and Harbour Research Zeng, X. & Steedman, R.S. 1998. Bearing capacity failure
Institute, Yokosuka, Japan: 362399. of shallow foundations in earthquakes. Geotechnique,
Takahashi, S., Tanimoto, K. & Shimosaka, K. 1994b. A Vol. 48(2): 235256.
proposal of impulsive pressure coefficient for design of

575

Copyright 2005 Taylor & Francis Group plc, London, UK


An interaction model for seismic stability analysis of
caisson type structure

H. Hazarika
Port and Airport Research Institute, Nagase, yokosuka, Japan

ABSTRACT: An interaction model was developed for a gravity type caisson with granular backfill sand-
wiching a protective compressible layer as cushion, whose function is to improve the seismic stability of the
structure. Interfaces on either side of the cushion were modeled as elements of finite thickness having different
stiffness and constitutive properties. Participation of each element of the interaction system was taken into the
account by incorporating appropriate factors at the respective interfaces. Seismic analyses were performed on a
gravity type caisson using the developed model. Comparison of the analysis results with a similar caisson with-
out any protective cushion showed that the use of cushion (made from a compressible recycled material) yields
a significant reduction of seismic load on the structure.

1 INTRODUCTION

Gravity type quay walls are widely used in many off-


shore and coastal facilities. These kinds of rigid struc-
tures are vulnerable to damages during large-scale
earthquakes. A comprehensive survey and analyses
results on the damages of caisson type quay walls dur-
ing the 1995 Hyogo-ken Nanbu Earthquake, Kobe,
Japan is reported in Inagaki et al (1996). In order to pro-
tect the structures from such devastating damages dur-
ing earthquakes, accurate estimation of the seismic load
acting on them is necessary, which in turn necessitates Figure 1. Cushion wrapped gravity type quay wall.
an appropriate load-deformation analysis. Despite
structural simplicity, the soil-structure interaction of technique, known as the compressible inclusion, was
such system is rather complicated, and the dynamic adopted by Inglis et al (1996) to improve the seismic
response of it has not yet been fully understood. stability of a rigid basement wall using foam type mater-
Damages suffered by many quay walls during the 1995 ial. The function of the compressible layer is to impart
Hyogo ken Nanbu Earthquake, has led to believe that a cushioning effect on the structure during the seismic
the quasi-static analysis based on the well known loading. Such technique has proven to be an effective
Mononobe-Okabe model is not applicable for the level method to reduce the seismic load on rigid yielding and
II design ground motion (JSCE 2001, Nozu et al 2004). non-yielding structures (Hazarika et al. 2003). In this
PIANC (2001) emphasized the need of performance- research, considering also the cost-benefit viewpoint, a
based design for the seismic performance of structure. recycled material was chosen as cushion.
Better understanding of the behavior of such soil- When a cushion of different stiffness characteristic
structure interaction system needs different perspec- is sandwiched between a structure and the backfill soil,
tives in the modeling. a hybrid system of interaction is generated. In such a
On the other hand, a cost-effective design alternative system, the sandwiched element possessing different
to such important and massive structures is to improve stiffness and compressible characteristic yields two
their stability by some other means, and thus enhancing different interfaces: structure-cushion, cushion-soil.
their seismic performance. This can be achieved by Proper modeling of these interfaces determines the
using some kind of compressible material sandwiching final outcome of analysis of such soil-structure inter-
the backfill soil and the structure (Fig. 1). This kind of action system.

577

Copyright 2005 Taylor & Francis Group plc, London, UK


The purpose of this research was to develop an II. The corresponding stiffness of the systems are given
interface model that can be used for analyzing a sys- by the following.
tem comprising of the backfill soil, the sandwiched
cushion and the structure. Hazarika & Okuzono (1)
(2004) developed an interface model, in which EPS
(Expanded Polystyrene) geofoam was used as a com- Where, [K]hi represents the sum of the stiffness of
pressible inclusion, for analyzing such hybrid nature the solid elements of the participating media at the
of interfaces. This paper describes an extension of that respective interfaces, and are given by the following
model when a recycled waste material is used as the equations.
cushion (Fig. 1) instead of EPS. Such recycled mater-
ial is preferred these days because of high material (2)
cost of EPS and also due to its impermeable nature.
The contributions towards the soil-structure inter-
action process, from each participating elements at the (3)
interfaces, were expressed in terms of some constants
called participation factors. Here, [K]S, [K ]C and [K]B are the stiffness of struc-
A numerical experiment was conducted on a caisson ture, cushion and backfill soil respectively. In Eq. (1),
type model quay wall using the described interface the stiffness of the interfaces ([K]Int i(i  I,II)) are
model. The analysis results were compared with a simi- given by the following equations.
lar caisson without any protective measures. It was
observed that the protective cushion yields a substantial (4)
reduction of the seismic load acting on the structure.
In the above equations [KS]Int i are the shear compon-
2 MODEL DESCRIPTION ents of the interface stiffness. They can be obtained
from the direct shear tests or similar interface testing.
When material of different stiffness and compressibil- The normal behavior of the thin interfaces can be
ity is sandwiched behind a structure, a hybrid inter- expressed by equations of the form given below.
active system is generated. In conventional analysis, it
is common to use interface element of zero thickness (5)
(Hazarika & Matsuzawa 1996, Day & Potts 1998) to
simulate the interface. However, in a situation depicted
in Figure 1, where sandwiched material is involved, the (6)
interfaces are not exactly in a planar surface. In such a
hybrid nature of interaction, inescapable gap exists In Eqs. (5) and (6), [Kn]Int I and [Kn]Int II are the
between the interactive media, where material particles stiffness matrices of the normal direction of the
can go in, producing a thin layer that participates in the Interface I and Interface II respectively. [Kn]S, [Kn]C
overall interaction process. The idea of including the and [Kn]B represents the stiffness of structure, cushion
finite thickness interface developed by Desai et al and backfill in the normal directions respectively. The
(1984) is, therefore, better suited for modeling such a terms 1, 2, 3, 1, 2 and 3 are the constants called
system. the participation factors. These factors represent the
Figure 2 shows the conceptual representation of the contribution of the respective media participating in
interface, which consists of two interactive systems, the interaction at each interface. They, therefore, satisfy
represented as interface system I and interface system the following relationship.

(7)

At each interface, the sandwiched element has a


finite zone of influence (Hazarika and Okuzono 2004)
within which it interacts. The influence zone was
assumed to be half the thickness of the adjoining inter-
face elements. When no cushion layer is present (which
represents a conventional soil-structure interaction prob-
lem), the stiffness, [K]Sys for the single interactive sys-
tem can be derived using Eq. (1) as follows.

Figure 2. Modeling concept. (8)

578

Copyright 2005 Taylor & Francis Group plc, London, UK


3 NUMERICAL ANALYSIS the foundation rubble was simulated by making use of
the Eq. (8). For thin layer interface, Zaman et al (1984)
3.1 Simulated model suggested the thickness of the interface element to be
0.05 times the dimension of the adjacent soil element
A model caisson with sandy backfill was selected for
so that the numerical stability can be maintained with
the numerical experiment, where the system was sub-
less margin of error. In the model described herein,
jected to an earthquake type loading. A plane strain
since two different interfaces are involved, it is difficult
FEM discretization of the model is shown in Figure 3.
to adopt any such general value for the thickness.
The model wall was 7 m in height with sandy backfill,
Parametric studies are required for proper estimation
which was assumed to be dry. The caisson rests on
of the respective thickness. However, for the sake of
foundation rubble made up of ballast overlying a
pure simplicity and also due to lack of any reliable
dense sandy layer with characteristics same as that of
data, it was assumed here that the recommendation of
the backfill. As shown in the figure, reflected bound-
Zaman et al (1984) is valid for such types of interfaces
aries (roller support) are allowed to have movement
as well. The following values of the participation fac-
only in vertical directions, while fixed boundaries are
tors were assumed for the caisson-cushion and cush-
restrained against both the movements. Viscous dampers
ion-backfill interfaces:
were introduced at the reflected boundaries.

3.2 Material modeling 1  1  0.75; 2  2  0.25; 3  3  0.00.

A 3 m thick cushion layer was placed behind the cais-


son of Figure 3. The cushion could reduce the static Table 1 lists the basic material parameter values that
load on the structure due to the controlled yielding were used in the simulation.
mechanism (Karpurapu & Bathrust 1992). Such inter-
action problem, therefore, needs to be analysed using
3.4 Analysis process
separate constitutive models for the respective mater-
ials involved in the interaction. For the sandy backfill, Analyses were conducted for two different backfill
the localization based constitutive law (Hazarika & conditions. One is the case in which the caisson was
Matsuzawa 1996) was adopted, while for the cushion without any protective cushion, and the other is the
material, the model proposed by Youwai & Bergado case in which a cushion layer was sandwiched between
(2003) was adopted. The selection of these two particu- the backfill soil and the caisson in order to enhance
lar models was purely due to proximity of the respect- its seismic performance.
ive materials constitutive behavior to those represented Performance observation showed that for quay
by the two models. wall effective stress analysis gives better results (Iai
1998). However, as a first step towards implementa-
tion of the developed model to a new class of light-
3.3 Interface modeling
weight geomaterials (Yasuhara 2002), only a very
The interface constitutive law was assumed to be ideal case comprising of dry backfill as well as
bi-linear elasto-plastic obeying the Mohr-Coulomb fail- foundation soils was considered here. Static analyses
ure criterion with zero cohesion. The parameters of the were first performed under gravity loading to calculate
interfaces were determined from the data based on the static earth pressure. Dynamic analyses were then
extensive direct shear tests performed on various inter- performed, by imparting an actual earthquake motion
face conditions involving cushion materials, soils and at the fixed boundary of the simulated model (Fig. 3).
structural materials (Hazarika et al. 2005, Karmokar The acceleration time history of the input seismic
et al. 2005). The interface between the caisson base and load is shown in Figure 4. This seismic motion was
recorded at the port of Hachinohe during the 1968

Table 1. Material parameters.

Backfill &
Foundation Cushion
Parameters Soil Material

Youngs modulus, E 26 MPa 2.6 MPa


Poissons ratio,  0.30 0.20
Unit weight,  15.0 kN/m3 7.0 kN/m3
Angle of internal friction,  40 30
Figure 3. FEM model (all dimensions are in meters).

579

Copyright 2005 Taylor & Francis Group plc, London, UK


250 400.00
200 300.00

Acceleration (Gal)
150
Acceleration (Gal)

200.00
100
50 100.00
0 0.00
-50
-100.00
-100
-150 -200.00
-200 -300.00
-250
0.00 5.00 10.00 15.00 20.00 25.00 30.00 -400.00
0.00 5.00 10.00 15.00 20.00 25.00 30.00
Time (sec)
Time (sec)
(a) Near the top of the caisson
Figure 4. Input earthquake motion.

300.00
Tokachi-Oki earthquake of magnitude 7.9. This is the
standard ground motion used frequently in the design 200.00

Acceleration (Gal)
of Port and Harbour facilities in Japan. Nozu (2004), 100.00
however, emphasized the need to adopt a design
ground motion reflecting the site characteristics. 0.00

-100.00

4 RESULTS AND DISCUSSION -200.00

-300.00
4.1 Acceleration on the caisson 0.00 5.00 10.00 15.00 20.00 25.00 30.00
Time (sec)
Figures 5 a, b show the response accelerations of the (b) Near the bottom of the caisson
caisson with sandwiched cushion at the elements
near the top and bottom of the caisson-cushion inter- Figure 5. Response accelerations.
face (Interface I in Fig. 2). It can be observed that
responses at the top and the bottom of the caisson are
quite different. While the top experiences a high accel-
values were obtained by summing up the nodal stresses
eration magnitude, the bottom experiences relatively
of the elements at the caisson-cushion interface. It can
low acceleration magnitude. Similar behavior was also
be observed that the use of cushion could significantly
observed in the case when no cushion was used.
reduce the earth pressure acting on the wall. While the
The results are reported and discussed elsewhere
caisson without any protective cushion experiences
(Hazarika et al. 2005).
high fluctuation of the earth pressure with a predom-
inant peak, the earth pressure on the protected caisson
4.2 Normal and shear force at the interfaces stabilizes soon after reaching the peak. The magnitude
of the latter is also much lower than the former, imply-
The response of the interactive system at the ing an increase in stability of the caisson.
caisson-cushion interface was also examined to see
how the developed model could interpret the soil-
structure interaction phenomenon. Figures 6 a, b show 4.4 Distribution of earth pressure
such responses. It can be seen that, both the normal The dynamic earth pressures at the maximum inertia
and the shear force exhibit higher values at the top force are plotted against the wall height to observe
part. The normal force at the top, however, does not how the pressure distributes along the height. Figure 8
increase much, and drops to zero as compared to at shows such distribution for the two cases considered
the bottom. This can be attributed to the debonding in the analyses. Only the values at some particular
(separation) tendency at the caisson-cushion inter- elements at the caisson-cushion interface were plot-
face. Such relative deformation at the interfaces, dur- ted. It can be seen that the distribution pattern is dif-
ing the dynamic loading, can be explained well by the ferent for the two cases. While that without cushion
interface model described in this paper. shows a nonlinear increase with the wall depth, the
one with the cushion shows a maximum increase in
the middle of the caisson height, and then a gradual
4.3 Resultant earth pressure
decrease. This tendency once again demonstrates that
Figure 7 shows a comparison of the resultant horizon- the compressibility effect becomes dominant with the
tal dynamic earth pressure acting on the caisson. The increase of depth, however, the dominancy wanes after

580

Copyright 2005 Taylor & Francis Group plc, London, UK


1.00 7.00
Top of Caisson
0.80 Bottom of Caisson 6.00
0.60
5.00
Normal Force (kN)

Wall Height (m)


0.40 Without cushion
0.20 4.00

0.00 3.00
With cushion
-0.20
2.00
-0.40
1.00
-0.60
-0.80 0.00
0.00 5.00 10.00 15.00 20.00 25.00
-1.00
0.00 5.00 10.00 15.00 20.00 25.00 30.00 Seismic Pressure (kPa)
Time (sec)
(a) Normal force Figure 8. Pressure distribution.

0.30
Top of Caisson a cushion layer behind the structure. This paper
Bottom of Caisson
0.20 describes an interface model that can simulate such
soil-structure interaction process involving hybrid inter-
Shear Force (kN)

0.10 faces (structure-cushion-soil) of finite thickness with


dissimilar stiffness and constitutive characteristics.
0.00 The model, that was proposed here, was applied to
simulate the response of an ideal caisson type quay
-0.10 wall subjected to an actual earthquake loading. The
normal and shear deformations at the interfaces could
-0.20 be evaluated well using the model. The sandwiched
cushion contributes to reduction of both the static and
-0.30
0.00 5.00 10.00 15.00 20.00 25.00 30.00 the dynamic load on such gravity type structures. The
Time (sec) results of the analyses also established that the use of
(b) Shear force a compressible layer renders substantial reduction (50
to 60%) of seismic increment against the structure,
Figure 6. Normal and shear forces at the caisson-cushion thereby, providing stability to the structure during the
interface. earthquake. Reduced load also leads to a slim struc-
ture with reduced material cost, and hence an eco-
60.00 nomic design.
Seismic Earth Pressure (kPa)

Without Cushion
40.00
With Cushion Experimental studies are in progress to evaluate
the proposed technique. The results are expected to
20.00 provide feedback to the refinement of the model so
0.00 that it can be applied to wide range of problems
involving offshore and coastal facilities.
-20.00

-40.00

-60.00
0.00 5.00 10.00 15.00 20.00 25.00 30.00 ACKNOWLEDGMENT
Time (sec)
The financial support for this research was provided
Figure 7. Resultant dynamic earth pressure. by Port and Airport Research Institute (PARI) under
the special grant for budding research. The author
gratefully acknowledges the support.
reaching a certain depth. That depth may depend upon
the height and rigidity of the structure as well as the
rigidity and the thickness of the cushion material itself.
REFERENCES
5 CONCLUSIONS Day R.A., and Potts, D.M. 1998. The Effect of Interface
Properties on Retaining Wall Behavior, International
An economic and cost-effective design alternative for a Journal for Numerical and Analytical Methods in
caisson type offshore and coastal structure is to use Geomechanics, 22: 10211033.

581

Copyright 2005 Taylor & Francis Group plc, London, UK


Desai C.S., Zaman, M.M., Lightner, J.G., and Siriwardane, Novel Expanded Polystyrene Foam Buffer Layer, Proc. of
H.J. 1984. Thin Layer Element for Interfaces and Joints, the 10th Annual Symp. on Earth Retention System,
International Journal for Numerical and Analytical Canadian Geotechnical Society, Vancouver, Canada: 110.
Methods in Geomechanics, 8: 1943. International Navigation Association (PIANC) 2001.
Hazarika, H., and Matsuzawa, H. 1996. Wall Displacement Seismic Design Guidelines for Port Structures: Balkema
Modes Dependent Active Earth Pressure Analyses Using Publishers, Rotterdam.
Smeared Shear Band Method with Two Bands, Japan Society for Civil Engineers (JSCE) 2001. The Third
Computers and Geotechnics, 19(3): 193219. Proposal on Earthquake Resistance for Civil Engineering
Hazarika, H., Kohama, E., Suzuki, H., and Sugano, T. 2005. Structures (In Japanese).
Enhancement of Earthquake Resistance of Structures Karpurapu, R., and Bathrust, R.J. 1992. Numerical
using Tire chips as Compressible Inclusion, Research Investigation of Controlled Yielding of Soil-retaining
Report of Port and Airport Research Institute (To be Wall Structures Geotextiles and Geomembranes,
Published). 11:115131.
Hazarika, H., and Okuzono, S. 2004. Modeling the Behavior Nozu, A. 2004. Current Status of Strong-motion Earthquake
of a Hybrid Interactive System Involving Soil, Structure Observation in Japanese Ports, Special Issue of Journal
and EPS Geofoam, Soils and Foundations, Japanese of Japan Association for Earthquake Engineering, 4(3):
Geotechnical Society, 44(5): 149162. 7983.
Hazarika, H., Okuzono, S., and Matsuo, Y. 2003. Seismic Nozu, A., Ichii, K., and Sugano, T. 2004. Seismic Design of
Stability Enhancement of Rigid Nonyielding Structures, Port Structures, Special Issue of Journal of Japan
Intl. Society of Offshore and Polar Engineers (ISOPE) Association for Earthquake Engineering, 4(3): 195208.
Transactions, 2: 697702. Yasuhara, K. 2002. Recent Japanese Experiences with
Iai, S. 1998. Seismic Analysis and Performance of Retaining Lightweight Geomaterials, Keynote Lecture, Proc. of the
Structures, Geotechnical Earthquake Engineering and Intl Workshop on Lightweight Geo-Material, (IW-LGM
Soil Dynamics, Geotechnical Special Publication No. 75, 2002), Tokyo, Japan: 3559.
ASCE: 10201044. Youwai, S., and Bergado, D.T. 2003. Strength and
Inagaki, H., Iai, S., Sugano, T., Yamazaki, H., and Inatomi, T. Deformation Characteristics of Shredded Rubber Tire-
1996. Performance of Caisson Type Quay Walls at Kobe Sand Mixture, Canadian Geotechnical Journal, 40:
Port, Special Issue of Soils and Foundations, Japanese 254264.
Geotechnical Society: 119136. Zaman, M.M., Desai C.S., and Drumm, E.C. 1984. Interface
Inglis, D., Macleod, G., Naesgaard, E., and Zergoun M. Model for Dynamic Soil-Structure Interaction, Journal
1996. Basement Wall with Seismic Earth Pressures and of Geotechnical Engineering, ASCE, 110(9), 12571273.

582

Copyright 2005 Taylor & Francis Group plc, London, UK


Progressive ocean wave modelling in drum centrifuge

F.P. Gao1,2 & M.F. Randolph2


1
Institute of Mechanics, Chinese Academy of Sciences, Beijing, P.R. China
2
Special Research Centre for Offshore Foundation Systems, The University of Western Australia,
Nedlands, Australia

ABSTRACT: Ocean wave modelling in the centrifuge is a novel technique for the study of seabed response
to waves and wave-seabed-structure interaction problems in offshore engineering. Based on the similarity
analysis of wave simulation in centrifuge, a paddle-type wave generation system was established in a drum cen-
trifuge. A variant of a double-beach wave absorber, made of porous metal plate covered with plastic mesh, was
adopted for the purposes of wave absorption. Instrumentation comprised a wave height monitor in the centre of
the water channel and pore pressure transducers in the base of the channel at two positions relative to the pad-
dle device. Results from a series of tests are presented and show that the wave generation system was able to
simulate progressive waves in shallow to intermediate water depths.

1 INTRODUCTION wave generation. Nevertheless, the efficiency of the


slotted partition wave absorber is relative to the wave-
Under severe ocean wave loading, sandy seabeds may length of the waves generated; only waves whose wave-
undergo shear failure or liquefaction, which may cause length is about 4 times the distance between the
instability of submarine pipelines, offshore platforms partition and the far end wall are dissipated with low
or other marine facilities. In the experimental investi- reflection coefficients, thus the wave frequency could
gation of the above problems, it is necessary to simu- not be changed during experiments in the beam cen-
late ocean waves in the laboratory. The most common trifuge (Sekiguchi et al. 1998). Compared with the
approach for wave simulation are flumes or tanks oper- beam centrifuge, the drum centrifuge offers more
ating under 1 g conditions, but very large flumes or advantages in many aspects, and in particular the longer
tanks would be required for most problems of practical testing area along its circumference, which is very help-
significance unless the problems are scaled down. This ful for water wave generation (Stewart et al. 1998).
can lead to difficulties in correctly simulating the In this paper, the principle of wave generation in
seabed response, which is stress-level dependent. centrifuge is presented first, based on which a new
Centrifugal experiments utilizing scaled models in progressive wave generation system is described, as
a high gravity environment are accepted as the best implemented in the drum centrifuge at the University
approach to create the stress conditions appropriate of Western Australia. A series of progressive wave
for prototype conditions (Schofield 1980, Taylor generation tests were conducted with the new facility
1995) and are particularly suitable for modelling and the measured results are compared with the theor-
gravity and time dependent problems. Wave simula- etical predictions.
tion in the centrifuge is a key technique for carrying
out the research mentioned above.
Tsunami-like solitary waves were first generated
2 MECHANISM OF WAVE GENERATION
in the drum centrifuge at Cambridge University, by
IN GEO-CENTRIFUGE
means of immersing a rectangular float into water
(Sekiguchi & Phillips 1991). Thereafter, standing
2.1 Similarity analysis for ocean wave modelling
waves and progressive waves were generated on the
in centrifuge
conventional beam centrifuge at Kyoto University
(Sekiguchi et al. 1994, Sekiguchi et al. 1998). Con- In the case of ocean waves with a free surface, gravi-
sidering the limited length of the test box in beam tational effects predominate while the effects of other
centrifuge, a slotted vertical partition was employed factors, such as viscosity, surface tension, etc, are
as a wave absorber for the purpose of progressive generally small and can be neglected. It is well known

583

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 1. Similarity laws for wave modelling in a centrifuge fluid can be described by a velocity potential flow and
based on Froude scaling. the velocity potential satisfies the Laplace equation.
Regular harmonic waves are commonly used for
Scale understanding the physics of fluidstructure inter-
Variable Unit Factor ()
action. According to wave generation theory (Hughes
Centrifugal acceleration (a) LT
2 N 1993), in the water region defined by 0 y
d as
Froude number (Fr) 1 1 shown in Figure 1, the wave elevation h(x,t) of the sinus-
Water depth (d ) L 1/N oidal wave generated with paddle-type wave maker is
Wave length (L) L 1/N
Wave period (T ) T 1/N
Wave velocity (c) LT
1 1
Water particle velocity (U ) LT
1 1
Fluid (water) density ( ) ML
3 1
Wave pressure on seabed (Po) ML
1T
2 1 (5)

that with water both Reynolds and Froude similitude


cannot be achieved concurrently. For gravitational wave y
simulation, it is reasonable to employ the Froude scal-
ing process with allowance made for variation in S L
Reynolds number (Chakrabrarti 1994). 0 x
The Froude number (Fr), which physically indi- paddle H
cates the ratio of inertial force to gravitational force, d
is defined as
ds water

(1)

where Um is the maximum value of wave-induced Figure 1. Sketch of paddle type wave maker.
water particle velocity, a is the centrifugal accelera-
tion, D is a characteristic dimension (e.g. the water
depth). According to Froudes law, the following scal- where S is the stroke of paddle at water surface,  is
ing law should be maintained: the angular velocity of the wave generated by the pad-
dle (  2 /T ), d is the water depth, ds ( 0.02 m) is
(2) the vertical distance from the seabed surface to the
paddle hinge, g is the acceleration of gravity, k
( 2 /L) is the wave number, and L is the wave
where  represents the ratio of the model parameters length. The wave dispersion equation is
to those of the prototype. In centrifuge tests, D  1/N
and a Ng, in which N is the centrifugal g-level and (6)
g is the gravitational acceleration. Therefore
(3) From Equation (5), the wave height H can be
expressed as
The wave-period scale can be deduced as
(4) (7)

where T is the wave period. The similarity laws for


ocean wave modelling in a centrifuge can be deduced
as listed in Table 1, in which c is the wave velocity 3 EXPERIMENTAL SETUP
(c  L/T ).
3.1 Drum centrifuge
The wave modelling experiments were conducted in
2.2 Theoretical background to the paddle-type
a drum centrifuge at the University of Western
wave maker
Australia. The drum centrifuge features a channel for
For gravitational ocean waves, the fluid can be sample containment, 300 mm in width (measured ver-
assumed to be homogeneous, inviscid, incompressible tically) and 200 mm in depth (measured radially), as
and its motion irrotational. As such, the motion of the shown in Figure 2. The channel has an outer diameter

584

Copyright 2005 Taylor & Francis Group plc, London, UK


Drum centrifuge channel
wave absorber
PTb

471
separating wall
wave probe

A A R60
0 PTa

wave maker

LDT

(a) Plan view of the wave generation system


Figure 2. View of drum centrifuge with protective enclosure
removed and control tool table in position.
paddle
D  1.2 m, and its maximum rotational speed is craft shaft
cam wheel
850 rpm, equivalent to an acceleration of 480 g at the DC servo motor
bottom of the channel. A novel and versatile tool Drum-centrifuge
300 motor housing
channel wall
table, which is located at the centre of the sample con-
tainment channel, can be stopped and started inde-
pendently of the channel. A fast sampling data
acquisition system, whose maximum sampling rate is
100 kHz, was designed for dynamic problems. The (b) A-A section view for the paddle-type wave maker (enlarged)
sample containment channel is also fitted with a fluid
distribution system, to enable the channel to be filled Figure 3. Arrangement of wave generation system in
with water or emptied easily (Stewart et al. 1998). drum-centrifuge (dimensions in mm).

3.2 Paddle-type wave maker in drum centrifuge


There are several methods to generate progressive A Maxon DC Servo Motor was chosen, with max-
waves on the ground. However, the design of the device imum continuous torque of 2.94 Nm, and maximum
for wave generation in the centrifuge must take account intermittent torque of 5.88 Nm. Thus the paddle can
of the high centrifugal acceleration environment and sway with a given frequency to obtain the desired
the corresponding high frequency of the waves to be wave height. The movement of the paddle is moni-
generated. For these reasons, a paddle-type wave maker tored with a Laser Displacement Transducer (LDT), as
was adopted, the key parts comprising a cam wheel shown in Figure 3(a).
driven by a servo DC motor, a paddle hinged at the bot-
tom of centrifuge channel, and a crank shaft to link the
paddle and cam wheel, as shown in Figure 3. 3.3 Wave absorber design
The lengths of the components of the wave maker
The incident waves generated by the wave maker must
were designed based on the crank-guide mechanism,
be dissipated in the drum centrifuge channel without
to achieve a maximum paddle-swaying angle of about
reflection in order to achieve high quality data. Many
18. The radius of the cam wheel is changeable from
different types of wave absorbers have been tried in
10 mm to 24 mm with an interval of 2 mm, by means
wave flumes and basins, with varying degrees of suc-
of screw holes with various offset distance on the cam
cess. They can be divided into two types: (1) caisson-
wheel. A closed-loop servo-controlled system, using
type and (2) beach-type.
a DC tacho mounted directly on the motor shaft, pro-
The caisson-type absorber consists of a slotted ver-
vides close control of the motor rotation speed. The
tical front partition and an internal chamber behind
servo-motor must provide sufficient torque to over-
the partition. In this type of absorber, transformation
come the following main loads:
of the wave energy into turbulence, produced by the
1. wave loading on paddle; narrowness of the front apertures, plays an important
2. inertia force on paddle and crank shaft; part in the wave dissipation phenomenon. In order to
3. centrifugal force on crank shaft. achieve high wave absorption efficiency, rectangular

585

Copyright 2005 Taylor & Francis Group plc, London, UK


vertical apertures are distributed over a quarter to a
third of the front partition.
Since one of the main advantages of this type of
absorber over a beach-type absorber is the very short porous plate
length, it was chosen for use in beam centrifuge tests,
where the box length was only 0.55 m (Sechiguchi
et al. 1998). It was found, however, that this kind
of absorber achieved an efficiency of 90% or more
only when the distance between the partition and
the back wall of the chamber was about a quarter
seperating wall
of the wavelength. This is a disadvantage if the
absorber is required to cope with a wide range of wave
Figure 4. Sketch of porous beach type absorber.
frequencies.
The beach-type absorber is simpler and more com-
mon, compared with the caisson-type absorber. The 3.4 Testing procedure
main working principle of this type of absorber is dis- The following testing procedure was adopted to gen-
sipation of the wave energy by the waves breaking on erate progressive water waves in the drum centrifuge:
a gently sloping plane of significant length. Wave
absorbing beaches have the advantage of simple 1. firstly, the wave maker, wave absorber and all trans-
design and of being very efficient for a wide range of ducers were installed in drum centrifuge channel;
wave frequencies. However, their large size prohibits 2. the centrifuge was then accelerated to a pre-set
use in a conventional beam centrifuge. g-level, such as 50 g;
Many solutions have been proposed to minimise 3. the centrifuge channel was filled with water to a
the size of the wave absorbing beaches under 1-g con- given height (e.g. 140 mm) through the fluid distri-
ditions. Constructions combining the slope of absorb- bution system;
ing beaches and porous walls increase the efficiency 4. thereafter, the DC servo-motor was started at a pre-
and also reduce the size (Ouellet & Datta 1987). As a determined frequency (e.g. 10 Hz) to generate the
general rule it has been found that combining several waves;
wave absorbing energy effects leads to an important 5. the fast data acquisition system was then employed
reduction in the required size. to sample the wave height and the wave-induced
Even though the channel of the drum centrifuge is pressures on the seabed, the latter measured with
much longer than the sample box for a beam cen- two pressure transducers, PTa and PTb, as shown
trifuge, its limited space still necessitates the use of in Figure 3.
an absorber with minimum length and slopes that are For the measurement of the wave height, a
steeper than optimum. The benefits of using perme- capacitance-type wave probe is employed, which is
able material in the sloping beach are well known in located above PTa as depicted in Figure 3. In addition,
both field and laboratory conditions. In order to maxi- the movement of the paddle was checked by a laser
mise the efficiency over a wide range of wave condi- displacement transducer (LDT), which was installed
tions, a special wave absorber was designed, based on on the centrifuge channel.
a double beach-type wave absorber made of porous
metal plate covered with plastic mesh (Figs. 3, 4).
Due to the swaying motion of the paddle, two sets
4 EXPERIMENTAL RESULTS AND
of waves are generated travelling in opposite direc-
DISCUSSIONS
tions around the channel. A separating wall is
mounted at the middle of the double beach, to avoid
4.1 Typical experimental results
the two series waves affecting each other. Only the
waves generated in the longer part of the channel are Typical experimental results are shown in Figure 5. The
used for the experiments. The purpose of using porous wave was generated under conditions: a  50 g,
metal sheet and plastic mesh was to increase the wave d  0.12 m. Note that the g level quoted is that at the
breaking phenomenon and the viscous dissipation bottom of the channel. The wave frequency is 8.6 Hz. It
process of the wave energy. The slope of each porous can be seen from Figure 5 that the pressure amplitudes
beach was variable to reduce the total required length at PTa and PTb are practically the same. The ratio of
of the beach. Besides the common advantages over wave propagation period between PTa and PTb (t) to
caisson-type absorbers, another main advantage of the wave period (T ) is t/T  0.08/0.12  0.67. Based on
new absorber is its shorter length than conventional the dispersion equation (6), the wavelength can be
beaches. The efficiency of the absorber will be dis- obtained as L  0.68 m. Thus the ratio of the distance
cussed in Section 4.2. between PTa and PTb (l) to the wavelength (L) is

586

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 2. Results of wave modelling in drum centrifuge.

Test d f H L P0
No. Nb (m) (Hz) (mm) (m) (kPa)

1 50 0.12 5.08 7.1 0.66 1.01


2 50 0.12 6.84 9.2 0.65 1.03
3 50 0.12 8.60 12.6 0.55 1.05
4 50 0.15 6.45 18.3 0.67 1.31
5 50 0.15 7.62 20.2 0.62 1.26
6 80 0.12 6.84 8.3 0.71 1.54
7 80 0.12 8.40 9.1 0.63 1.50
8 80 0.15 7.62 15.4 0.60 2.10
9 80 0.15 8.02 15.6 0.52 2.31

Figure 5. Typical wave generation experiment results. Figure 6. Regions of validity for available wave theories.

l/L  0.47/0.68  0.69. The agreement between the frequency and g levels. The measured wave param-
time and distance proportions therefore confirms that eters in the experiments are listed in Table 2, with meas-
the water waves generated are consistent with a pro- ured wave heights of up to 1.2 m. Guided by the two
gressive pattern. governing dimensionless groups above, the appropri-
ate theory can be chosen from those available. As indi-
4.2 Comparison with theoretical results cated in Figure 6, Stokes 2nd-order theory is the
appropriate one for the present tests, so that the
Several theories exist for computing wave-induced expression for the amplitude of wave-induced pres-
pressures on the seabed. These include the Airy (linear) sure on an impermeable rigid horizontal bottom is
wave theory, the second order Stokes wave theory and
higher order wave theories. Stokes theory gives a
multi-order approximation to the exact solution of the
differential equations describing wave behaviour and is
a finite amplitude wave theory. The choice of an appro- (8)
priate theory for a particular situation is determined by
the wave characteristics: wave height H, wave period T,
water depth d and gravitational acceleration. In the in which N is chosen as the effective g level, which is
centrifuge tests, gravitational acceleration is replaced related to the g-level at the base of drum centrifuge
by the centrifugal acceleration a. The relevant character- channel Nb by N  (D  2d/3)Nb. The theoretical
istics are grouped into two governing dimensionless wave-induced pressures corresponding to the measured
parameters: d/aT2 and H/aT2 (Dean 1994). wave height, wavelength and wave frequency are cal-
The wave modelling experiments were conducted culated from Equation (8). With the measured pres-
at various water depths (up to 12 m prototype), wave sure as ordinate value and the theoretical prediction

587

Copyright 2005 Taylor & Francis Group plc, London, UK


established on a drum centrifuge. Comparison of the
phase shift in measured pore pressure changes and the
distance between the transducers confirmed the pro-
gressive nature of the waves. A novel double beach-
type wave absorber was found to provide an efficient
solution for the absorption of waves over a range of
frequencies. The experimental results show that the
wave generation system is capable of modelling pro-
gressive ocean waves in intermediate to shallow water
zones, with wave heights of up to 10% of the water
depth. The device can be employed for further appli-
cations in experimental research on wave-structure-
seabed interaction.

ACKNOWLEDGEMENTS

Figure 7. Comparison of measured wave pressures with Funding from the Australian Research Council (Grant
theoretical predictions. No. A00104092) is gratefully acknowledged. We are
also grateful to Bart Thompson who provided general
assistance with assembly, testing and operation, and
Clem Ryan carried out the electronics work on the
servo-motor control and other parts of the drum cen-
trifuge. The first author is also grateful for the support
from Tenth Five-year Plan of Chinese Academy of
Sciences (Grant No. KJCX2-SW-L03).

REFERENCES

Chakrabarti, S.K., 1994. Offshore Structure Modeling. JBW


Printers & Binder Pte. Ltd.
Dean, R.G., Dalrymple, R.A., 1994. Water Wave Mechanics
for Engineers and Scientists. World Scientific, Singapore.
Goda, Y. and Suzuki, Y., 1976. Estimation of incident and
reflected wave in random wave experiments with reflect-
ing coastal structures. Proceedings, Fifteenth Coastal
Engineering Conference, ASCE, Vol.1, 828845.
Figure 8. Wave reflection coefficients. Hughes, S.A., 1993. Laboratory wave generation. In:
Physical Models and Laboratory Techniques in Coastal
as abscissa, the data points are then plotted in Figure 7. Engineering, World Scientific Publishing Co., 333457.
Ouellet, Y., Datta, I., 1987. A survey of wave absorbers.
As indicated in this figure, most of the test data points
Journal of Hydraulic Research, 24 (4): 265280.
lie within about 15% error range from the 45 solid Schofield, A.N, 1980. Cambridge geotechnical centrifuge
line. The measured values and the theoretical predic- operations. Gotechnique, 30(3), 227268.
tions are in good agreement. Sekiguchi, H., Kita, K., Okamoto, O., 1994. Wave-induced
In addition, the wave reflection coefficients can be instability of sand beds. Proceedings, International
obtained using the method of Goda et al. (1976), as Conference CENTRIFUGE 94, Singapore, 1994: 295300.
shown in Figure 8. Unlike the vertical partition wave Sekiguchi, H., Kita, K., Sassa, S., and Shimamura, T., 1998.
absorber used in the beam centrifuge, the wave reflec- Generation of Progressive Fluid Waves in a Geo-Centrifuge.
tion coefficients of the beach-type absorber are nor- Geotechnical Testing Journal, ASTM, 21(2): 95101.
Sekiguchi, H., Phillips, R., 1991. Generation of water waves
mally less than 20%, indicating that the wave absorber
in a drum centrifuge. Proceedings, International
is efficient over a wide range of wave frequencies. Conference CENTRIFUGE 91, Colorado, 343350.
Stewart, D.P., Boyle, R.S. and Randolph, M.F., 1998. Experi-
ence with a new drum centrifuge. In: Proceedings, Inter-
5 CONCLUSIONS national Conference CENTRIFUGE 98, Tokyo, 3540.
Taylor, R.N, 1995. Centrifuges in modelling: Principles and
Based on a similarity analysis for modelling progres- scale effects. Geotechnical Centrifuge Technology, Blackie
sive waves, a new wave generation system was Academic & Professional, 1933.

588

Copyright 2005 Taylor & Francis Group plc, London, UK


Numerical analysis of dynamic response of seabed under random
wave loading

Zhongtao Wang1, 2, Maotian Luan1, 2, 3, Zhange Liu1, 2 & Dong Wang1, 2


1
State Key Laboratory of Costal and Offshore Engineering, Dalian University of Technology, Dalian, China
2
Institute of Geotechnical Engineering, School of Civil and Hydraulic Engineering, Dalian University of
Technology, Dalian, China
3
Institute of Rock and Soil Mechanics, The Chinese Academy of Sciences, Wuhan, China

ABSTRACT: The analysis of dynamic response of seabed due to wave loading is of practical significance in
design and construction of marine structures and offshore installations. The purpose of this paper is to discuss
the influence of the feature of randomness of wave loading on the dynamic response of seabed under stochastic
wave loading. Comparative study is principally made between the presented analysis considering randomness
of wave loading and conventional deterministic analysis based on linear theory of regular wave. Based on the
generalized Biots dynamic theory of consolidation of soil as a linear elastic system, a FEM-based numerical
formulation in the form of uU is established and is solved in time domain. The analyses considering characteris-
tics of randomness of wave loading are formulated in a stochastic framework. Indications from the numerical results
which are given in graphical form will be helpful to evaluate the construction safety and seabed instability.

1 INTRODUCTION Some effective procedures based on theory of


dynamics of random vibration which are successfully
The stability of seabed constitutes an important con- adopted in earthquake engineering offer a sound basis
sideration in planning and design of various marine for random analyses of dynamic response of seabed
structures and offshore facilities. The significant under stochastic wave loading. Recently, numerical cal-
dynamic water pressures generated in the sea floor dur- culations and field measurements on this subject have
ing ocean wave propagation will induce stresses and been done by Rahman et al. (1986, 1991), Madsen
associated pore water pressure fluctuations within the (1994), Myrhaug et al. (1998), Liu (2002), Pilotto
seabed, which may cause shear failure or liquefaction in (2003). However, these investigations at present rarely
sea floor sediments. Considerable efforts for this prob- can link the actual feature of sea waves with charac-
lem have been made with growing interest partially teristic parameters of seafloor sediments for practical
because a vast number of offshore installations is being constructions.
or will be built and partially because indeed a number In this paper, wave spectrum is employed to simu-
of marine or offshore facilities or sites have been late random characteristics of waves which is able to
damaged due to wave-induced seabed instability. It is consider a number of influence factors in different gen-
obvious that wave loading plays a significant role in the eration phases of waves. The seabed is considered as
evaluation of construction safety and seabed instability. isotropic homogeneous elastic medium with a finite
However, in the conventional analyses of seabed dynam- thickness. The generalized Biots dynamic theory of
ics, wave loading is basically regarded as a determin- consolidation is used in conjunction with FEM leaving
istic process and is usually taken into consideration the displacement vector of soil skeleton u and the
by using linear theory of regular wave. In fact, ocean total displacement vector of porous water U as the
wave is of intrinsic randomness in both time sequences primary variables. Hence numerical formulation in
and spatial distribution. The random nature of both the form of uU is established and is solved in time
wave and wave-induced loading will subsequently affect domain. Based on numerical results, the difference
dynamic response behavior of seabed. Therefore the sto- between the proposed analysis for random wave and
chastic feature of wave loading has to be duly taken into the conventional analysis for linear regular wave is
account in the analysis on dynamic response of seabed. examined.

589

Copyright 2005 Taylor & Francis Group plc, London, UK


2 RANDOM WAVES 3 FINITE ELEMENT FORMULATIONS OF
THE GENERALIZED BIOTS EQUATION
The wave motion is postulated to be a stationary
Gaussian narrow-band process and can be represented 3.1 Weak formulation of the governing equation
as a superposition of numerous linear regular waves
The generalized Biots dynamic equations of consoli-
with different heights, frequencies and random initial
dation of soil have been popularly employed in geo-
phase angles. The statistical properties of wave surface
mechanics. They have been used in conjunction with
are completely specified by the spectral density
the finite element method by Zienkiewicz and et al.
function, which is achieved from the autocorrelation
(1980) to establish their numerical formulations in
function of the Fouriers transform of wave surface
uU form. Its general weak form can be given by the
fluctuations. As first approximation, a linear random
Garlerkins weighted residual approach (Luan and
wave, being homogeneous and stationary Gaussian
et al., 2001) which is rewritten as following
process with a zero mean, is used here. The transient
height of the surface of wave relative to static water line,
(x, t), may be represented by

(4)

(1)

(5a)

where ai  wave amplitude; ki  wave number; i 


wave frequency; Li  wave length; Ti  wave period;
d  water depth; i  random initial phase angle,
which is evenly distributed in the range of (0, 2 );

i  representative wave frequency, which is evenly (5b)
distributed in the second half of (i
1, i); the sub-
script i represents the sequence number of compositive
waves; g  acceleration of gravity; M  50100.
S() is the frequency spectrum of the surface fluc-
tuation. The average JONSWAP spectrum (Hasselmann,
1973) as following is used here
(5c)

(2)

where   scaling factor of energy;   rising factor (5d)


of spectrum; m  wave frequency corresponding to
peak value of spectrum;   parameter of shape of peak
value.
It is assumed that waves propagate over a rigid where M, K and C are the mass, stiffness and damping
seabed. Therefore, the resulting wave-induced pres- matrices of the system, respectively, and these three
sures p on the seabed surface are obtained with no matrices remain invariable; F is the matrix of force
consideration of interaction effect of waves and seabed vector; D  matrix of elastic coefficient, B  matrix
and are given as below of geometry; f  density of pore fluid; kx, kz  coef-
ficients of permeability respectively in x- and z-
directions; Q  kf/n, kf  compressibility modulus,
(3) n  porosity; Nku, NkU  interpolation functions of u
and U, pn  perpendicular pore water pressure on the
surface of seabed, n  total perpendicular stress on
where  the total density of sea floor sediments. the surface of seabed.

590

Copyright 2005 Taylor & Francis Group plc, London, UK


3.2 Solution procedure for finite element equations
The following composite variable of the system is 1 3
introduced,

(6)
2 4
The Newmark- scheme of time integration as given
below is used

Figure 1. The demonstrative layout of schemes of meshes


(7) generation in seabed and serial numbers for nodes and
elements.

the natural division of elements, the vertical depth of


elements is approximately to be equal to the horizontal
(8) length of element. Shown in Figure 1 is a demonstrative
layout of the typical finite element meshes consisting
of 2  2 elements. In this figure, the numbers with
circle mark represent sequence numbers of elements,
Substitution of these two equations into Equation 4 the numbers without circle mark represent sequence
leads to numbers of the nodes.

3.5 Boundary conditions


On the surface of seabed, the total normal stress n
and pore pressure p equal to wave-induced pressure.
(9)
On the bottom of seabed, all displacements are assumed
to vanish, i.e., ux, uz , Un equal to zero. On both lateral
The resulting displacement at t  t can be solved
sides of seabed, un, Un are assumed to equal to zero.
from this equation and thereafter the acceleration and
Although this type of boundary condition would lead
velocity at t  t can be determined from Equation 7.
to a certain energy accumulation in the finite domains
In order to assure the stability of numerical computation
which is included in computational model, a wider
and improve the precision and convergence, specific
range for computation is chosen enough to eliminate
values of the integral parameter,  and are chosen by
the effect of artificial boundary on numerical precision
within the region interested.
(10)

4 NUMERICAL RESULTS AND ANALYSES


3.3 Interpolation functions
Based on available experience and theoretical justifi- In this paper, two-dimensional numerical computations
cation, interpolation function for the total displace- are made to illustrate the proposed method.
ment vector of porous water, U, should be one order
lower than that of interpolation for the displacement 4.1 Simulation and verification of random waves
vector of soil skeleton, u. Therefore 8-node quadratic
quadrilateral element for u and 4-node linear quadri- The physical properties of seafloor sediments and
lateral element for U are adopted in order to avoid the wave parameters used in the analyses are shown in
spurious oscillation of numerical results. Table 1. Average JONSWAP wave spectrum is used and
a storm with duration of 1 h is assumed. From the
available data, random waves are simulated, and com-
3.4 Solution to FEM meshes
parison of (x, t) between random waves and linear
The seabed is assumed to be horizontally layered with waves is shown in Figure 2.
different soil parameters. The seabed is divided equally In order to assure the precision and efficiency of
to elements along the horizontal direction. The horizon- numerical simulation, the characteristic parameters for
tal length of element is adopted between 2040% of diagnosis of time sequence of random waves simulated
wave length. In addition to the layer interfaces which are are worked out as shown in Table 2. Here, cv and cs

591

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 1. Physical parameters used in numerical examples. Table 2. The statistic characteristics of simulated random
waves and theoretical solutions.
Coarse
Parameter sandy Fine sandy Result Theoretical
Parameter simulated solution
Coefficient of permeability, kx 10
2 (m/s) 10
4 (m/s)
Coefficient of permeability, kz 10
2 (m/s) 10
4 (m/s) H 3.67 (m) 3.75 (m)
Water depth, d 20 (m) 20 (m) H1/3 5.66 (m) NA
Distance of wind, X 220 (km) 220 (km) H1/3/H 1.54 1.598
Poisson ratio,  0.333 0.333 H1/10 7.53 (m) NA
Shear modulus, G 10 (MPa) 10 (MPa) H1/10/H 2.05 2.032
Porosity, n 0.3 0.4 Hmax 10 (m) 9.729.87 (m)
Thickness of seabed, h 30 (m) 30 (m) cv of wave height 0.523 0.522
Length of seabed, L 750 (m) 750 (m) cs of wave height 0.828 0.635
Density of pore fluid, f 1.0 (t/m3) 1.0 (t/m3) T 8.88 (s) NA
Total density of soil, 1.7(t/m3) 1.8 (t/m3) Tmax 13.23(s) NA
Significant wave height, Hs 6 (m) 6 (m) TH1/3 9.36 (s) NA
Significant wave period, TH 1/3 10 (s) 10 (s) cv of wave period 0.242 0.283
cs of wave period 0.732 0

5
4 Linear regular wave
Random waves simulated 12
3
2 Predicted spectrum
10
Target spectrum
h (t) (m)

1
S (v) (m2..s)

0 8
-1 6
-2
-3 4
-4 2
-5
0 20 40 60 80 100 120 0
0 10 20 30 40 50
Time (s)
v (rad/s)
Figure 2. Comparison of time history of linear regular
wave and random wave simulated. Figure 3. Comparison of target spectrum and predicted
spectrum.

represent respectively the coefficient of discrete-


ness and coefficient of aberrance; H  and T represent
respectively the average wave height and wave period;
NA represents that there is no theoretic result. Mean-
while, the frequency spectrum of auto-correlation,
S(), is acquired from the results simulated. The
predicted spectrum is compared with the target spec-
trum in Figure 3. It is shown that the simulation of ran-
dom wave is effective and can be used for the analyses
of the dynamic response of seabed.

4.2 Dynamic response of sandy seabed


Numerical computations of dynamic response of fine
sandy seabed under random waves are carried out by
the proposed finite element method. The computed
effective stresses and pore water pressure is given in
Figure 4. It is demonstrated that the distribution of Figure 4. Distributions of normalized amplitudes of pore
normalized amplitudes of pore water pressure and pressures and effective stresses along depth in fine sandy
effective stresses along depth in fine-sandy seabed seabed.

592

Copyright 2005 Taylor & Francis Group plc, London, UK


| p | (kPa) | z | (kPa)
z/h

z/h
Figure 5. Distributions of amplitudes of pore pressures along Figure 7. Distributions of amplitudes of vertical effective
depth in fine sandy seabed. stresses along depth in fine sandy seabed.

| xz | (kPa)
| x| (kPa)
z/h
z/h

Figure 6. Distributions of amplitudes of horizontal effect- Figure 8. Distributions of amplitudes of shear stresses along
ive stresses along depth in fine sandy seabed. depth in fine sandy seabed.

seabed, different feature of dynamic response to ran-


subjected to random wave loading is similar to the dom and deterministic wave loadings is examined as
numerical results obtained under linear regular wave following. By varying the random initial phase angle, a
loading. In Figure 4, the dash lines represent the results series of random waves can be generated according to
of linear regular wave, and the tagged lines represent the simulation principle as proposed above. For a
the results of random wave. given number of wave generation, i.e., 100, a series of
dynamic analyses of seabed are made to the different
simulated random waves which are corresponding to
4.3 A comparative study
the same spectrum. Based on the numerical results, the
Although the distributions of normalized results under mean and standard deviation of amplitudes of dynamic
random waves are similar to those under linear regu- response can be solved from the statistic analyses. Com-
lar waves, the actual amplitudes under these two dif- parisons are made in Figures 58 between the varying
ferent patterns of wave loading are rather different. In ranges of dynamic response from random analysis with
fact, mechanisms of liquefaction and criterions of the numerical results gained from linear theory of
evaluation of instability of seabed or offshore building regular wave. It is noted that the results obtained by
are always directly relevant to the actual amplitudes random analysis and deterministic analysis based on
of pore water pressure and effective stress fluctuations linear theory of regular wave are respectively by solid
in the computed domains. Therefore, for the fine-sandy lines and dash line in the Figures 58. Furthermore, the

593

Copyright 2005 Taylor & Francis Group plc, London, UK


solid line in the middle represents the mean of dynamic is a CASs member, for his continuing support and
response under random wave loading. The possible invaluable advice for the investigation. The financial
range of dynamic response is given between the left and support for this study through the grants 50179006 and
right solid lines, which represent the mean plus/minus 50439010 from National Natural Science Foundation
one time of standard deviation of variable respectively. of China is mostly grateful.
It can be indicated that the mean dynamic response
under random wave loading is approximately same as
the results under linear regular wave loading, and the
REFERENCES
possible scattered range covers all the results given
based on the conventional method. Hasselmann, K. 1973. Measurements of wind wave growth
and swell decay during the Joint North Sea Wave Projected
(JONSWAP). Deutsches Hydrograph. Inst., Vol.12: 95.
5 CONCLUDING REMARKS Liu, Haixiao. 2002. Frequency domain analysis of dynamic
wave pressure on deeply embedded large cylinder struc-
The ocean waves during a storm may cause shear tures due to random waves. Proceedings of the Inter-
failure or liquefaction of seafloor, leading to damage national Offshore and Polar Engineering Conference,
of marine construction or instability of seabed. In Vol.12, pp. 701708.
this paper the intrinsic randomness of sea waves is Luan, Maotian, Wang Dong & Guo Ying. 2001. Numerical
analyses of dynamic response of porous seabed under
considered in the dynamic response of seabed. A wave loading (in Chinese). Ocean Engineering. Vol.19,
comparative study is made for the dynamic response of No.4, pp. 4045.
seabed subjected to random wave and linear regular Madsen, O.S. 1994. Spectral wave-current bottom boundary
wave through numerical computations. The main con- layers flows. Proceeding of the 24th Conference On Coastal
clusions can be stated as following. (1) The distributions Engineering, ASCE, Kobe, Japan, Vol.1, pp. 384398.
of normalized amplitudes of pore water pressure and Myrhaug, D., Slaattelid, O.H. & Lambrakos, K.F. 1998.
effective stresses under random wave are similar to Seabed shear stress under random waves: predictions vs.
those under linear regular wave. (2) The actual ampli- estimates from field measurements. Ocean Engineering,
tudes of dynamic response of seabed for random wave Vol.25, No.10, pp. 907916.
Pilotto, Micaela & Ronalds, Beverley F. 2003. Dynamic behav-
and linear regular theory are considerably different iour of minimum platforms under random seas. Proceeding
although the mean dynamic responses are approxi- of the International Conference on Offshore Mechanics and
mately identical. The scattered range of dynamic Arctic Engineering-OMAE, Vol.1, pp. 441449.
response covers all the numerical results gained from the Rahman, M.S. & Jaber, W.Y. 1986. A simplified drained
conventional deterministic method. (3) Stochastic char- analysis for wave-induced liquefaction in ocean floor
acteristics must be taken into consideration. Otherwise sands. Soils and Foundations, Vol.26, No.3, pp. 5768.
the numerical results calculated from the conventional Rahman, M.S. & Jaber, W.Y. 1991. Submarine landslides:
method will over- or under-estimate the risk level in the elements of analysis. Marine Geotechnology, Vol.10,
stability evaluation of seabed or marine construction. pp. 97124.
Zienkiewicz, O.C. & Chang, C.T. 1980. Drained, undrained,
consolidating and dynamic behaviour assumptions in
soils. Geotechnique, Vol.30, No.4, pp. 385395.
ACKNOWLEDGEMENTS

The authors wish to express their gratitude to Professor


Dahong Qiu of Dalian University of Technology, who

594

Copyright 2005 Taylor & Francis Group plc, London, UK


Pipelines

Copyright 2005 Taylor & Francis Group plc, London, UK


The performance of pipeline ploughs in layered soils

M.F. Bransby & G.J. Yun


The University of Dundee, Dundee, UK

D.R. Morrow & P. Brunning


Stolt Offshore, Aberdeen, UK

ABSTRACT: Plough performance has been investigated through a series of 1/50th-scale laboratory model
tests. Of particular interest has been how thin coarse grained soil layers modify the required tow force. The soil
conditions investigated have included dry sand, saturated fine medium dense sand, and uniform sand with thin
layers of well-graded sand. The relationship between plough tow force and distance is presented for the model
plough for different loading and soil conditions. It is shown when appropriate scaling laws are applied the tow
forces in medium dense uniform saturated sand are approximately in line with the values measured in the field
and that rate effects are consistent with previous research. However, the addition of a thin, horizontal layer of
well graded soil either at share tip depth or at mid-depth along the share both increase the required tow force, with
the largest increase for the case with the coarse grained deposit at mid-share depth. Further work is required to
determine the reason for the tow force difference and to formulate appropriate predictive models for layered soils.

1 INTRODUCTION curves for which plough performance predictions have


found to be reliable based on operational experience
For many sub-sea infrastructure projects around the and in house research. There are also difficulties when
world pipeline burial forms an integral part of the soil conditions are more complex then those which the
installation process for submarine pipelines providing empirical models are based such as where interbeded
improved stability, protection and thermal insulation. layers with markedly different geotechnical properties.
One method of pipeline burial is by forming a trench A program of laboratory tests was performed to
into which the pipeline is lowered. This is commonly model the pipeline plough-soil interaction for different
achieved by using a pipeline plough. The pipeline soil conditions. Ploughing tests were carried out for
plough is towed behind a support vessel that provides uniform soil conditions and soil conditions with thin
the required bollard pull as well as the equipment and layers of deposits. Comparison between results pre-
personnel for launch, recovery and operation of the sented here allowed determination of the effect of the
plough. additional soil layers.
When ploughing in granular soils partial drainage
and dilatancy effects result in increasing tow force with
increasing speed. The principle factors governing this
2 EXPERIMENTAL METHODS
speed tow force relationship are plough/trench geom-
etry and the geotechnical properties of the seabed soils.
2.1 Introduction
A knowledge of the speed/tow force relationships for
a particular set of seabed conditions is essential to esti- Reduced scale physical model tests were conducted
mate ploughing duration and hence realistic schedul- to measure the performance of a plough. A 1/50th-
ing of a pipeline burial project. However, this is not scale model plough was translated through dry and
straightforward because only empirical predictions of saturated medium-dense sand and saturated sand con-
pipeline performance exist (e.g. Cathie & Wintgens taining a thin layer of coarse-grained material.
2001). Furthermore there may be additional difficul- Preliminary tests were carried out in dry sand to opti-
ties when seabed soils fall outside the range of geo- mise the kinematics of the plough so that it behaved as
technical conditions on which the empirical ploughing on the seabed. Next, six tests were carried out to inves-
coefficients are based. For example Finch et al (2000) tigate the effect of a thin silty-sand layer on plough
give upper and lower bound particle size distribution performance. For all the tests, the target trench depth

597

Copyright 2005 Taylor & Francis Group plc, London, UK


selected was 36 mm, representing a 1.8 m depth at full 2.3 Material properties
scale. Slow fully-drained and faster partially drained
The sand selected for the uniform sand deposit was
tests were performed to investigate the effect of tow
fine Congleton sand. This was a poorly graded sand
rate for each soil condition.
with D50  0.22 mm and D10  0.16 mm. Direct
shear tests revealed that crit  31, but no tests were
performed to measure the peak friction angle of the
2.2 Test geometry soil at the appropriate relative density and effective
Model testing was carried out in the laboratory. The tests stress levels for the model tests reported here. Standard
consisted of towing a 1/50th-scale model (see Fig. 1) laboratory testing revealed that Gs  2.627 and
of an offshore plough through approximately 500 mm max  1.986 t/m3 and min  1.324 t/m3.
of soil representing the seabed. During testing the tow A well graded red gravelly sand was selected for
force and plough position were continuously measured. the additional soil layer. This has D50  0.44 mm and
The plough was constructed as a 1/50th-scale model. D10  0.16 mm. Soil element testing is needed to
All dimensions of the full-scale plough were reduced measure its mechanical properties, but provisional
by 50 for the model and the mass was reduced approxi- results from a series of large shear box tests revealed
mately 503 times. Consequently, the model plough that,   42 for loose sand at low effective stress
(shown in Fig. 1) was of length 344 mm (excluding levels (n  3 kPa).
the mould boards) and of mass 1.406 kg (submerged
weight, W  12.0 N). 2.4 Sample preparation
The soil was constrained within a box of dimen-
sions 450 mm wide by 1030 mm long and 295 mm A layer of dense gravel was first placed in the base of
deep. The box contained a drain for fixing the water the box to act as a base drain. This was covered with a
table level. A linear actuator and pulley arrangement geo-membrane to ensure that the soil sample above did
was used to pull the plough laterally. not mix and ensure that the drainage layer maintained
Figure 2 shows the set-up for a typical plough test. its high permeability. The base layer was considered to
The pulley height was selected so that there was a be incompressible, but allowed free drainage of water
slight upwards inclination of the tow rope representa-
tive of a tow line to a plough support vessel. The lin-
ear actuator comprised of a screw jack and was driven
by a computer controlled 3-phase motor. The actuator
was fixed on a mounting plate positioned above the
soil box. A load cell was placed on the lower end of
the actuator. The pulley wire was connected to the
lower size of the load cell so that any towing force on
the line was measured by the load cell. The towing
line led vertically downwards to a pulley in a fixed
position where the line was then attached to the plough.
The pulley was positioned just above the soil surface
so that the tow line angle was 510 above horizontal
representing operational conditions. A potentiometer
was positioned to measure the translation of the
plough during the tests.

Figure 1. The 1/50 th-scale model of the pipeline plough


(prior to testing; mould boards up). Figure 2. Testing apparatus.

598

Copyright 2005 Taylor & Francis Group plc, London, UK


into the soil sample above. Fine Congleton sand was and pulley system described above until the plough
then pluviated dry from a height of 750 mm at a fixed approached the end of the box. For fast tests, the motor
pouring rate to form a sand layer of thickness 100 mm. was replaced by manually controlled actuation. Load
The base drainage layer was then flooded and the was measured using a load cell and position using
sand was saturated slowly from the base by raising the a potentiometer. The velocities used for the tests are
water table incrementally. This was continued until given in Table 1.
there was free water over the whole sand surface and Plough tow forces will vary due to the velocity
took up to 10 hours to perform. The free water surface (e.g. Palmer 1999, Cathie & Wintgens 2001) and it is
was then carefully raised until it was 40 mm above the often assumed that the tow force varies linearly with
soil surface and the sample was left for at least 1 hour plough velocity unless cavitation occurs (Palmer 1999).
to ensure that the water conditions were fixed.
The dry sand pluviated as above was measured
to have a repeatable unit weight,   15.1 kN/m3. Table 1. Summary of the testing programme (N.B. all
This gave the soil a relative density of 42.2%.When loads and dimensions are given at model scale).
saturated this bulk weight was increased to  
19.2 kN/m3, giving an effective unit weight, Test conditions
  9.4 kN/m3.
For the tests with the thin coarse-grained deposit the Trench Tow
Vel Depth force
initial base layer of uniform fine sand was prepared as Test ID Model bed (mm/sec) (mm) (N)
above. The red gravely sand was then pluviated from
30 cm using a 2 mm sieve until the target layer thickness D0 Dry sand 4.2 23
(normally 5 mm) was achieved. The remaining sand on
D1 Dry sand 4.17 35 21
the 2 mm sieve was scattered on surface directly by
hand to leave a surface with a slightly coarser material. S3 Dense saturated sand 4.2 34 17
(uniform sand bed)
Following placement of the gravely sand layer, the
fine sand was pluviated above it as described previ- S1 Dense saturated sand 0.94 36 19.2
ously. The layered sample was saturated from the base S2 Dense saturated sand 35.69 27 25.5
in the same manner as for the uniform sand sample. ML1 Dense saturated sand. 4.17 33 26.1
Layer thickness 
2.5 Test arrangement 5 mm; 17 mm beneath
soil surface
The pipeline plough was placed on the far side of the (layer at mid share
soil box from the actuator and pulley, but on the centre depth)
line with the tow line pointing towards the pulley TL1 Medium dense 4.21 34 21.1
(Fig. 2). The share tip was partially buried manually saturated sand Layer
to the expected final depth to be achieved after transi- thickness  5 mm;
tion with the skids resting on the soil surface (see 2530 mm beneath
Fig. 1). This share tip depth was selected after several soil surface
calibration tests had been carried out for the final (layer at share tip
trench depth selected. However, the initial embed- depth)
ment depth was not found to be critical as the plough ML2 Medium dense 30.2 31 32.1
still went through a (reduced) transition phase to saturated sand. Layer
reach its steady-state share depth and tow force. thickness, 5 mm
Tow force and plough position were measured dur- at 15 mm
(layer at mid share
ing each ploughing test. A 10 kg capacity load cell
depth)
was connected to the towing line between the actuator
and the plough to measure the force on the towing TL2 Medium dense 56.4 27 34.2
saturated sand.
line. A 500 mm stroke potentiometer was positioned
710 mm thick layer
on the centre-line axis of the box connected to an at 2730 mm depth.
extension above the plough beam to measure the (layer at share tip
plough movement. Both instruments were logged on depth)
a computer during each test. TL3 Medium dense 4.47 31 19.1
saturated sand. 6 mm
2.6 Plough testing thick layer at
31 mm depth
2.6.1 Testing method
(layer at share tip
The plough was displaced laterally through the soil depth)
at a constant velocity (v) using the linear actuator

599

Copyright 2005 Taylor & Francis Group plc, London, UK


The degree of drainage occurring during a mono- tip depth. The trench depth and tow force are given at
tonic event will be due to the plough velocity, v, a model scale.
characteristic length of water flow, B and the coeffi-
cient of consolidation, cv. Palmer (1999) obtained a
dimensionless group (vD/cv) from a mathematical 3 RESULTS
study of flow processes around a steady-state translat-
ing simplified plough, where D is the trench depth. The following sections present the results from indi-
He therefore suggested that for model scale tests con- vidual plough tests in dry sand and saturated sand.
ducted in the same soil conditions as the full-scale Because of space limitations the results from the laye-
test (when cv is identical), vD should be the same in the red soil profiles will be collated to show the effect of
model as the full-scale plough i.e. vm  Dm  vp  Dp, the soil layering, the velocity and the trench depth in
where the subscripts m and p represent the scale a later section.
model and the full-scale prototype respectively. Thus
the model speed, vm  vp Dp/Dm  50 vp for the scale
3.1 Dry sand
factor of 50. This argument relies on the fact that the
characteristic length for drainage is proportional to The soil was uniform dry medium dense sand. The test
the trench depth. was carried out at a plough velocity of 4.17 mm/sec,
Using Palmers (1999) results suggests that to recre- but will have provoked fully drained conditions as the
ate the exact drainage conditions experienced in the sand was dry. The test will therefore provide the
field plough at a speed of 200 m/hr would require the purely frictional response of the plough, but for slightly
model plough to move at 50  200 m/hr  2.78 m/sec. higher effective stress conditions than for the satu-
This is impractical for the reported tests. It was there- rated sand because of the higher effective (buoyant)
fore decided to carry out tests at two different veloci- unit weight.
ties for each soil condition tested to ascertain both the The plough tow load-distance data as measured in
zero velocity frictional plough resistance and the the test is shown in Figure 3. After an initial load of
rate effect. Given that the rate effect for ploughs in 14 N is required to start significant plough movement,
sand has been observed to be linearly dependent on there is a transition period requiring about 100 mm of
velocity prior to cavitation (e.g. Grinstead 1985, movement before a maximum constant peak load of
Palmer 1999) the faster tests will allow identification approximately 21 N is mobilised. Minor fluctuations
of the gradient of the increased tow force divided by of load are measured during the following 360 mm of
velocity (the rate effect) when compared to the slower movement until the plough is stopped.
(drained) one. Cavitation would not have occurred in It is clear that a steady-state condition was reached
the tests performed because of the relatively low peak by the plough after a tow length of approximately
velocity and so the two test results would have been in 100150 mm. Examination of the plough kinematics
the linear range. revealed that a steady trench depth (D  35 mm) was
Following the plough tests, measurements were achieved and the kinematics were stable.
taken of the final soil height across the trench cross- In another test in dry sand, measurements of the
section at different distances from the initial plough soil surface position were taken manually perpen-
position. For some of the tests containing the coarse- dicular to the trench axis (e.g. Fig. 4) and along the line
grained deposit, the water table was lowered below of the trench (e.g. Fig. 5). The accuracy of the meas-
the base of the soil sample after the test so that the urements are estimated to be 
1.5 mm. Figure 4
sand was experiencing capillary suction. Excavations shows the soil profile before and after trenching after
could then be made in the sand to examine the trench 40 cm of plough movement. Although the original soil
cross-section behind the plough and the soil condi- surface was clearly not quite flat in this preliminary
tions around the main plough share.
25
2.7 Test programme 20
Load, N

The following sets of tests were carried out as shown 15

in Table 1. Many of the tests were carried out to 10


ensure that the plough kinematics were correct, to 5
verify the instrumentation, and to select the appropri-
0
ate soil conditions for modelling. In the first column 0 100 200 300 400 500
in Table 1, D represents a test carried out in dry sand, Displacement, mm
S represents uniform saturated sand, ML represents a
test with a coarse-grained layer at mid-share depth Figure 3. Tow load against position for the plough in uni-
and TL symbolises the coarse grained layer at share form dry sand (D  35 mm). Test D1.

600

Copyright 2005 Taylor & Francis Group plc, London, UK


test, the post-ploughing profile reveals a trench with For the faster test, there was considerably less data
soil on either side as expected. The slope angle of the collected because of the fixed data logging rate (Fig. 6).
trench was measured as 34. It can be seen that there is no transition; a force of
Figure 5 shows the profile along the base of the approximately 30 N is required to start the plough
trench before and after the ploughing operation. It can movement. After this initial peak, there is an approxi-
be seen clearly that the depth of the trench initially mately constant force of 25.5 N required to sustain
increases as the plough goes through a limited transi- plough movement at the proscribed rate (36 mm/sec).
tion phase. A trench depth of approximately 20 mm is However, there is a tendency for the tow load to
achieved after 100 mm of plough movement which reduce with further movement, which might reflect
remains almost constant with further movement. reaching steady-state water flow conditions after a
certain time or distance or an initial tow force increase
due to inertial effects. Excavation revealed that a steady
3.2 Saturated medium-dense sand
trench depth, D  27 mm was achieved. This is shal-
The soil was a saturated medium dense sand. One test lower than for the slow test.
was carried out at 0.94 mm/sec. This will have pro-
voked fully drained conditions due to its slow vel-
3.3 Summary of the results
ocity. The test will therefore provide the purely
frictional response of the plough in submerged soil Summarised results from the plough tests are given in
conditions. A second test was performed at a velocity tables 2 to 4. The results are discussed in more detail
of 36 mm/sec. This will have provoked partially in the following sections.
drained conditions thereby allowing some measure-
ment of the rate effect for submerged soil conditions
when compared to the slower test. 35
The plough tow load-distance data as measured in 30
both tests are shown in Figure 6. For the slower test, 25
Load, N

after an initial load of 12 N is required to start signifi- 20


cant plough movement, there is a transition period 15
requiring about 150 mm of movement before a max- 10 0.94 mm/sec
imum constant peak load of approximately 19.2 N is 5 35.7 mm/sec
mobilised. Fluctuations of load after reaching the 0
plateau stage were later attributed to a kink in the load 0 100 200 300 400 500
Displacement, mm
line going though the pulley. Excavation revealed that
a steady trench depth, D  36 mm was achieved.
Figure 6. Tow load against distance for the plough in uni-
form saturated sand.
-30
soil surface position, mm

-20
-10 Table 2. Tests in uniform saturated sand: effect of velocity.
0
10
Steady-state
20
After trenching Velocity, v Trench depth, tow force,
30
Before trenching Test (mm/sec) D (mm) F (N)
-200 -150 -100 -50 0 50 100 150 200 250
Position from trench centre, mm S1 0.94 36 19.2
S2 35.7 27 25.5
Figure 4. Example trench cross-section after transition. S3 4.2 34 17

plough movement, mm
0 200 400 600 800
0
Table 3. Tests in uniform saturated sand with layer at share
5
Drag direction
Original soil surface position
tip depth: effect of velocity.
section depth, mm

10
15 Steady-state
origin
20 Trench depth Velocity, v Trench depth, tow force,
25 after pulling Test (mm/sec) D (mm) F (N)
30
35 Position of final trench base TL1 4.21 34 21.1
40
TL2 56.4 27 34.2
TL3 4.47 31 19.1
Figure 5. Trench base profile.

601

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 4. Tests in uniform saturated sand with layer at mid- where W is the submerged unit weight of the plough;
share depth: effect of velocity.  is the effective unit weight of the soil; D is the
trench depth; and Cw and Cs are dimensionless tow
Steady-state
Velocity, v Trench depth, tow force, force coefficients.
Test (mm/sec) D (mm) F (N) For the soil conditions in this test,   0.954 t/m3,
W  153.4 t, D  1.8 m. Cathie and Wintgens (2001)
ML1 4.2 33 26.1 suggested that Cw  0.4 (for all soils) and Cs  10
ML2 30.2 31 32.1 (for medium dense sand). This gives a tow force of
117 tonnes or 1147 kN. For very dense sand, Cs  20
which gives a tow force, F  1693 kN. The equation
using either Cs coefficient under-predicts the towing
4 DISCUSSION force measured (2400 kN).
The under-prediction may be because the Coulomb
4.1 Comparison with full-scale behaviour envelope is not appropriate to describe the soil strength
To convert the small-scale plough test data to the full- in the scaling equations. The shear stress-normal
scale equivalents for comparison with monitored field stress failure envelope is not linear in reality because
plough performance it is necessary to consider the dilation is suppressed by increasing mean effective
scaling laws. stress (which requires the addition of an apparent
For the tests at 1/50-th scale, all physical lengths are cohesion, c, to the failure criterion). Consequently,
reduced by a factor of 50, areas (such as the projected there may be additional dilation (with correspond-
area of the share) by a factor of 502 and volumes by a ingly larger peak angles of friction) in the small-scale
factor 503. Because the tests were carried out in model tests where the effective stress is 50 times
Earths gravity field, the unit weight of the soil will be lower than in the full-scale prototype. This will lead
identical to that on the seabed. However, the effective to proportionately higher tow forces in the small-
stresses in the soil at any depth (i.e. at share tip depth) scale model tests, and will increase the values of Cs
will be reduced by a factor of 50 compared to that in and Cw. This hypothesis could be tested by carrying
the full-scale event (prototype). If it is assumed that a out model tests at other scales, by adding additional
Coulomb failure envelope (with c  0) describes the weight to the plough (to determine Cw directly) or by
soil failure envelope, this suggests a reduction of using centrifuge modeling. Steel-sand interface test-
shear strength of 50 times in the model tests (i.e. ing in the shearbox at the appropriate stress levels
f,model  f,p/50). In addition, the weight of the soil would also allow further examination of the coeffi-
above the share in the model test will be reduced by a cient Cw.
factor of 503 because of the volume ratio. Despite the under-prediction by the empirical
The tow forces measured will be a result of an area models, the overall tow force is of the approximate
(e.g. the projected area of the plough, or the surface magnitude experienced when ploughing offshore. This
area) multiplied by a shear stress (which may be suggests that approximately the correct soil deform-
dependent on the plough self-weight) and also a soil ation mechanisms are occurring in the model (notwith-
self-weight term. The area will be reduced by a factor standing potentially higher dilation angles) and so
502 as discussed above and the shear stresses also by changes due to different soil conditions and tow rates
a factor 50 (if it is assumed that a Coulomb failure will modify the tow force appropriately in the remain-
envelope describes the soil failure envelope). Thus der of the tests.
the model tow force due to friction/stress will equal
(1/50)2  (1/50)  1/503 times the field (prototype) 4.2 Rate effect
tow force. This also agrees with the scaling factor for
the self-weight terms (1/503). The tow force from the saturated uniform sand tests
In summary, model forces should be multiplied by (test S1 (slow) and S2 (faster) and an intermediate
503 to recreate the full-scale equivalent and distances slow speed (S3)) are summarised in Figure 7, where tow
and trench depths should be multiplied by 50. force (at full-scale) is plotted against plough velocity.
The results from test S1 (slow test in uniform sat- There is a clear rate effect, with the fastest velocity
urated sand) can be converted to full scale. The steady provoking the largest tow force. The linear regression
state tow force becomes 19.2 N  503  2,400 kN line on the Figure takes the form F  2200  7.533 v,
(245 t) for a trench depth, D  0.036  50  1.8 m. where the units are as shown in Figure 7.
Reese & Grinstead (1986) suggested that the fric- Some care is required to translate the model scale
tional tow force, velocity to a prototype velocity. Palmer (1999) rec-
ommends that vp  vm/50 and the resulting graph is
shown as Figure 8 with the bottom axis being the
(1) scaled plough velocity.

602

Copyright 2005 Taylor & Francis Group plc, London, UK


3500 40
F = 7.533v + 2201.7
3000 35
Tow force, kN

2500 30

Tow force, N
2000 25
Eqn. 2 (model velocity)
1500 Equation 2 20
1000 15 Uniform sand
10 Mid-share layer
500
5 Share tip layer
0
0 20 40 60 80 100 120 140 0
Velocity, m/hr 0 10 20 30 40 50 60
Velocity, mm/sec

Figure 7. Prototype tow force against model velocity in


uniform medium dense sand. Figure 9. Tow force against velocity for varying soil con-
ditions (model scale).

3500
3000 However, if the unscaled velocity is used in the
equation 2 (i.e. the actual rate of the model plough),
Tow force, kN

y = 376.65x + 2201.7
2500
2000 the dotted line on Figure 7 results. This captures the
1500 rate effect observed in the model tests more satisfac-
1000 Eqn. 2 (scaled velocity) torily. This suggests that (i) no scaling of model
500 velocities is required, and (ii) furthermore that the
0 characteristic drainage path length is not proportional
0 0.5 1 1.5 2 2.5 3
to the plough dimensions (for example, this might be
Velocity, m/hr
proportional to a local shear plane thickness).
Figure 8. Tow force against velocity in uniform medium
dense sand (velocity scaled using Palmer (1999)).
4.3 Effect of layered soil deposits
Figure 9 shows a graph of tow force against velocity
Cathie & Wintgens (2001) extended Reese & (in model scale) for the tests in uniform saturated
Grinsteads (1986) tow force relationship to add an sand and the tests with the coarse grained layer. There
extra term to allow for the velocity. This gave tow force, is a small increase in tow force compared to the uni-
form sand when there is a coarse-grained layer at share
(2) tip depth. There is a considerably larger increase
when the deposit is positioned at mid-share depth.
where v  plough velocity (m/hr); Cd  a dynamic Examination of the gradients of the regression lines
force coefficient with dimensions t/(m3/hr). They of tow force against plough velocity reveals that the rate
suggested that Cd would increase slowly with density, effect is similar for each soil condition. The above two
but quickly with reducing permeability. They calcu- effects suggest that the coarse-grained layer has an
lated Cd by back-calculation from field plough data influence only on the static frictional resistance of
(after the use of the first two static terms of eqn. 2) the plough, but not on the rate term.
and gave recommendations of variation of Cd with d10
particle size and soil density from their data set.
For the tests performed here, D10  0.16 mm and 5 DISCUSSION
the soil was medium dense. Because of the very low
effective stresses and the consequently higher dilation A disparity was observed between predictions of
rates in the model tests, it is believed that the dilation the model test tow forces from empirical equations
behaviour of the soil will be more like a very dense (Cathie & Wintgens 2001) using recommended coef-
sand at full-scale. Therefore, the value of Cd  ficients, and those measured in the model tests. It is
0.15 t/(m3/hr) for very dense soil is used in Equation believed that this is because of the sand will have
2 with Cw  0.4 and Cs  20 to calculate the higher peak angles of friction at the lower effective
expected tow force. Full-scale parameters are used stress levels in the model tests. This will lead to
(W  0.001227 503  153.4 tonnes, D  0.031  proportionately higher tow forces in the small-scale
50  1.55 m,   0.954 t/m3) together with the model tests, and mean that the empirical coefficients,
scaled velocity v  vm/50. This method gives the dot- Cs and Cw need to be increased to represent the
ted line shown in Figure 8. The equation underesti- results. Further investigation of this phenomenon
mates the tow force at zero velocity, and significantly could be examined by additional model tests at other
underestimates the rate effect. scales or by using centrifuge modeling. Alternatively,

603

Copyright 2005 Taylor & Francis Group plc, London, UK


Cw could be determined more directly by adding
additional weight to the plough.
Despite the above, it is believed that the general
performance of the small-scale plough in uniform
sand is similar to that of the full-scale plough and so
variations in performance due to changing soil prop-
erties should be reflected in the model tests.
The model test results suggest that the thin soil
layer significantly increases the tow force required to
translate the plough. However, there seems to be little
difference in the rate effects for the ploughs in uni-
form soil or with the layering.
One reading of the test results would suggest that
the layer (or possibly the developing layer-sand mix)
Figure 10. Photograph of excavated soil cross-section per-
has significantly greater frictional strength than the pendicular to plough axis after test TL3.
uniform fine sand (for example an enhanced angle of
friction which might be accompanied by a larger dila-
tion angle). This would simply increase both coeffi-
cients Cw and Cs in equation 1 for the static plough
resistance and leave the dynamic coefficient, Cd in eq. 2
unaffected. This hypothesis could be tested with a
range of element tests on the layer deposit and the
sand-layer mix and further plough tests with the aim
of finding the link between soil properties and the
static coefficients Cw and Cs.
It is possible that the thin soil layer with its poten-
tially higher frictional properties and increased dila-
tion, might have significantly changed the soil
deformation mechanism around the plough. Further
study of the soil deformations could be made by
examination of the soil profiles around the plough
after excavation. For the final plough tests, post-test Figure 11. Photograph of excavated soil cross-section par-
excavation allowed photography of the soil around allel to plough axis after test TL3.
the plough after the test. Figure 10 shows the cross-
section perpendicular to plough axis after excavation
of soil after test TL3 and it initially appears that the 6 CONCLUSIONS
thin layer is below the trench base level. However,
there is evidence of some entrainment of the layer 1/50th-scale plough model tests were conducted to
material by the plough immediately below the share investigate plough performance. Of particular interest
tip because of the indistinct layer in this zone. was how a coarse grained soil layer within an other-
Figure 11 shows the excavated soil in elevation along wise uniform fine sand would affect the tow force.
the line of the plough. From this figure it is clear that the The model plough was pre-embeded in the soil and
darker soil layer affected the soil around the plough. then displaced beyond its transition length at a fixed
There is contact between the layer and the plough share velocity. During each test the movement of the plough
and a volume of the well-graded soil was entrained and the tow force were measured. Tests were repeated
into the soil zone above the share blade. This explains for different soil conditions and at different rates to
why even the layer at the share tip depth may increase quantify the plough performance for uniform sand
the ploughing force. Additional tests could be con- deposits, and sand deposits containing thin layers of
ducted with uniform sand containing dyed sand layers granular deposit.
to see if the soil movement through the plough for The results showed that the plough behaves approxi-
uniform conditions varies to that observed here. mately as expected in uniform saturated sands: the
In summary, considerably more work is required to plough maintained a stable tow force after a reduced
understand fully plough performance. There are no transition zone due to pre-embedment; trench cross-
empirical models to predict the effect of heterogenous sections appeared credible; the scaled tow forces appear
soil deposits on plough tow forces, and the empirical approximately in line with field measurements, but
models to predict plough performance in uniform somewhat higher than those predicted using existing
sands did not model the test results particularly well. empirical models; increased trench depth causes

604

Copyright 2005 Taylor & Francis Group plc, London, UK


increasing tow force; and larger velocities produced Offshore Technology Conference OTC 13145, Houston,
larger tow forces. May 2001.
When there was a thin soil layer present, the tow Finch, M., Fisher, R., Palmer, A.C., Baumgard, A. 2000. An
force increased, particularly when the horizontal layer integrated approach to pipeline burial in the 21st century.
Deep Offshore Technology 2000.
was at mid-share depth. There seems to be consistent Grinstead, T.W., 1985. Earthmoving in submerged sands.
differences in towing force at different rates between PhD Dissertation, University of Newcastle-upon-Tyne.
the mid-share depth deposit, the share tip deposit and Palmer, A.C., 1999. Speed effects in cutting and ploughing.
the uniform deposit, so that the rate effect seemed Geotechnique 49, No. 3, pp. 285294.
unaffected by the soil layer deposit. Reese, A.R. and Grinstead, T.W. 1986. Soil mechanics of
submarine ploughs, OTC 5341 (May 1986)

REFERENCES

Cathie, D.N., Wintgens, J-F. 2001. Pipeline trenching using


plows: performance and geotechnical hazards. Proc.

605

Copyright 2005 Taylor & Francis Group plc, London, UK


Physical and numerical modelling of lateral buckling of a
pipeline in very soft clay

J.R.M.S. Oliveira, M.S.S. Almeida & M.C.F. Almeida


COPPE Federal University of Rio de Janeiro, Brazil

R.G. Borges, C.S. Amaral & A.M. Costa


CENPES PETROBRAS, Rio de Janeiro, Brazil

ABSTRACT: This work discusses soil-structure interaction applied to thermal snaking of shallowly buried
pipelines embedded in very soft clay. The main motivation of this research was the accident that occurred in
January 2000 in Rio de Janeiro where more than 1 million liters of crude oil has been spilled into the Guanabara
Bay. In that way, a set of comprehensive centrifuge tests has been undertaken in order to assess the lateral resist-
ance of the soil. The results were compared with numerical simulations of the same centrifuge scenarios using
a software developed by PETROBRAS.

1 INTRODUCTION

In Rio de Janeiro, Brazil, a 17 km long pipeline had


been in use since early nineties to transport heated
(95C) oil from ships to the main local refinery.
In January 2000 a large-scale environmental accident
occurred when this pipeline failed and 1 million liters
of oil spilled into Guanabara Bay. Detailed studies
confirmed the integrity of the steel used and no
evidences of corrosion. In situ piezocone and vane
tests were carried out, as well as laboratory studies
included triaxial and oedometer tests and soil char-
acterization, showing low values of Su increasing
with depth.
It was concluded that the cause of the failure was
local buckling of the free pipeline supported at the
offshore extremity by the silted soil layer and at the Figure 1. View of the deformed pipeline at failure scenario.
onshore extremity by the remaining soft clay cover.
The lateral buckling occurred as a result of the tem-
perature increase causing a large displacement of
2 PHYSICAL MODELLING
4.1 m laterally along a length of 44 m (Fig. 1). Further
information on that subject is described in Almeida
2.1 Soil properties and strength
et al. (2001).
This accident motivates a new research interest on Site investigations and laboratory tests have been car-
pipe-soil interaction. Thus, a set of centrifuge tests ried out to provide geotechnical parameters for the
has been planned using the local Guanabara Bay soft analyses. Laboratory tests included soil index tests,
clay in order to assess the lateral resistance of the soil UU and CIU triaxial tests as well as oedometer tests.
when loaded laterally in plane strain condition. In situ tests included piezocone and vane tests.
These curves were developed to check a well known Index tests and some geotechnical parameters are
algorithm developed and widely used by PETROBRAS summarized in Table 1. The soil data obtained is typical
on pipeline design. of the region around Guanabara Bay (Futai et al. 2001).

607

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 1. Summary of some soft clay properties (Almeida
et al. 2001).

Soft clay properties Data

Plastic limit (wp)  50%


Liquid limit (wl) 140180%
Plasticity Index (Ip ) 90120%
Water content (w) 150200%
Voids ratio (e) 3,64,5
CR  Cc/(1  e0) 0.36
OCR  1.3

Su (kPa)
0 2 4 6 8 10
0

2
Borehole 3
3
Depth (m)

4
Figure 3. COPPE geotechnical mini-drum centrifuge.
5
Vane - Deltageo
constructed by G-Max Scotland in 1995 with 90 g-ton
6 Vane - COPPE full load capacity. It comprises 20 slip rings, 16 data
Laboratory Vane acquisition channels and 2 independent actuators: a
7 Triaxial UU linear one and a turntable where the first one is
mounted on. The linear actuator is a step motor drive
Su = 0,126 + 1,373 z and servo-controlled system and shows high accu-
8
racy. On the other hand, the turntable consists of a DC
Figure 2. Undrained strength profiles for a representative motor and a tachometer system controlled through a
borehole (Almeida et al. 2001). mini maestro driver. This kind of set is mainly suit-
able for fast movements where high precision is not
required. This limitation allowed only monotonic
It consists of lightly overconsolidated and highly com- tests. However, some improvement in velocity control
pressible soft clay. has been done changing the analogical reference volt-
The clay undrained strength Su has been measured age for a digital one, via a digital-analogue card.
by a number of procedures including electric vane The centrifuge is also capable of a 90 movement
borer tests (two tipes of equipment), triaxial UU tests changing the rotation axis from the vertical to the hor-
and laboratory vane tests on samples not suitable for izontal position (Fig. 3).
UU tests. A representative profile of Su is shown in For the application related to this work it was
Figure 2. Data of vane tests have been corrected for decided not to use the whole centrifuge channel once
each depth with the corresponding Ip value. Data of that procedure would require a large amount of natu-
in situ and lab vane and also of UU tests show a sim- ral soil. Instead, a 260 mm length, 210 mm wide and
ilar range of values and therefore a regression line has 178 mm high strongbox was adopted.
been drawn through the whole data set.
2.3 Centrifuge tests setup
2.2 COPPE geotechnical centrifuge
A 1:30 scaled pipeline with 15.2 mm diameter and
The COPPE geotechnical centrifuge (Gurung et al. 3 different possible length sizes (92.5 mm, 112.5 mm
1998) is a 1.0 m diameter mini-drum designed and and 125 mm) adapted to a rigid shaft was used as a

608

Copyright 2005 Taylor & Francis Group plc, London, UK


Figure 4. The 1:30 pipe model attached to the instrumented
shaft.

plane strain model as shown in Figure 4. The choice for


dragging the pipe sideways using a rigid shaft instead of
wires was based on a criterion to assess soil reaction
independently at both directions through an imposed
horizontal movement. Zhang et al. (2001) used a sim-
ilar approach to measure soil reaction in sand.
The horizontal force was measured using four strain
gages glued on the shaft surface. To enhance moment
sensibility and therefore horizontal force sensibility, a Figure 5. Centrifuge tests setup.
cut reduced the shaft thickness creating a flat area
where the strain gages were mounted on. The vertical
force was measured using an Entran 50 N traction The slurry was placed inside the channel whilst spin-
and compression load cell, compensated for bending ning through a specially designed rotating joint. This
moments and thermal variations. One Druck Pore- process produced a smooth and regular surface ade-
pressure transducer was positioned inside the soil quate for the shallow tests.
layer for consolidation monitoring (Fig. 5). Each test was divided in two phases: consolidation
It was decided to use the natural Guanabara Bay at 100 g followed by actuation at 30 g. All samples
clay, taking into account two different alternatives to reached around 90% consolidation in 10 hours flight.
place the clay inside the channel. The first one, using Enough time was dispensed for pore-pressure dissi-
clay lumps, needed the equipment to be steady while pation during the slowing down process. After that,
the layer was slowly being built. This process was not the pipe was driven into the soil until it reached the
adequate for the swipe tests because the soil surface, pre-established target position.
even though highly consolidated, was quite irregular, The planned set of tests included burial depths
featuring small hills and valleys which directly inter- ratio H/D from 17% to 124%, where H is the distance
fered with the measurements. Thus, an alternative between the soil surface and the bottom of the pipe,
solution was adopted increasing the water content, and D is the pipe diameter. The velocity used for these
mixing the sample and making slurry from the clay. tests was 0.12 D/s which represents a normalized

609

Copyright 2005 Taylor & Francis Group plc, London, UK


velocity v  D/cv  580. House et al. (2001) present a
reference value of 100, and therefore values greater
than that, as indicative of undrained behavior.

2.4 T-Bar penetration tests


As the pipe is penetrating the soil whilst being placed
in position, the whole set (Figure 4) can be easily con-
verted into a T-Bar penetrometer (Stewart & Randolph
1991) and the load cell measurements used to esti-
mate the undrained shear strength.
Nevertheless, some considerations must be done
before using these data. Considerations about the Figure 6. Horizontal projection of the pipe bottom contact
T-Bar factor Nb are presented in Stewart & Randolph area with soil.
(1991), with a 10.5 mean recommended value at
intermediate roughness. However, this value is appli-
cable only when the bar is completely submerged in into the soil until it reaches fully plastic behavior. No
soil. Shallow depths require different T-Bar factors softening or hardening effects was considered
for each depth. Also the contact area increases as the throughout the analysis. Three different approaches
pipe is being pushed into the soil, magnifying the were used to simulate the soil-pipe interface: the first
amount of material involved in the failure process. considering a maximum mobilized friction of 100%
When the bar is just touching the surface of the layer, Su, and the second considering 50% and 20%.
the cylindrical shell can easily be assumed as a flat plate The soft clay soil was simulated using Eu  300 Su
foundation. This solution, after Terzaghi, is 5.14 for a (Almeida & Marques 2002),   0.5 and a unitary
pure cohesive material. However, it is still important strength profile of Su  1 kPa. These conditions intent
to evaluate the depth at which the T-Bar factor reaches to isolate the parameter of interest from the other ones.
its full value and how this variation develops. To assess the T-Bar factor, the following equation,
proposed by Stewart & Randolph (1991), was used.

3 NUMERICAL MODELLING (1)


3.1 The AEEPECD/SIGMA program
For the simulation of the non-linear physical behavior The D* value in equation (1) has to be adapted
of the soil a continuous medium is considered, which once the contact area between pipe and soil varies
demands the application of interactive incremental with depth, and has a straight relationship with
integration algorithms. strength. To take this variation into account, the fol-
The software SIGMA (2004) version 4.22, was lowing relation was proposed to express the horizon-
used for the bi-dimensional finite element discretiza- tal projection of the pipe bottom half contact area
tion and model post-processing. with soil (Fig. 6)
The 2.1 version of the numerical simulator AEEP-
ECD developed by Costa (1984) was used to solve the (2)
equilibrium of continuum differential equations. This
software allows linear and non-linear analysis of where D is the pipe diameter and H is the distance
plane strain, plane stress and axi-symmetric models. between the soil surface and the bottom of the pipe.
The program uses constitutive laws for plastic and Figure 7 presents the T-Bar factor variation for three
elastic-plastic rheologies. The adopted plastic criteria adhesion factors:   0.2,   0.5 and   1.0. The
are Von Mises and Mohr-Coulomb incorporating geo- initial value is around 5.24, for   0.5, which is
metric and physical non-linearities. close to the expected Terzaghis 5.14. In the same way,
the final value for the same adhesion factor is around
10.5 which is exactly the same value proposed by
3.2 Modelling the penetration phase Stewart & Randolph (1991). The present analysis also
In order to quantify the variation of the T-Bar factor, indicates that the T-Bar factor can be considered sta-
from the first contact with the soil surface to full fail- bilized after H/D ratio of about 300%. The numerical
ure mechanism development, 20 numerical simula- variation interval for a fully buried penetrometer with
tions have been carried out with depths varying from respect to roughness is between 10.0 (  0,2) and
17% to 500%. In each case, the T-Bar has been pushed 11.3 (  1,0).

610

Copyright 2005 Taylor & Francis Group plc, London, UK


14 12

12
10
10
8
8

P (N)
Nb

6
6

4 4
 = 0,20
2  = 0,50 2
 = 1,00
0
0% 100% 200% 300% 400% 500% 600% 0
0 100 200 300 400 500 600 700 800 900
H/D (%)
Time (s)

Figure 7. T-bar factor variation during early penetration


Figure 8. Typical force decay curve after bar penetration.
process.

The Nb variation presented in Figure 7 was used 10

Normalized Horizontal Force Ph/(Su x D)


to calculate Su by means of equation (1), with force 9
data obtained from the centrifuge tests (pipe placing 8
phase). 7

6 H/D=124%

5 H/D=81%
3.3 Modelling lateral drag phase 4

3 H/D=57%
Numerical simulation of centrifuge tests adopted the
2
same burial depths used for the physical tests. The H/D=27%
1 H/D=17%
strength prototype profile assumed was obtained
from centrifuge penetration tests with T-Bar factor 0
0 0,5 1 1,5 2 2,5 3 3,5 4 5,5 5
corrections, and is presented in Equation (3). Normalized displacement (x/D)

Figure 9. Horizontal force centrifuge tests results for vari-


(3) ous H/D ratios.

4 PHYSICAL AND NUMERICAL RESULTS


5
Normalized Vertical Force Pv/(Su x D)

Between the penetration phase and the lateral drag 4


phase, the pipe was left steady in position for 15 min, H/D=17%
whilst vertical force was stabilizing enough time not 3 H/D=27%

to interfere with the test measurements. Figure 8


2 H/D=57%
shows a typical force decay curve. This behaviour is
mainly related to relaxation and excess pore-pressure 1 H/D=81%
dissipation at the clay layer.
H/D=124%
Figures 9 and 10 show the centrifuge tests results 0
using normalized forces and normalized displacement.
-1
Force normalization was obtained through the same 0 0,5 1 1,5 2 2,5 3 3,5 4 4,5 5
Equation (1), and displacement normalization was Normalized Displacement (x/D)

obtained dividing the values by the pipe diameter (D).


All forces were set to zero before the lateral drag Figure 10. Vertical force centrifuge tests results for various
H/D ratios.
phase begins. Figure 9 and 10 present horizontal and
vertical force results for five different burial depths,
respectively.
Figure 9 shows an expected behavior of horizontal Actually, the horizontal force is related to the amount
force, increasing as the burial depth also increases. of failure surface mobilization which is more signifi-
On the other hand, Figure 10 shows a decreasing ver- cant as the pipe goes further into the soil. However, the
tical force as the burial depth increases. vertical force presents an opposite behavior, being more

611

Copyright 2005 Taylor & Francis Group plc, London, UK


10 subjected to lateral pipe displacements. Five different
Normalized Horizontal Force Ph/(Su x D)

9 H/D ratios have been adopted covering only shallow


8
H/D=600%
depths behaviour.
7
H/D=500%
H/D=400%
The pipe-shaft set has been used as a T-bar pen-
6 H/D=300% etrometer during vertical actuation. This procedure
H/D=200%
5 H/D=124% should allow undrained shear strength assessment
H/D=100%
4 H/D=81% whilst positioning the pipe. Nevertheless, some con-
3 H/D=57% siderations has to be done about T-bar factor variation
2
H/D=27%
for very shallow depths. Numerical analysis has been
1 H/D=17% done to identify the T-bar factor variation from an
0
H/D=5%
unburied situation up to a completely buried one.
0,00 0,10 0,20 0,30 0,40 0,50 0,60 0,70
Limiting values of shallow and deep foundation
Normalized Displacement (x/D)
agreed well with consolidated analytical solution.
Figure 11. Numerical analysis results for normalized hori- Numerical analysis has also been done to simulate
zontal forces. the lateral centrifuge tests. The results show good
agreement with physical data for a shallowly buried
situation but might slightly underestimate soil
1,8
response for fully buried conditions. This indicates
that other phenomena not considered in numerical
Normalized Vertical Force Pv/(Su x D)

1,6 H/D=200% H/D=27%


H/D=124% analysis, such as suction behind the pipe, might be
1,4 H/D=100% related to this behaviour.
H/D=57%
1,2
H/D=81%
H/D=5%
1,0

0,8
ACKNOWLEDGEMENTS
0,6

0,4 The authors would like to gratefully acknowledge


0,2 FINEP and PETROBRAS for sponsoring this research
0,0 throughout CTPETRO program and also acknowl-
0,00 0,10 0,20 0,30 0,40 0,50 0,60 0,70
Normalized Displacement (x/D)
edge CENPES for the various professionals involved
in the studies reported herein, and also for allowing
Figure 12. Numerical analysis results for normalized ver- the publication of the present data.
tical forces.

important when the pipe is not completely buried. REFERENCES


Hence for H/D ratios, i.e. greater than 100%, the ver-
tical force is negligible. Almeida, M.S.S., Costa, A.M., Amaral, C.S., Benjamin, A.C.,
Figure 11 presents the results of the numerical analy- Noronha, Jr. D.B., Futai, M.M. & Mello, J.R. 2001.
sis for normalized horizontal forces, also obtained from Pipeline failure on very soft clay. 3rd International
AEEPECD/SIGMA program. This analysis assumes Conference on Soft Soil Engineering: 131138. Hong
intermediate roughness value of   0.5. Additional Kong: Balkema.
Almeida, M.S.S. & Marques, M.E.S. 2002. The behaviour of
considerations about  and the theoretical influence on
Sarapu soft organic clay. International Workshop on
T-Bar factor can be obtained from Randolph (2004). Characterisation and Engineering Properties of Natural
Values for H/D 50% seems to agree well with Soils 1: 477504. Singapore: Balkema.
measured data. Nevertheless, values for H/D  50% Costa, A.M. 1984. An application of computational methods
appears to be slightly underestimated in numerical and rock mechanics principles in excavation analysis of
analysis. underground mines (in Portuguese). D.Sc. Thesis COPPE
Figure 12 shows the numerical results for the nor- Federal University of Rio de Janeiro.
malized vertical forces. Most values seem to be close Futai, M.M., Almeida, M.S.S. & Lacerda, W.A. 2001.
to a narrow band between 1.0 and 1.6 which does not Geotechnical Properties of Rio de Janeiro Clays (in
Portuguese). Brazilian Meeting of Soft Clays Properties
agree well with physical data.
(in Portuguese): 138165. Rio de Janeiro.
Gurung, B., Almeida, M.S.S. & Bicalho, K.V. 1998.
Migration of zinc through sedimentary soil models. Proc.
5 CONCLUSIONS of The International Conference Centrifuge 98 1:
589594. Tokyo.
A set of centrifuge tests has been undertaken to assess House, A.R., Oliveira, J.R.M.S. & Randolph, M.F. 2001.
the lateral resistance of Guanabara Bay soft clay when Evaluating the coefficient of consolidation using

612

Copyright 2005 Taylor & Francis Group plc, London, UK


penetration tests. International Journal of Physical Stewart, D.P. & Randolph, M.F. 1991. A new site investiga-
Modelling in Geotechnics 1(3): 1726. tion tool for the centrifuge. International Conference on
Randolph, M.F. 2004. Characterization of soft sediments Centrifuge Modelling: 531538. Boulder.
for offshore applications. Geotechnical and Geophysical Zhang, J., Stewart, D.P. & Randolph, M.F. 2001. Centrifuge
Site Characterization: 209232. Rotterdam: Millpress. modelling of drained behavior for pipelines shallowly
SIGMA 2004. Software for bidimensional finite element embedded in calcareous sand. International Journal of
pre and post processing. PUC Rio de Janeiro and Physical Modelling in Geotechnics 1(1): 2539.
PETROBRAS.

613

Copyright 2005 Taylor & Francis Group plc, London, UK


Bearing capacity and large penetration of a cylindrical object
at shallow embedment

E.R. Barbosa-Cruz & M.F. Randolph


Centre for Offshore Foundation Systems, The University of Western Australia, Perth, Australia

ABSTRACT: The penetration of a cylindrical object into soft clay starting from a very small embedment
has been investigated using a large deformation finite element approach, the Remeshing and Interpolation
Technique with Small Strains (RITSS). The study has application to pipeline penetration of the seabed, and also
to the interpretation of cylindrical penetrometers such as the T-bar at very shallow depths. The results of the
analyses show the evolution of bearing capacity factor and soil flow mechanisms as the cylindrical object is
penetrated from 0.5% of the diameter up to nearly 5 diameters. Comparisons of results are made for fully
smooth and fully rough interfaces and for homogeneous and non-homogeneous soil profiles.

1 INTRODUCTION penetration resistance will reduce to the classical


result of 5.14Bsu where B is the contact width between
The penetration resistance of an infinite cylindrical pipeline and soil.
object at shallow embedment into soft soils is of A numerical study has been undertaken using large
importance to offshore pipeline design and the inter- deformation finite element analysis in order to follow
pretation of T-bar penetration tests close to the sur- the evolution of the bearing resistance and deformation
face. Previous studies of pipeline penetration have mechanism during pipe penetration from the surface
made use of bound solutions (Murff et al. 1989, Aubeny to full burial. The analyses have covered fully smooth
et al. 2005) but these do not account of factors such to fully rough interface conditions and both uniform
as: local heave; variation of shear strength with depth; soil and soil with a significant strength gradient.
and disturbance of the strength profile due to pipe
penetration. For T-bar tests (Stewart & Randolph 1991,
1994), a constant bar factor is generally adopted for
estimating the shear strength from the bearing resist- 2 NUMERICAL ANALYSIS
ance, regardless of depth. However, at very shallow
depths this factor should be reduced. 2.1 Details of the analysis
At deep penetration, the limiting effective force P Deep penetration of objects into the soil from the sur-
per unit length of cylinder may be related to the shear face implies very large strains and deformations within
strength su of the soil by a bearing factor Nb: the soil, and thus cannot be simulated using conven-
tional small strain finite element analysis. Hu &
(1) Randolph (1998a) proposed a process that they termed
Remeshing and Interpolation Technique with Small
Strain or RITSS, which falls in the category of
where D is the diameter of the cylinder. Randolph & Arbitrary Lagrangian-Eulerian (ALE) finite element
Houlsby (1984) presented a plasticity solution for this techniques, to analyse this kind of problem. This tech-
factor, with a lower bound increasing from 9.14 for nique combines a series of small-strain analysis fol-
fully smooth to 11.92 for fully rough interface condi- lowed by a complete remeshing of the domain and then
tions. Subsequent upper bound solutions give a very interpolation of the stress field and material param-
similar range, from 9.21 to 11.92 (Randolph 2004). eters between the Gauss points of the new and old
For interpretation of T-bar tests, a bearing factor of meshes. Details of the technique and strategies for the
10.5 has been suggested. size of displacement increments and frequency of
Near the surface, however, the bearing factor will remeshing may be found in Hu & Randolph (1998a,
reduce and in the limit at very shallow penetration the 1998b) and Lu et al. (2000, 2001, 2004).

615

Copyright 2005 Taylor & Francis Group plc, London, UK


The RITSS technique can accommodate remeshing and strength profiles (homogeneous and non-
the domain totally or partially. In the present study the homogeneous) were modelled.
domain has been remeshed partially, wherever the The soil was represented as a simple elastic, per-
nodal displacements exceeded 10% of the maximum fectly plastic material with Tresca failure criterion.
displacement; in the remainder of the domain the Poissons ratio was taken as 0.49, and the modulus
position of the nodal coordinates were updated ratio, E/su, was taken as 500. A shear strength profile
according to the displacements. Additionally a min- of su  5  1.5 z/D kPa was adopted for the non-
imum zone for remeshing was adopted, bounded homogeneous soil analyses, where z/D is the depth
approximately by a circular line centred on the cylin- normalised by the cylinder diameter. For the homoge-
der centre and with a radius equal to 5 times the cylin- neous soil the shear strength was taken as su  9 kPa.
der radius. Partial remeshing is useful to minimise Initial geostatic stresses were calculated based on an
any error in the interpolated stress field close to the effective unit weight for the soil of D  6 kPa (typ-
domain boundaries, where the mesh is coarse. ical for submerged soft soil with a pipe diameter of
Taking advantage of the vertical axis of symmetry about 1 m) and K0 of unity. The quantities su and 
only half of the domain was modelled, using six- have been non-dimensionalised with respect to D in
noded triangular elements with three Gauss points; order to allow application of the results for D values
nodal joint elements were used to model the soil- close to 1 m, with appropriate adjustment to .
cylinder interface. Four different combinations of The cylinder was modelled as a cylindrical bound-
cylinder roughness (fully smooth and fully rough), ary that was advanced into the soil. The size of the
finite element mesh was chosen to ensure that the
boundaries were far from the plastic zone. The hori-
zontal and vertical extends of the domain were taken
as 15 times the cylinder diameter.
The large penetration analyses were initiated from
a point where the cylinder was very slightly embed-
ded (by 0.5% of the diameter). Figure 1 shows an
example of an initial mesh used in the analysis. As a
check, analyses were performed from an initial pre-
embedment of half a diameter, to confirm that the
response resistance curves from the two different pre-
embedments would merge at deep penetration.
The details of the analysis presented here are sum-
marised in Table 1. In the case of the slightly pre-
embedded fully smooth cylinder penetrating into
non-homogeneous soil, two different analyses were
performed to evaluate the effect of the minimum
mesh size on the results. The combined effect of the
minimum mesh size and remeshing interval was
Figure 1. Initial finite element mesh (distance scale in assessed for the same cylinder and soil conditions,
cylinder diameters). but for pre-embedment by half a diameter.

Table 1. Details of finite element analyses.

Initial Minimum Displacement Remeshing Cylinder Soil strength


Analysis embedment (m) mesh size (m) increment (m) interval (m) roughness profile*

c1 0.005 0.05 0.00005 0.001 Smooth Non homogeneous


c2 0.005 0.01 0.00005 0.001 Smooth Non homogeneous
c3 0.005 0.01 0.00005 0.001 Rough Non homogeneous
c4 0.005 0.01 0.00005 0.001 Smooth Homogeneous
c5 0.005 0.01 0.00005 0.001 Rough Homogeneous
c6 0.5 0.01 0.00005 0.001 Smooth Non homogeneous
c7 0.5 0.05 0.00005 0.005 Smooth Non homogeneous
c8 0.5 0.05 0.00005 0.005 Rough Non homogeneous
c9 0.5 0.05 0.00005 0.005 Smooth Homogeneous
c10 0.5 0.05 0.00005 0.005 Rough Homogeneous

* Strength profile: Non homogeneous, su  5  1.5 z/D kPa; Homogeneous, su  9 kPa; (E  500su; v  0.49 for both).

616

Copyright 2005 Taylor & Francis Group plc, London, UK


It should be noted that this type of analysis is For penetration greater than D/2 (  /2), B
computationally intensive. Using a computer with a becomes equal to the diameter, D. Note that since
2.81 GHz Intel Pentium Processor and 2.0 GB of some heave will occur adjacent to the cylinder, the
RAM the analysis times were between 1 to 30 days, actual contact width, B, will initially exceed the the-
mainly depending on the density of the mesh and the oretical contact width. This is shown in Figure 4, with
roughness of the cylinder. For the 10 analyses pre- results from the finite element analyses c1 and c2
sented here a total computing time of 111 days were compared with the theoretical relationship.
required. It was found necessary to use a very fine Since the cylinder is modelled as a hollow cylin-
mesh and frequent remeshing, particularly during the drical zone penetrating soil with a non-zero effective
initial penetration of the cylinder, where the cylinder- unit weight, a buoyancy force will act on the cylinder.
soil contact region expanded rapidly. This buoyancy force is calculated from the (theoret-
The need for the finer mesh and smaller remeshing ical) volume of cylinder below the soil surface, ignor-
interval at shallow embedment is shown in Figure 2, ing local heave, multiplied by the effective unit
which compares analyses c1 and c2 up to an embed- weight of the soil. This volume may be related to the
ment of 0.2D. The two analyses converge for dis- angle, , by
placements greater than 0.1D, but the coarser mesh,
c1, shows significant jumps at certain remeshing
(4)
stages at shallow embedment.
Once the embedment exceeds the cylinder diameter,
2.2 Processing of results the buoyancy force per unit length of cylinder stays at
The computed loads are reported as a bearing capa- R2.
city factor, Nc, calculated in broad terms as the cylin- The average contact pressure resisted by the soil is
der load (per unit length) divided by the contact width then calculated as the force per unit length of cylinder
(maximum of 1 cylinder diameter) and then by a char- (from the finite element results), less the buoyancy
acteristic soil strength. Details of this processing of force, divided by either the theoretical contact width, B,
the data are given below.
As the cylinder penetrates the soil, the theoretical
contact width is given by (Figure 3)
cylinder
(2)
R heave

where R is the radius of the cylinder (D/2), and  is
seabed
the semi-angle subtended at the cylinder centre by the
theoretical contact width, which is related to the nom- nominal
inal embedment depth, z, by B embedment,
B
(3)
Figure 3. Geometry of cylinder penetrating seabed soil.

1.2
20 Theoretical
18 Minimum element size 0.01 m
1 c1
16 c1
c2
Contact widths

14 c2 0.8
B/D or B'/D
Load kN/m

12
10
Minimum element size 0.05 m 0.6
8
Analysis Minimum Step Remeshing
6 0.4 mesh size size every
4 mm mm mm
Smooth cylinder
2 Soil properties su = 5 + 1.5 z/D kPa, E = 500 su 0.2 c1 50 0.05 1.0
0 c2 10 0.05 1.0
0 0.05 0.1 0.15 0.2 0
Embedment, z/D 0 0.1 0.2 0.3 0.4 0.5
Embedment, z/D
Figure 2. Comparison of load-displacement responses
from analyses c1 and c2. Figure 4. Variation of contact width with embedment.

617

Copyright 2005 Taylor & Francis Group plc, London, UK


7 Table 2. Bearing capacity factors.
Nc max = 5.89 c1
6 Minimum element size 0.05 m
c2 Max. Final
5 Case Initial Nc Minimum Nc Nc Nc
4
Nc Nc max = 5.13 Nc computed P/Bsuo P/Bsuo P/Bsuo P/Bsuo P/Dsuo P/Dsuo
3 Minimum element size 0.01 m contact width
c1 5.48 5.89 4.80 4.23 9.10 8.96
2 c2 5.52 5.13 4.81 4.18 9.01 8.97
Smooth cylinder c3 6.12 5.43 6.01 5.24 10.83 10.72
1 Soil properties su = 5 + 1.5 z/D kPa, E = 500 su c4 5.15 5.02 4.50 4.03 9.26 9.25
0 c5 5.54 5.25 5.42 4.96 11.99 11.97
0 0.05 0.1 0.15 0.2 c6 5.17 9.04 8.97
Normalised embedment, z/D c7 5.14 9.48 9.40
c8 6.63 11.49 11.46
Figure 5. Detailed variation of bearing capacity factor, Nc. c9 4.49 9.28 9.27
c10 5.86 12.01 12.01

12
Minimum element size 0.05 m
10 re-meshing every 0.005 m Nc = 9.40
Nc = 8.97 10
8 c2, Nc = 8.97
9 Nc theoretical
Minimum element size 0.01 m contact width, c2
Nc 6 8
re-meshing every 0.001 m c6, Nc = 8.97
c6 7
4 c7 6
Nc 5 Nc computed
2 Smooth cylinder contact width, c2
Soil properties su = 5 + 1.5 z/D kPa, E = 500 su 4
0 3
0 1 2 3 4 5
2
Normalised embedment, z/D Smooth cylinder
1 Soil properties su = 5 + 1.5 z/D kPa, E = 500 su
Figure 6. Overall variation of bearing capacity factor, Nc. 0
0 1 2 3 4 5
Normalised embedment, z/D
or the actual contact width, B. In order to calculate a
bearing capacity factor, Nc, the contact pressure is Figure 7. Bearing capacity factor, Nc for smooth cylinder.
then divided by a characteristic soil strength, suo,
taken as that at the widest (theoretical) contact width
of the cylinder with the soil. Thus suo is taken as the for initial embedments of 0.005D and 0.5D for a fully
surface soil strength, sum, for embedment depths less smooth cylinder in non-homogeneous soil (analyses
than D/2, and as c2 and c6). For the former case, results are presented
in terms of the theoretical contact width, B, and the
(5) computed contact width, B, allowing for soil heave.
Details of the corresponding Nc variation for shallow
where k is the strength gradient. embedment are shown in Figure 8.
The bearing capacity factor Nc for a smooth cylin- As commented above, the value of Nc increases
der may be expected to increase from around 2  from just over 5 at very shallow embedment to just
initially to 9.2 (average of upper and lower bound solu- over 9 once it becomes fully embedded. The com-
tions) once the cylinder is fully embedded. Figures 5 puted asymptotic Nc factor at large embedment is
and 6 show the need for finer mesh and smaller remesh- about 2% below the theoretical lower bound of 9.2,
ing intervals for shallow and deep pre-embedment. which is excellent accuracy considering the complex-
Using the finer mesh size of 1% of diameter and the ity of the analysis. The curves in Figure 7 show a
smaller remeshing interval equal to 0.1% of diameter slight jump at embedment greater than 2.5D. This
from the analysis, Nc varies from 5.13 at shallow jump is associated with closure of the soil around the
depth to 8.97 after deep penetration. top of the cylinder; a small region of water appears to
be trapped immediately above the cylinder before clo-
sure, but the remeshing strategy ignores this and
3 BEARING CAPACITY FACTOR replaces this zone with soil (see later, Figure 11). The
close agreement of the occurrence of this closure,
The bearing capacity factors, Nc, are summarised in with that for an initial embedment of 0.005D occur-
Table 2. The overall variation of Nc is shown in Figure 7 ring just ahead of that for an initial embedment of

618

Copyright 2005 Taylor & Francis Group plc, London, UK


8 14
Nc theoretical contact width c2,Nc = 8.97
7 Nc max = 5.52 12 c3,Nc = 10.72
6 10 c4,Nc = 9.25
Nc min = 4.81
c5,Nc = 11.97
5 8

Nc
Nc 4 6
Nc max = 5.13 Nc min = 4.18 c2, Non homogeneous soil, smooth cylinder
3 Nc computed contact width 4 c3, Non homogeneous soil, rough cylinder
2 c4, Homogeneous soil, smooth cylinder
2
c5, Homogeneous soil, rough cylinder
1 Smooth cylinder 0
Soil properties su = 5 + 1.5 z/D kPa, E = 500 su 0 1 2 3 4 5
0
Normalised embedment, z/D
0 0.1 0.2 0.3 0.4 0.5
Normalised embedment, z/D
Figure 9. Summary of overall variations in Nc.
Figure 8. Detailed variation of Nc at shallow embedment
for smooth cylinder.

8
su = 5 + 1.5 z/D kPa
0.5D (with additional soil heave present in the former
analysis) is further confirmation of the robustness of 6
Nominal Nc

the analyses.
At shallow embedment (Figure 8) the initial peak in
4
Nc based on the computed contact width is very close c2
(within 0.2%) to the theoretical value of 5.14. The su = 9 kPa c3
bearing capacity factor then reduces slightly reaching 2 c4
a local minimum of 4.18 (based on computed contact E = 500 su c5
width) or 4.81 (theoretical contact width) for embed-
ments of 0.18D to 0.24D. This is due to the smooth 0
cylinder-soil interface condition adopted, so that the 0 0.1 0.2 0.3 0.4 0.5
cylinder surface represents a weakness introduced into Normalised embedment, z/D
the soil, reducing the bearing capacity below the the-
oretical value for a flat strip. Figure 10. Nominal bearing capacity factor, P/Dsuo.
The reverse would occur for a fully rough cylinder
placed into non-homogeneous soil (analysis c3),
although in this case soft soil from the surface level
will be dragged down more with the cylinder, so there It is also useful to see the variation of the nominal
is a combined effect of the cylinder geometry and bearing factor, P/Dsuo (where P is the net cylinder
distortion of the soil stratigraphy (Figure 9). In this force, allowing for buoyancy effects), which is plotted
case the Nc based on computed contact width, B, is in Figure 10. This factor increases monotonically,
5.43 at shallow penetration then reduces slightly to reaching Nc values between 5.00 and 6.64 depending
5.24 and reaches an asymptotic value of 10.72 at deep on the interface roughness and the soil profile once
penetration. the pipe embedment is 0.5D.
For the homogeneous strength profile (analyses c4 Examples of the patterns of incremental displace-
and c5) the bearing capacity factor Nc of the cylinder ment vectors for the fully smooth cylinder and non
calculated from the computed contact width varies homogeneous soil (analysis c2) are shown in Figure 11
from 5.02 (fully smooth) to 5.25 (fully rough) at shal- for cylinder embedments of 0.014D, 0.205D, 1.505D,
low embedment (Figure 9). Then the Nc decreases to 2.524D (before closure), 2.525D (after closure), and
4.03 (fully smooth) and 4.96 (fully rough) and reaches 2.555D. The incremental displacements show the
limiting values at deep displacement between 9.25 classical rotational velocity field around the fully
(fully smooth) and 11.97 (fully rough). At shallow embedded cylinder, extending out to about 2.2 cylin-
embedment the values are very close to the theoret- der radii from the centre of the cylinder. The plastic
ical lower value 5.14 and for deep penetration the cal- zones (not shown here) were somewhat larger than
culated Nc values are in agreement with the plastic anticipated from upper bound solutions, but this did
solution of 9.14 for a fully smooth cylinder and 11.94 not seem to affect the overall load displacement
for a fully rough cylinder. response of the flow patterns.

619

Copyright 2005 Taylor & Francis Group plc, London, UK


z/D = 0.014 z/D = 0.205

z/D = 2.525 (just after enforced closure)

z/D = 1.505

z/D = 2.555

z/D = 2.524 (just before closure)

Figure 11. Incremental displacement vectors at increasing penetration of cylinder (analysis c2: smooth cylinder in non-
homogenous soil).

4 CONCLUSION of the bearing factor for a cylindrical object penetrat-


ing from very small initial embedment. The analyses
This paper has summarised the results of large dis- involved a 1 m diameter cylinder displaced into soft
placement finite element analysis of a cylindrical soil, from an initial embedment of 0.005D, to a final
object penetrating soft clay. The work used the RITSS embedment from 3 to 5D.
algorithm developed by Hu & Randolph (1998a). The It was found that the bearing capacity factor (ex-
aim of the study was to evaluate the gradual evolution pressed as force normalised by contact width times

620

Copyright 2005 Taylor & Francis Group plc, London, UK


soil strength at maximum cylinder contact width) Hu, Y. & Randolph, M.F. 1998a. A practical numerical
depends on the combination of the interface rough- approach for large deformation problem in soil. Int. J.
ness and soil strength profile. The calculated Nc val- Num. and Anal. Methods in Geomech., 22(5): 327350.
ues are close to those predicted from plasticity theory. Hu, Y. & Randolph, M.F. 1998b. Deep penetration of shal-
low foundations on non-homogeneous soil. Soils and
Analyses run from an initial embedment of 0.5D Foundations, 38(1): 241246.
showed essentially identical resistance to that for the Lu, Q., Hu, Y. & Randolph, M.F. 2000. FE analysis for T-bar
initial embedment of 0.005D, with the curves merg- and ball penetration in cohesive soil. Proc. 10th Int.
ing by a penetration of 1.5D. Offshore and Polar Engineering Conf. ISOPE 2000,
The analyses showed that the mechanism gradually Seattle, USA, 2: 617623.
evolved from a classical Prandtl field at shallow Lu, Q., Hu, Y. & Randolph, M.F. 2001. Deep penetration in
embedments to a rotational mechanism similar to soft clay with strength increasing with depth. Proc. 11th
confined flow of soil around a laterally loaded pile. Int. Offshore and Polar Engineering Conf. ISOPE 2001,
Closure of the soil over the top of the cylinder Stavanger, Norway, 2: 453458.
Lu, Q., Randolph, M.F., Hu, Y. & Bugarski, I.C. 2004.
occurred (for the particular soil properties modelled) A numerical study of cone penetration in clay.
at an embedment from 2.4D to 4.2D. Geotechnique 54(4): 257267.
Murff, J.D., Wagner, D.A. & Randolph, M.F. 1989. Pipe
penetration in cohesive soil. Geotechnique 39(2):
ACKNOWLEDGMENTS 213229.
Randolph, M.F. 2004. Characterisation of soft sediments for
This work was carried out while the first author was offshore applications, Keynote Lecture, Proc. 2nd Int.
an International Postgraduate Research Scholar in Conf. on Site Characterisation, Porto, 1: 209231.
Australia. Part of the work was undertaken in collab- Randolph, M.F. & Houlsby, G.T. 1984. The limiting pressure
on a circular pile loaded laterally in cohesive soil.
oration between the Norwegian Geotechnical Institute Geotechnique 34(4): 613623.
(NGI) and the Centre for Offshore Foundation Systems Stewart, D.P. & Randolph, M.F. 1991. A new site investiga-
(COFS), University of Western Australia. tion tool for the centrifuge. Proc. Centrifuge 91, Boulder,
Balkema, 531538.
Stewart, D.P. & Randolph, M.F. 1994. T-bar penetration test-
REFERENCES ing in soft clay. J. Geotechnical Engineering, ASCE,
120(12): 22302235.
Aubeny, C.P., Shi, H. & Murff, J.D. 2005. Collapse loads for
a cylinder embedded in trench in cohesive soil. Int. J. of
Geomechanics, ASCE (in press).

621

Copyright 2005 Taylor & Francis Group plc, London, UK


Stability design of untrenched pipelines geotechnical aspects

J. Zhang & C.T. Erbrich


Advanced Geomechanics, Nedlands, Australia

ABSTRACT: The mechanics of pipeline-soil interaction are discussed from a fundamental geotechnical per-
spective. Theoretically based models for assessing the ultimate lateral soil resistance for both drained and
undrained soil conditions are presented. The additional lateral resistance due to pipe embedment is also evalu-
ated. Examples are given to demonstrate that there is no unique friction factor which is of general applicabil-
ity. Model predictions are compared with experimental data to show model accuracy.

1 INTRODUCTION The basic mechanics of pipeline-soil interaction is


first investigated. Theoretical models for the assess-
Offshore pipelines must be geotechnically stable ment of the ultimate lateral resistance due to bearing
under the environmental conditions that will be failure are then presented for both drained and
imposed during their design life, such as wave and undrained soil behaviour, which are typical for
current loading and thermal effects. This is normally pipelines on sand and clay respectively. Additional
achieved by following the design approaches outlined soil resistance due to pipe embedment is evaluated.
in standard industry guidelines such as those pre- The models are demonstrated through examples and
sented by American Gas Association (AGA) or Det comparison with published experimental data.
Norske Veritas (DNV). Most of these guidelines are The effect of cyclic loading, scouring and sedi-
based on empirical approaches, which lack a sound ment deposition on pipeline self burial are not
theoretical basis, but are calibrated to test results from covered by this paper although this is also an impor-
a limited range of soils. While these empirical methods tant factor in the assessment of the long term lateral
may be conservative in certain cases, this is not guar- stability for an untrenched pipeline.
anteed. Industry is increasingly moving to more
sophisticated but inherently less conservative design
approaches such as full 3D dynamic pipeline model-
2 MECHANICS OF PIPELINE-SOIL
ling, where the pipelines are permitted to move sig-
INTERACTION
nificant distances under the design events. Under
such conditions the limitations of these empirical
An untrenched pipeline is a unique seabed structure
approaches means that unconservative designs can
supported by the surficial sediment. The mechanics
arise and hence obtaining a proper understanding of
of the interaction between this special structure and
the basic mechanics of pipeline-soil interaction
the soil is a fundamental aspect in the evaluation of
becomes increasingly important.
the geotechnical stability of a pipeline.
Most pipeline design engineers focus on the struc-
tural analysis of the pipeline, and request values for sim-
ple parameters such as the pipeline-soil friction factor
2.1 Pipeline and strip footing
and pipeline embedment. However, as will be discussed
in this paper, the simplicity of the required parameters Because the length of a pipeline is much greater than
belies the complexity of the problem at hand and the its diameter, a pipeline resting on a flat seabed may be
non-uniqueness of these parameters. Nevertheless, treated as a plain strain problem, similar to a strip
through consideration of basic soil mechanics some footing. However, a pipeline has several unique prop-
quite general behavioural principles can be estab- erties compared to a strip footing, as shown in Table 1.
lished leading to design methods that can provide the One of the most important differences that needs to
most appropriate values for the simple pipeline be fully appreciated in order to understand what fol-
design parameters, as demonstrated in this paper. lows, is that a pipeline will inherently induce a soil

623

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 1. Comparison of pipeline and skirtless strip footing.

Skirtless strip footing Pipeline


V M V
H
H
2B 2B

Footing width large, constant small, variable


Weight high low
FoS on laydown %1 1
Lateral movement not allowed %2B
Embedment minimal significant
Moment load significant negligible

Note: FoS  Factor of safety.

Figure 2. Yield surface for drained bearing failure.


V
H
l3
Seabed resistance in accordance with the least energy criteria.
Passive Failure This will be illustrated in Section 3.3.
Sliding Failure The ratio between the total soil resistance, which is
l1 2l2 in equilibrium with the lateral load FH, and the net
vertical load (difference of submerged weight Ws and
Bearing Failure uplift force FL) V  Ws
FL, FH/(Ws
FL), is
defined as the friction factor ; this is the parameter
Figure 1. Failure planes for a pipeline under combined required by most pipeline structural designers.
loading.

bearing failure during laydown due to its basic geom- 3 LATERAL RESISTANCE DUE TO
etry; only in the strongest soils will this not occur. BEARING FAILURE

The combined bearing and sliding response of a


2.2 Failure mechanism
pipeline is best described using yield surfaces similar
The geotechnical failure of a pipeline may occur in to the interaction diagrams used for shallow founda-
different ways depending on the soil and loading con- tion design. Depending on the soil and loading condi-
ditions. For example, both bearing failure and horizon- tions encountered, drained or undrained conditions
tal sliding failure, accompanied by a passive wedge may apply.
failure, could take place, as illustrated in Figure 1.
These failures can occur under fully drained, partially
3.1 Yield surface under combined loading
drained or undrained conditions, depending on the
soil type and the loading rate. Drained failure occurs 3.1.1 Drained conditions
if the soil is very permeable and the loading is rela- Under drained conditions, the yield surface proposed
tively slow, while undrained failure occurs if the soil by Zhang et al. (2001) based on model testing is appro-
has low permeability and/or the loading is relatively priate to describe the ultimate capacity of a pipeline.
fast. It is also possible that while deep bearing failure As illustrated in Figure 2, this yield surface has a para-
may exhibit an undrained mode, the sliding failure bolic shape and has an intercept at zero vertical load
and the passive failure may exhibit drained behaviour with the horizontal load axis, indicating a certain lat-
due to the much shorter seepage paths that apply for eral capacity even when the vertical load acting on the
these conditions (l2 and l3 compared to l1). pipe drops to zero. This intercept is due to the passive
Different failure planes, either sliding failure  pas- failure of the soil wedge on the advancing side of the
sive failure or bearing failure  passive failure, and pipeline, which is treated separately in this paper.
different failure modes, drained or undrained, lead to Hence all yield surfaces pass through the origin.
different soil resistances. However, the failure would Another important finding of the work of Zhang
always take place on the plane that yields the least et al. is that the pipe penetrates further into the soil

624

Copyright 2005 Taylor & Francis Group plc, London, UK


under lateral loading if the vertical load acting on the
pipe is more than about 10% of the ultimate vertical
capacity of the pipeline, as illustrated by the orienta-
tion of the displacement vectors in Figure 2. The yield
surface therefore expands until an ultimate state is
obtained where the pipeline does not embed any fur-
ther under the imposed loading. Pure sliding therefore
occurs when the pipe is pushed laterally in combin-
ation with a vertical load equal to about 10% of the
maximum vertical capacity. Conversely, if for some
reason the vertical load is actually lower than this
characteristic value then the pipe will move up
through the soil when subject to lateral load, until
once again the applied vertical load is about 10% of
the maximum vertical capacity. Experimental evi-
dence suggests that a unique ultimate state exists for a
pipeline under the same vertical load regardless of its
preloading history. The direction of pipeline move-
ment can be described using plasticity theory through Figure 3. Yield surface for undrained bearing failure.
the concept of a plastic potential; a surface similar to
the yield surface but which is always normal to the
direction of the displacement vectors for any given related to pipeline behavior. For example, Bransby
combination of load. For drained loading a non-asso- and Randolph (1998) demonstrated normality for
ciated flow rule is required, which means that the strip footings on undrained clay through numerical
shape of the yield surface and plastic potential are dif- analysis. This is essentially supported by the experi-
ferent, as illustrated in Figure 2. mental results obtained by Martin and Houlsby
(2000) for spudcan foundations on undrained clay
(Strictly speaking Martin and Houlsby (2000) found a
3.2 Undrained conditions non-associated flow rule. However, the specific form
For undrained conditions, we are not aware of any of their flow rule leads to the same conclusions with
specific pipeline test data presented in terms of a respect to pipeline friction as would be obtained using
yield surface approach. However, extensive research an associated flow rule).
on the combined loading of shallow foundations and
spudcans (e.g. Bransby & Randolph 1998, Ukritchon 3.3 Effect of loading path
et al. 1998, Martin & Houlsby 2000, Gourvenec &
A pipeline is usually required to be hydrotested prior to
Randolph, 2003) demonstrate the general applicabil-
operation. The submerged weight of the pipeline dur-
ity of this approach. Based on this research, a yield
ing hydrotest (Ws-flooded) is normally higher than that
surface as shown in Figure 3 is suggested for the
during operation (Ws-operation). In addition it appears
undrained response of pipelines.
to be not uncommon for the maximum vertical load
From a perusal of Bransby & Randolph (1998) and
imposed on the soil during pipelay to be significantly
Houlsby & Martin (2001), it is evident that the shape
greater than the pipe weight alone. Therefore almost
of yield function is not unique although there appear
all pipelines experience a reduction in vertical load at
to be only a relatively small range of possibilities. We
the start of operation. During a design storm, the waves
estimate that where the pipeline has sufficient rough-
apply a drag force, an inertial force and an uplift force
ness to mobilize the full undrained shear strength
to the pipeline. The uplift force leads to a further net
across the pipeline-soil interface, the range of min-
vertical load reduction while the drag force and the
imum true friction factor for other possible yield
inertial force lead to lateral loads that are generally of
functions varies between 0.32 and 0.42, and hence the
similar amplitude to the uplift force. Because the
mean of these two values (i.e. 0.37) is adopted in this
degree of vertical unloading between hydrotest and
paper.
operation varies for different pipelines, a variety of
The plastic potential defining the movement vector
loading paths may result (remember that all start from
is also shown on Figure 3. For undrained loading an
a state of pure vertical bearing failure under the lay-
associated flow rule has been assumed, which means
down load):
that the vector of movement is always perpendicular
to the yield surface. This is a critical assumption but For a typical drained bearing failure as shown in
we believe that sufficient evidence exists in the litera- Figure 4, all loading paths lead to the same ultimate
ture to support it, although none of this is explicitly state which is defined by the drained sliding cut-off

625

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 2. Ultimate lateral resistance for five cases.

Loading path Resistance (kN/m) Friction factor

1 0.82 0.7
2 0.49 0.7
3 1.12 0.37
4 0.97 0.48
5 0.90 0.7

sliding cut-off before it reaches the undrained yield


surface thus leading to a drained sliding failure.

3.4 Ultimate lateral resistance


The ultimate lateral resistance and corresponding
friction factor calculated for the five loading cases
discussed above are summarized in Table 2. For this
Figure 4. Loading path effects on drained loading. assessment, a drained friction factor of 0.7 was
adopted for the pipeline-soil interface, which is typ-
ical of a concrete coated pipeline on sand.
It can be seen from Table 2 that for drained bearing
and sliding failures, a constant friction factor of 0.7
was achieved for loading paths 1, 2 and 5, although
the lateral resistance varies from 0.49 to 0.9 kN/m.
For the undrained cases, different friction factors
were obtained for the different load paths. Although
the exact shape of the yield envelope for a pipeline on
clay is not known at this stage, we believe that a value
of about 0.37 is appropriate if the reduction in vertical
load after pipe laydown is small, but with higher val-
ues obtained when the reduction in vertical load is
increased. However, even this minimum undrained
friction factor is much greater than often recom-
mended (e.g. 0.2) for pipelines on clay.
In some cases the soil beneath a pipeline will con-
solidate under the pipeline weight and will increase in
strength before lateral loading is applied. Under these
conditions even higher undrained friction factors are
initially possible than computed here, although the
Figure 5. Loading path effects on undrained loading. maximum value is unlikely to exceed the drained
value and the response is also likely to be quite brittle,
since any lateral movement of the pipe will move it
away from the zone of consolidated soil.
line, although loading path 1 results in yield surface
expansion (further burial of the pipeline) while load-
ing path 2 does not. 4 ADDITIONAL LATERAL RESISTANCE
For a typical undrained bearing failure as shown in DUE TO EMBEDMENT
Figure 5, different loading paths lead to different
results. Loading path 3 leads to yield surface In addition to the ultimate lateral resistance due to
expansion until the ultimate state is achieved and a bearing failure or sliding failure, the soil on the
minimum friction factor of 0.37 is obtained. advancing side of the pipeline may provide a passive
Loading path 4 intersects the yield surface at a resistance component, FP. This resistance is depend-
lower vertical load and hence there is no yield sur- ent on the embedment of the pipeline only and there-
face expansion, but the friction factor is deter- fore could increase the apparent friction factor
mined from the intersection between the load path significantly when the net vertical load on the pipeline
and the initial undrained yield surface. On the reduces to a small value, corresponding to the peak
other hand, loading path 5 intersects the drained uplift force.

626

Copyright 2005 Taylor & Francis Group plc, London, UK


Figure 6. Lower and upper bounds passive resistance.

4.1 Lower bound value


A lower bound assessment of the lateral resistance
due to passive failure may be evaluated directly from
the classical Rankin formulae. The results for a Figure 7. Comparison of model prediction and test data.
pipeline embedded in sand with an effective unit
weight of 8 kN/m3 and an internal friction angle of Rankin formula is believed to reflect the effect of
45 are plotted in Figure 6. mounding in front of the pipeline. For the model test
interpretation, an effective unit weight of 8 kN/m3
4.2 Upper bound value was adopted and an internal friction angle of 45 was
used to assess the passive earth pressure coefficient, KP.
Experimental results presented in Zhang (2001) were
utilised for assessing an appropriate upper bound for
FP. These experiments were conducted in a centrifuge 5 COMPARISON WITH PUBLISHED DATA
at the University of Western Australia using a 20 mm
diameter model pipeline and an acceleration of 50 g A comparison between model predictions and published
(g is the ground acceleration). At this acceleration, experimental data is shown in Figure 7. A drained
the model pipeline is considered to represent a proto- friction factor of 0.7 was assumed for the pipe-sand
type pipeline of 1.0 m diameter. The soil used was interface in the model calculations. Soil parameters and
sampled from the North West Shelf of Australia and pipeline submerged weights for all cases considered are
was classified as calcareous silty sand. summarised in Table 3. It should be noted that the soil
The experimental data showed that when the pipe resistances obtained from the model predictions are for
was pushed laterally under a constant vertical load, the ultimate state and the lower bound value passive
the pipe embedment increased from its initial value. resistance component was assumed. Nevertheless, for
The maximum embedment achieved at large lateral a wide range of soil conditions pipe diameters and
displacement was, on average, three times the initial pipe submerged weights, the calculated results agree
embedment under the same vertical load. reasonably well with the test data.
The interpreted FP values from the model test
results are plotted in Figure 6 versus the pipeline
embedment (relative to the original mudline). The 6 CONCLUSIONS
model pipe was pushed laterally about 0.1 m during
the swipe tests and from 1 m to 3 m for the probe tests. For untrenched pipelines, undrained or drained bear-
An upper bound trendline fit to the data is also pre- ing and sliding failure may occur depending on the
sented on this figure, along with a modified version of soil and loading conditions. For drained bearing or
the standard passive pressure equation. The enhanced sliding failure, a constant pure friction factor deter-
capacity relative to that obtained from the classical mined purely from the pipe/soil interface friction

627

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 3. Soil parameters and pipe submerged weight.

OTC paper Soil condition Pipeline weight


reference
Type1 Strength (kN/m3) Ws-flooded (kN/m) Ws-operation (kN/m) Load2

OTC3477 Clay 10 kPa 7 0.050.24 0.050.24 M


OTC5504 SFS 32 9 0.72.8 0.51.6 M
0.52.8 0.31.8 C
LMCS 32 8.5 0.83 0.51.6 M
0.83 0.51.5 C
DMCS 36 9.5 0.83 0.61.9 M
0.83 0.51.8 C
OTC5504 SFC 1 kPa 6 12.8 0.92.1 M
12.8 0.61.9 C
STC 70 kPa 7.5 1.56 1.25.1 M
1.53 1.12.8 C
OTC5853 DMS 37 10.5 0.11.5 0.060.9 M

Note 1: SFS  silty fine sand; LMCS  Loose medium/coarse sand; DMCS  Dense medium/coarse sand;
DMS  Dense medium sand; SFC  Soft clay. STC  Stiff clay.
Note 2: M  Monotonic load; C  Cyclic load

angle may be assumed for design calculations. combined loading of strip and circular foundations on
However, for undrained bearing failure, the pure fric- clay. Geotechnique 53 (6): 575586.
tion factor is dependent on the specific load path Martin, C. and Houlsby, G. 2000. Combined loading of
applied but is bounded between a minimum value of spudcan foundations on clay: Laboratory Tests.
Geotechnique, 50 (4): 325338.
about 0.37 and the drained friction factor. Martin, C. and Houlsby, G. 2001. Combined loading of
In addition to the bearing/sliding failure pure fric- spudcan foundations on clay: numerical modeling.
tion factor, passive resistance may also be mobilised Geotechnique, 51 (8): 687700.
on the advancing side of the pipeline. This force is Palmer, A. C., Steenfelt, J. S., Steensen-Bach, J. O. and
dependent on the pipeline embedment and therefore Jacobsen, V. 1988. Lateral resistance of marine pipelines
may lead to a significant increase in the apparent on sand. OTC 5853, Houston.
total friction factor when the net vertical load is low. Ukritchon, B., Whittle, A. J. and Sloan, S. W. 1998.
Equations for evaluating lower and upper bound esti- Undrained limit analyses for combined loading of strip
mates of the passive soil resistance have been proposed. footings on clay. ASCE J. of Geotechnical and
Geoenvironmental Engineering, 124 (3): 265276.
Predictions made with the proposed drained and Wagner, D. A., Murff, J. D., Brennodden, H. and
undrained pipeline-soil interaction models are shown Sveggen, O. 1987. Pipe-soil interaction model. OTC5504,
to agree reasonably well with experimental data Houston.
obtained for a wide range of pipe and soil conditions. Wantland, G. M., ONeil, M. W., Reese, L. C. and Kalajian,
E. H. 1979. Lateral stability of pipelines in clay. OTC
3477, Houston.
Zhang, J., Stewart, D. P. and Randolph, M. F. 2001.
REFERENCES Modelling of shallowly embedded offshore pipelines
in calcareous sand. ASCE J. of Geotechnical and
Bransby, F. M. and Randolph, M. F. 1998. Combined loading Geoenvironmental Engineering, 128 (5): 363371.
of skirted foundations, Geotechnique, 48 (5): 637655. Zhang, J. 2001. Geotechnical stability of untrenched off-
Gourvenec, S. and Randolph, M. F. 2003. Effect of strength shore pipelines in calcareous sand. PhD Thesis, The
non-homogeneity on the shape of failure envelopes for University of Western Australia.

628

Copyright 2005 Taylor & Francis Group plc, London, UK


Pipeline-seabed interaction analysis subjected to horizontal cyclic loading

T. Takatani
Department of Civil Engineering, Maizuru Natl College of Technology, Maizuru, Kyoto, Japan

ABSTRACT: An advanced finite element analysis of liquefaction process for pipeline-seabed interaction
problem is carried out in order to simulate pore pressure build up response in seabed around a pipeline sub-
jected to horizontal cyclic loading under a constant vertical load. In two-dimensional dynamic non-linear finite
element method based on an effective stress theory, the cyclic mobility model is adapted to simulate excess pore
water pressure. The displacement behaviour of pipeline, the effective stress path and the shear stressstrain rela-
tionship in seabed are investigated through some numerical examples by this liquefaction analysis. Pore pres-
sure accumulation and pipeline settlement are greatly affected by not only the constant vertical load during
horizontal cyclic loading and its cyclic frequency.

1 INTRODUCTION the viewpoint of the vertical and horizontal loading


ratio and joint element stiffness. The joint element is
It is very important in pipeline design to make an evalu- used in this numerical analysis to simulate a slip phe-
ation of the pore pressure build up in the soil around a nomenon between pipeline and seabed. The displace-
pipeline on the seabed due to direct wave action and ment behaviour of pipeline, the effective stress path and
induced currents. The pore pressure accumulation will the shear stressstrain relationship in seabed around a
reduce the effective strength of the soil and degrade pipeline during horizontal cyclic loading under a con-
the bearing response of a pipeline. The reduction of stant vertical load are obtained by the liquefaction
bearing capacity can lead to large vertical and horizon- analysis. In addition, the effect of the frequency of hor-
tal displacements of a pipeline. In particular, an increase izontal cyclic loading on the pore pressure build up
of horizontal displacement of the pipeline may lead to response is numerically investigated in this paper.
sudden break-out, which has a serious influence on the
safe operation of the pipeline. It is therefore very impor-
tant for design engineers to evaluate the pore pressure 2 PIPELINE-SEABED INTERACTION
accumulation in the soil around a pipeline under hori- ANALYSIS
zontal cyclic loading conditions. This can be achieved
by carrying out numerical analyses (Taiebat & Carter An advanced numerical analysis used in this paper is
2000, Zhang et al. 2001a, Takatani & Ogawa 2004) and two-dimensional dynamic non-linear finite element
simulations (Zhang et al. 2002, Takatani & Randolph method based on an effective stress theory to simulate
2003) based on experimental data of pore pressure liquefaction process of saturated sandy soil under
build up. The geotechnical centrifuge tests were undrained condition. In this finite element analysis, a
conducted for the partially drained cyclic loading of non-linear relationship between shear stress and shear
pipelines on uncemented calcareous soil (Zhang & strain of soil element is accurately expressed by a
Randolph 2001b). In centrifuge tests the pore pressure multi shear spring model (Towhata & Ishihara 1985)
accumulation and the pipeline movement behaviour and the Masing rule for loading and unloading curves
were measured for 12 loading conditions. is employed so as to adjust the amplitude of hyster-
In this paper, an advanced finite element analysis of esis damping for the multi shear spring model. Also
liquefaction process for the pipeline-seabed interaction the cyclic mobility model (Iai et al. 1990), which is of
problem is carried out as a first step in order to simu- a generalized plasticity-multiple mechanism type, is
late pore pressure build up response and pipeline move- adapted to simulate excess pore water pressure. Pore
ment. Not only the pore pressure build up response fluid is assumed to be non-compressibility, and also
in seabed but also the displacement behaviour of the viscous boundary technique is used to create the
pipeline on seabed due to horizontal cyclic loading is infinite of seabed ground in this analysis. There are
numerically investigated through some examples from three governing equations of a kinematic equation

629

Copyright 2005 Taylor & Francis Group plc, London, UK


between soil and pipeline, pore water input/output of joint element for liquefaction analysis, and both
balance equation and dynamic water pressure wave values Kn and Ks shown in Table 1 are unit tangential
propagating equation for pore fluid. Pore water pres- stiffness for normal direction and shear direction,
sure can be expressed by an increment of volumetric respectively. Three cases of Soft, Medium and
strain of soil skeleton because of undrained condition, Hard are used in this numerical analysis.
and also dynamic water pressure wave propagating Material properties for sand layer are shown in
equation for pore fluid can be represented by the tech- Table 2. The liquefaction analysis in this paper uses a
nique that the effect of pore fluid can be taken into strain space plasticity approach for cyclic mobility in
consideration by using an additional mass to the soil- order to represent the realistic hysteretic damping factor
structure kinematic equation. under cyclic loading. In this approach, actual cyclic
Figure 1 shows a finite element mesh for liquefac- shear mechanism is decomposed into a set of one
tion analysis. The pipeline diameter D is 1 m and the dimensional virtual simple shear mechanism. Material
initial depth of pipeline zo is 0.1 m as shown in Figure 1. properties of dilatancy S1, w1, c1, p1 and p2 shown in
The analytical domain is 3.5 m  10 m and is assumed Table 2 are five parameters to define the cumulative
to be a sand layer with thickness 3.5 m. The joint elem- volumetric strain of plastic nature for representing
ent is used at contact area between pipeline and sand cyclic mobility. These parameters define the correlation
layer. Before the cyclic loading of pipeline, the self- between the liquefaction front parameter (Iai et al.
weighted analysis for pipeline-seabed interaction 1990) and the normalized shear work. The liquefaction
problem is carried out under the completely drained front parameter is given by a function of shear work,
condition to obtain the initial effective stress of each and Towhata & Ishihara (1985) obtained the correlation
soil element. between the shear work and the excess pore pressure,
and also concluded that the correlation is independent
of the shear stress paths with or without the rotation
3 NUMERICAL RESULTS of principal stress axes.
Figure 2 shows pipeline behaviour during 100 hori-
In the liquefaction analysis of a pipeline on seabed zontal cyclic loading in Medium case for cyclic fre-
quencies f  0.5, 1.0 and 2.0 Hz. These horizontal
surface subjected to horizontal cyclic loading, the verti-
cal load V is assumed to be maintained a constant value
during horizontal cyclic loading H. Before cyclic load- Table 1. Joint element stiffness.
ing, the static analysis subjected to a constant vertical
load Vmax is carried out, and then both a constant ver- Friction angle
tical load V and horizontal cyclic loading H operate at Kn (kN/m) Ks (kN/m) (degree)
the centre of pipeline as shown in Figure 1. The joint
Soft case 1.0  104 1.0  103 25
element is employed in this numerical analysis to simu- Medium case 1.0  106 1.0  105 25
late a slip phenomenon at contact area between pipeline Hard case 1.0  108 1.0  107 25
and seabed. Table 1 indicates the stiffness conditions

D=1m
Joint Element V, Vmax
Element A
Element B H

3.5m

10 m

Figure 1. Fem mesh for pipeline-seabed interaction analysis.

630

Copyright 2005 Taylor & Francis Group plc, London, UK


cyclic frequencies are employed according to the exper- increasing the number of cycles, and also the final
imental data by Zhang & Randolph (2001b). It can be displacement has a tendency to increase with increas-
seen from these figures that both horizontal and verti- ing the cyclic frequency. This is because that the pore
cal displacements of pipeline gradually increase with pressure accumulation in seabed under a pipeline due
to horizontal cyclic loading is larger and also the ampli-
Table 2. Parameters for sand layer material. tude of horizontal cyclic movement of pipeline is much
larger with increasing the cyclic frequency f.
Initial shear modulus, Gma (kPa) 112,400 Figure 3 illustrates the pipeline behaviour during 100
Bulk modulus of pore water, Kf (kPa) 2.2  106 horizontal cyclic displacement H  0.05 m under a
Mass density, (t/m3) 1.95 constant vertical load V  4 kN/m and the maximum
Limiting value of damping factor, Hm 0.3 vertical load Vmax  4 kN/m in Medium case. The
Elastic tangent bulk modulus of
soil skeleton, Kma (kPa) 299,700
final settlement of pipeline is larger with increasing the
Porosity of soil skeleton, n 0.45 cyclic frequency f, and also approaches a stable value
Poissons ratio, v 0.33 for each cyclic frequency. Although the effects of both
Friction angle,  (degree) 35 joint element stiffness and vertical load on the pipeline
Material parameters for dilatancy settlement are not included in this paper due to the
p1 0.5 limited space, it is found that the pipeline settlement is
w1 7.0 larger with the increase of both joint element stiffness
c1 1.5 and vertical load. This is because that the pipeline
p2 1.3 movement due to horizontal cyclic displacement is
S1 0.005 directly transmitted to the seabed surface as both the
joint element stiffness and vertical load increase.
2.5 2.5 2.5
2.0 2.0
Horizontal Load (kN/m)

2.0
Horizontal Load (kN/m)

Horizontal Load (kN/m)


1.5 1.5 1.5
1.0 1.0 1.0
0.5 0.5 0.5
0.0 0.0 0.0
-0.5 -0.5 -0.5
-1.0 -1.0 -1.0
-1.5 -1.5 -1.5
-2.0 -2.0 -2.0
-2.5 -2.5
-0.10 -0.08 -0.06 -0.04 -0.02 0.00 0.02 0.04 0.06 0.08 0.10 -0.10 -0.08 -0.06 -0.04 -0.02 0.00 0.02 0.04 0.06 0.08 0.10 -2.5
-0.10 -0.08 -0.06 -0.04 -0.02 0.00 0.02 0.04 0.06 0.08 0.10
Horizontal Displacement (m) Horizontal Displacement (m) Horizontal Displacement (m)
f=0.5Hz f=1.0Hz f=2.0Hz
(a) Horizontal displacement and horizontal load
0.00 0.00 0.00
Vertical Displacement (m)
Vertical Displacement (m)

Vertical Displacement (m)

-0.05 -0.05 -0.05

-0.10 -0.10 -0.10

-0.15 -0.15 -0.15

-0.20 -0.20 -0.20

-0.25 -0.25 -0.25

-0.30 -0.30 -0.30

-0.35 -0.35 -0.35


-0.10 -0.05 0.00 0.05 0.10 -0.10 -0.05 0.00 0.05 0.10 -0.10 -0.05 0.00 0.05 0.10
Horizontal Displacement (m) Horizontal Displacement (m) Horizontal Displacement (m)
f=0.5Hz f=1.0Hz f=2.0Hz
(b) Horizontal displacement and settlement behaviour
0.00 0.00 0.00
Vertical Displacement (m)

Vertical Displacement (m)


Vertical Displacement (m)

-0.05 -0.05 -0.05

-0.10 -0.10 -0.10

-0.15 -0.15 -0.15

-0.20 -0.20 -0.20

-0.25 -0.25 -0.25

-0.30 -0.30 -0.30

-0.35 -0.35 -0.35


0 20 40 60 80 100 120 0 20 40 60 80 100 120 0 20 40 60 80 100 120
Number of Cycles Number of Cycles Number of Cycles
f=0.5Hz f=1.0Hz f=2.0Hz
(c) Vertical displacement and number of cycles

Figure 2. Pipeline behaviour during horizontal cyclic loading (Medium, H  2 kN/m, V  4 kN/m, Vmax  4 kN/m).

631

Copyright 2005 Taylor & Francis Group plc, London, UK


0.00 0.00 0.00

Vertical Displacement (m)

Vertical Displacement (m)


Vertical Displacement (m)

-0.05 -0.05 -0.05

-0.10 -0.10 -0.10

-0.15 -0.15 -0.15

-0.20 -0.20 -0.20

-0.25 -0.25 -0.25

-0.30 -0.30 -0.30

-0.35 -0.35 -0.35


-0.10 -0.05 0.00 0.05 0.10 -0.10 -0.05 0.00 0.05 0.10 -0.10 -0.05 0.00 0.05 0.10
Horizontal Displacement (m) Horizontal Displacement (m) Horizontal Displacement (m)
f=0.5Hz f=1.0Hz f=2.0Hz
(a) Horizontal displacement and vertical displacement
0.00 0.00 0.00

Vertical Displacement (m)


Vertical Displacement (m)

Vertical Displacement (m)


-0.05 -0.05 -0.05
-0.10 -0.10 -0.10
-0.15 -0.15 -0.15
-0.20 -0.20 -0.20
-0.25 -0.25 -0.25
-0.30 -0.30 -0.30
-0.35 -0.35 -0.35
0 20 40 60 80 100 120 0 20 40 60 80 100 120 0 20 40 60 80 100 120
Number of Cycles Number of Cycles Number of Cycles
f=0.5Hz f=1.0Hz f=2.0Hz
(b) Vertical displacement and number of cycles

Figure 3. Pipeline behaviour during horizontal cyclic loading (Medium, H  0.05 m, V  4 kN/m, Vmax  4 kN/m).

10 10 50
Pore Water Pressure (kPa)
Shear Stress, xy (kPa)

45
Shear Stress, xy (kPa)

40
5 5
35
30
0 0 25
20
15
-5 -5
10
5
-10 -10 0
-30 -20 -10 0 10 20 30 0 5 10 15 20 0 20 40 60 80 100
Shear Strain, xy (%) Mean Effective Stress, - m' (kPa) Number of Cycles
Shear stress and shear strain Shear stress and mean effective stress Pore pressure response
(a) Hard case
10 50
Pore Water Pressure (kPa)

10
Shear Stress, xy (kPa)

45
Shear Stress, xy (kPa)

40
5 5 35
30
0 0 25
20
15
-5 -5
10
5
-10 -10 0
-30 -20 -10 0 10 20 30 0 5 10 15 20 0 20 40 60 80 100
Shear Strain, xy (%) Mean Effective Stress, - m' (kPa) Number of Cycles
Shear stress and shear strain Shear stress and mean effective stress Pore pressure response
(b) Soft case

Figure 4. Stressstrain relationship (Element A, H  2 kN/m, V  2 kN/m, Vmax  4 kN/m, f  1.0 Hz).

Figure 4 indicates stressstrain curve, stress path strain extremely increases with increasing the number
and pore pressure response at Element A as shown in of cycles. This is because that there exists a peak in
Figure 1 during 100 horizontal cyclic loading H  the pore pressure response at the initial cyclic stage and
2 kN/m under a constant vertical load V  2 kN/m a liquefaction phenomenon occurs at the same time.
and the maximum vertical load Vmax  4 kN/m in In addition, the amplitude of pore pressure response
Hard and Soft cases. It can be seen from these fig- after a peak in Hard case is much larger than that in
ures that the shear stress rapidly decreases and the shear Soft case, because the pipeline movement during

632

Copyright 2005 Taylor & Francis Group plc, London, UK


60 40
f=1Hz

Excess Pore Pressure (kPa)


Excess Pore Pressure (kPa)

f=1Hz f=2Hz
50 f=0.5Hz 30
40
20
30
10
20

10 0
f=0.5Hz
f=2Hz
0 -10
0 10 20 30 40 50 0 10 20 30 40 50
Time (s) Time (s)
(a) Element A (b) Element B

Figure 5. Excess pore pressure response (Medium, H  2 kN/m, V  2 kN/m, Vmax  4 kN/m).

60 40
f=1Hz
f=0.5Hz Excess Pore Pressure (kPa)
Excess Pore Pressure (kPa)

50 30
40
20
30
10
20 f=2Hz
f=0.5Hz
10 0
f=2Hz f=1Hz
0 -10
0 10 20 30 40 50 0 10 20 30 40 50
Time (s) Time (s)
(a) Element A (b) Element B

Figure 6. Excess pore pressure response (Medium, H  2 kN/m, V  4 kN/m, Vmax  4 kN/m).

cyclic loading in Hard case is more directly transmit- phenomenon occurs at seabed due to rapid build up of
ted to seabed in comparison with Soft case. This pore pressure.
implies that the joint element stiffness plays an import- Figure 7 illustrates the deformation of seabed and
ant key role to make an accurate estimation of pore pipeline movement after horizontal cyclic loading
pressure build up response in seabed due to horizon- for three cyclic frequencies f  0.5, 1.0 and 2.0 Hz in
tal cyclic loading of pipeline. Medium case. The liquefaction analysis can deal with
Figures 5 and 6 show excess pore pressure responses a large strain problem by an automatic re-meshing
at Elements A and B during horizontal cyclic load technique based on the increment stress due to the
H  2 KN/m for V  2 KN/m and 4 kN/m in self-weighted analysis in accordance with the dis-
Medium case. It can be seen from these figures that placement amount at each loading step. It can be seen
the excess pore pressure response at each Element for from these figures that the seabed around pipeline is
V  4 KN/m is larger than that for V  2 KN/m, deeply excavated by the horizontal cyclic movement
because the excess pore pressure greatly depends on a of pipeline with increasing the horizontal cyclic fre-
constant vertical load V during horizontal cyclic load- quency. Although there is insufficient space to indi-
ing of pipeline. Excess pore pressure seems to be sensi- cate the effect of the ratio H/V on the deformation of
tive to the ratio V/Vmax under the same horizontal cyclic seabed in detail in this paper, the settlement of pipeline
loading H. It also can be observed from Figures 4 and 5 increases and the seabed surface is deeply excavated
that a peak in pore pressure response occurs at initial by the horizontal cyclic movement of pipeline as the
cyclic stage with increasing the cyclic frequency and ratio H/V increases under a constant vertical load and
the pore pressure response after a peak for each cyclic also the vertical load V increases under a constant
frequency gradually decreases with increasing the horizontal cyclic loading H. In particular, the larger the
number of cycles. This is because that the liquefaction joint element stiffness is, the more widely and deeply

633

Copyright 2005 Taylor & Francis Group plc, London, UK


Deformation Mesh scale
20cm 50cm

(a) f = 0.5Hz

Deformation Mesh scale


20cm 50cm

(b) f = 1.0Hz

Deformation Mesh scale


20cm 50cm

(c) f = 2.0Hz

Figure 7. Deformation of seabed and pipeline movement. (Medium, H  2 kN/m, V  4 kN/m, Vmax  4 kN/m).

the seabed surface will be excavated. It should be noted pore water pressure build up response, displacement
that the settlement of pipeline during horizontal cyclic behaviour of pipeline and shear stressshear strain
loading is so sensitive to not only the vertical load V relationship during horizontal cyclic loading under a
but also the ratio H/V under a constant vertical load. constant vertical load. The numerical examples are
also given to support the liquefaction analysis for a
pipeline-seabed interaction problem. In summary, the
following conclusions can be made based on the
4 CONCLUSIONS
results presented in this paper.
The advanced finite element analysis for pore water 1 The pore pressure build up response greatly depends
pressure build up response and displacement behav- on not only the vertical load V during horizontal
iour of pipeline due to horizontal cyclic loading has cyclic loading but also the horizontal cyclic fre-
been investigated to simulate the geotechnical cen- quency f.
trifuge tests. The liquefaction analysis based on the 2 Shear stressstrain relationship in seabed is so sen-
cyclic mobility approach has been carried out to obtain sitive to the joint element stiffness. Although the

634

Copyright 2005 Taylor & Francis Group plc, London, UK


joint element is used to simulate a slip phenom- Education, Culture, Sports, Science and Technology
enon between pipeline and seabed in this numerical in Japan. This support is gratefully acknowledged.
analysis, it is very important to evaluate a non- The author would also like to thank Dr. Y. Kato at
linear relationship between pipeline and seabed Maizuru National College of Technology for some
surface instead of joint element when making an significant comments on the liquefaction analysis of
accurate estimation of pore pressure accumulation pipeline-seabed interaction problem.
during horizontal cyclic loading of pipeline.
3 The settlement of pipeline greatly depends on both
the cyclic frequency f and the ratio H/V of horizon- REFERENCES
tal to vertical load, and also is slightly affected by
the joint element stiffness. The larger value of the Iai, S., Matsunaga, Y. and Kameoka, T. 1990. Strain space
ratio H/V, the more widely and deeply the seabed plasticity model for cyclic mobility, Report of Port and
surface will be excavated by the horizontal cyclic Harbour Research Institute, 29(4), 2756.
movement of pipeline. Taiebat, H. and Carter, J.P. 2000. A semi-empirical method
for the liquefaction analysis of offshore foundations, Int.
Although the comparison of the analytical result J. Num. Anal. Meth. Geomech., 24, 9911011.
with the experimental data in geotechnical centrifuge Takatani, T. and Randolph, M.F. 2003. Experimental study
tests is not presented in this paper due to the limited of pore pressure build up due to horizontal cyclic loading
space, it is necessary for an intensive study on the com- of pipeline, Proc. 13th Int. Offshore Polar Eng. Conf.,
parison because it is the most important role in the Honolulu(USA), 2, 3137.
design of offshore pipeline. Particularly, a non-linear Takatani, T. and Ogawa, F. 2004. Pore pressure build up
response due to horizontal cyclic loading of pipeline,
relationship between pipeline and seabed surface may Proc. 3rd Int. Conf. Advances Struc. Eng. Mech.,
be expressed by other mechanism such as a non- Seoul(Korea), S2C.
linear spring model instead of the joint element used Towhata, I. and Ishihara, K. 1985. Modelling soil behaviour
in this liquefaction analysis. There seems to be a need under principal stress axes rotation, Proc. 5th Int. Conf.
for the comparison study concerning the non-linear Num Method Geomech., Nagoya, 523530.
spring model and the joint element model in future. Zhang, J., Stewart, D.P. and Randolph, M.F. 2001a. Centrifuge
On the other hand, although the analytical results modelling of drained behaviour for pipelines shallowly
for the initial depth of pipeline zo  0.1 m and the effect embedded in calcareous sand, Int. J. Phys. Model.
of the ratios H/V and V/Vmax on the pore pressure Geotech., 1(1), 2539.
Zhang, J. and Randolph, M.F. 2001b. North west shelf trunk-
response are not sufficiently indicated due to the limited line system expansion project, cyclic loading of pipelines
space, the simulation results for not only the ratios H/V Centrifuge modelling, GEO: 01257, The University of
and V/Vmax but also the other initial depth of pipeline Western Australia.
will be described in another article in near future. Zhang, J., Stewart, D.P. and Randolph, M.F. 2002. Vertical
load-displacement response of untrenched pipelines,
J. Offshore Polar Eng., 12(1), 7480.
ACKNOWLEDGEMENTS

This study is supported by Grants-in-Aid for Scientific


Research (No.16560441) from the Ministry of

635

Copyright 2005 Taylor & Francis Group plc, London, UK


A numerical model of onset of scour below offshore pipelines
subject to steady currents

D. Liang & L. Cheng


School of Civil and Resource Engineering, The University of Western Australia, Crawley, Australia

ABSTRACT: This paper presents a numerical model of onset of local scour below offshore pipelines subject
to steady currents. Based on experimental evidence, onset of scour occurs when sediments immediately down-
stream the pipeline become unstable due to the seepage flow in the underlying sediment bed. The criterion for
onset of scour is derived through examining the forces acting on a small volume of sediment grains, which
include the hydraulic pressure gradient force induced by the seepage flow, the submerged weight of soils, and
the normal and shear forces imposed by the underlying sands. The pressure gradient that governs the seepage
flow below the pipeline is determined by solving the two-dimensional Reynolds-averaged continuity and
Navier-Stokes equations with the standard k- turbulence closure in a general curvilinear coordinate system. The
free water surface is tracked in the model. The critical incoming flow velocity for the onset of scour is then cal-
culated and the results are compared well with experimental data available in literature.

1 INTRODUCTION when the exit pressure gradient in the soil exceeds the
flotation gradient of the sand. The existence of the high
Offshore pipelines laid directly on the sandy seabed are pressure upstream and the low pressure downstream
often subject to local scour under severe environmental the pipeline induces a seepage flow through underlying
conditions. The scour poses a risk to the safe operation sediments and thus seepage force on soil particles.
of pipelines. Therefore understanding of critical onset Sumer et al. (2001) studied the onset of scour in cur-
condition of scour is important for pipeline design. rents and in waves. The pressure was measured on the
Over the last two decades, considerable research surface of a slightly buried pipeline at two points in the
efforts have been devoted to understand the maximum sand, one at the upstream side and the other at the down-
scour depth and its time development. For example, stream side of the pipe. It was shown that the excessive
the experimental work can be found in Mao (1986) seepage flow and the resulting piping are the major
and Chiew (1991), while the numerical study can be factors causing the onset of scour. In the same study,
found in Li & Cheng (1999) and Brrs (1999). There the critical condition for the onset of scour (expressed
have also been some experimental studies on the as the non-dimensional incoming flow velocity) was
onset of scour below pipelines (Mao 1986, Chiew measured for several embedment depths.
1990, Sumer et al. 2001). When a pipe is placed on a Study on the onset of scour requires multidiscipli-
flat bed without any gap in between, the flow is dis- nary examination of the flow around the pipeline and
turbed and vortices are formed in the neighborhood of the seepage flow in the soil. There are still some ambi-
the pipeline. Piping and stagnation eddy may combine guities in understanding the onset mechanism, e.g.
to undermine the pipeline. Early experiments (Mao whether the direction of the pressure gradient in the
1986) emphasized the role of the two small vortices soil needs to be considered, how to express the pressure
close to the pipeline: one in front of and the other just gradient appropriately, what is the effect of water
behind the pipeline. They were thought of being depth, etc. This paper quantitatively analyzes the
responsible to carry sand particles away from the onset of scour and proposes a numerical method to
footing area of the pipeline until a small opening is evaluate the critical onset condition. The critical con-
formed underneath the pipeline, which then leads to dition is determined from the pressure drop across the
tunnel scour. Whereas, through a series of experi- pipeline solved from a hydrodynamic model and the
ments conducted by Chiew (1990), it was found that stability analysis of the sand grains. In the latter
the piping is the dominant cause for the initiation of analysis, the direction of the hydraulic gradient force
scour. Conclusion was drawn that scour takes place acting on sand particles next to the pipe (downstream

637

Copyright 2005 Taylor & Francis Group plc, London, UK


side) is taken into account, and the average pressure
gradient along the buried part of the pipe surface is
taken as the representative hydraulic gradient. It is
noticed that the non-dimensional critical flow velocity
for the onset of scour is dependent on the water depth
(at least for shallow water) and the pipeline embedment
depth. The predicted results agree well with the
experimental data. Until now, no such studies have
been reported in literature. Figure 1. Sketch map of the flow past a pipeline.

2 THEORETICAL BACKGROUND

2.1 Overview
Figure 1 shows the sketch map of the flow after a
pipeline is placed on a plane seabed. The incoming wall
boundary layer has the thickness of  and free stream
velocity of u0. One large and two small vortices are
often formed around the pipe (Mao 1986). In Figure 1,
the subscript D stands for detached, while R
stands for reattached. In front of the pipeline, xD is
where the flow detaches from the bottom wall, and R
is the position on the pipeline surface where the Figure 2. Forces on the sand grains behind the pipe.
incoming flow reattaches to the pipelines front side.
This reattached shear layer separates again at around grains is totally balanced by the vertical pressure gradi-
the top of the pipeline, D, due to the adverse pressure ent. Therefore sand particles behave like fluid and are
gradient. A large recirculation zone is formed behind free to move. The second is that the friction force acting
the pipeline, up to x  xR where the flow reattaches to on sand particles can not resist the horizontal com-
the bottom wall. At the same time, a small downstream ponent of the pressure gradient force, so that sand
vortex at the corner of the pipeline and the bottom wall particles begin to slide on the bed. Close examination
is induced within this primary recirculation zone. of the two situations indicates that the second situation
The pressure is high in front of the pipe where the always happens before the first. For this reason, only
flow becomes partially stagnant. The pressure is low the second situation is considered in the following
in the wake region. This pressure difference will derivations.
induce a seepage flow and exert a seepage force on Force equilibrium on the soil volume at point A
soil particles if the underlying soil is permeable. shown in Figure 2 gives,

2.2 Stability analysis of sand grains (1)


It is observed in experiments (Sumer et al. 2001) that
scour always initiates immediately behind the pipeline. (2)
Considering a small volume (with the volume of l3)
of sands in the sand-water interface just downstream where  is soil internal friction angle. When unstable,
the pipe, the forces acting on it are shown in Figure 2.
Assume that the sand particles are very loosely
connected with one another in this horizontal layer (3)
and there are no forces among them. In the figure,
Fs  hydraulic pressure gradient force;  water so,
density; g  acceleration due to gravity; s  the spe-
cific weight of sediment; n  the bed porosity; N and (4)
f  the normal and the shear forces exerted by the
underlying soils. The hydraulic gradient force acting Thus, piping occurs when the pressure gradient along
on soil particles directly downstream the pipeline is in the pipeline satisfies:
the tangential direction of the pipeline at that point
because the pipe surface is a streamline. (5)
There are two situations that sand particles may
become unstable. The first is that the weight of sand

638

Copyright 2005 Taylor & Francis Group plc, London, UK


Fs/l 3 is the hydraulic gradient force per unit volume, 2
i.e. the pressure gradient at point A.

y/D
It is noted that the average pressure gradient along 1
the pipeline surface in the soil is 0
-2 0 2 4 6 8
(a) x/D
(6)
2 0.

-0.10
40

30
20 0
-0.

0.1
-0.5

-0.4
-0.
0

-0
0.3

y/D

.5
0

0
1 -0.60
where p1, p2  pressure acting on the seabed directly

-0.20
0.40

-0.30
0.50
upstream and downstream the pipeline;  is angle 0
0.
60

defined in Figure 2; the pressure coefficient is -2 0 2 4 6 8


defined by (b) x/D

2 0.06 0.08 0.10


(7) 0.0
4 0.14 0.18
0.22

y/D
6 0.18 0.26
1 0.0
0.2 0.24

0.12
.08 0.1 0.18 2

10
0 0.20

0.0
0.
2 0.14
0. 0. .06

0.

10
0.16

8
0.08 0044
0
where p0  reference pressure. The pressure gradient -2 0 2 4 6 8
at point A can be related to the average pressure (c) x/D
gradient along the buried pipeline surface by a
coefficient A. Figure 3. Calculated flow past a surface-mounted cylinder.
(a) Streamlines; (b) Pressure coefficient; (c) Square root of
turbulent energy (k/u0)1/2.
(8)
The computation of unbounded flow past a surface-
Therefore the piping criterion can be written as: mounted cylinder is presented in the following as
another benchmark test case for the flow model, which
is also very relevant to the onset of scour. The flow
(9) field sketch is demonstrated in Figure 1. In the veri-
fication, e  0.0 and the computational domain
extends from 15 D upstream to 40 D downstream the
cylinder horizontally and covers 10 D above the bottom
It can be seen from Equation (9) that the critical wall vertically. The Reynolds number, Re, based on u0
current velocity required for onset of scour depends and D is 45,000 and the boundary layer thickness is
on pipeline diameter, embedment depth (in terms of 1 D. Smooth wall-function boundary condition is speci-
), specific weight, internal friction angle and poros- fied on the cylinder surface and at the bottom wall.
ity of the sediments, and pressure gradient coefficient
ACP. The following two sections will focus on the
numerical method to determine CP. A will be dis- 3.2 Verification results
cussed in Section 5. The predicted flow field is shown in Figure 3. As can
be seen in Figure 3a, the major flow features observed
by Mao (1986) are reproduced numerically. Figure 3b
3 VERIFICATION OF THE FLOW MODEL illustrates the distribution of the pressure coefficient,
which confirms the existence of the pressure drop
3.1 Model and flow descriptions between the upstream and downstream sides of the
The flow model used in this paper to calculate CP has pipe. Figure 3c shows the distribution of the turbulent
been thoroughly validated against experimental results kinetic energy. The maximum value of about k  0.07
on flow around a circular cylinder 0.37 D above a rigid u02 is predicted in the wake region. This is in good agree-
wall (Liang & Cheng 2005). The governing equations ment with that predicted by Brrs (1999) using an
are the two-dimensional Reynolds-averaged continuity independent model.
and Navier-Stokes equations with standard k- turbu- Figure 4 plots the pressure coefficient along the
lence closure. The pressure and velocity are coupled by bottom wall. It can be seen that the predicted wall pres-
pressure Poisson equation. The differential equations sure is compared well with the measurements of
are solved using finite difference method. For details of Tsiolakis (1982). This pressure distribution can be
the numerical implementation and boundary condi- used to calculate the seepage flow in the soil. It is found
tions, readers are referred to Liang & Cheng (2005). from the predicted result that the pressure remains

639

Copyright 2005 Taylor & Francis Group plc, London, UK


almost constants closely in front of and behind the 4 CALCULATION OF THE PRESSURE DROP
sudden drop at x/D  0, indicating that the seabed can
be roughly treated as the iso-pressure line in the region The pressure drop over the pipeline is crucial in
very close to the cylinder. This characteristic will be studying the onset of scour. According to the authors
used in analyzing the seepage force below the pipeline. computational experiment, the bed pressure distribu-
Some key parameters used and predicted by the tion is not very dependent on the Reynolds number.
flow model are compared with the laboratory experi- The flow concerned in this study may be similar to flow
ments and some other numerical simulations in Table 1. around an isolated circular cylinder where the hydro-
The drag and lift force coefficients are defined by dynamic properties are insensitive to the Reynolds
number in the sub-critical regime. Therefore, for each
geometric configuration (a combination of e/D and
(10) h/D, where h is water depth), only one representative
flow condition is simulated and the predicted CP is
assumed to be universal for all Reynolds numbers in the
where FD and FL are the forces acting on the cylinder sub-critical regime. In the computation, the thickness of
by the flow in the longitudinal and transverse direc- incoming boundary layer is set the same as the water
tions respectively. The definitions of other symbols depth; D is 100 mm; u0 is 250 mm/s; the bed is assumed
can be found in Figure 1. Compared with the experi- to be hydraulically rough and the Nikuradse rough-
mental data, the upstream vortex is predicted shorter ness is equal to 2.5d50  0.45 mm. These parameters
and the main downstream vortex is predicted smaller. are chosen according to the experimental conditions
The overall agreement is quite acceptable. reported by Sumer et al. (2001).
The free surface effect is considered. The pressure
on the free surface is specified constant in solving
N-S equations. The position of the free surface is
0.5 determined by the following kinematic boundary
condition on the free surface.
CP

0
-0.5
(11)
-1
-5 0 5 10 15
x/D
where ys is the free surface position, u and v are the flow
Figure 4. Pressure coefficient along the bottom wall. Calcu- velocity components in x and y directions respectively.
lated result; Experimental results by Tsiolakis (1982) for Figure 5 shows some typical flow fields simulated
Re  9000 () and Re  18,000 (). by the model for the water depth of 3.0 D, with two

Table 1. Comparison of some key parameters in different researches.

Reference Re (104) /D CD CL

Experiments 0.615 0.25.2 0.51.0 0.170.8


Solberg & Eidsvik 1989 4.5 0.8 0.85 0.77
Solberg 1992 4.5 1.0 0.63 0.65
Brrs 1999 1.5 1.0 0.72 1.16
Present 4.5 1.0 0.69 0.65

Upstream Downstream
vortex primary vortex

Reference
xD/D R* D* xR/D hs/D

Experiments 1.01.5 6090 180210 810 1.6


Solberg & Eidsvik 1989 1.20 66.5 201.2 5.9 1.02
Solberg 1992 0.98 75.0 190.0 9.0 1.28
Brrs 1999 0.77 69.5 208.9 7.2 1.08
Present 0.69 68.7 194.0 7.45 1.22

* in degree.

640

Copyright 2005 Taylor & Francis Group plc, London, UK


different pipe embedment depths. The comparison of pressure drop. For a given pipeline, larger embedment
Figure 5a with Figure 3a indicates that the size of pri- is effectively the same as larger water depth in decreas-
mary downstream vortex decreases with the water ing the blockage ratio of the pipeline to the flow. For
depth. In addition, the length and height of this main h/D  2.5, comparison is made between the compu-
vortex decrease when the pipeline is partially embed- tational results and the experimental data (Chiew
ded in the seabed (Figure 5b). 1990). Good agreements are observed.
The purpose of this section is to determine the
hydraulic drop for different combinations of water
depth and pipeline embedment depth. Figure 6 shows 5 DETERMINATION OF THE CRITICAL
the variance of CP versus e/D with h/D as a third CONDITION FOR THE ONSET OF SCOUR
variable. Three water depths of 2.5 D, 3.0 D and 15.0 D
are examined. It shows that the pressure drop is Coefficient A appears on the right hand side of
highly dependent on both e/D and h/D. In both water Equation (9), which reflects the amplification of the
depths examined here, hydraulic jump does not exist pressure gradient at the exit point of the seepage flow
behind the pipeline. The larger pressure drop pre- behind the pipe relative to the average pressure gradient
dicted for the smaller water depth is mainly caused by along the pipe surface. It is anticipated that A will be
the larger blockage effect. It is not surprising to find generally greater than 1 due to the concentration of
that the larger the embedment depth, the smaller is the equal-potential lines near the seepage exit point.
However A  1.0 is assumed in the actual calculations.
A justification for this is discussed in the following.
3 Figure 7 shows the predicted pressure distribution
in the water and in the soil. The numerical labels in
2 the figure indicate the pressure coefficients. The pre-
y/D

sent study computes the water flow above and the


1
seepage flow below the bed separately. The flow field
0 around the pipeline is calculated first so the pressure
-2 -1 0 1 2 3 4 5 6 7 distribution on the bed is known. Then this pressure
(a) x/D distribution is used as the boundary condition for cal-
culating the seepage flow in the soil. The seepage
3 flow is governed by the Laplaces equation, which is
also solved by the finite difference method.
2 It can be seen from Figure 7 that the pressure actually
y/D

does not decrease uniformly along the pipeline surface


1 in the soil. Mathematically, the two corner points in
front of and behind the pipeline in the soil-water inter-
0 face are singular points in the seepage flow model.
-2 -1 0 1 2 3 4 5 6 7
(b) x/D At these two points, the streamline, which follows
the pipe surface in the soil, is not perpendicular to the
Figure 5. Calculated streamlines for h/D  3.0. iso-pressure lines, which roughly coincide with the
(a) e/D  0.0; (b) e/D  0.15. seabed. It is therefore difficult to accurately calculate
the hydraulic pressure gradient near these two points.
The other consideration is that the piping is a macro
0.3 phenomenon. Thus, it may not be worthy to spend a
h/D=2.5 (Exp.)
h/D=2.5 (Comp.)
lot of effort to determine the exact local pressure gra-
h/D=3.0 (Comp.) dient at one point. In the present study, the average
0.2 h/D=15 (Comp.) pressure gradient is adopted to calculate the critical
condition for the onset of scour.
e/D

Bearing in mind that A  1.0 in Equation (9), the


0.1 non-dimensional critical flow velocity is dependent
on , CP and . Among these parameters, angle 
can be derived through the embedment depth e/D;
0 CP can be related to h/D and e/D as stated in Section
0 1 2 3 4 4; the soil internal friction angle  is taken to be 33
Cp at present. Therefore, the non-dimensional critical
flow velocity should be the function of e/D and h/D.
Figure 6. Variance of the pressure drop coefficient with The solid curve in Figure 8 illustrates the predicted
h/D and e/D. critical flow velocity for h/D  3.0, in comparison

641

Copyright 2005 Taylor & Francis Group plc, London, UK


2

-0.10
5
20
30

0.

5
-0.1
-0.

0.0
0.00
-0.35

.2
0.1

-0.05
10

-0.
-0
5
1.5

5
0.20 -0.40

0.25 -0.45
1 1

y/D
0.30
y/D

5
0.3
0.
40 Water 0.5
0 0.35
-0.45
-0.40 Soil
0.30 -0.3
5
0
-0.3
0.25 -0 0
-0 .25 -1 0 1 2
-0.05

-0 .
-0.1
0.05
0

.2
0.00
0.1

0 15
0
15

0.2 0. (a) x/D


0

-1
-2 -1 0 1
x/D

Figure 7. Pressure distribution in the water and in the soil. 1.5

1
y/D
1
u0 2

gD(1-n)(s-1) 0.5

0.1 0

-1 0 1 2
(b) x/D

0.01 Figure 9. Initial computational meshes for predicting


0.01 0.1 1 seabed deformation. (a) When scour will onset; (b) When
scour will not onset.
Figure 8. Predicted and measured critical conditions for
the onset of scour. Solid and dashed lines indicate the pre-
dicted results for h/D  3.0 and 15.0 respectively; Symbols force equilibrium acting on the sand grains immedi-
are the measured data by Sumer et al. 2001 (h/D  3.0). ately behind the pipeline. The hydraulic gradient
force is approximated by using the average pressure
gradient along the pipe surface in the soil. The pres-
with the experimental measurement. The experimental sure distribution along the bed is computed from a
results of Sumer et al. (2001) are denoted in the figure 2-D hydrodynamic model. The hydrodynamic model
as symbols. For different water depths, this curve will is verified against a benchmark flow problem. The
change. It is expected that the curve will shift down- predicted pressure drop over the pipe and the critical
ward for smaller water depth, so that the scour is more flow velocity for the onset of scour are compared with
likely to happen. On the other hand, it will move upward the experimental results of Chiew (1990) and Sumer
as the water depth increases. The calculated critical et al. (2001). Good agreements are found. Therefore,
condition for h/D  15.0 is shown in the figure as the the present study confirms the conclusion drawn
dashed line to represent the situation in deep waters. from the experiments that the seepage flow and pip-
Figure 8 can be used to determine the likelihood of ing are the dominant causes for the initiation of scour.
scour with a certain embedment depth and incoming The seepage flow is driven by the pressure drop
flow velocity. When the real flow velocity is larger than between the stagnation pressure upstream and wake
the critical value scour will happen, and vice versa. pressure downstream the pipeline.
Linking with a morphological model (e.g. Liang
et al. 2005), the present study makes it possible to
6 FINAL REMARKS build an integrated numerical package to predict the
seabed deformation automatically. When the onset of
This paper proposes a numerical method to predict scour is predicted by the method proposed in this paper,
onset of scour for a specific incoming flow velocity, a typical topology of the initial computational mesh
water depth and pipe embedment depth. The criterion such as the one shown in Figure 9a should be employed
for the onset of scour is established by analyzing the in the scour simulation. There is an arbitrary small

642

Copyright 2005 Taylor & Francis Group plc, London, UK


gap beneath the pipeline so that the tunnel scour will Li, F. & Cheng, L. 1999. A numerical model for local scour
develop using this kind of mesh. The topology of the under offshore pipelines. Journal of hydraulic engineer-
mesh will not change as the scour develops. When the ing 125(4): 400406.
onset of scour is not predicted, there are two possibil- Liang, D. & Cheng, L. 2005. Numerical modeling of scour
below a pipeline in currents. Part I: Flow simulation.
ities. The seabed will not change if the flow velocity Coastal engineering 52(1): 2542.
is smaller than the threshold velocity for the sediment Liang, D., Cheng, L. & Li, F. 2005. Numerical modeling of
motion. Otherwise, the sediment buildup around the scour below a pipeline in currents. Part II: Scour simula-
pipeline will take place as the result of ambient sedi- tion. Coastal engineering 52(1): 4362.
ment transport being blocked by the pipeline. The ini- Mao, Y. 1986. The interaction between a pipeline and an
tial mesh topology like Figure 9b should be chosen to erodible bed. PhD thesis, Technical University of
predict this morphological process. Denmark, Lyngby, Denmark.
Solberg, T. & Eidsvik, K.J. 1989. Flow over a cylinder at a
phase boundary A model based on (k-) turbulence.
Journal of fluids engineering 111: 414419.
REFERENCES Solberg, T. 1992. A numerical study of laminar and turbulent
separated flows over a circular cylinder at a plane wall.
Brrs, B. 1999. Numerical modeling of flow and scour at PhD thesis, University of Trondheim, NTH, Trondheim,
pipelines. Journal of hydraulic engineering 125(5): Norway.
511523. Sumer, B.M., Truelsen, C., Sichmann, T. & Fredse, J. 2001.
Chiew, Y.M. 1990. Mechanics of local scour around submar- Onset of scour below pipelines and self-burial. Coastal
ine pipelines. Journal of hydraulic engineering 116(4): engineering 42: 313335.
515529. Tsiolakis, E.P. 1982. Reynoldische spannungen in einer mit
Chiew, Y.M. 1991. Prediction of maximum scour depth at einem kreiszylinder gestrten turbulenten plattengren-
submarine pipelines. Journal of hydraulic engineering zschicht. Dissertation, Dortmund University, Dortmund,
117(4): 452466. Germany (in German).

643

Copyright 2005 Taylor & Francis Group plc, London, UK


Arctic seabed ice gouging and large sub-gouge deformations

A.C. Palmer
Cambridge University Engineering Department, Cambridge, England

I. Konuk
Geological Survey of Canada, Ottawa, Canada

A.W. Niedoroda
URS Corporation, Tallahassee, FL, USA

K. Been
Golder Associates, Houston, TX, USA

K.R. Croasdale
K.R. Croasdale and Associates, Calgary, Canada

ABSTRACT: Seabed ice gouging is a potential problem in Arctic hydrocarbon production. Large ice masses
ground in shallow water and gouge into the seabed, pushed along by wind, current and the pressure of other ice
masses. Gouging occurs in the Beaufort Sea, off the coast of Sakhalin, in the Canadian Arctic Islands, in the
Russian Arctic, and in the northern Caspian. It has been of much concern to pipeline engineers. Severe damage
would occur if the cutting force required to cut one of the observed gouges were applied to a marine pipeline,
and is possible even if a pipeline is buried below the maximum gouging depth, because the ice drags the seabed
soil beneath the gouge and distorts the pipeline. New developments using ALE methods make it possible to
determine deformations confidently. The results are compared with those from centrifuge model tests.

1 INTRODUCTION The seabed conveniently divides into three zones


(Palmer et al. 1990). Zone 1 is the highest: a structure
Sea ice in the Arctic and sub-Arctic drifts into shallow in this zone is contacted by ice directly. Zone 2 is the
water, grounds on the seabed, and is then pushed fur- next highest: a structure in zone 2 is not contacted by
ther. As it moves under the force of wind, current and the ice itself, but the large deformations that occur
other masses of ice, it gouges into the bottom. Often beneath the gouge drag the pipeline forward, and may
repeated fracture between sheets of ice creates pres- damage it. Zone 3 is the lowest: there the deform-
sure ridges that reach much deeper than the ice sheet ations of the soil are small, and a pipeline at that level
as a whole. Darwin (1855) and Lyell noted this phe- is not loaded to the extent that it reaches a limit state.
nomenon. Gouging has been observed in many parts The potential threat to Arctic pipelines and subsea
of the world, as far south as the northern Caspian, structures was identified a long time ago. Much work
Lake Erie and offshore Sakhalin, and repetitive map- has been done, though the level of research has ebbed
ping confirms that it is contemporary. and flowed with oil companies engagement with the
Gouge breadths up to 100 m and depths to 7 m have Arctic, and therefore in phase with the oil price. Two
been observed, in water down to 45 m or more deep. Arctic marine pipelines have been built, but they are
True icebergs are relatively rare, but they can be much in relatively sheltered areas of the Arctic Archipelago
larger still. A back-of-the-envelope calculation shows (Palmer et al. 1979) and the Beaufort Sea where deep
that the force that must be applied to the seabed to cut a gouges do not occur.
deep gouge can easily reach 100 MN. If this level of The emphasis of gouging research has shifted with
force were applied to a seabed pipeline, substantial dam- operators moves into geographical areas where the
age would occur. Protection against it is extremely diffi- environmental conditions are different. The principal
cult, and scarcely likely to be economically practicable. outstanding questions centre on the maximum depth

645

Copyright 2005 Taylor & Francis Group plc, London, UK


of gouges and the extent of subgouge deformation. depth that corresponds to an acceptably low exceedance
Pipelines are routinely trenched to trench depths probability. This is the method most commonly used.
between 1 and 2 m, in order to protect them against It is often found that gouge depths follow an expo-
fishing gear, dragging anchors, hydrodynamic forces nential distribution, though there is no physical reason
and seabed level changes. That is not enough to why this should be so. A seductive attraction of the
secure a high level of protection in severely gouged exponential distribution is that it is characterised by
areas. Trenching to much greater depths is rarely neces- one parameter, the mean gouge depth. Extrapolation
sary, except in shallow water less than about 20 m needs to be carried out very cautiously, and the confi-
deep, and there conventional dredgers can reach the dence limits are disappointingly far apart. A recurrent
bottom and cut deep trenches, although the cost is high. difficulty is that the depths of the large number of
New types of trenching equipment may be needed if shallow gouges fit an exponential distribution quite
long deep trenches are to be excavated economically. well, but that the few deep gouges do not: it is of
In the Beaufort Sea the seabed is generally clay or course the deepest gouges that interest us.
silt, and loose sands are only found close to shore. A more fundamental difficulty with the first strat-
Seabed currents are very small, with under-ice speeds egy is the effect of infill caused by wave-induced sedi-
less than 0.1 m/s. In general, there is no sign of sedi- ment transport (Niedoroda & Palmer 1986). In the
ment transport features such as megaripples and Beaufort Sea this is thought not to be a severe prob-
sandwaves. Open water seldom extends very far from lem, except in very shallow water, because the open-
shore, so that the size of storm waves is generally water season is short, the survey is necessarily carried
limited by the small fetch. Because the multi-year out within the same open-water season, and there is a
polar ice is present all the year round, and because good chance that the time interval between gouging
there is an intense shear zone not far from the coast, and survey does not include storms. Repetitive map-
the ice includes well-consolidated and strong multi- ping then confirms that gouges survive through the
year ridges. Those ridges are driven by the wind and winter and remain clean with minimal infill. In other
the moving pack, and they cut the wide and deep areas infill is much more effective, and a single storm
gouges found in water depths between 20 and 30 m can wipe out gouges altogether or partially fill them
and occasionally in deeper water. in, and then the measured depth statistics will be mis-
In contrast, the Sakhalin environment is totally dis- leadingly low. Section 4 quantifies this.
tinct in the following respects: An alternative second strategy is to use keel depth
statistics rather than gouge statistics. Upward-looking
1 the seabed is largely composed of sand;
sonar on the seabed measures the underwater profile
2 longshore currents are much larger, up to 1 m/s;
of the ice as it drifts overhead, a technique that has
3 there appear to be many features indicative of
also be extensively applied from submarines to meas-
active sediment transport, such as sandbanks ori-
ure ice in the Arctic Ocean. One difficulty is that the
ented parallel to the shore, and the coastal lagoon-
sonar can only be placed in deep water, beyond the
and-spit topography;
depth of significant ice contact with the seabed, since
4 there is no multi-year ice;
otherwise ice would destroy it. The measured keel
5 the climate is much less severe, and the ice is thinner;
depths are not necessarily those that are reached when
6 the open-water season is much longer, and in the
the ice begins to interact with the seabed and the
open water season the fetch to the east and north is
shoreline. A second difficulty is that the maximum
essentially unlimited, so that large storm waves can
depth of a keel that is observed before it grounds is
occur;
not the same as the maximum depth it gouges to.
7 first-year ice keels are less consolidated and more
Direct observation by divers confirms that a first-year
fragile than multi-year keels.
keel has an open structure of ice fragments, loosely
These factors have a major influence on interpret- held together by buoyancy and some sintering. It may
ation and assessment. be quite fragile, especially at the bottom, and the
forces induced when they come into contact with the
bottom can easily break the fragments away. Moreover,
2 DETERMINATION OF MAXIMUM the vertical component of the force exerted by the
GOUGE DEPTH seabed on the ice is at least as large as the horizontal
component, and the ice can therefore lift signifi-
Three strategies have been suggested, and there may cantly. This has been observed in the Canadian Arctic,
well be other possibilities. Each strategy encounters and is consistent with calculations.
difficulties. A third strategy is to use geotechnical evidence of
The first strategy is to measure gouge depths by the maximum gouging depth (Palmer 1998). A gou-
echosounding bathymetry, or from side-scan sonar, ging ice mass remoulds the seabed soils, very exten-
and to apply extremal statistics to estimate the gouge sively in zone 1 and to a lesser extent in zone 2.

646

Copyright 2005 Taylor & Francis Group plc, London, UK


Repeated remoulding would be expected to bring the pattern of intersecting clean gouges, and the overwrit-
soil to a critical state, whereas deeper soil would ing sequence can be deduced by looking at the inter-
remain unremoulded in a condition much closer to the sections. A loose analogy is graffiti on a subway car
one in which it was deposited. The current maximum that is hardly ever washed. Repetitive gouge mapping
gouge depth ought therefore to show up as a break suggests that parts of the Canadian Beaufort Sea
in the depth profile of properties such as cone pene- approach this limit. The gouge record gives a clear
tration resistance, vane shear strength, and water con- indication of the statistics of gouge depths and breadths.
tent, measurements that might in any case be made Imagine now the other extreme case. Blocks of
for other reasons such as trenching and platform floating ice gouge the bottom in shallow water, but the
foundation design. As far as is known, this strategy sea is always rough, and any gouge is almost immedi-
has not been deployed in practice, but it is being ately infilled and erased by wave-induced bottom
investigated currently. Potentially confusing factors velocities. A survey shows no evidence of gouging: it
are the original stratigraphy and secular changes of tells us nothing about gouge depths, and there is a risk
sealevel and ice climate. that it might be misinterpreted as an indication that
there is no gouging. Continuing the analogy, that cor-
responds to a subway car that is washed as soon as a
graffito is written, and that might lead an observer to
3 GOUGE STATISTICS FROM KEEL
a mistaken conclusion that no one is writing graffiti.
DEPTH MEASUREMENTS
Active environments correspond to an intermediate
case, the subway car that is washed frequently but
Upward-looking sonar (ULS) makes a continuous
incompletely. Observation can find the number of
record of the depth z(s) of the bottom of the ice, where
gouges and the frequency distribution of gouge depths.
s is the distance moved by the ice. That record can be
However, the engineer who designs a pipeline is not
used in two ways. The method of extremes identifies
interested in the gouge depth for its own sake. He or she
local draft maxima, and uses the maxima to estimate
is interested in the probability density function (pdf)
a keel draft exceedance function G(z). An alternative
for gouge depth because the pdf partially reflects the
is to use the whole depth record, not just the local
underlying distribution of new gouging incidents.
draft maxima, and to use the statistics of depth pro-
Denote by N the total number of gouges in a
files to infer the extreme depths: that alternative
defined area, and by f(x) the pdf for gouge depth x, so
extracts more information from the same ULS record.
that Nf(x)dx is the number of gouges whose depths lie
Similar questions arise in the mechanics of surfaces
between x and x  dx.The number of new gouges
(see, for example, Longuet-Higgins 1957, Nayak
with depths between x and x  dx formed in unit time
1971, Greenwood 1984, Johnson 1985), though much
is n(x)dx. Sediment transport is infilling the gouges,
of that work is founded on an assumption that the sur-
and the rate of gouge infill i(x) of a single gouge is
face is Gaussian, which is not the case for the ice
idealised as depending only on gouge depth, so that in
problem. Nayak emphasises the distinction between
a small time interval dt the depth of a gouge reduces
the statistics of a profile which is what a ULS
from x to x  i(x)dt.
measures and the statistics of a surface. He calls
If the pdf is stationary with time
high points on a surface summits and high points on
a profile peaks. Much more often than not, a profile
will cross the shoulder of a high region, and will iden- (2)
tify peaks rather than summits. The area density of
summits (number of summits per unit area) s is
If the gouge depth distribution can be measured, and
found to be related to the density of peaks p (number
if a sediment transport model determines the rate of
of peaks per unit length of profile) by
gouge infill, then the new gouging rate can be found.
The rate that an ice gouge fills is influenced by the
(1) magnitude of the hydrodynamic forcing, the sediment
properties, the water depth and the geometry of the
gouge. This was explored with the Channels Model
(Reed et al. 2004), a time-dependent 2-D coupled
4 INFILL AND THE GOUGE RECORD process-based hydrodynamic, sediment transport and
morphodynamic model based on a numerical solution
Imagine first an environment in which there is no sedi- to the Reynolds-Averaged-Navier-Stokes (RANS)
ment transport by waves and currents, and ice gou- equations with the shallow water assumptions, the
ging is the only effect that modifies the seabed. Any scalar transport equation and the Exner equation for
gouge retains its original form until it is overwritten bed evolution. It includes representation of turbulence
by another gouge. A survey of the seabed reveals a generation and mixing, wave-current interactions,

647

Copyright 2005 Taylor & Francis Group plc, London, UK


density stabilization, multiple grain sizes, and bed once much deeper. On the other hand, many small
armoring, and also simulates sediment erosion, trans- gouges fill quickly and join other older features that
port, deposition and bed elevation and composition. are too subtle to be reliably counted. A second paper
The model was set up for a sea floor made up of on this issue is in press (Palmer & Niedoroda 2005).
fine sand (D 0.125 mm). Hydrodynamic forcing con- Numerical experiments have been carried out to
sisted of 1 m high 8 second waves acting in conjunc- investigate the sensitivity of these conclusions. The fill
tion with a 0.1 m/sec steady current superimposed on rate is a strong function of the original gouge depth, and
a 0.2 m/sec oscillating tidal current. All the simula- total fill at a fixed time is an approximately linear func-
tions applied the same hydrodynamic and sediment tion of initial depth, and so fractional fill (defined as
conditions. The gouge cross-section shape has a fill/initial depth) is approximately independent of initial
major effect on the rate that the gouge fills. However, depth. Fill rate decreases with increasing water depth.
there is a relatively limited range of shapes, and they These results demonstrate that a good numerical
could be grouped into a system of classes, the sim- model of marine sedimentary processes is an important
plest a V-shaped trough. One shape is sufficient to tool for understanding the way gouges infill, and
illustrate the general relationships and trends. for using this information to determine the gouging
Figure 1 shows a simulation for a gouge that was rate n(x).
originally 2 m deep and 12 m wide. The biggest
change occurs in the first time interval where the
shape changes from its original chiseled form to one 5 SUBGOUGE DEFORMATION
adapted to the dynamics of sediment transport and
deposition. After this initial adjustment that the rate A realisation that subgouge deformation is an impor-
of filling decreases, because sediment is transport tant aspect of the problem came relatively late. Some
right across the gouge, and only a fraction of the total model-scale experiments (Woodworth-Lynas et al.
remains in the gouge. As the gouge continues to fill 1996) indicate that large subgouge deformations
the fraction that is deposited in the gouge gets smaller extend nearly one gouge depth below the visible base
and smaller. The biasing current is from left to right. of a gouge. The explanation is that an ice keel is a
The deepest part of the gouge (thalweg) migrates to blunt cutter, much more so than a device such as a
the right as the filling progresses, something that has plough or a scraper blade, which is designed to cut
also been seen in pipeline trenches. soil as smoothly as possible. Plainly it would be ser-
The sedimentation patterns shown in Figure 1 have ious if subgouge deformation implied that in a region
several implications for analyses of ice gouge statis- where the maximum gouge depth is 5 m, a pipeline
tics. First, it is clear that the sharp original relief is has to be trenched to 10 m.
quickly modified and the true original depth is It is not possible to make full-scale measurements on
obscured, at least for this common gouge shape. The contemporary gouges, though some measurements
filling of the gouge depends on the difference between have been made on relic gouges: there the difficulty is
the rate that sediment is transport to and away from to know the dimensions of the original gouging ice
the gouge. This difference becomes smaller as the mass, and to be certain of the original bed level. Most
gouge shape becomes subtler. Such subtle features investigations have relied on models, sometimes on 1-g
can be difficult to detect with typical sounding and models at modest scales (with gouge depths of the
acoustic imagining instrumentation. order of 0.3 m) and sometimes in centrifuge models at
Competition between two factors distorts gouge smaller scales (0.05 m) (Woodworth-Lynas et al. 1996,
measurement statistics. There is a disproportionate Lach 1996, Allersma et al. 2005). Uncertainty per-
number of small gouges. A small gouge can be either sists, some of it to do with questions about centrifuge
a newly-cut gouge or a partially filled gouge that was modelling of shear zones (Palmer 1991, 2003).
The problem can also be attacked computationally.
The difficulty is that most finite-element (FE) soft-
horizontal distance (m)
ware for geotechnical applications has been designed
20 40 60 80 and validated for problems in which the displace-
19.5
ments are small by comparison with the linear dimen-
20
initial sions of the system. The gouging problem is inherently
depth (m)

20.5 three-dimensional, the displacements are of the same


15 days
21 order as the gouge depth, and a soil particle within
30 days
21.5 zone 1 follows a complicated path which takes it first
60 days
22 upwards into a mound in front of the keel, and then
22.5 transversely until it comes to rest in a berm on either
side of the gouge. It has happened more than once that
Figure 1. Infill of a 12 m wide 2 m deep trench. FE analysts have attacked the problem confidently,

648

Copyright 2005 Taylor & Francis Group plc, London, UK


but have then been dismayed to discover that conver- gouge base, and within one gouge depth from the base,
gence fails after displacements of a small fraction of the particles have very large displacements. Further
the gouge depth, long before the gouge has reached down, the displacements are relatively small but decay
its final form and a mound has developed. It is hard to more slowly. Below two gouge depths below the gouge
have confidence in subgouge deformations estimated base, the displacements for a 1000 mm gouge depth are
in this way. about 40 to 50 mm, and the particle displacements return
Recent work has much improved the position. almost to their original positions after the keel has
Konuk & Gracie (2004) describe the application of passed: this corresponds to the boundary between zone
an Arbitrary-Lagrangian-Eulerian (ALE) technique 2 and zone 3.
implemented in LS-DYNA software (1988). The Figure 3 plots the calculated final positions of soil
Eulerian version is chosen, so that the mesh remains particles that initially lay in a vertical line under the
stationary and the soil moves through the mesh. The centre of the gouge, and compares them with the
soil is modelled by the MAT_GEOLOGICAL_CAP empirical displacement profiles given by the cen-
model (Simo 1988) which is essentially a generalised trifuge models in the PRISE program (Woodworth-
Drucker-Prager cone-cap model with nonlinear kin- Lynas et al. 1996). The PRISE results gave an empirical
ematic hardening. Soil 1 corresponds to soft clay, and relationship for the soil displacement at the base of
soil 2 to a stiffer clay. The deformation is taken to be the keel, as 0.6 times the square root of the product of
undrained. Marker elements allow the displacements the gouge depth and breadth, together with an expo-
to be followed. nential decay law for horizontal displacements below
Numerical experiments with a soil model that cor- the gouge base. It can be seen that the ALE calculation
responds to soft clay show that his model allows the gives larger movements at the immediate base of the
horizontal movement of the keel to continue to sev- gouge, but much smaller movements further down. If
eral times the gouging depth, without the intervention this is correct, it is hugely important for practical
of convergence problems. A mound and berms form decisions about trenching depths.
in the way observed in tests. In principle the ice Another aspect of the gouge problem relates to the
mechanical properties could be modelled as well, so geometry and mechanical properties of the ice keel.
that the base of the ice could erode and the ice mass An advantage of the ALE approach is that the ice ridge
could lift, roll and pitch, but those possibilities have geometry be changed to other geometrical shapes, to
not been included so far. model an irregular keel, and to take account of the
Figure 2 plots particle trajectories. The soil particles limited strength of the keel. That would allow valid-
under the gouge move slightly downwards first and then ation of the results against large-scale tests such as
move upwards. The formation of the mound induces the those carried out by Liferov & Hyland (2005).
initial downward movement, but the soil is then forced Thus far, the numerical analysis has treated the
down by the large vertical forces under the keel. The par- deformation as undrained. It would be highly desir-
ticle displacements exhibit a boundary layer behaviour able to adopt a two-phase model, so that the pore
(as they do close to ploughs). Immediately below the

Figure 2. Particle trajectories. PRISE function refers to Figure 3. Horizontal subgouge displacements. PRISE
Woodworth-Lynas et al. (1996). refers to Woodworth-Lynas et al. (1996).

649

Copyright 2005 Taylor & Francis Group plc, London, UK


water could drain within the soil in response to effect- Johnson, K.L. 1985. Contact mechanics. Cambridge,
ive stress changes. That is technically feasible, but Cambridge University Press.
comes with a substantial computational cost. Konuk, I. and Gracie, R. 2004. A 3-dimensional Eulerian finite
Gouging deformations in nature may often approach element model for ice scour. Proceedings, International
Pipeline Conference, Calgary, paper IPC04-0075.
undrained conditions. The relevant dimensional group Lach, P.R. 1996. Centrifuge modelling of large soil deform-
is VH/ (Palmer 1999), where V is the ice velocity, H ation due to ice scour. Unpublished PhD dissertation,
the gouge depth and  the diffusivity for soil water Memorial University of Newfoundland, St. Johns.
(more often called the Terzaghi coefficient of consoli- Liferov, P. and Hyland, K.V. 2004. In-situ ice ridge scour
dation). If VH/ is large, the deformation is fast tests: Part I: experimental set-up and basic results.
enough to be very nearly undrained. If V is 0.1 m/s, H Submitted to Cold Regions Science and Technology.
is 5 m (for a deep gouge) and  is 2  10
7 m2/s in a Longuet-Higgins, M.S. 1957. Statistical properties of an
silty glacial clay, VH/ is 2.5  106: the deformation is isotropic random surface. Philosophical Transactions of
then undrained (and would still be undrained if both the Royal Society, series A, 250: 157174.
LS-DYNA 1988. Theoretical Manual, Livermore Software
the gouging depth and the ice velocity were only a tenth Technology Corporation.
as large). In a silty sand with  0.5 m2/s, and with the Nayak, P.R. 1971. Random process model of rough surfaces.
same velocity and gouge dimensions, VH/ would Transactions, American Society of Mechanical Engineers,
be 1: partial drainage effects might then be significant, Journal of Lubrication Technology, 93: 398407.
as they are for ploughs. In coarse sand, the deformation Niedoroda, A.W. and Palmer, A.C. 1986. Subsea trench
would be drained and the velocity effect negligible. infill, Proceedings, Eighteenth Annual Offshore Technology
Because of these velocity effects, the vertical and Conference, Houston, OTC 5340, 4:445452.
horizontal gouging forces are both higher when the Palmer, A.C., Baudais, D.J. and Masterson, D.M. 1979.
ice is moving than when it stops. If the force from Design and installation of a submarine flowline in the
Canadian Arctic Islands, Proceedings, Eleventh Annual
neighbouring ice diminishes, the gouging ice mass Offshore Technology Conference, Houston, 2: 765772.
stops, the pore pressures dissipate and the vertical Palmer, A.C., Konuk, I., Comfort, G. and Been, K. 1990. Ice
reaction falls, and the ice sinks and creates the pit gouging and the safety of marine pipelines. Proceedings,
often seen at the end of a gouge. Twenty-second Offshore Technology Conference,
Houston, 3: 235244.
Palmer, A.C. 1991. Centrifuge modelling of ice and brittle
6 CONCLUSIONS materials. Canadian Geotechnical Journal, 28: 896898.
Palmer, A.C. 1998. Alternative paths for determination of
The new insights suggested by recent work are that minimum burial depth to safeguard pipelines against ice
gouge infill is important in active areas, and that there gouging. Proceedings, Ice Scour and Arctic Marine
is increasing evidence that subgouge deformation Pipelines Workshop, Thirteenth International Symposium
may extend less far than previously supposed. ALE on Okhotsk Sea and Sea Ice, Mombetsu, Japan, 916.
methods appear to escape the convergence difficul- Palmer, A.C. 1999. Speed effects in cutting and ploughing.
ties that have dogged previous finite-element ana- Geotechnique, 49: 285294.
Palmer, A.C., White, D.J., Baumgard, A.J., Bolton, M.D.,
lyses of the problem. Barefoot, A.J., Finch, M., Powell, T., Faranski, A.S. and
Baldry, J.A.S. 2003. Uplift resistance of buried submar-
ACKNOWLEDGEMENTS ine pipelines, comparison between centrifuge modelling
and full-scale tests. Geotechnique, 53: 877883.
Palmer, A.C. and Niedoroda, A.W. 2005. Ice gouging and
Andrew Palmer thanks the Jafar Foundation for con- pipelines: unresolved questions. Proceedings, 18th
tinued support. The authors thank several colleagues, International Conference on Port and Ocean Engineering
and in particular Ralf Peek, for helpful discussions. under Arctic Conditions, Potsdam, NY, in press.
Reed, C.W., Das, H. and Niedoroda, A.W. 2004. Application
of a predictive shoaling and migration model, M3D,
REFERENCES to St. Marys Entrance, Florida. In press. Proceedings,
ICCE conference.
Allersma, H.G.B. and Schoonbeek, I.S.S. 2005. Centrifuge Simo, J.C., Ju, J.W., Pister, K.S. and Taylor, R.L. 1988.
modelling of scouring ice keels in clay. Proceedings, Assessment of cap model: consistent return algorithms
15th International Offshore and Polar Engineering and rate-dependent extension. American Society of Civil
Conference, Seoul, Korea, in press. Engineers, Journal of Engineering Mechanics, 114:
Darwin, C.R. 1855. On the power of icebergs to make recti- 191218.
linear uniformly-directed grooves across a submarine Woodworth-Lynas, C.M.L., Nixon, J.D., Phillips, R. and
undulatory surface. London, Edinburgh and Dublin Palmer, A.C. 1996. Subgouge deformations and the
Philosophical Magazine, 10: 9698. security of Arctic marine pipelines. Proceedings, Twenty-
Greenwood, J.A. 1984. A unified theory of surface rough- eighth Annual Offshore Technology Conference,
ness. Proceedings of the Royal Society. series A, 393: 133. Houston, 4: 657664.

650

Copyright 2005 Taylor & Francis Group plc, London, UK


Model tests to simulate riser-soil interaction in touchdown point region

E.C. Clukey & L. Haustermans


BP, Houston & Baku Azerbaijan

R. Dyvik
Norwegian Geotechnical Institute, Oslo

ABSTRACT: An experimental testing program was performed to investigate the pipe-soil response under
loading conditions that can result in fatigue damage for deepwater risers. The tests were primarily performed on
152 mm pipe sections subjected to a series of load- and displacement-controlled cyclic loads and displacements.
The soil used in the tests was overconsolidated to try and replicate similarly consolidated soil and shear strengths
found at some deepwater sites in the Gulf of Mexico. The results are presented in terms of a non-dimensional-
ized stiffness ratio and compared with predictions from two different hyperbolic models. The results show that
for loads dominated by compressive loads, the measured stiffness ratios were reasonably predicted by the hyper-
bolic model that simulated unloading-reloading conditions. However, with additional and more robust loading
the soil stiffness ratios decreased appreciably to values below what either model would predict. The impact of
the separation of the pipe from the soil had a significant impact on the measured stiffnesses. Soil sensitivity,
resulting in the complete soil remolding was not sufficient to explain these reductions in soil stiffness. It
appears that soil-water mixing, caused by a combination of dilation and jetting action, were the primary causes
for this reduction. A continued understanding of the loads over the life of the riser will help to better resolve
issues not specifically addressed in this study.

1 INTRODUCTION based on the test results, of the major factors that


influence the pipe-soil response in the touchdown
To better understand the response of very soft to soft point region.
clayey soils to the long-term riser motions in a fatigue
analysis, an experimental testing program was per-
formed. Although the soils are considered soft because 2 SIGNIFICANCE OF SHORT & LONG TERM
they have little overburden, they were overconsolidated SOIL BEHAVIOR FOR RISER FATIGUE
with resulting mudline shear strengths of about 4 kPa.
These shear strengths were intended to replicate some The fatigue life of a Steel Catenary Riser (SCR) is
of the soils encountered in deepwater Gulf of Mexico highly dependent on the soil response in the region
(GoM), seaward of the Sigsbee Escarpment. This pro- where the riser pipe touches down on the seafloor
gram involved a series of tests aimed at determining the (TDP). The riser motions contributing to riser fatigue
soil response to both downward and upward vertical are quite complex, which likewise make the assess-
motion of a short pipe section. Seven separate tests were ment of the soil response complex. Generally, how-
performed. Within each test several different loading ever, it appears that the most important contribution
phases were performed with consolidation allowed to riser fatigue occurs with the downward vertical
between the various phases. Tests were performed motion of the pipe. This motion is most critical
either with load or displacement control. Separation because as the riser pipe penetrates into the TDP, the
of the pipe from the soil occurred in several tests. curvature and resulting moments increase.
This paper will begin with an overview of some The downward vertical response of the soil can,
of the key soils related issues that contribute to riser however, also be affected by the motions in the other
fatigue. This will be followed by a description of directions (upward vertical, lateral, and longitudinal)
the testing program performed to determine the soil and their resulting interaction effects. Pipe-soil separ-
response to riser loads. The results of the testing ation, re-penetration and trench formation are influ-
program are then presented along with a discussion, encing factors on the short term. These interaction

651

Copyright 2005 Taylor & Francis Group plc, London, UK


effects tend to produce a softer soil response which Riser TDP Displacements
6
increases the fatigue life.
If separation occurs, the soil has been loaded to a 5
Uplift
failure state which can cause significant cyclic damage
with repeated loading. The soil can potentially tend to 4

Displacements, ft
the remolded state and a reduced stiffness. Separation Note: 1ft = .305 m.
3
of the pipe from the seafloor is most likely where the
riser initially touches down on the seafloor. However, 2
further away from the initial TDP the soil response may Compression

be much different. For motions that do not exceed the 1

uplift (suction) capacity of the soil, pipe-soil separa-


0
tion will not occur. Further along the riser the cyclic 8150.0 8200.0 8250.0 8300.0 8350.0 8400.0 8450.0 8500.0 8550.0 8600.0
motions may not be sufficient to cause uplift forces -1
on the soil, resulting in even less cyclic damage. Distance from Hang-off pt, ft

The reloading process that occurs after separation


can be equally as significant as soil remolding. As the Figure 1. Deepwater riser displaced shapes.
pipe moves back toward the soil, the water underneath
the pipe is pushed downward. This jetting action can Time History of Force
lead to soil-water mixing, erosion and eventual trench 800
t = 3766 s
formation. As the pipe further penetrates into the soil, 700 t = 3908 s
the downward water velocities are further enhanced 600
t = 4627 s
t = 4818 s
since water flow is restricted. The mixing process can
Maximum force, lb/ft

t = 5166 s
500
reduce the soil strength and stiffness to less than the t = 5814 s

remolded values. 400

In addition to the potential impacts on the soil stiff- 300


ness, the trench formation can additionally influence 200
the fatigue life by causing reverse curvature of the riser. Note: 1 ft. = 0.305
100
This reverse curvature occurs as the pipe transitions
0
out of the trench back to the seafloor. 8150.0 8200.0 8250.0 8300.0 8350.0 8400.0 8450.0 8500.0 8550.0 8600.0
Soil response is also very history dependent and -100
Disatnce from hang-off point
can be significantly influenced by a number of ongoing
processes that occur over the design life of the riser. Figure 2. Time histories of maximum force.
Perhaps most important among these process would
be the consolidation effects from cyclic loading. Con-
solidation will tend to increase the soil stiffness, thereby For the portion of the pipe initially in contact with
reducing the fatigue life of the riser. However, if the the soil the upward displacements are about 75% of
soils are overconsolidated, they may tend to dilate and the pipe diameter. It is unlikely that the uplift resist-
become softer as water is added to the soil as a result ance of the soft soils would be sufficient to resist this
of the dilation process. type of movement and separation of the pipe from the
A second important long-term effect involves the seafloor would be expected. For this case the soil
riser response during very large loading events. These would likely be significantly reworked and ultimately
larger events can influence the soil response if they remolded as a result of cyclic loading.
are sufficient to relocate the riser to a different TDP Figure 2 show a series a force profiles for the same
location, thereby starting the entire loading process simulation. In this figure the maximum force changes
and corresponding soil response again. position over about a 30 m. region as a result of the
To illustrate some of the effects described above, loads incurred during this sea state. This shifting pos-
Figures 1 and 2 show positions and force profiles for ition for the maximum implies that the touchdown
a pipe in pipe (0.6 m OD) riser subjected to sea states point position is not static. Rather, it is dynamically
that cause significant fatigue damage. Figure 1 illus- shifting during the course of the loading event, mak-
trates the potential for separation from the displaced ing the separation more likely over a broad region.
configuration of a riser as a function of distance for Therefore, the soil curves throughout this broader
the riser hang-off point for a spar in over 1700 m. water region need to consider this potential impact on the
depth. In this simulation the soil is modeled with a overall riser fatigue response.
downward acting vertical spring. However, the upward The testing program designed for this study was
spring caused by suction is assumed to be negligible. especially aimed at investigating cyclic and pipesoil
The simulation shows that with uplift loading, signifi- separation effects on the pipe-soil response. Earlier
cant pipe displacements occur. efforts (Bridge et al. 2004) to characterize the pipe-soil

652

Copyright 2005 Taylor & Francis Group plc, London, UK


response have focused on using a hyperbolic relation- Shear strength (kPa)

ship to represent soil springs. Two different types of 0 2 4 6 8 10 12


0
models were proposed. The first model focuses on the Fall Cone
initial downward loading of the soil and assumes that
5
the peak load is achieved at a displacement of 10% of
pipe diameter. The second model assumes that the

Approximate depth (cm)


10 Hand held vanes
unload-reload behavior of the soil-pipe system should
Before
be used to characterize the soil springs. With this model Before
15
the displacement from the unload to reload position Before
Before
at the ultimate has been found to be 2.5% of the pipe Before
20 Before
diameter. During
Both these models, however, did not consider the 25
During
After
effects of cyclic loading and soil separation on the After
overall response. The results obtained from this study 30
After
After
will, therefore, be compared to the stiffnesses obtained After
from these previous models to determine the potential 35
After

impacts of these additional effects.


Figure 3. T-bar measurements, before, during and after
model tests.
3 TESTING PROCEDURES
results the cone was not sufficiently embedded to pre-
3.1 Bin preparation vent being impacted by surface effects until it had
The bin used in the model testing was 360 cm by penetrated about 5 cm.
175 cm with a depth of 60 cm. A kaolinite freshwater The shear strengths down to about 15 cm appear to
slurry was prepared and placed inside the bin to a be slightly less for the T-bar tests taken during and after
depth of 48.5 cm. A filter was then placed on top of the model testing. This reduction in shear strength
the clay, followed by a steel plate and a second filter. may be related to additional swelling of the clay, but
A membrane was then placed on top. also could be the result of the T-bar test being located
Vacuum was then applied to the bin through outlets only about one pipe diameter away from the model test
in the membrane. At the maximum the vacuum applied locations. Therefore, the soil in the upper 15 cm may
a stress to the clay of 85 to 89 kPa. Volume change have been slightly disturbed during these later tests.
measurements indicated that primary consolidation
was complete in about two weeks.
At the end of consolidation the membrane, filters 3.3 Model tests
and steel plate were removed and freshwater was placed A series of tests was performed to investigate the
on top of the clay. The clay was then allowed to swell stiffness of the pipe during vertical loading. The pro-
for a few weeks prior to model testing. gram consisted of seven tests (Table 1).
During testing there was sufficient water over the
clay to keep the pipe submerged. The turbid nature of
3.2 Shear strength
the water prevented detailed visual observations of
T-bar measurements in the corners and sides of the the tests. However, in some tests, there was obvious
bin were made to determine the soil shear strength pipe-soil separation, as suspended soil particles were
before testing. Measurements were also taken after observed being displaced out of the soil trench as the
Model Test 3, Model Test 7, at the conclusion of all pipe penetrated downwards.
the testing. The results of the T-bar tests are show on As shown on Table 1, the tests were either performed
Figure 3. The shear strengths were obtained by divid- in load or displacement control. The load-controlled
ing the T-bar force by 10.5 (Stewart and Randolph, tests had the advantage of free pipe movement, thereby
1994). Also shown on Figure 3 are fall cone (1 cm simulating more realistically the pipe embedment under
below mudline) and hand-held vane measurements field conditions. The disadvantage was that because
taken at 10 cm below the mudline. of the large displacements, the tests were terminated
The agreement between the different measure- shortly after the initial large excursions of the pipe to
ments is considered very good. The lower shear prevent damaging equipment and/or disturbing the
strengths in the top few centimeters obtained with the soil in the bin prior to subsequent tests. With dis-
T-bar are likely due to an overestimation of the bear- placement-controlled tests, the cycles could continue
ing capacity factor (10.5) at shallow depths. The T-bar for significant time after separation, although the pipe
was 20 mm in diameter by 125 mm long. Based on the was not free to move and embed deeper into the soil.

653

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 1. Testing Summary.

Test 1 Test 2 Test 3 Test 4 Test 5 Test 6 Test 7


Day 1 Pentr. 0.28D Pentr. 0.26D Pentr. 0.27D Pentr. 0.27D, Pentr. 0.50D, Pentr. 0.27D, Pentr. 0.26D
Set load to300 N Set load to Slow pullout
400 N
Pentr. 0.50D,
Slow pullout
Pentr. 0.50D
Overnight Depth 0.28D Depth 0.26D Depth 0.27D Load 300 N Load 400 N Depth 0.50D Depth 0.26D
Day 2 Displ. Load, N Load, N Load, N Load, N Displ. Load, N
1  1 mm 732  100 860  400 300  800 400  600 4  5 mm 1230  580
350  600 12  13 mm Cycles
Cycles Cycles Cycles Cycles 250  600 Load, N 500
1000 1000 1000 11* 300  650 200  300 Pentr. to 0.5D
Cycles 300  500
333 300  750
300 (2) Cycles
100 200
325
250 (3)
Overnight Depth 0.28D Depth 0.26D Depth 0.34D Load 300 N Load 300 Depth 1.06D Depth 0.50D
Day 3 Pentr. 0.50D Pentr. 0.50D Pentr. 0.50D Load to 400 N Load, N Depth 1.06D Displ.
400  750 10  13 mm
250  800 Load, N
200  900 580  1080
0  1100 Cycles
Set load 400  600 500
150 N 100  400 100
Cycles
300 Pullout at
175 cyclic rate
60
30*
5
50
Overnight Depth 0.50D Depth 0.50D Load 150 N Load 400 N Load 100 N Depth 1.06D
Day 4 Pullout slow Pullout slow Load, N Load, N Load, N Displ. FOOTNOTE:
150  400 400  750 100  400 2  5 mm * is failure by
Displ 10  13 mm model pipe
Cycles Cycles 7 8 mm 22  25 mm lifting off
1000 48* 24  25 mm clay
Cycles Load, N
Pullout at Pullout at 75 200  500
cyclic rate cyclic rate 100 200  750
150 200  800 N
200  900
Pentr. 1.32D 200  1000
200  1100
200  1200
Pullout slow 200  1300
200  1400
150  1400
Cycles
250 (3)
125
50 (8)
Pullout slow

654

Copyright 2005 Taylor & Francis Group plc, London, UK


Figure 4. Stiffness ratios for load controlled tests. Test designations are described in text.

4 RESULTS pipe diameter while the unload-reload stiffness is


achieved at 2.5%.
4.1 Load-controlled tests As discussed by Bridge et al. (2004) the maximum
values of the stiffness ratio for the initial loading and
The stiffness can be represented in a non-dimesional-
unload-reload models are 68 and 267 respectively.
ized form (stiffness ratio) by dividing the measured
These values correspond to the very small initial dis-
stiffness (cyclic load/cyclic displacement) per unit
placements in each case. The stiffness ratios at the peak
length by the bearing capacity factor, Nc, times the
loads are 10 and 40 for the corresponding models.
undrained shear strength, su. Su is obtained from T-bar
The data presented on Figure 4 shows a series of data
results (Figure 3) while the bearing capacity factor is
points for the various tests. The first number in
from Skempton (1951):
the series represents the test number (Table 1). The
second number represents the loading phase (Table 1)
while the third number represents the number of
cycles. If 1000 cycles were applied, then the sequen-
where ZD is the embedment depth and B is the width tial numbers (third number in designation) represent
of pipe in contact with the soil. When the pipe has 1, 100 and 1000 cycles respectively. If more than 100
penetrated a half pipe diameter or more, B equals D, cycles but less than 1000 cycles were applied then the
the pipe diameter. third number in the test designation represents 1, 10,
The stiffness ratios for the load-controlled tests are 100 and the final number of applied cycles respect-
shown on Figure 4. Also shown on Figure 4 are pre- ively. If less than 100 cycles were applied then the
dicted values of the stiffness ratio using the hyperbolic designated numbers represent 1, 10 and the final num-
models proposed by Audibert et al. (1984) and Bridge ber of cycles. Therefore, for example, test designation
et al. (2004). Both these models rely on a hyperbolic 5.2.2 represents Test 5, the second load controlled
fit. However, the model proposed by Audibert et al. loading phase (350 N  600 N, Table 1) and the secant
attempts to predict the secant stiffness from the initial stiffness ratio at 10 cycles.
downward loading, whereas the model proposed by The test results show that for the first cycle of load-
Bridge et al. uses the unload-reload secant stiffness. ing the unload-reload model reasonably predicts the
The initial loading model assumes that the peak load stiffness ratio. However, except for the initial phases
is achieved at a vertical displacement of 10% of the of Test 2 and 7, the stiffness ratios begins to significantly

655

Copyright 2005 Taylor & Francis Group plc, London, UK


decrease as a result of repeated cycles. Test 2 did not -1200 -600 0 600
Force (N)
1200 1800 2400 3000
incur any degradation to the stiffness ratio, probably -0.10

because the forces were strongly biased toward the 0.00


compressive portion of the cycle (732  100 N).

Displacement/Diameter (-)
0.10 Soil separation
Similarly, Test 7 actually had a slight increase in stiff-
0.20
ness ratio with number of cycles for loading largely Uncontrolled displacements

dominated by the compressive cycle (1230  580 N). 0.30

Neither of these tests had a substantial increase in the 0.40

embedment depth during the cycling. Therefore, the 0.50

lack of a reduction in the stiffness ratio can not be 0.60


attributed to deeper pipe embedment.
All the other load-controlled tests had significant Figure 5a. Force-displacement plot showing pipe-soil
reductions in the stiffness ratio as the tests proceeded. separation in Phase 1 of Test 4.
Tests 3, 4 and 5 all started with stiffness ratios about
equal to those predicted by the unload-reload model, Force (N)
but ended with stiffness ratios from slightly above to -800 -400 0 400 800 1200 1600 2000
significantly below those predicted with the initial 0.55

loading model. Test 6 was either slightly above or sig- 0.56


Uncontrolled displacement

Displacement/Diameter (-)
nificantly below the initial loading model for the 0.57 Final full
load-controlled tests. However, this test had signifi- 0.58
hysterisisloop

cant amounts of displacement-controlled loading


0.59
prior to the onset of load-controlled loading (Table 1).
The results from Test 4 illustrate the change in stiff- 0.60

ness as the pipe separates from the soil. In Figure 5a, the 0.61
pipe is initially embedded to about 0.40D. Over the first 0.62
ten loading cycles the stiffness ratio decreases steadily
from 76 to 13. On the 11th cycle the soil response Figure 5b. Force displacement plot showing pipe-soil
becomes more unstable and the hysteresis increases. separation in Phase 2 of Test 4.
During the following and final cycle, separation of the
pipe occurs. A dramatic change in the soil-pipe
response is observed as the force initially drops to zero The displacement-controlled tests were used to
before recovering as the pipe contacts the soil again. For observe in more detail the pipe-soil response for
this test the hydraulic control actuators were turned cycles subsequent to the soil separation. The displace-
off once this unstable condition was encountered. ment-controlled tests, however, were not able to prop-
A similar type of response is also shown on Figure 5b. erly model the additional pipe embedment during the
In this second loading phase of Test 4 the pipe is cyclic loading process and, therefore, the stiffnesses
embedded deeper into the soil, therefore requiring may be somewhat low because this additional pipe
about 45 cycles for separation. The loading for this penetration was not permitted.
test phase (400  750 N) was only slightly less than The small displacement tests (Test 1.1, Test 6.1 and
the loading (300  800 N) for the initial test phase Test 6.6) had stiffness ratios ranging from 38 to 9 for
shown on Figure 5a. The response for this test shows the first loading cycle. For Test 1.1.1, which had just
a very dramatic change in the pipe-soil behavior with 2 mm of displacement, the first cycle stiffness ratio
a significant reduction in the stiffness as the pipe was about half the predicted using the unload-reload
begins separation from the soil. hyperbolic model and equal to the stiffness ratio
using the initial loading model. However, the stiffness
ratio decreased significantly with cyclic loading. With
4.2 Displacement-controlled tests a 100 cycles of loading the range for the stiffness ratio
decreased to between 12 and 3.
A series of nine displacement-controlled test seque-
The intermediate displacement test (5.12) had an
nces were performed (Table 2). These tests can be
initial stiffness ratio of 4, but did not experience much
separated into the following:
additional cyclic degradation. At 100 cycles the stiff-
1. Small displacements between 2 mm to 10 mm, less ness ratio had only decreased to 3. However, prior to
than 1% of pipe diameter performing this test phase a significant number of
2. Intermediate displacement, 16 mm, approximately load-controlled tests had been performed. These load-
equal to 10% of pipe diameter controlled tests penetrated the pipe to embedment
3. Large displacements, 26 mm to 50 mm, displace- depths greater than the depth (0.12 m) at which this dis-
ments sufficient to cause pipe- soil separation placement-controlled tests was performed. Therefore,

656

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 2. Displacement-controlled test results.

Pipe Offset Cycl. 2-way Meas. stiffness Unload-reload Initial load


Test # # of cycles embed., m Load N ampl, mm ampl, mm ratio stiffness ratio stiffness ratio

1.1.1 1 0.043 1623


1 2 38 67 38
1.1.2 10 0.043 876
1 2 21 67 38
1.1.3 100 0.043 487
1 2 12 67 38
1.1.4 1000 0.043 422
1 2 10 67 38
5.12.1 1 0.122 1352
7 16 4 11 10
5.12.2 10 0.122 1311
7 16 4 11 10
5.12.3 100 0.122 1082
7 16 3 11 10
5.13.1 1 0.102 2500
24 50 2.2 4 3
5.13.2 10 0.102 875
24 50 0.8 4 3
5.13.3 100 0.102 400
24 50 0.4 4 3
5.13.4 153 0.102 375
24 50 0.3 4 3
6.1.1 1 0.073 1900
4 10 9 17 14
6.1.2 10 0.073 1178
4 10 5 17 14
6.1.3 100 0.073 722
4 10 3 17 14
6.1.4 200 0.073 665
4 10 3 17 14
6.2.1 1 0.066 1500
12 26 2.7 7 6
6.2.2 10 0.066 1200
12 26 2.1 7 6
6.2.3 100 0.066 225
12 26 0.4 7 6
6.2.4 320 0.066 180
12 26 0.3 7 6
6.6.1 1 0.158 4100
2 10 18 17 14
6.6.2 10 0.158 2378
2 10 10 17 14
6.6.3 100 0.158 1312
2 10 6 17 14
6.6.4 250 0.158 1107
2 10 5 17 14
6.7.1 1 0.152 2830
10 26 5 7 6
6.7.2 10 0.152 2604
10 26 4 7 6
6.7.3 100 0.152 1783
10 26 3 7 6
6.7.4 250 0.152 1472
10 26 2 7 6
6.8.1 1 0.138 2300
22 50 2.0 4 3
6.8.2 10 0.138 1725
22 50 1.5 4 3
6.8.3 100 0.138 437
22 50 0.4 4 3
6.8.4 250 0.138 345
22 50 0.3 4 3
7.2.1 1 0.101 6800
10 26 12.1 10 9
7.2.2 10 0.101 3128
10 26 5.5 10 9
7.2.3 100 0.101 884
10 26 1.6 10 9
7.2.4 500 0.101 272
10 26 0.5 10 9

most of the cyclic degradation likely occurred during Figure 6 illustrates the typical response observed
this previous cyclic loading. with large displacements. The top part of this figure
The large displacement-controlled tests were per- 6a, shows the time histories of displacement and force.
formed to examine the pipe-soil response during sep- The displacement is referenced to the axis on the right
aration the pipe and soil. Separation occurred in all while the force is referenced to the axis on the left.
but one of these tests. The one test where separation The bottom part of this figure 6b shows the force ver-
did not occur was Test 6.7 which had a displacement sus the normalized displacement. From this plot the
amplitude of 26 mm. The stiffness ratio for the first force appears to have approached zero (separation)
cycle of this test was only 5, suggesting the prior after about 100 to 200 cycles (frequency  0.1 Hz).
load- and displacement-controlled test had (Test 6.6) Once the pipe makes contact with the soil the response
already reduced the stiffness. At 250 cycles the stiffness is initially soft, but becomes progressively stiffer as
ratio was reduced to 2 with a continuing reduction. the pipe continues to penetrate into the soil.
When the displacement increased to 50 mm, separation There is a factor of 24 reduction in the stiffness
occurred in 1520 cycles. The stiffness ratio on the ratio from the first to last cycle in this test. If the
first cycles for all the large displacement tests ranged reduction in stiffness is the result of the remolding of
from 2 to 12. The stiffness ratio decreased to below 0.5 the soil, then the stiffness reduction would be limited
after 250 cycles of loading for all the tests that exper- by the soil sensitivity. For the kaolin used in the testing
ienced soil separation. the soil sensitivity would be about 1.5 to 2. Therefore,

657

Copyright 2005 Taylor & Francis Group plc, London, UK


10000 0.30 Force (N)
-4000 -3000 -2000 -1000 0 1000 2000 3000 4000
8000 0.40

Displacement/Diameter (-)
0.39
6000 0.50
0.41

Displacement/Diameter (-)
Force (N)

4000 0.60
0.43
2000 0.70
0.45
0 0.80
0.47
-2000 0.90
0.49
-4000 1.00
0.51
0 1000 2000 3000 4000 5000
Time (sec) 0.53

Figure 6a. Force and displacement time history for Test 7, Figure 6b. Force-displacement record for Test 7, with pipe-
with pipe soil separation. soil separation.

the stiffness reduction cannot be explained from predicted with a hyperbolic model that simulates
the soil sensitivity. This suggests the possibility of some initial loading and a displacement of 10% of pipe
soil mixing and the addition of water to the soil as a diameter at peak load.
result of the separation process. Two factors could have For the load-controlled tests, after a very signifi-
influenced the addition of water to the soil. First the soil cant amount of robust loading, the stiffness ratios
was prepared in an overconsolidated condition to decreased to values of less than 5. This is an order
achieve the shear strength profiles shown on Figure 1. of magnitude difference between the stiffness ratio
Overconsolidated soils tend to dilate and add water predicted (40) with the unload-reload model and a
during the dilation process. Secondly, the separation factor of 2 difference for the initial loading model.
of the pipe from the soil creates hydraulic jetting The displacement-controlled tests similarly showed
action as the pipe moves back toward the soil surface. stiffness ratios of 5 or less. Values lower than 1
This jetting action, as observed in the tests, causes were observed for tests that ultimately resulted in
scour and resuspension of the clay particles resulting pipe-soil separation.
in higher water content soils at the pipe-soil interface. For tests with pipe-soil separation, the characteris-
The potential for this erosional type of action was tic shape of the force-displacement curve changed
observed in Test 5, when the clay was observed to appreciably. It is unlikely that a hyperbolic model
have sufficiently scoured about a 10 cm cavity that can reasonably predict this behavior. Soil sensitiv-
extended to the bottom of the pipe. The combination ity cant account for the large reductions in the
of soil dilation and scouring for the tests with pipe- stiffness. It is likely that the reduction is related to
soil separation appears to have resulted in very large some mixing and addition of water to the soil.
and significant reductions in the soil stiffness.

REFERENCES
SUMMARY & CONCLUSIONS
Audibert, J.M.E., Nyman, D.J. ORourke, T.D. 1984.
A series of vertically loaded tests was performed to Differential Ground Movement Effects on Buried
investigate the soil response in the TDP region of a Pipelines. Chapter 5 Guidelines for the Design of Oil and
SCR. The soil used was prepared in an overconsoli- Gas Pipeline Systems, prepared by ASCE Council on
dated condition, to replicate some soils found in deep- Lifeline Earthquake Engineering.
water GoM. Tests were performed in both load- and Bridge, C., Laver, K., Clukey, E.C. and Evans, T.R. 2004.
displacement-controlled modes to investigate a range Steel Catenary Riser Touchdown Point Interaction Model.
of displacements. Special emphasis was placed on Proceedings Offshore Technology Conf., OTC16628,
Houston, May, 23p.
understanding the soil response under cyclic loading Fontaine, E, Nauroy, J.F., Foray, P., Roux, A. and Gueveneux, H.
and where the pipe separated from the soil. The pri- 2004. Pipe-Soil Interaction of Soft Kaolinite: Vertical
mary conclusions are: Stiffness and Damping. Proceedings 14th Intl. Offshore
and Polar Engineering Conf., Toulon, France, May,
For tests dominated by downward vertical forces, the
pp 517524.
soil stiffness was reasonably modeled with an unload- Skempton, A.W. 1951. The Bearing Capacity of Clays.
reload hyperbolic model, assuming the pipe displace- Proceedings Building Research Congress.
ment at peak load is 2.5% of the pipe diameter. Stewart, D.P. and Randolph, M.F. 1994. T-Bar Penetration
With more robust loading sequences, the soil stiff- Testing in Soft Clay, J. Geot. Eng. Div., ASCE, Vol. 120,
ness degraded to values closer to or below those No. 12, pp 22302235.

658

Copyright 2005 Taylor & Francis Group plc, London, UK


Capacity of piles in sand

Copyright 2005 Taylor & Francis Group plc, London, UK


Results from axial load tests on pipe piles in very dense sands: the
EURIPIDES JIP

H.J. Kolk & A.E. Baaijens


Fugro Engineers B.V., The Netherlands

P. Vergobbi
Fugro France SA, France

ABSTRACT: The paper presents test results from the EURIPIDES Joint Industry Project. This project had
the objective to obtain reliable soil and axial resistance data for offshore type test piles in sands in order to
improve offshore pile design criteria. Main soil data and detailed pile load test results from this testing pro-
gramme are presented.

1 INTRODUCTION Table 1. Soil stratigraphy.

An extensive load testing programme was conducted in Location I Location II


1995 on a 0.76 m outer diameter pipe pile in very dense Geotechnical Depth Depth
layer description bgl [m] bgl [m]
sands at Eemshaven, The Netherlands. A highly instru-
mented pile was driven and tested at one location, FINE SAND, medium 0.0 to 5.3 0.0 to 5.3
extracted and redriven and tested at a second location. dense to very dense
Static compression and tension tests were performed at FINE SAND, silty, 5.3 to 15.8 5.3 to 16.0
three penetration depths (30.5 m, 38.7 m and 47.0 m) at medium dense
the first location and at one penetration depth (46.7 m) Alternating layers 15.8 to 22.0 16.0 to 21.8
at the second location. The latter series of tests was of soft CLAY and
repeated after 1.5 years. loose sandy SILT and
The tests were performed to obtain reliable data FINE SAND
FINE to MEDIUM 22.0 to 24.7 21.8 to 25.0
on offshore type piles in sands in order to improve SAND, locally silty,
offshore pile design criteria. It was a Joint Industry medium dense
Project, initiated, managed and performed by a Joint FINE to MEDIUM 24.7 to 29.0 25.0 to 27.5
Venture of Fugro Engineers B.V. (The Netherlands) SAND, locally silty,
and Geodia S.A. (France). medium dense to
This paper presents main soil data and pile load test very dense
results from this testing program. Reference is made to FINE to MEDIUM 29.0 to 42.8 27.5 to 41.5
Zuidberg & Vergobbi (1996) for details on load testing SAND, slightly
equipment and instrumentation. to very silty,
very dense
SILT 42.8 to 43.4 41.5 to 42.8
FINE to MEDIUM 43.4 to 49.0 42.8 to 50.4
SAND, silty to very
2 SOIL CONDITIONS
silty, very dense
Sands of interest for this program lie 22 m below
ground level (bgl) and consist of Pleistocene materials, Soil conditions at the two test locations were defined
which have been subjected to two glaciations. These by 6 Cone Penetration Tests (CPT) and 1 borehole (BH)
glaciations resulted in sands being overconsolidated at the first location and by 3 CPTs and 2 BHs at the
similar to many offshore areas in the southern part of second location. These conditions are summarised in
the North Sea. Table 1.

661

Copyright 2005 Taylor & Francis Group plc, London, UK


Cone resistance [MPa] Cone resistance [MPa]
0 20 40 60 80 100 0 20 40 60 80 100
20 20

25 25 Upper Bound
Upper Bound
Average
30 30.5 m 30

Depth [m]
Depth [m]

35 35
38.7 m Average
40 40

45 45
47.0 m 46.7 m
50 50
Lower Bound
55 55 Lower Bound

60
Location I Location II
65

Figure 1. Cone resistance data at test locations.

Made ground and Holocene sands with an average


CPT cone resistance of about 5 MPa were found from E
ground level to about 22 m depth. These sands were of
no interest and the pile load test program was designed
to provide pile-soil interaction data on the test pile TPC
below 22 m bgl.
115 mm
N S
CPT cone resistances were in the order of 40 MPa
to 80 MPa in the zone of interest for these pile load
tests (i.e. the Pleistocene sands from 22 m to 50 m
below ground level, Fig. 1).
Ring shear tests, using steel of roughness Rmax W
25 m similar to the test pile, revealed an interface Position of Channels
NSEW seen from toe of pile upwards 35 mm
friction angle in the order of 31 degrees in the fine to Channel Dimensions
medium sands above 44 m depth and about 27 degrees
in the medium to coarse sands below 44 m.
26 mm

Pile OD 762 mm
wt 36 mm

3 TEST PILE AND INSTRUMENTATION 115 mm

The test pile had an outer diameter of 0.76 m. The lower Figure 2. Cross-section of instrumented pile section.
27 m of the pile was highly instrumented and had a
wall thickness of 36 mm. The instrumentation in this
section was to measure the soil-pile interaction in the each other. Two (opposite) channels were on the out-
sands of interest, i.e. below 22 m depth. The add-on side and two were on the inside (Fig. 2).
section above this instrumented section had no instru- Instrumentation was mounted at various levels along
mentation (except near the pile head) and had a wall the instrumented section with a spacing of 0.5 pile
thickness of 42 mm. diameter near the pile toe increasing to 4 pile diameters
The instrumented section was provided with four at the top of the instrumented section. The amount of
instrumentation cable channels 90 degrees apart from instrumentation near the pile toe was enhanced in order

662

Copyright 2005 Taylor & Francis Group plc, London, UK


Details on design and performance of this test rig
Add-on or Top Section are provided by Zuidberg and Vergobbi (1996).

Upper Table
5 PILE INSTALLATION AND REMOVAL
Jacks

The pile was driven to subsequent testing depths using


Lower Table
an IHC S-90 hydraulic hammer with a rated energy of
90 kJ. Strain and acceleration in the pile head was
measured during driving together with measurement
of soil level inside the pile.
After completion of the third (deepest) test at loca-
Reaction Piles
tion I, the pile was extracted using the jacking system
Test Pile
and a crane. Subsequently, the pile instrumentation was
Bottom Frame checked at the roughness of the pile wall (both inside
and outside) was measured.
Figure 3. Load test rig. Comparison of the roughness measurements before
installation and after pile extraction revealed that:
to determine wall end bearing and pile wall inner and
outer friction. the outer wall roughness had reduced from approxi-
The key instrumentation included: mately 32 m to about 12 m;
the inner wall roughness had increased from
axial and limited tangential and shear strain gauges approximately 17 m to about 27 m.
total pressure cells
pore pressure cells The reduction in roughness of the outer surface is
toe load cells. attributed to abrasion of the outer pile wall by sand
during pile installation and removal. The increase in
Details on design and performance of these (and roughness of the inner surface is attributed to the long
other) instruments are given by Zuidberg & Vergobbi time (three weeks) between pile retrieval and rough-
(1996). ness measurement. The inner pile wall remained wet
The instrumented pile section was subjected to for a long time prior to and after soil plug removal,
repeated axial loading after the instruments were during which time the inner pile wall has probably
mounted and hooked up to a data acquisition system. rusted significantly.
This served two purposes: Upon retrieval a cemented crust was found on the
outer surface of the pile up to 16 m from the pile tip.
removal of residual stresses in the test pile; This crust consisted for 70% of quartz and some other
calibration of both pile and instruments. particles in a dense matrix of iron-hydroxide with a
laminated texture (30%).
Of importance was that during plug removal a 0.60
4 LOADING SYSTEM to 0.65 m thick disk of wood was encountered at the
first pile load test location. It is our best estimate that
A special test rig was designed and fabricated in order this was cut from a horizontal tree trunk lying at about
to facilitate changing over from compression to ten- 22 m depth. Its strength/toughness was similar to that
sion testing (and vice versa) as well as changing over of fresh hard wood.
from the testing mode to the pile installation/removal Check of the instrumentation read-outs after retrieval
mode (Fig. 3). proved that negligible zero shifts occurred. This con-
The reaction system of this test rig consisted of six firmed the success of the pre-installation pile loading
0.76 m diameter, 35 m penetration pipe piles, equally and calibration effort.
spaced at six metres distance from the test pile. Load Unfortunately, significant parts of the instrumenta-
was applied from the rig to a force ring on the add- tion channels and instruments near the pile toe were
on pile by three double acting hydraulic jacks. Each broken off probably because of high pile toe shear
jack had an effective stroke of 1 metre and a capacity stresses during driving.
of 10 MN in compression and 5 MN in tension. Following repair of part of the instrumentation, the
The force ring for each pile load test was left on the pile was driven to 46.7 m penetration at the second
add-on pile. Thus, the soil resistance on this section location whereafter pile load tests were performed
during the three deeper tests (38.7 m, 47.0 m and 6 days after installation. Another series of pile load tests
46.7 m) was affected by these force rings. was performed 533 days thereafter.

663

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 2. Overview of pile load tests.

Test Penetration Set-up period


no. depth [m] Sequence Date [days]

Location I I30.5 30.45 C-T April 1995 7


I38.7 38.70 T-R-C-T April 1995 2
I47.0 46.95 C-T-R-C May 1995 12
Location II II46.7 46.65 C-T-R-C June 1995 6
Location II R47.0 46.95 C-T January 1997 533
retest

Table 3. Summary of test results.

Tension at maximum Compression at 0.1 D At maximum Maximum toe


Test no. displacement [MN] toe displacement [MN] displacement [MN] displacement [mm]

I-30.5 1.7 7.3 10.6 242


I-38.7 8.4 12.5 15.5 225
I-47.0 12.5 19.1 22.7 208
II-46.7 9.7 18.8 20.5 143
R-47.0 15.9 30.3 4

6 PILE LOAD TESTS Typical test results are presented in Figures 4


through 6. The following comments apply to these
Pile load tests at each penetration consisted of an alter- graphs:
nating series of compression (C) and tension (T) tests.
Figure 4 The curve labelled 22 m is the axial force
If a compression test was preceded by a tension test,
derived from the axial strain gauges closest to this
the pile was first repositioned to the depth prior to the
level. The curve labelled toe is the axial force com-
tension test by means of a reloading test (R) to obtain
puted from axial (and other) strain gauges closest to
a virgin compression load-displacement curve.
the pile toe (i.e. 0.38 m above the pile tip), since the
A constant rate of pile head displacement of
toe load cells were damaged during pile driving.
1 mm/min was applied. The compression tests were
Figure 5 The axial loads were computed from
generally performed to a pile head displacement in
axial strain gauges and tangential and/or shear
the order of 250 mm (0.3 D) and the tension tests to
strain gauges (if available). Axial forces deter-
a pile head displacement of about 75 mm (0.1 D).
mined from combined axial and tangential (or
The retest at location II was performed to smaller pile
shear) strain gauge data are more accurate than
head displacements since the maximum capacity of
from axial strain gauge data only, particularly near
the loading system was reached both in tension and in
the pile tip. Unfortunately, limited tangential and
compression. An overview of the tests is given in
shear gauge data were available, partially since
Table 2.
these were not installed and partially because of
increased damage to the instrumentation channels
with increased pile driving resistance. A larger
7 TEST RESULTS source of uncertainty on axial load data near the
pile toe is that often no proper correction for pile
The resistance at the end of the tension tests was vir- bending effects could be made because strain gauge
tually constant, indicating that unit friction values data were only available on one side of the pile. In
along the test pile had reached constant values. summary, it is estimated that axial load data at 3 m
However, a significant increase in resistance per unit (4 D) or more above the pile toe are accurate but
of pile head displacement was still observed at the that the accuracy reduces nearer the pile tip.
end of the compression tests. This is primarily attrib- Figure 6 The total unit skin friction was deter-
uted to end bearing still increasing with increasing mined by taking the difference in axial load at two
pile toe displacement. subsequent strain gauge levels and dividing this by
The maximum resistances at about 22 m depth in the outer pile wall area. The resulting unit friction
the first tension and the first compression test of all value is plotted between these two strain gauge levels.
5 test series are summarised in Table 3. This value is probably in the order of the average

664

Copyright 2005 Taylor & Francis Group plc, London, UK


Location I -Test 30.5 m (C-T) Location II -Test 46.7 m (C-T-C)
14 30
12 pile head 25
pile head
10 20
8 22 m bgl 15 22 m bgl
Force [MN]

Force [MN]
6 10
4 5
2 pile tip 0
50 0 -50 --100 -150 --200 -250 -300 -350 -400
0 -5
0 -50 -100 -150 -200 -250 -300
-2 -10
-4 -15
Pile head displacement [mm] Pile head displacement [mm]

Location I -Test 38.7 m (T-C-T) Location II -Retest 46.7 m (C-T)


20 40
pile head
15
30
10 22 m bgl
20 pile head
Force [MN]

5
Force [MN]

pile tip 10
0
100 50 0 -50 -100 -150 -200 -250 -300
0
-5 100 50 0 -50 -100
-10 -10

-15 -20
Pile head displacement [mm]
Pile head displacement [mm]
Location I -Test 47.0 m (C-T-C)
30
pile head
25
20
15 22 m bgl
pile tip
Force [MN]

10
5
0
0 -50 -100 -150 -200 -250 300 350 -400
-5
-10
-15
-20
Pile head displacement [mm]

Figure 4. Force at pile head, 22 m bgl and at toe versus pile head displacement.

outer friction at levels in excess of 3 m (4 D) above significant, particularly during compression loading.
the pile toe since it is unlikely that there is significant Hence, the average unit friction on the inner and
friction between the soil plug and the inner pile wall. the outer wall of the pile near the pile toe may be in
However, inner wall friction nearer to the pile toe is between the presented values and half these values.

665

Copyright 2005 Taylor & Francis Group plc, London, UK


Axial Load [MN]
-20 -15 -10 -5 0 5 10 15 20 25 30 35
20

I - 30.5 I -30.5
25
I - 47.0

I - 38.7 30
Depth [m]

I - 38.7 R - 47.0

35
R - 47.0 II - 46.7
II - 46.7 40

45
I - 47.0

50
Tension Tests Compression Tests
(10% OD tip displacement) (10% OD tip displacement)

Figure 5. Distribution of axial load in pile for 0.1 D pile toe displacement.

Total Unit Skin Friction [kPa]


-1500 -1000 -500 0 500 1000 1500
20
Location I

25
I - 30.5
I - 30.5
30
Depth [m]

I - 38.7
35
I - 38.7

40

45 I - 47.0
I - 47.0

50

-1500 -1000 -500 0 500 1000 1500


20
Location II
II - 46.7
25

30 R - 47.0
Depth [m]

35
R - 47.0
40
II - 46.7
45

50
Tension Tests Compression Tests

Figure 6. Distribution of total unit skin friction for 0.1 D pile toe displacement.

666

Copyright 2005 Taylor & Francis Group plc, London, UK


8 CONCLUSIONS Community DG XVII Thermie Programme and the
following Participants and Sponsors:
The following conclusions can be drawn from the
EURIPIDES pile load test program: Participants:
Agip
A wealth of high quality soil and pile instrumenta-
Amoco
tion data has been obtained.
Elf
Four instrumentation channels are essential in view
Health and Safety Executive UK
of required data redundancy and/or correction for
Institut Francais du Petrole
bending moments.
Ifremer
The outer surface of the pile has become smoother,
NAM (Dutch combine of Shell and Exxon)
probably as a result of abrasion by sand during pile
Saudi Aramco
driving. Conversely, the inner surface has become
Geodia
rougher. However, this is probably an artefact result-
Fugro
ing from rusting during the long periods between
pile retrieval, plug removal and measuring pile
Sponsors:
roughness.
The Netherlands Ministry of Economics Affairs
Axial strain gauges need to be complemented by
The French Fonds de Soutien des Hydrocarbures
tangential strain gauges for an accurate assessment
Mobil Netherlands.
of axial load in a pile.
Stress conditions at a pile toe are highly complex
The management of Fugro is thanked for their
and cannot be assessed from strain gauge data only.
financial support and management guidance to bring
Internal and external radial pressure data facilitate
the project to such a good end and for their permission
interpretation of test results although interpretation
to publish this paper.
of all data remains complex.
Unfortunately a tree trunk was encountered at the
first pile load test location. It is highly uncertain if
this trunk has affected the subsequent pile load REFERENCE
tests at this location.
Zuidberg, H.M. & Vergobbi, P. 1996. EURIPIDES, Load
Tests on Large Driven Piles in Dense Silica Sands. 28th
ACKNOWLEDGEMENTS Offshore Technology Conference, Houston, Texas, 1(7977):
193206.
The pile test programme has been made possible by
a grant from the Commission of the European

667

Copyright 2005 Taylor & Francis Group plc, London, UK


CPT-based design method for steel pipe piles driven in very dense silica
sands compared to the Euripides pile load test results

P.Y. Foray
University of Grenoble, Grenoble, France

J.-L. Colliat
Total Exploration Production, Pau, France

ABSTRACT: A CPT design method, based on a modification of the original UWA and IC-MTD methods, is
compared to the results of the large-scale Euripides pile load tests, representative of the very dense silica sands
of the southern North Sea area. The comparison with the currently applied API RP 2A design model shows that
the recommended criterion with a maximum limiting friction resistance is the main cause for the conservative
bearing capacities calculated in very dense silica sands. The application of the proposed design method is illus-
trated with a case history from the Dutch sector of the North Sea.

1 INTRODUCTION North Sea area shows that there are important impli-
cations, in particular regarding the life extension of a
For about two decades, the API RP 2A design method large number of existing platforms.
for axially loaded piles driven in silica sands has been
the subject of large debate within the offshore geo-
2 MODEL PILE TESTS IN LABORATORY
technical specialists. It is now widely accepted that the
API RP 2A criterion, relatively unreliable due to a large
An extensive experimental program has been carried
scatter in the database of load tests generally per-
out at the INPG-L3S laboratory at the University of
formed on small-scale piles, is conservative for piles
Grenoble, with the objective to better understand the
driven into very dense silica sands. It is also recog-
behavior of steel pipe piles driven into dense normally
nized that less conservative pile capacities can be cal-
consolidated (NC) and over-consolidated (OC) sands.
culated by direct correlation of the pile shaft friction
It included the performance of tension and compression
resistance and end bearing capacity with the in-situ
pile load tests on model-scale piles (70 mm diameter)
cone resistance from Cone Penetration Tests (CPTs).
in a large calibration chamber. The research program,
The CPT design method presented in the paper,
focusing on direct correlation of pile end bearing and
is based on a modification of the original UWA and
shaft friction resistance with the CPT cone resistance,
IC-MTD design methods (Randolph et al. 1994,
has allowed the identification of the basic parameters
Jardine & Chow 1996), completed by extensive test-
for a new design method, in particular (Foray et al.
ing of model piles driven into dense silica sands in the
1998):
INPG-L3S calibration chamber at the University of
Grenoble. This method, compared to the Euripides End bearing: Open-ended steel pipe piles, typically
pile test results, is applicable to steel pipe piles of 0.75 m to 1.5 m in diameter, generally coring dur-
usual size (i.e. 0.75 m to 1.5 m of pile diameter) ing driving, behave as partially to fully plugged
driven into very dense silica sands, as those encoun- (or closed-ended) during static load testing, and
tered in the southern North Sea area. The comparison pile end bearing capacity continuously increases
with the currently applied API design criterion shows with pile toe displacement for a range of practical
that the RP 2A maximum limiting friction resistance deformations
is the main cause for the conservative bearing capaci- Shaft friction resistance: High peak friction resist-
ties calculated for steel pipe piles driven into very dense ance is observed near the pile toe, with exponential
silica sands. The application of the proposed design decay behind the pile tip. This behavior, contra-
method to a typical case study from the southern dicting the API RP 2A model with a maximum

669

Copyright 2005 Taylor & Francis Group plc, London, UK


constant limiting friction value, is in general agree-
ment with the alternative design methods proposed
by both UWA (Randolph et al. 1994) and IC-MTD
(Jardine & Chow 1996).
In terms of ultimate pile capacity, the conservatism
of the API RP 2A method in very dense silica sands
has been discussed by a number of authors, but the
comprehensive definition of a new design method
needs to be validated by comparison to the results of
large-scale pile load tests.

3 EURIPIDES PILE LOAD TESTS

In 1993, the Euripides JIP was launched by Fugro


Engineers in Holland and Geodia in France, including
the performance of an extensive pile load testing pro-
gram at an onshore site in the Netherlands. In 1995,
compression and tension load tests were carried out on
a highly instrumented 0.762 m (30 inch) diameter steel
pipe pile driven into very dense silica sands similar to
those encountered in the southern North Sea. A total of
twelve tests were performed at two locations (18 m
apart) and three penetration depths at 30.5 m, 38.7 m
and 47 m. Figure 1 shows the typical CPT cone resist-
ance profile at the Euripides test site, with the three pile
load test depths. One can see that dense to very dense
sands occur from about 25 m below ground level, and
the CPT profile shows qc values varying between 50
Figure 1. Euripides CPT profile and pile test depths.
and 85 MPa beyond 30 m depth. Alternative compres-
sion and tension load tests were performed at each test
depth (Zuidberg & Vergobbi 1996, Kolk et al. 2005).
With the Euripides pile load test results, recently In dense NC sands:
released in the public domain (Fugro Engineers 2004, f(v) 
0.001v  1.2, with fmin  0.7
Kolk et al. 2005), the offshore industry has access to a(D)  0.5D  0.2
accurate measurements allowing the key points of In dense OC sands:
debate regarding the ultimate capacity of piles driven f(v)  0.63
into dense silica sands to be addressed, including in a(D)  0.5D  0.1
particular (Kolk et al. 2005):
The ultimate capacity of the pile is defined at a pile
Distribution of friction resistance along the pile head displacement equal to 10% of the pile diameter
shaft, and variation with pile penetration (i.e. w/D  0.1). One should notice that this criterion
Measurement of end bearing capacity with pile toe is slightly more conservative than defining the ultim-
displacement ate pile capacity at 0.1D of pile toe displacement.
Correlation with in-situ CPT measurements. The same criterion is also easier and more practical to
apply when dealing with the interpretation of pile
4 PROPOSED DESIGN METHOD load tests.
The effect of in-situ vertical and horizontal stresses
4.1 End bearing resistance v and h on the pile end bearing capacity is not
defined explicitly but considered as being included
The pile end bearing resistance qb is given by: into the CPT cone resistance qc.
In the North Sea area, very dense silica sands have
(1) been over-consolidated by the weight of glaciers. It has
been shown that the CPT cone resistance measured into
where qc  CPT cone resistance (kPa); v  vertical such OC sands will be higher than that measured into
effective stress at pile toe (kPa); w  pile head dis- NC sands, and piles driven into OC sands will exhibit a
placement (m); and D  pile diameter (m). significantly higher driving resistance and ultimate

670

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 1. Classification of dense NC and OC silica sands. 25

Relative density DR
Sand type (%) Cone resistance qc
20
Dense NC 6585 20 MPa at 20 m
30 MPa at 60 m

Pile head or tip load (MN)


Euripides pile head load
Dense OC 6585 30 MPa at 20 m Euripides pile tip load
INPG-L3S pile head load
45 MPa at 60 m INPG-L3S pile tip load
15
Very dense NC 85100 35 MPa at 20 m API RP 2A pile head load
45 MPa at 60 m API RP 2A pile tip load
IC-MTD pile head load
Very dense OC 85100 45 MPa at 20 m IC-MTD pile tip load
60 MPa at 40 m 10

loading capacity (Foray et al. 1998). This behavior,


related to the in-situ effective stresses, and more partic- 5
ularly to the horizontal stress h, can not be described
properly by the relative density of the sand DR, as pro-
posed in the API RP 2A design method. Table 1 gives a
proposal for a classification of dense NC and OC silica 0
sands on the basis of the in-situ CPT cone resistance qc. 0 50 100 150 200 250 300
For typical pile sizes, the pile end bearing capacity is Pile head displacement (mm)
calculated over the full pile tip area, with the pipe pile
considered as being closed-ended (plugged behav- Figure 2. Euripides compression load test at 47 m at location
iour). For large piles with D larger than 1.5 m, one 1. Pile head and pile tip load-displacement curves.
should consider a reduction of the base capacity under
static compression loading, as proposed in the IC-MTD
An exponential decay of the friction resistance
design method for example (Jardine & Chow 1996).
along the pile shaft is defined, in accordance with the
As shown by the Euripides pile tests, one can
UWA and IC-MTD models. The friction resistance
hardly differentiate between end bearing and shaft
under tension loading is taken as 85% of the friction
friction near the pile toe. The friction resistance along
under compression loading.
the bottom 1.3D of pile length is then included into
As shown in Figures 2 and 3, the complete Euripides
the pile end bearing capacity.
pile load test curves are modeled by the INPG-L3S
method. However, this is beyond the scope of the
4.2 Shaft friction resistance present paper and not detailed further.
The maximum peak shaft friction resistance &max(z),
defined at 1.3D above the pile toe, is given by: 5 COMPARISON WITH THE EURIPIDES
PILE TESTS
(2)
The objective of having three pile loading depths at
(L)  0.21-0.003L, with min  0.06 the first Euripides test location was to measure the
1/  0.012 variation in friction resistance with pile penetration,
in order to check the friction decay models proposed
&max(z) in tension  0.85&max(z) in compression
by both UWA and IC-MTD. This could then be com-
pared with the results from the second Euripides test
where L  total embedded pile length (m); and location where the pile was driven to the deepest test
z  depth under consideration along the pile shaft (m). penetration at once.
The effect of the in-situ horizontal stress h on the The INPG-L3S design method is compared to the
shaft friction capacity is not defined explicitly but Euripides results from the compression and tension
considered as being included into the CPT cone pile load tests carried out at 38.7 m and 47 m at the
resistance qc. For typical pile sizes, with D ranging two test locations.
between 0.75 m and 1.5 m, the small change in hori-
zontal effective stress along the pile shaft is neglected, 5.1 Compression pile load-displacement curves
i.e. the dilation at the soil-pile interface under pile
loading in very dense sands (as proposed in the The load-displacement curves from the two compres-
IC-MTD model) is not taken into account. sion load tests performed at 47 m depth at the two test

671

Copyright 2005 Taylor & Francis Group plc, London, UK


30 0
0 200 400 600 800
-5
Euripides friction 30.5 m test
25 Euripides friction 38.7 m test
-10 Euripides friction 47 m test
INPG-L3S friction 38.7 m test
Pile head or tip load (MN)

20 -15 INPG-L3S friction 47 m test

Penetration depth (m)


IC-MTD friction 47 m test
Euripides pile head load
Euripides pile tip load API RP 2A friction profile
INPG-L3S pile head load
-20
INPG-L3S pile tip load
15 API RP 2A pile head load
API RP 2A pile tip load -25
IC-MTD pile head load
IC-MTD pile tip load
-30
10
-35

5 -40

-45
0
0 50 100 150 200 250 300 -50
Pile head displacement (mm) Unit shaft friction (kPa)

Figure 3. Euripides compression load test at 47 m at Figure 4. Euripides compression load tests at location 1.
location 2. Pile head and pile tip load-displacement curves. Friction resistance profiles versus depth.

locations are given in Figures 2 and 3. In each of 1 (Figure 4) and one single deepest test depth at
Figure 2 and 3 are given: location 2 (Figure 5)
The same friction profiles, as calculated with both
The pile tip and pile head loads as a function of pile the INPG-L3S and IC-MTD models.
displacement, measured at the two Euripides test
As shown in Figures 4 and 5, the exponential distri-
locations
bution of the friction resistance along the pile shaft
The same pile tip and head load curves, as calcu-
yields an extremely high peak friction near the pile toe
lated with the proposed INPG-L3S model
(up to more than 600 kPa) and a rapid decay in friction
The ultimate pile capacities at a displacement of
resistance behind the pile tip with increased pile pene-
0.1D (76 mm), as calculated with both the API RP
tration. The trend of the Euripides pile test results is in
2A and IC-MTD methods.
general agreement with both the INPG-L3S and
A much stiffer response of the pile tip load curve is IC-MTD design methods, but is completely different
observed at the second Euripides test location (Fig. 3), from the API RP 2A criterion which utilizes a constant
which is actually not captured by the proposed INPG- limiting friction resistance equal to 115 kPa. As a key
L3S model. One key result observed is that the meas- result, the shaft friction capacity calculated with the
ured pile capacity is about 40% higher than that API RP 2A model is about only 50% of that obtained
calculated with the API RP 2A, but the measured end from the Euripides pile load test results.
bearing capacity is close to that given by the API RP
2A when applying 12 MPa of base resistance. Both the
5.3 Tension load tests
INPG-L3S and IC-MTD design methods are in general
agreement with the Euripides pile load test results. Similar results are obtained with the Euripides tension
load tests, with the only difference that the shaft friction
5.2 Compression friction resistance curves capacity under tension loading is about 15% lower
than that measured under compression loading (see
The friction resistance profiles from the same com- Equation 2). This is attributed to principle stress rota-
pression pile load tests are shown in Figures 4 and 5. tion effects, radial contraction/dilation of the pile-soil
In each of Figure 4 and 5 are given: interface during pile loading, and/or effects of residual
The friction profiles, as measured at the two Euripi- stresses generated by driving (Jardine & Chow 1996,
des test locations, with three test depths at location Foray et al. 1998).

672

Copyright 2005 Taylor & Francis Group plc, London, UK


0 0
0 200 400 600 800 0 10 20 30 40 50
-5
-5 INPG-L3S Compression capacity
Euripides friction 47 m test
IC-MTD Compression capacity
-10
INPG-L3S friction 47 m test API RP 2A Compression capacity
-10
IC-MTD friction 47 m test -15
API RP 2A friction profile

Penetration depth (m)


-15
-20
Penetration depth (m)

-20 -25

-25 -30

-35
-30
-40
-35
-45
-40
-50

-45 -55
Compression pile capacity (MN)
-50
Unit shaft friction (kPa)
Figure 6. North Sea case study. Compression pile capacities.

Figure 5. Euripides compression load test at 47 m at location


2. Friction resistance profiles versus depth. counts were predicted with the MRBS 8000 hammer,
and pile driving monitoring was used to control the
final driving conditions and pile acceptance criteria.
6 SOUTHERN NORTH SEA CASE STUDY For three of the four piles, the blow counts reached
the driving refusal criteria (defined at 250 blows for
6.1 Site conditions 250 mm of pile penetration) between 45 m and 49 m
of penetration depth when the piles were driven into
In order to illustrate the implications for design, a case the very dense OC sands. At this time, the MRBS
study is presented, where the use of both the INPG- 8000 hammer worked with a global efficiency ran-
L3S and IC-MTD methods is compared with the cur- ging between 50 and 60% (defined as the ratio of the
rently applied API RP 2A model for a jacket platform driving energy transmitted to the pile, calculated from
from the Dutch sector of the North Sea. the impact stresses measured by the pile monitoring
Typically, the soil conditions at this southern North system, to the rated hammer energy). For these three
Sea site are composed of medium dense (qc  10 to piles, driving operations had to be resumed with the
25 MPa) to dense (qc  30 to 40 MPa) fine silica sands MRBS 12500 hammer to reach the target penetration.
in the top 35 meters, becoming very dense (qc  50 to
70 MPa) fine to medium silica sands beyond that
depth (i.e. OC sands according to the classification 6.2 Pile capacities
given in Table 1). The calculated pile capacities are given in Figure 6
The platform under consideration is a four-legged (design condition under compression loading for two
steel jacket, located in about 40 m of water depth, piles) and Figure 7 (design condition under tension
founded on four 1.22 m (48 inch) diameter piles. The loading for the two other piles), showing the pile
design penetration of the piles was equal to: capacities calculated with the API RP 2A model, the
49 m for two piles under storm compression load- IC-MTD method, and the proposed INPG-L3S method.
ing condition with a design capacity of 35 MN From the API RP 2A capacities given in Figures 6
51 m for two piles under storm tension loading and 7, the design penetrations of the piles should have
condition with a design capacity of 18.5 MN. been equal to 54 m (for the two piles with 35 MN of
design compression load) and about 60 m (for the two
Pile driving was performed with Menck steam other piles with 18.5 MN of design tension load). On
hammers MRBS 8000 (ram weight 80 tons, rated driv- the basis of some earlier model pile tests, less conser-
ing energy 1200 kNm) and MRBS 12500 (ram weight vative parameters were applied in the API model, using
125 tons, rated driving energy 2150 kNm). High blow maximum limiting values of 135 kPa and 15 MPa for

673

Copyright 2005 Taylor & Francis Group plc, London, UK


0 of the friction capacity with time (Fugro Engineers
0 10 20 30 40 50 2004). In addition to the conservatism of the originally
-5 applied API RP 2A design model, a complete re-assess-
INPG-L3S Tension capacity

-10
IC-MTD Tension capacity ment of the driven piles from this case study jacket
API RP 2A Tension capacity would require that other aspects related to the pile
-15 capacities are taken into account, such as the increase in
Penetration depth (m)

shaft capacity with time and the possible detrimental


-20 effects from cyclic loading.
-25

-30
7 CONCLUSION
-35

-40
A relatively simple CPT-based design method, applic-
able to steel pipe piles driven into very dense silica
-45 sands, is proposed. This design method is compared
to the results of the Euripides pile load tests, which have
-50 shown much larger ultimate pile capacities than cal-
-55 culated with the currently applied API RP 2A model,
Tension pile capacity (MN) with two key features:
End bearing capacity continuously increases with
Figure 7. North Sea case study. Tension pile capacities. pile toe displacement, for a range of practical tip
deformations, but the ultimate end bearing resist-
ance, defined at 0.1D of pile head displacement, is
the shaft friction and end bearing resistances, respect- close to that given by the API RP 2A model
ively (versus 115 kPa and 12 MPa in the original API Friction resistance is distributed exponentially
model). This allowed the acceptance of reduced pile along the pile shaft, with peak friction values at the
penetrations of 49 m (compression piles) and 51 m pile toe (600 kPa) which are much higher than
(tension piles). the maximum limiting friction proposed in the API
For comparison, the ultimate pile capacities, defined RP 2A document (i.e. 115 kPa).
at a pile head displacement ratio w/D  0.1, calcu-
lated with both the INPG-L3S and IC-MTD methods
are given in Figures 6 and 7 for a pile penetration of
50 metres. One can see that the actual ultimate pile ACKNOWLEDGEMENTS
capacities could be:
The definition of the proposed INPG-L3S pile design
About 25% higher under compression loading, method was supported by the CLAROM, with the
ranging between 42 and 45 MN (see Figure 6) participation of BV, ELF (now TOTAL), IFREMER
In the range of 30% to 95% higher under tension and SAGE GEODIA. The reporting work performed
loading, ranging between 24 and 36 MN as calcu- by C. Jaeck from SAGE GEODIA (now FUGRO
lated with the IC-MTD and INPG-L3S method, Engineers) was highly appreciated. Both TOTAL and
respectively (see Figure 7). CLAROM are acknowledged for granting the permis-
The jacket platform under consideration has been sion to publish this paper.
installed in the Dutch sector of the North Sea more than
ten years ago. According to Chow et al. (1996), another
key aspect for the foundations of such structures would REFERENCES
be the effect of time on the pile capacity, with a possible
85% increase in shaft capacity during the interval API RP 2A 2000. Recommended practice for planning,
between six months and five years after installation. designing and constructing fixed offshore platforms
The vast majority of pile load tests, generally performed Working stress design. 21st edition, American Petroleum
a few days to a few weeks after pile driving, do not Institute (ed.), Washington DC.
Chow, F.C., Jardine, R.J., Brucy, F. & Nauroy, J.F. 1996. The
cover this point. It should be noted that all ultimate pile
effects of time on the capacity of pipe piles in dense
capacities given in the present paper refer to relatively marine sand. Proc. Offshore Technology Conf., Houston,
short term pile loading in the order of a few days after 69 May 1996, OTC paper 7972.
driving. This effect of time has been partly addressed by Foray, P., Balachowski, L. & Colliat, J.-L. 1998. Bearing
the Euripides JIP, by re-testing one pile about 18 months capacity of model piles driven into dense overconsolidated
after driving, which did confirm a significant increase sands. Canadian Geotechnical Journal 35: 374385.

674

Copyright 2005 Taylor & Francis Group plc, London, UK


Fugro Engineers 2004. Axial pile capacity design method Euripides JIP. Proc. intern. symp. on Frontiers in Offshore
for offshore driven piles in sand. Report to the American Geotechnics, ISFOG, Perth, 1921 September 2005.
Petroleum Institute, Issue 3, August 2004. Randolph, M.F., Dolwin, J. & Beck R. 1994. Design of
Jardine, R.J. & Chow, F.C. 1996. New design methods for driven piles in sand. Gotechnique 44(3): 427448.
offshore piles. Imperial College and the Marine Technical Zuidberg, H.M. & Vergobbi, P. 1996. EURIPIDES, load tests
Directory Ltd (ed.), London, Publication MTD 96/103. on large driven piles in dense silica sands. Proc. Offshore
Kolk, H.J., Vergobbi, P. & Baaijens, A. 2005. Results from Technology Conf., Houston, 69 May 1996, OTC paper
axial load tests on pipe piles in very dense sands: the 7977.

675

Copyright 2005 Taylor & Francis Group plc, London, UK


Bearing capacity of driven piles in sand, the NGI approach

C.J.F. Clausen, P.M. Aas & K. Karlsrud


Norwegian Geotechnical Institute, Oslo, Norway

ABSTRACT: Based upon comparisons between calculated and measured axial capacity of driven piles in sand,
it is concluded that the present API RP2A recommendations should be revised to better reflect the measured
capacities. The authors propose a new empirical calculation method called NGI-99. This method was calibrated
against well documented pile tests with results from CPTs in a database established by NGI. Based upon this
method a best-fit conversion between SPT and CPT was established. Results of comparisons between calculated
and measured pile capacities for the NGI-99 and two other methods are presented.The proposed new method
gives a good agreement between calculated and measured capacities for the most relevant tests in NGIs database.

1 INTRODUCTION

The calculation of axial bearing capacity for offshore


piles has traditionally been carried out as recom-
mended by the American Petroleum Institute, API
RP2A, API (1993). After high quality test results
from driven piles in dense sand became available to
the participants in the Euripides project, Zuidberg &
Vergobbi (1996), several authors have pointed out the
important difference between the capacities calcu-
lated by the API RP2A method and those actually
measured. Over the last 10 years, the Norwegian
Geotechnical Institute (NGI) has built a database with
well-documented results from tests on driven piles in
sand, and carried out comparative calculations with
different methods.
In the following, results of such comparisons are
presented, together with a new calculation method,
called NGI-99, that leads to an improved agreement
between predicted and observed capacities. An
accompanying paper by Karlsrud et al (2005) deals Figure 1. Range of the 85 pile tests in NGIs database for
with the capacity of driven piles in clay. piles in sand.

scale pile tests were included in the database, Fugro


2 DATABASE (2004) and Tvedt & Fredriksen (2003).
Figure 1 indicates the range of all the pile tests in
The NGI database for pile tests in sand includes 85 NGIs database in terms of tip depth and the average
tests from 35 different locations, NGI (2001). These relative density Dr along the pile shaft.
tests are all in the public domain. Results from CPTs
(see below) are available for 56 of these tests. The 3 SOIL PARAMETERS
database includes the detailed soil layering and key
soil parameters at each of the pile test locations. To calculate the capacity of a driven pile in sand, the
Recently the Euripides and the Drammen E18 full following soil parameters are needed for each of the

677

Copyright 2005 Taylor & Francis Group plc, London, UK


sand layers (the term sand is used herein for any friction. However, the authors decided not to include a
non-cohesive and non-carbonate soil type): time correction on the measured pile capacities, since
Unit weight the time between driving and load testing is not known
Porewater pressures for the majority of the pile tests in the database.
Relative density
Estimates of the relative density must be based 5 THE NGI-99 CALCULATION METHOD
upon results from some type of in situ test, for example
SPT (standard penetration test), CPT (cone penetrom- Several authors have previously pointed out that the
eter test), PMT (pressuremeter test) or observed pile API RP2A method does not correctly predict the
driving resistance. For most offshore projects, CPTs capacity of driven piles in sand, e.g. Toolan et al
are routinely carried out as a part of a normal site (1990), Randolph et al (1994) and Jardine & Chow
investigation program. (1996). Studies by Clausen & Aas (2001) confirmed
For some of the older pile tests in NGIs database these earlier findings. NGI therefore opted to develop
the site investigations only include SPT results. Since a simple CPT-based method for the calculation of
the calculation method proposed in the following is axial capacity of piles in sand. This method is referred
linked to CPT results, one needs a conversion to as NGI-99 in the following.
between SPTs and CPTs. The approach used to obtain For the majority of the pile tests loaded in com-
this conversion was to: pression, only the total capacity, and not the separate
contributions from skin friction and tip resistance, are
Calibrate the NGI-99 method against the high known. To develop the NGI-99 method, the piles
quality pile tests with CPT results. loaded in tension were therefore considered first :
Find the SPT to CPT conversion that gives the best
agreement between the observed capacity, and the Find an expression for the average skin friction that
capacity calculated by the NGI-99 method for the acts upon piles loaded in tension.
pile tests with SPTs only. Assume that the skin friction in compression is a
constant times the skin friction in tension.
This resulted in the following simple expression: Assume that the skin friction in a homogeneous
(1) deposit has a triangular distribution with depth.
Determine the tip resistance in compression as
some function of the cone tip resistance.
(2)
Based upon calculations with the NGI sand data-
where NSPTcorr is the corrected SPT N-value, Peck base, and guided by the detailed results from instru-
et al (1974), NSPTmeas the measured SPT N-value, vo mented pile tests, the following expressions were
the vertical effective stress, atm the atmospheric ref- reached.
erence pressure (100 kPa), and qc the calculated cone The local unit skin friction on a driven pile in sand,
tip resistance. skin, is given by:
Section 6 includes a comparison between the scat-
ter in the ratio of calculated to measured capacity for
SPT- and CPT-based pile tests. (4)
The NGI-99 calculation method (Section 5) uses
the following relationship between CPT tip resistance z  Depth below the ground surface
qc and the sand relative density Dr: ztip  Pile tip depth
atm  Atmospheric reference pressure  100 kPa
(3)
(5)
This expression is a compromise between the two
diagrams shown on Figure 5.47 of Lunne et al (1997).
(6)
It should be noted that for very dense sands at shallow
depth, equation (3) may result in Dr  1.0. Such a Ftip  1.0 for a pile driven open-ended, 1.6 for a
result is not considered as unrealistic, and should be close-ended pile
used in the calculations. Fload  1.0 for tension, 1.3 for compression
Fmat  1.0 for steel, 1.2 for concrete.
4 EFFECT OF TIME BETWEEN PILE The relative density Dr to be used for FDr is the
DRIVING AND TESTING value calculated from equation (3) at the depth z. It
should be noted that the ratio z/z tip leads to a friction
Chow et al (1998) and Fugro (2004) present results that fatigue effect, i.e. as the pile is driven deeper the
indicate an important effect of time upon the pile skin local unit skin friction at depth z goes down.

678

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 1. Comparison between calculated and meas-
ured capacities for tests on steel piles at sites with CPT
data and the highest quality rating.

Calculated/measured pile capacity

Method Average C/M CoV

Piles loaded in tension, 8 tests


API-93 0.57 0.33
MTD-96 0.96 0.15
NGI-99 0.95 0.20
Fugro-04 0.79 0.43
Piles loaded in compression, 20 tests
API-93 0.67 0.39
MTD-96 0.85 0.27
NGI-99 0.95 0.23
Fugro-04 0.94 0.39

Figure 2. Measured average skin friction on piles in sand NGIs sand database. Four calculation methods were
subjected to tension loading. included:
API RP2A from 1993, API-93
The tip resistance acting against a pile driven Jardine & Chow (1996), MTD-96
close-ended is given by: NGIs method as presented above, NGI-99
Fugro (2004), Fugro-04
(7)
For each of these methods the average value of the
The tip resistance acting against a pile driven open- ratio between calculated and measured capacity (called
ended is taken as the smallest of the coring and the C/M below) was found, together with the coefficient of
plugged tip resistance. The coring tip resistance is cal- variation (CoV), i.e. the standard deviation divided by
culated assuming a stress against the pile wall of qc, and C/M. The comparative calculations were carried out
an internal pile/plug unit skin friction taken as 3 times for tests on tubular steel piles, with the highest quality
the external skin friction. This higher inside friction is rating and with CPT data, 28 tests met these criteria.
caused by arching near the pile tip. The plugged tip As one would expect, the NGI-99 method has little
resistance of an open-ended pile is calculated as: bias and a relatively small CoV in this comparison,
since that method was calibrated against many of the
(8) same pile tests as those used for the Table 1 compari-
son. On Figure 3 the C/M ratios calculated by the
From the above it follows that the skin friction is NGI-99 method for the 28 pile tests in Table 1 are
mainly governed by the relative density, while the ver- plotted against pile tip depth.
tical effective stress has only a modest influence. This The high CoV values found for Fugro-04 in Table
is supported by the observed skin friction on piles 1 are puzzling since this method has also been cali-
subjected to tension loading, Figure 2, where the brated against essentially the same tests, in particular
average vertical effective stress along the pile is given the Euripides tests.
for each data point. Table 2 compares calculated and measured capaci-
The limiting skin friction values recommended by ties for these high quality tests and the three CPT-
the API RP2A for non-silty sands are included on based calculation methods.
Figure 2 for comparison. For long piles in loose sand For these tests the main difference between the
the API-93 limiting skin frictions values are too opti- methods is that Fugro-04 gives somewhat lower ten-
mistic, whereas they can be too pessimistic in very sile capacity. The CPT profiles used by the authors
dense sands. for this comparison are shown on Figure 4.
In an attempt to explain the high CoV values found
6 COMPARISON BETWEEN CALCULATED for the Fugro-04 method, measured and calculated
AND MEASURED CAPACITY unit skin friction for the two Euripides tensile tests at
47 m tip depth were plotted against depth on Figure 5.
The following presents a comparison between calcu- It is observed that the Fugro-04 method matches
lated and measured capacities for the pile tests in the low skin friction measured between 22 m and

679

Copyright 2005 Taylor & Francis Group plc, London, UK


Figure 4. Idealisation of the CPT profiles used in NGIs
Figure 3. Comparison between calculated and measured
calculations for the two Euripides test locations.
pile capacity for the pile tests in Table 1 and the NGI-99
method.

Table 2. Measured and calculated capacities for six of the


Euripides piles.

Calculated/measured capacity
Pile Measured
test (MN) MTD-96 NGI-99 Fugro-04

Tension loading
L138.7 m 8.4 0.85 0.85 0.83
L147.2 m 12.5 0.85 0.87 0.73
L246.9 m 9.7 1.13 1.15 0.97
Average C/M 0.94 0.96 0.84
Compression loading
L138.7 m 12.5 1.08 1.04 1.19
L147.0 m 19.1 1.00 0.96 0.94
L246.7 m 18.8 1.06 1.00 0.99
Average C/M 1.05 1.00 1.04

37 m quite well. However, in order to obtain the meas-


ured total capacity as well, the Fugro method needs to Figure 5. Comparison between measured and calculated
introduce very high unit skin friction values near the local skin friction for the Euripides tension piles with 47 m
pile tip, higher than 1000 kPa. tip depth.
This means that the properties of the soil layers
close to the pile tip will have a dominating influence used for calibration of the MTD-96 and Fugro-04
upon the calculated capacity. The authors believe that methods. It is therefore of interest to check how the
this could explain the high CoV values found for the different CPT-based methods predict the Drammen
Fugro-04 method in Table 1. tests. Results of such comparisons are shown in Table 3.
The Euripides tests were carried out in very dense MTD-96 gives too low capacity for these four
sands. Four pile load tests with open-ended tubular piles, probably as a result of a too low calculated tip
steel piles of OD  813 mm in loose to medium resistance. The two other methods give a good aver-
dense sand were recently carried out for the E18 age C/M, but there is a wide scatter between the indi-
motorway bridge in Drammen, Norway, Tvedt & vidual results. Part of the difference between the
Fredriksen (2003). These tests are not part of the data calculation methods for these tests is due to the

680

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 3. Measured and calculated capacities for the ACKNOWLEDGEMENTS
Drammen test piles, loaded in compression.
The work presented herein was supported by Norsk
Calculated/measured capacity Hydro, Statoil and NGI. The authors gratefully
Pile position Measured
acknowledge their generous support. The Fugro (2004)
and tip depth (MN) MTD-96 NGI-99 Fugro-04
report was sponsored by the API Subcommittee on
Pier 2515 m 1.89 0.44 0.79 1.29 Offshore Structures.
Pier 2525 m 3.45 0.43 0.81 0.77
Pier 1611 m 1.30 0.61 1.54 1.78
Pier 1617 m* 1.30 0.84 0.67 0.53 REFERENCES
Average C/M 0.58 0.95 1.09
American Petroleum Institute 1993. Recommended Practice
* This pile has its tip in clay for Planning, Designing and Constructing Fixed Offshore
Platforms Working Stress Design. API RP 2A-WSD,
assumed tip behaviour, plugged or coring. For all four 20th Edition, Washington, 1 July 1993.
tests it was reported that the soil plug followed the Chow F.C., Jardine, R.J., Brucy F. & Nauroy J.F. 1998.
Effects of Time on Capacity of Pipe Piles in Dense
pile as the load was applied. Marine Sand. ASCE, JGGE, Vol. 124, No. 3, March 1998.
The pile tests in the database with SPT data (29 Clausen C.J.F. & Aas P.M. 2001.Capacity of Driven Piles in
tests) were analysed by the NGI-99 method only and Clays and Sands on the Basis of Pile Load Tests.
the proposed SPT to CPT conversion, equations Proceedings of the 11th (2001) International Offshore
(12). This resulted in a C/M of 0.97 and a CoV of and Polar Engineering Conference, ISOPE, Volume II,
0.32. The same numbers from Table 1 for the piles Stavanger June 2001. p. 581586.
with CPT data are 0.95 and 0.22. This difference in Fugro Engineers B.V. 2004. Axial Pile Capacity Design
CoV most likely reflects the improved quality of the Method for Offshore Driven Piles in Sand. Report P1003
soil data obtained by the CPT. presented to the American Petroleum Institute, issue 3, 5
August 2004.
Jardine R.J. & Chow F.C. 1996. New Design Methods for
7 CONCLUSIONS Offshore Piles. Marine Technology Directorate Ltd.,
Publication MTD 96/103, London 1996. ISBN 1 870553
31 4.
NGI established a comprehensive database with results
Karlsrud K., Clausen C.J.F. & Aas P.M. 2005. Bearing
from load tests on driven piles in sand. A new empirical Capacity of Driven Piles in Clay, the NGI Approach.
calculation method, referred to as NGI-99, was devel- Proc., International Symposium on Frontiers in Offshore
oped and calibrated against the results in this database. Geotechnics, Perth Sept. 2005, A.A. Balkema Publishers.
Trial calculations with different methods and the NGI Lunne T., Robertson P.K. & Powell J.J.M. 1997. Cone
database lead to the following conclusions: Penetration Testing in Geotechnical Practice. Blackie
Academic & Professional, ISBN 0 751 40393 8, London
1 The API-93 method should be replaced by other 1997.
methods that better reflect the results from recent Norwegian Geotechnical Institute 2001. Bearing Capacity
high quality pile tests, e.g. the Euripides and the of Driven Piles, Piles in Sand. Internal report no. 525211-
Drammen tests. 2, 21 January 2001.
2 The MTD-96 method seems to work well for piles Peck R.B., Hansen W.E. & Thornburn T.H. 1974. Foundation
in dense sand. Full scale tests on piles in loose to Engineering. John Wiley & Sons, New York 1974.
medium sand in Drammen, Norway, do however Randolph M.F., Dolwin J. & Beck R. 1994. Design of Driven
Piles in Sand. Geotechnique, Vol. 44, No. 3, September
show capacities significantly higher than those
1994, p.427.
predicted by MTD-96. Toolan F.E., Lings M.L. & Mirza U.A. 1990. An Appraisal
3 The new method proposed by Fugro (2004) has of API RP2A Recommendations for Determining Skin
been calibrated against the Euripides tests and there- Friction of Piles in Sand. OTC Paper 6422, Houston,
fore predicts these tests quite well. However, May 1990.
when this method is used for other well documented Tvedt G. & Fredriksen F. 2003. E18 Ny motorvegbru i
tests in NGIs database, the scatter in the calculated Drammen. Prvebelastning av peler. Proceedings from
results becomes higher than for the MTD-96 and the the Conference on Rock Blasting and Geotechnics, Oslo,
NGI-99 methods. This effect could be caused by the November 2003.
Zuidberg H.M. & Vergobbi P. 1996. EURIPIDES, Load
very high local unit skin friction calculated near the
Tests on Large Driven Piles in Dense Silica Sands. OTC
pile tip by the new Fugro method. paper 7977, Houston, May 1996.
4 The NGI-99 method presented herein results in a
good agreement between calculated and measured
capacities for the most relevant pile tests in NGIs
database.

681

Copyright 2005 Taylor & Francis Group plc, London, UK


The UWA-05 method for prediction of axial capacity of driven piles
in sand

B.M. Lehane, J.A. Schneider & X. Xu


The University of Western Australia (UWA), Perth

ABSTRACT: This paper describes a new method for evaluating the axial capacity of driven piles in siliceous sand
using CPT qc data. The method is shown to provide better predictions than three other published CPT based meth-
ods for a new extended database of static load tests. The design expressions incorporate the most important fea-
tures currently accepted as having a controlling influence on driven pile capacity at a fixed time after installation
(e.g. The effects of soil displacement, friction fatigue, sand-pile interface friction, dilation at the shaft and load-
ing direction) and are seen to reduce to a simplified form for typical (large diameter) offshore piles.

1 INTRODUCTION AND BACKGROUND


assume a near-proportional relationship between
local shaft friction (f) and qc and allow for the
The authors, at the request of the American Petroleum
degradation of f with distance above the pile tip
Institute (API) piling sub-committee, recently con-
(h) due to friction fatigue.
ducted a review of methods for the assessment of the
3 The ICP-05 method indicated the lowest coefficient
axial capacity of driven offshore piles in siliceous sand.
of variation (COV) for calculated to measured
The review, which is described in detail in Lehane et al.
capacities (Qc/Qm) of 0.32, when an equal weighting
(2005a) and involved the development of an extended
is given to each pile test in the database. However,
database of static load tests, evaluated the existing API
the relative performance of each method for vari-
recommendations (API-00) and three Cone Penetration
ous categories within the database is less clear. For
Test (CPT) based methods namely: Fugro-04 (Fugro
example, NGI-04 predictions appear best for open-
2004), ICP-05 (Jardine et al. 2005) and NGI-04
ended piles in compression while Fugro-04 and
(Clausen et al. 2005). A new design method, referred to
ICP-05 provide comparable predictive accuracies
as UWA-05, emerged following the evaluation exer-
for open-ended piles in tension.
cise and is the described in this paper and in Lehane
4 When account was taken of the relative reliability
et al. (2005b).
of the pile test data (using a carefully designed
The assessment of the predictive performance of
weighting procedure), the methods listed below for
API-00, Fugro-04, ICP-05 and NGI-04 against the new
each category of pile lead to the lowest probability
UWA pile test database indicated the following trends
of failure:
(which are described in detail in Lehane et al. 2005a):
API-00: closed-ended piles in compression
1 All three CPT based design methods considered Fugro-04: closed-ended piles in tension
(Fugro-04, ICP-05 & NGI-04) had significantly ICP-05 & NGI-04: open-ended piles in com-
better predictive performance than the existing API pression
recommendations, which were seen to lead large ICP-05 & Fugro-04: open-ended piles in tension
under-predictions in dense sands and become pro- 5 API-00 gives the lowest probability of failure for
gressively non-conservative as the pile length (L) closed-ended piles in compression partly because
or aspect ratio (L/D) increased. the method generally under-predicts the capacity of
2 Despite the CPT based methods having a broadly the database piles to a significant degree. However,
similar predictive performance against the new data- while the same average level of under-prediction
base of load tests, their formulations relating the also applies to API-00 predictions for closed-
pile end bearing with the cone tip resistance (qc) ended piles in tension, the estimated probability of
are notably different. Formulations for shaft fric- failure is larger than the three alternative CPT
tion also differ significantly in detail, although all design methods.

683

Copyright 2005 Taylor & Francis Group plc, London, UK


6 The ICP-05 method displays a tendency to under- A simplified (and conservative) means of determi-
predict pile base capacities (when assuming cap- nation of the Dutch q-c value is provided in Lehane
acity solely from annular end bearing) and to et al. (2005b), which may be more practical when
become potentially non-conservative for tension using CPT data collected offshore, which are often
capacity as the pile aspect ratio (L/D) increases. The not continuous.
Fugro-04 method indicates a tendency to under- The values of qb0.1 for driven piles are less than q-c
predict compression capacities for long piles and because the displacement of 0.1D is insufficient to
to over-predict base capacities in loose sand. mobilise the ultimate value (of q-c).
The findings of Randolph (2003), White & Bolton
The examination of the three CPT based methods
(2005), and others, are consistent with the UWA-
coupled with a review of their various deficiencies
05 proposal to adopt a constant ratio of qb0.1/q-c for
and a careful examination of the new extended data-
driven closed-ended piles.
base of static load tests prompted the authors to pro-
pose the UWA-05 method presented here. This The UWA-05 design equation for the end bearing
method is believed to represent a significant improve- of a closed-ended pile, with diameter D, is given as:
ment on Fugro-04, ICP-05 and NGI-04 methods.
Particular comparisons are made with ICP-05, which (1)
Lehane et al. (2005a) adjudged to have a marginally
better predictive performance than the other two CPT
based methods.
2.1.2 Open-ended piles
Salgado et al. (2002), Lehane & Gavin (2001,
2004), and others, have shown that a relatively con-
2 THE UWA-05 DESIGN METHOD FOR sistent relationship between qb0.1 for a pipe pile and
PILES IN SAND the CPT qc value becomes apparent when the effects
of sand displacement close to the tip during pile
2.1 End bearing driving are accounted for. This installation effect is
best described by the incremental filling ratio (IFR)
Factors that were considered in the development of
measured over the final stages of installation and is
the UWA-05 proposals for base capacity evaluation
referred to here as the final filling ratio (FFR). As
of closed and open-ended piles are listed in the fol-
the FFR approaches zero, qb0.1 approaches that of a
lowing. These proposals are based on the analyses
closed-ended pile with the same outer diameter.
reported in Xu & Lehane (2005) and Xu et al. (2005).
The displacement induced in the sand in the vicinity
The base capacity is defined as the pile end bearing
of the base is most conveniently expressed in the
resistance at a pile base movement of 10% of the pile
terms of the effective area ratio Arb* , defined in
diameter, qb0.1.
Equation 2c. This ratio depends on the piles D/t
(diameter to wall thickness) ratio and the FFR value,
2.1.1 Closed-ended piles varying from unity for a pile installed in a fully
The strong direct relationship between the end plugged mode to about 0.08 for a pile installed in
bearing resistance of a closed-ended driven pile coring mode with D/t of 50.
and the cone tip resistance, qc, has been recognised Lehane & Randolph (2002), and others, have shown
for many years and arises because of the similarity that, if the length of the soil plug is greater than 5
between their modes of penetration. internal pile diameters (5Di), the plug will not fail
Given the difference in size between a pile and a under static loading, regardless of the pile diameter.
cone penetrometer, a correlation between qb0.1 and Experimental data and numerical analysis indicate
qc requires use of an appropriate averaging tech- that the resistance that can develop on the tip annu-
nique to deduce an average value of q-c. Xu & lus at a base movement of 0.1D varies between about
Lehane (2005) show that, for many stratigraphies 0.6 and 1.0 times the CPT qc value (e.g. Bruno
encountered in practice, q-c may be taken as the 1999, Salgado et al. 2002, Lehane & Gavin 2001,
average qc value taken in the zone 1.5 pile diam- Paik et al. 2003, Jardine et al. 2005).
eters (D) above and below the pile tip. Lehane & Randolph (2002) suggest that the base
Xu & Lehane (2005), however, also show that resistance provided by the soil plug for a fully cor-
when qc varies significantly in the vicinity of the ing pile (with FFR  1) is approximately equiva-
pile tip (i.e. within a number of diameters), the lent to that of a bored pile.
Dutch averaging technique (Van Mierlo & Recommended values of qb0.1/qc for bored piles
Koppejan 1952, Schmertmann 1978) provides the range from 0.15 to 0.23 (Bustamante & Gianeselli
most consistent relationship for end bearing and 1982, Ghionna et al. 1993). These ratios are not
should be employed to calculate q-c. dependent on the pile diameter.

684

Copyright 2005 Taylor & Francis Group plc, London, UK


The value of q-c should be evaluated in the same ratio, Ars*, which is unity for a closed ended pile
way as that employed for closed-ended piles, but and, for a pipe pile, includes displacement due to
using an effective diameter (D*) related to the the pile material itself and the additional displace-
effective area ratio, Arb* i.e. D*  D Arb*0.5. ment imparted when the pile is partially plugging
There are relatively few documented case histories or fully plugged during driving. White et al. (2005)
that report the incremental or final filling ratios. In use a cavity expansion analogy to deduce that the
the absence of FFR measurements, a rough esti- equalized lateral effective stress is likely to vary
mate of the likely FFR may be obtained using with the effective area ratio raised to a power of
equation 2d (see Xu et al. 2005). between 0.30 and 0.40.
The incremental filling ratio (IFR) is a measure of
The UWA-05 proposal for end bearing of driven
soil displacement near the tip of a pipe pile and
pipe piles is provided in Equation (2). This proposal is
depends on a number of different parameters, includ-
developed in Xu et al. (2005) and shown to compare
ing soil layering, pile inner diameter, pile wall thick-
favourably with the existing database of base capacity
ness, plug densification or dilation, and installation
measurements for open-ended piles.
method. For the (limited) database of IFRs reported,
the mean IFR over the final 20 D of penetration
(2a) (where most friction is generated) can be reasonably
approximated using Equation (3e) for relatively uni-
form dense to very dense sands in the database.
(2b) After displacement of the sand near the tip in a
given soil horizon and as the tip moves deeper, the
radial stress acting on the pile shaft (and hence the
available f value) in that horizon reduces. This
(2c) phenomenon, known as friction fatigue, is now an
accepted feature of displacement pile behaviour
(e.g. see Randolph 2003).
The rate of radial stress and f reduction with height
(2d) above the tip (h) depends largely on the magnitude
and type of cycles imposed by the installation
method. White & Lehane (2004) show that the rate
where Di is the inner pile diameter. of decay is stronger for piles experiencing hard
driving and much lower for jacked piles, which are
typically installed with a relatively low number of
2.2 Shaft friction (one-way) installation cycles.
White & Lehane (2004), and others, also show that
Factors that were considered in the development of the rate of degradation with h is greater at higher
the UWA-05 method for shaft friction are discussed levels of radial stiffness (4G/D) and therefore f at
in Schneider & Lehane (2005) and Lehane et al. a fixed h value (i.e. after a specific number of
(2005a). These are now summarised as follows: installation cycles) in a sand with the same oper-
Local shaft friction (f) shows a strong correlation ational shear modulus (G) reduces as D increases.
with the cone tip resistance (qc). This correlation, The foregoing, plus the tendency for hammer selec-
which has been observed directly in instrumented tion to be such that the number of hammer blows is
field tests has been employed successfully in well broadly proportional to the pile slenderness ratio
known design methods, such as that proposed by (L/D), suggest that f may be tentatively considered
Bustamante & Gianiselli (1982). a function of h/D. This approximation is supported
The shaft friction that can develop on a displace- by field measurements such as those provided in
ment pile is related to the degree of soil displace- Lehane et al. (2005a), and is also compatible with the
ment imparted during pile installation. The higher occurrence of a critical depth at an embedment
capacity developed by the new generation of screw related to a fixed multiple of the pile diameter (such
piles compared to that of a bored and continuous as 20D proposed by Vesic 1970 and a number of
flight auger piles is just one example of this effect. workers). The same approximation is employed by
The degree of displacement imparted to any given the ICP-05 and Fugro-04 design methods.
soil horizon is related to the displacement experi- Based on the former point, the ICP-05 method pro-
enced by that horizon when it was located in the poses that f varies in proportion to (h/D)
c, where
vicinity of the tip. This level of displacement can c  0.38. However, given that this value of c was
conveniently be expressed for both closed and estimated on the basis of field tests with jacked
open-ended piles in terms of an effective area piles (Lehane 1992 and Chow 1997) where the

685

Copyright 2005 Taylor & Francis Group plc, London, UK


32 The shaft friction that can develop on a pile in ten-
Employed for
sion is smaller than that which can be mobilised by
database evaluation
30 a pile loaded in compression for the reasons
Interface Friction Angel, cv

described by Lehane et al. (1993), de Nicola &


tan 0.55
28
Randolph (1993) and Jardine et al. (2005).
UWA-05 Because of the shortage of high quality measure-
recommendation
ments of f very close to the tip of a driven pile and
26
the variable and inconsistent trends shown by the
available measurements, one simplifying option is
24 to assume f is constant over the lower two diam-
eter length of the pile shaft for both closed and
22 open-ended piles in tension and compression.
Shaft capacity increases with time as shown by
20 Axelsson (1998), Jardine et al. (2005a), and others.
0.01 0.1 1 10 Lehane et al. (2005a) show, however, that rate of
Median Grain Size, D50 (mm) increase over the period 3 days to 50 days is not
statistically significant for the UWA database of
Figure 1. cv variation with D50 (modified from ICP-05 load tests. A design time of 10 to 20 days is con-
guidelines). sidered appropriate for shaft friction calculated
using UWA-05.
The UWA-05 design equations for shaft capacity
type and number of cycles imposed is less severe
of driven piles arose from the foregoing consider-
than is typical of driven piles, a higher value of c is
ations and are expressed as follows:
considered more appropriate for offshore pile.
Strong indirect evidence in support of this obser-
vation is also apparent in Lehane et al. (2005a), (3a)
which shows that the Fugro-04, ICP-05 and NGI-
04 progressively under-predict the shaft capacity
of jacked piles as the pile length increases.
The radial effective stress acting on a driven pile (3b)
increases during pile axial loading and its magni-
tude (when f is mobilised and dilation has ceased)
increases as the pile diameter reduces, the sand shear
stiffness around the pile shaft increases and the (3c)
radial movement during shear (dilation) of the sand
at the shaft interface increases. These increases are
not significant for offshore piles (with large D) but
need to be considered when extrapolating from (3d)
load test data for small diameter piles in a data-
base. The recommendations of the ICP-05 method
are considered reasonable for assessment of the
increase in radial stress (rd), but with a modi-
fied expression for the shear stiffness derived from (3e)
the CPT data.
f varies in proportion to tan cv (where cv is the con-
stant volume interface friction angle between the
(3f)
sand and pile); this cv value, which should be meas-
ured routinely, increases as the roughness normal-
ized by the mean effective particle size (D50) where
increases. Verification of the dependence of f on tan cv  constant volume interface friction angle
cv has been provided by Lehane et al. (1993), Chow rf  radial effective stress at failure
(1997), and others. In the absence of specific labora- rc  radial effective stress after installation and
tory measurements of cv. UWA-05 recommends the equalization
trend shown on Figure 1, which is the same as that rd  change in radial stress due to loading stress
employed by ICP-05 but with an upper limit on path (dilation)
tan cv value of 0.55 (due to the potential for changes f/fc  1 for compression and 0.75 for tension
in surface roughness during pile installation). G/qc  185  qc1N
0.75 with qc1N  (qc/pa)/(v0/pa)0.5

686

Copyright 2005 Taylor & Francis Group plc, London, UK


pa  a reference stress equal to 100 kPa Table 1. Sensitivity of pipe pile capacity to A*r (A*rb
v0  in situ vertical effective stress and A rs*).
r  dilation (assumed for analyses  0.02 mm,
Method for calculation
as for ICP-05) of A*r Mean Qc/Qm COV for Qc/Qm

Open-ended piles in compression


Using Equations 2d & 0.99 0.23
3 PREDICTIVE PERFORMANCE OF UWA-05 3e for all tests
Assuming IFR  1 0.81 0.24
The UWA database of static loads tests, as discussed
in Lehane et al. (2005a & b), was employed to assess Using measured IFR 0.98 0.19
the predictive performance of the proposed UWA-05 when available
method. The predictions described employed equa- Open-ended piles in
tions (1), (2) and (3) with the following additional tension
considerations: Using Equations 2d & 0.97 0.26
3e for all tests
Measured interface friction angles, when available, Assuming IFR  1 0.77 0.22
were adopted. Figure 1 was used in the absence of Using measured IFR 0.91 0.23
measured cv values. when available
When the incremental filling ratio (IFR) was
recorded, Arb* was assessed using the mean IFR
value measured over the final 3D of pile penetra-
pile test categories (except for closed-ended
tion while the value of Ars* was assessed from the
piles in compression where UWA-05 and ICP-
mean IFR value recorded over the final 20D of
05 have the same COV for Qc/Qm).
penetration. In the absence of IFR data, Arb* and
(iii) The COV of 0.19 for Qc/Qm of the UWA-05
Ars* were evaluated using Equations 2d & 3e.
method for open-ended piles in compression is
The database included 74 load tests at sites where significantly lower than the corresponding COV
CPT qc data were measured. Pile test data at sites con- of 0.25 of ICP-05.
taining micaceous, calcareous and residual sands (iv) UWA-05 shows no apparent bias of Qc/Qm with
were excluded from consideration as were sites for pile length (L), pile diameter (D), pile aspect
which only Standard Penetration Test data were avail- ratio (L/D) and average sand relative density.
able. The database included substantially more pile One of the factors giving rise to the superior per-
tests than used for verification of the Fugro-04, ICP- formance of the UWA-05 method for pipe piles is the
05 and NGI-04 design methods and was sub-divided inclusion of the effective area ratio terms in the expres-
into the following four categories: sions for base and shaft capacities of open ended
piles. This is not surprising given the acknowledged
(a) Closed-ended piles tested in compression
importance of soil displacement on capacity and the
(b) Closed-ended piles tested in tension
fact that many of the database piles showed evidence
(c) Open-ended piles tested in compression
of partial plugging. However, given that the incre-
(d) Open-ended piles tested in tension
mental filling ratio (IFR) is not commonly measured
A detailed presentation and discussion of this stat- in practice, the sensitivity of the predictive perform-
istical analysis, which was conducted for API-00, ance to the IFR parameter employed was re-examined
Fugro-04, ICP-05 and NGI-04, as well as for UWA- and a summary of this exercise is provided in Table 1.
05 is presented in Lehane et al. (2005a & b) and may It is clear from Table 1 that the estimation of IFR
be briefly summarized as follows: using the empirical equations 2d & 3e, rather than
direct use of the measured IFRs to deduce Ar* values,
(i) For the database taken as a whole (i.e. including has only a minimal impact on the COV values for
all pile categories), the UWA-05 method pre- Qc/Qm. It may also be inferred that the assumption in
dicts a mean ratio of calculated to measured UWA-05 of a fully coring pile (i.e. IFR  1) for the
capacity (Qc/Qm) of 0.97 and the lowest overall database piles (most of which had diameters less than
coefficient of variation (COV) for this ratio of 800 mm) will lead, on average, to a 20% under-
0.29; this compares well with the respective prediction of capacity. Such an under-prediction is in
COVs of 0.32, 0.38, 0.43 and 0.6 for ICP-05, keeping with observed levels of partial plugging of
Fugro-04, NGI-04 and API-00. (smaller diameter) database piles and suggests that
(ii) The UWA-05 method has the lowest COV for other design methods, such as ICP-05, which may
Qc/Qm of all five methods for each of the four provide a good fit to the existing database of load tests,

687

Copyright 2005 Taylor & Francis Group plc, London, UK


but do not include an appropriate soil displacement (ii) provides better predictions for a new extended
term (such as Ar*), will over-predict the capacity of database of load tests than the ICP-05, Fugro-04
full scale offshore piles. and NGI-04 CPT based design approaches;
(iii) employs soundly based formulations that draw
on the considerable recent developments in our
4 PREDICTIONS FOR OFFSHORE PILES understanding of displacement piles in sand;
(iv) provides formulations that enable a rational
The UWA-05 method simplifies to the following extrapolation beyond the existing database base
form for full scale offshore piles, as IFR  1 and the of load tests.
dilation term (rd) can be ignored.

(4a) REFERENCES

Axelsson G. 1998. Long-term increase in shaft capacity of


driven piles in sand. Proc., 4th Int. Con. on Case Histories
(4b) in Geotechnical Eng., St. Louis, Missouri: 125.
Bruno, D. 1999. Dynamic and static load testing of driven
piles in sand. PhD Thesis, University of Western Australia.
(4c) Bustamante, M. & Gianeselli, L. 1982. Pile bearing capacity
by means of static penetrometer CPT. Proc., 2nd European
Symp. on Penetration Testing, Amsterdam: 493499.
Clausen, C.J.F., Aas, P.M. & Karlsrud, K. 2005. Bearing
(4d) capacity of driven piles in sand, the NGI approach. in
Proc., ISFOG, Perth.
Chow, F.C. 1997. Investigations into the behaviour of dis-
placement piles for offshore foundations. PhD thesis,
Univ. of London (Imperial College).
(4e) de Nicola, A. & Randolph, M.F. 1993. Tensile and compres-
sive shaft capacity of piles in sand. J. Geotech. Engrg.
Div., ASCE 119(12): 19521973.
Fugro Engineers B.V. (Fugro) 2004. Axial pile capacity
Lehane et al. (2005b) examined the implications of design method for offshore driven piles in sand, P-1003,
equation (4) and assessed its performance against Issue 3, to API, August 2004.
existing API recommendations and ICP-05 (the best Ghionna, V.N., Jamiolkowski, M., Lancellotta, R. &
performing of the three CPT based methods con- Pedroni, S. 1993. Base capacity of bored piles in sands
sidered). This examination indicated that equation (4) from in situ tests. in Deep Foundation on Bored and
provides a more conservative extrapolation than ICP- Auger Piles, ed. V. Impe, Balkema, Rotterdam: 6775.
05 for shaft capacity from the existing database (of Jardine, R.J., Chow, F.C., Overy, R. & Standing, J. 2005. ICP
relatively small diameter piles with a mean D of design methods for driven piles in sands and clays.
Thomas Telford, London.
about 0.7 m) to typical offshore piles used in practice. Jardine, R.J., Standing, J.R. & Chow, F.C. 2005a. Field
Equation (4) also predicts higher base capacities than research into the effects of time on the shaft capacity of
ICP-05 because of its assumption that a pile plug with piles driven in sand. Proc., ISFOG, Perth.
a length greater than 5 diameters will not fail under Lehane, B.M., Jardine, R.J., Bond, A.J. & Frank, R. 1993.
static loading. Mechanisms of shaft friction in sand from instrumented
It is also noteworthy that Equation (4) tends to pro- pile tests, J. of Geotech. Engrg., ASCE, 119 (1): 1935.
vide lower capacities than API-00 in loose sands, but Lehane, B.M. & Gavin, K.G. 2001. Base resistance of
higher capacities for dense sands in compression. jacked pipe piles in sand. J. of Geotech. and Geoenv.
API-00 and UWA-05 predictions for tension capacity Engrg, ASCE 127(6): 473480.
Lehane, B.M. & Gavin, K. 2004. (Discussion). Determination
in dense sands are broadly similar for pile lengths in of bearing capacity of open-ended piles in sand. J.
excess of 20 m. However, the UWA-05 method, Geotech. & Geoenv. Engrg. ASCE 130 (6): 656658.
unlike API-00, does not show any prediction bias Lehane, B.M. & Randolph, M.F. 2002. Evaluation of a min-
with L, D, L/D and Dr. imum base resistance for driven pipe piles in siliceous
sand. J. of Geotech. and Geoenv. Engrg., ASCE 128(3):
198205.
5 CONCLUSIONS Lehane, B.M. 1992. Experimental investigations of pile
behaviour using instrumented field piles. PhD thesis,
This paper has shown that the UWA-05 method: Univ. of London (Imperial College).
Lehane, B.M., Schneider, J.A. & Xu, X. 2005a. Evaluation
(i) is a significant improvement on existing API of design methods for displacement piles in sand. UWA
recommendations; Report, GEO: 05341.1.

688

Copyright 2005 Taylor & Francis Group plc, London, UK


Lehane, B.M., Schneider, J.A. & Xu, X. 2005b. CPT based Vesic, A.S. 1970. Tests on instrumented piles. Ogeechee
design of driven piles in sand for offshore structures. River site. Journal of the Soil Mechanics and
UWA Report, GEO: 05345. Foundations Division ASCE SM2: 561584.
Paik, K., Salgado, R., Lee, J. & Kim, B. 2003. Behavior of White, D.J. & Bolton, M.D. 2005. Comparing CPT and pile
open- and closed-ended piles driven into sands. J. of base resistance in sand. ICE, Geotechnical Enginnering
Geotech. and Geoenv. Engrg., ASCE. 129(4): 296306. 158: 314.
Randolph, M.F. 2003. Science and empiricism in pile foun- White, D.J. & Lehane, B.M. 2004. Friction fatigue on dis-
dation design. Geotechnique 53(10): 847875. placement piles in sand. Geotechnique 54(10): 645658.
Salgado, R., Lee, J. & Kim, K. 2002. Load tests on pipe piles White, D.J., Schneider, J.A. & Lehane, B.M. 2005. The
for development of CPT-based design method. Final influence of effective area ratio on shaft friction of dis-
report, FHWA/IN/JTRP-2002/4. placement piles in sand, Proc., ISFOG, Perth.
Schneider, J.A. & Lehane, B.M. 2005. Correlations for shaft Xu, X. & Lehane, B.M. 2005. Evaluation of end-bearing
capacity of offshore piles in sand. In Proc., ISFOG, Perth. capacity of closed-ended pile in sand from CPT data. in
Schmertmann, J.H. 1978. Guidelines for cone test, perform- Proc., ISFOG, Perth.
ance, and design. U.S. Federal Highway Administration, Xu, X., Schneider, J. A. & Lehane, B.M. 2005. Evaluation of
FHWATS-78209. end-bearing capacity of open-ended piles driven in sand
Van Mierlo, W.C. & Koppejan, A.W. 1952, Lengte en from CPT data. in Proc., ISFOG, Perth.
draagvermogen van heipalen, Bouw, January.

689

Copyright 2005 Taylor & Francis Group plc, London, UK


An updated assessment of the ICP pile capacity procedures

R.J. Jardine
Imperial College, London

F.C. Chow, J.R. Standing, R.F. Overy, E.Saldivar-Moguel, C. Strick van Linschoten &
A. Ridgway
WorleyParsons Pty Ltd, Perth (formerly Imperial College); Imperial College; Shell UK Ltd; Jacobs (formerly
Imperial College); Arup Geotechnics (formerly Imperial College) and Imperial College London respectively

ABSTRACT: This paper summarises a recent re-assessment of the reliability of the Imperial College Pile
design procedures, identifying and discussing a substantial set of new and modified database entries, covering
over 40 piles driven in nine widely different geotechnical settings. Shaft capacity is the main focus and it is
shown that the ICP approach leads to far more reliable predictions than the current API RP2A recommendations
for offshore driven piles driven in sands and clays.

1 INTRODUCTION tension and only a minority of the compression tests


involved accurate base capacity measurements. For
Jardine & Chow (1996) offered a new set of procedures this reason, and to save space, we focus on shaft cap-
for assessing the axial capacity of offshore piles driven acity alone. The new database regarding end resist-
in sand and clay, drawing on the series of research pro- ance is less substantially different to that reviewed by
jects described by Jardine (1985), Bond (1989), Lehane Jardine and Chow (1996).
(1992) and Chow (1997). Their Imperial College
Pile, or ICP, approach has been used widely since in
onshore and offshore projects, often being referred to 2 NEW DATABASE ENTRIES
as the MTD method.
Jardine et al (2005a) update the ICP procedures. The New load test and site investigation data are constantly
static capacity sections for cylindrical piles in silica being gathered and published. Tables 1 and 2 sum-
sands and clays involve only minor changes to the end marise nine sites where new tests on piles driven in
bearing rules and interface friction angles for sands. sands, gravels and clays have been evaluated by the
New sections are included on pile ageing; cyclic load- authors and their co-workers. In each case outline
ing; group action; pile shape and seismic effects, as details are offered of the soil and pile types and the
well as applications in calcareous and micaceous capacity predictions obtained following the ICP and
sands, diatomaceous clays and clay-silts. More detailed API (1993) procedures. The cases cannot be reviewed
recommendations are made on parameter selection and here, but references are given to more detailed reports
practical implementation. The choice of appropriate that describe the tests and how the calculation param-
Factors of Safety for WSD (Working Stress Design) or eters were derived.
Resistance Factors for LRFD (Load and Resistance Tables 1 and 2 exclude three data sub-sets of spe-
Factor Design) approaches is explored in terms of reli- cial cases that are discussed separately by Jardine et al
ability theory. (2005a): the Jamuna Bridge pile load tests involving
Jardine et al (2005a) also present a re-assessment micaeous sands (Tomlinson 1996 and CUR 2001);
of the ICP procedures ability to predict the capacity the calcareous sand cases assembled by Thompson &
of single piles, adding to Chows (1997) database and Jardine (1997) and the square and H pile databases
referring to other independent studies. This paper assembled by Cowley (1998). The Liestranda clay-silt
summarises some of the new and modified database case history reported by Karlsrud et al (1993) is also
entries and presents the principal reliability findings. excluded from the main clay database, as described
Almost half the tests considered were performed in later.

691

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 1. Additional sand case histories. Table 2. Additional clay case histories.

Hound Point Jetty: Williams et al (1997) Mexico City Database: Saldivar Moguel (2002)
Very soft clay over medium dense/dense gravel and cobbles
Marine structure tests (1 compression, 2 tension) Diatomaceous high plasticity low YSR clays; low unit
weights and high ring shear  angles. Partial pilot pre-
Piles 2 steel pipes, 1.22m OD, 3 tests bores (tests mostly in compression, one test per pile)
Pile Lengths, m 26, 34 & 41
Piles 26 concrete (square), 0.3 to
Predictions ICP API 0.5 m wide
Qc/Qm mean 1.02 1.68 Pile Length, m 10 to 38
Qc/Qm range 0.89 to 1.19 1.17 to 2.25
COV 0.15 0.32 Predictions ICP API
Qc/Qm mean 1.06 0.91
Sungai Perak Bridge: Williams et al (1997) Qc/Qm range 0.69 to 1.28 0.51 to 1.31
Medium dense gravelly sand (3 compression tests) COV 0.16 0.21
Piles 2 steel pipes, 1.5 m OD, 4 tests
West Delta 58A: Gulf of Mexico; Bogard & Matlock
Pile Lengths, m 33 & 38 (1998); Jardine & Saldivar (1999); Saldivar Moguel
Predictions ICP API (2002); Underconsolidated plastic clays; offshore pile test
Qc/Qm mean 1.11 2.08 (in tension)
Qc/Qm range 0.82 to 1.50 1.64 to 2.89
COV 0.31 0.34 Pile Steel pipe, 0.762 m OD, 1 test

EURIPIDES: Zuidberg & Vergobbi (1996); CUR (2001); Pile Length, m 71.3
Jardine et al (2005a); Loose sand and clayey sand over Predictions ICP API
very dense sand (4 compression 4 tension tests) Qc/Qm 0.97 0.96

Piles 2 steel pipes, 0.763 m OD, 8 tests Onsy; Karlsrud et al (1993); Clausen & Aas (2001);
Pile Lengths, m 30.5, 38.7, 47 & 46.7 Ridgway (2004)
Medium plasticity low YSR clay. Pre-bored, cased
Predictions ICP API starter-holes, soil core removed for pipe pile (all tension
Qc/Qm mean 0.97 0.58 tests)
Qc/Qm range 0.78 to 1.12 0.43 to 0.89
COV 0.13 0.26 Piles 4 closed steel piles, 0.219 m
OD, 6 tests
Dunkirk 19981999: Jardine & Standing (2000), Pile Length, m 10 to 32.5
Medium dense/dense marine sand (1 comp., 2 tension Pile Steel pipe, 0.812 m OD, 1 test
tests) Pile OD, m 0.812
Pile Length, m 10
Piles 3 steel pipes, 0.457 m OD
Closed ICP API
Test C1 R1 R6 Qc/Qm mean 1.43 1.37
Age, days 68 9 80 Qc/Qm range 1.26 to 1.63 1.16 to 1.76
Pile Lengths, m 10 19 19 COV 0.09 0.16
Qc/Qm ICP 0.57 0.91 0.54 Open prediction ICP API
Qc/Qm ICP Age 0.85 1.00 1.00 Qc/Qm 1.19 1.03
adjusted
Qc/Qm API 0.39 1.01 0.54 Kinnegar: Lehane et al (2004); Strick van Linschoten
(2004); Low YSR clay-silt Belfast Sleech (Vane Su
Leman BD, North Sea; Jardine et al (1998) tests); (1 compression, 1 tension test)
Medium dense to dense marine sand; conductor tension
test Piles type 2 concrete square, 0.25 m,
2 tests
Pile Steel pipe, 0.66 m OD, 1 test Pile Length, m 6

Pile Length, m 38.1 Predictions ICP API


Predictions ICP API Qc/Qm comp 1.08 1.64
Qc/Qm 1.05 0.94 Qc/Qm tens 0.86 1.30

692

Copyright 2005 Taylor & Francis Group plc, London, UK


3 SAND CASE HISTORIES Table 3. Overall results for shaft resistance in sand; total
database of 81 tests.
Points to note relating to the sand cases include:
No pile No Mean
1. The Leman case entry refers to the updated analy- Method tests sites Qc/Qm COV
sis given by Jardine et al (1998).
2. The recent tests at Dunkirk focused on the strong ICP, all piles 81 27 0.99 0.28
effects of time on shaft capacity. The entries given ICP, all open-ended piles 40 16 1.05 0.28
in Table 1 consider only the tests conducted at rela- ICP, open-ended piles,
dense North Sea sand 19 4 0.99 0.17
tively early ages. Jardine et al (2005b) discuss the
ICP, first time tests on
aged piles in more detail. young piles (all types) 41 22 0.98 0.31
The results from the EURIPIDES tests in dense API RP2A(1993); all piles 81 27 0.87 0.60
sand are considered as worked calculations by Jardine
et al (2005a), who note that several independent work- capacity components. Table 3 summarises the com-
ers have reported ICP calculations that match the parisons between the calculated and measured shaft
EURIPIDES entries well, demonstrating that the ICP capacities Qcalculated/Qmeasured (Qc/Qm) for piles driven
calculations are not operator sensitive when CPT and in sand. Statistics are given for both the ICP approach
unambiguous interface shear test data exist. and the more routinely applied API (1993) methodology.
The Coefficient of Variation (COV) is defined as the
standard deviation, s, divided by the mean, . Ideally the
4 CLAY CASE HISTORIES mean of Qc/Qm should be close to unity and the COV
(or s) should be as low as possible; the ICP method
One factor that has restricted the size of the available meets these aims far better than the API procedures.
clay data set summarised in Table 2 is the lack of site- Jardine et al (2005b) emphasise that ageing affects
specific ring shear interface tests, along with reliable capacity in sands very strongly and that allowances
YSR and sensitivity data. Points to consider in con- must be made when considering piles significantly
nection with the tabulated entries include: older or younger than the nominal 10 day test age.
Similarly, pre-testing to failure is shown to degrade
1. Saldivar-Moguel (2002) performed ring-shear inter-
capacity markedly and retard the beneficial ageing
face, oedometer and other tests in his evaluation of
processes. Table 3 presents the statistics found by con-
the ICPs applicability to square concrete piles driven
sidering a restricted data base of piles that have been
in Mexico City clay, covering a typical range of plas-
(i) installed to just one depth, (ii) tested only once, and
ticity indices (150 to 240%). His combined pile test
at an age between 1 day and one month after driving.
data set comprised 26 piles driven at several sites.
The average age (10 days) contrasts with the 25 day
2. Partial pre-boring to reduce driving disturbance is
average associated with the full data set. However, the
common in Mexico City and Saldivar-Moguel used
mean values of Qc/Qm are similarly close to unity. It
model tests to estimate the effects on shaft capacity.
appears that in this case the effects of prior testing
Site-specific pile load tests are recommended when-
accidentally compensate for the greater age of the
ever such non-standard pile installation techniques
re-tested piles.
are employed.
The mean values quoted in Table 3 are similar to
3. Interface ring-shear and other tests have been car-
those quoted by Jardine & Chow (1996), as are the
ried out recently at Imperial College to help apply
COVs. The method was derived independently of the
the ICP approaches to pile load tests conducted at
database used to test it and it is very encouraging that
the Gulf of Mexico (WD58A), Norwegian (Onsy)
the results did not change significantly when a sub-
and Northern Irish (Kinnegar) sites. Further details
stantial volume of new data was introduced.
of these test programmes are given in the cited PhD
Variability (represented by COV) is likely to reduce
Thesis and MSc Dissertations.
when considering sub-sets of similar piles and ground
conditions, as shown by the North Sea sand, open-
5 COMBINED DATA SET AND SHAFT ended pile entry in Table 3. Jardine et al (2005a) note
RELIABILITY STATISTICS FOR SAND that another variant of the method is needed when
dealing with calcareous sands, and that the ICP
The new and revised information contained in Table 1 capacities may need to be downgraded when dealing
has been combined with Chows original data set. To with micaceous sands.
avoid biasing the statistics it was decided to limit the Figures 1 to 4 show scatter diagrams of (Qc/Qm)
number of tests from any single geotechnical deposit to against Dr, and normalised pile length for the updated
10 statistically representative entries, giving a total of database. The ICP method eliminates the strong skew-
81 pile tests in sands. We focus here on just the shaft ing produced by the API method and is equally reliable

693

Copyright 2005 Taylor & Francis Group plc, London, UK


Pile type & test direction 3.5
3.5 Steel, closed-ended, tension
Steel, closed-ended, compression
Concrete, closed-ended, tension 3.0
3.0 Concrete, closed-ended, compression
Steel, open-ended, tension
Steel, open-ended, compression 2.5
Concrete, open-ended, tension
2.5
H

Qc /Qm
SP 2.0
2.0
Qc /Qm

SP
H 1.5
1.5 SP SP H
EUDK
SP EU EU BD
DK 1.0 H SP H EU
EU
EU
1.0 HBD DK EU

EU
EU 0.5
0.5 EU EU EU
EU
DK
G
0.0
0.0 0 10 20 30 40 50 60 70 80 90 100
0 10 20 30 40 50 60 70 80 90 100
L/D
Relative density, Dr (%)

Figure 4. Distribution of Qc/Qm with respect to slender-


Figure 1. Distribution of Qc/Qm with respect to relative ness ratio L/D: ICP shaft procedure for sands.
density Dr: API (1993) shaft procedure for sands.
ICP end bearing procedures may be conservative for
3.5
offshore piles driven in sand and any such trend would
compensate any slight over-estimate of shaft capacity.
3.0 Overall, the re-assessment confirms that the ICP
2.5
approach can be applied with much greater confidence
SP
H
than the conventional API (1993) recommendations.
Qc /Qm

2.0
SP
H
SP
1.5
DK 6 COMBINED DATA SET AND SHAFT
1.0 H BD

EU
RELIABILITY STATISTICS FOR CLAY
HO EU
EU
0.5 EU EU
EU
EU
DK
Adding the cases listed in Table 2 to Chows data set and
0.0
0 20 40 60 80 100 eliminating Lierstranda leads to a new total of 84 tests
L/D on piles driven in clay. As with sands, the number of
tests from any single geotechnical deposit has been
Figure 2. Distribution of Qc/Qm with respect to pile slen- limited to 10 statistically representative entries. The
derness ratio, L/D: API (1993) shaft procedure for sands.
reliability parameters assessed for the resulting data-
base of 68 tests are given in Table 4.
3.5
The ICP results are generally close to those quoted
3.0 by Jardine and Chow (1996), with the mean Qc/Qm and
COV both rising very slightly with this expanded data-
2.5
base. The API (1993) results are as before, indicating
no overall bias, but giving a COV far higher than that
Qc /Qm

2.0

1.5
found for the ICP.
SP
EU
SP
H
BD
EU DK
Figures 5 to 8 present plots of Qc/Qm against YSR
EU HH
1.0 SP
EU
EU
EU DK and pile slenderness ratio (L/D) drawn from the new
0.5 combined database. Unlike the current API RP2A pro-
cedures, the ICP approach shows no sign of skewing
0.0 with YSR or L/D and appears to be equally reliable
0 10 20 30 40 50 60 70 80 90 100
Relative density, Dr (%) under a wide range of circumstances.
The new ICP database does not show any system-
Figure 3. Distribution of Qc/Qm with respect to relative atic skewing with regard to plasticity index, as noted
density Dr: ICP shaft procedure for sands. also by Aldridge (2004) in an independent study of the
API database. However, Karlsrud et al. (1993) and
Clausen & Aas (2001) report non-conservative pre-
for open-ended and closed-ended piles. The slightly dictive bias for both the API and ICP procedures when
non-conservative bias indicated for open-ended piles applied to piles driven at the low plasticity, low YSR,
in Table 3 is not considered to be very significant, and Pentre and Lierstranda sites.
could well reverse if only first time tests on fresh piles Jardine et al (2005a) found that the ICP approach
are considered, working with tests performed at an age applied well to the large-scale LDP tests run on steel
of ten days. Several authors have suggested that the pipe piles at Pentre (see Clarke 1993). But they were

694

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 4. Overall results for peak shaft resistance in clay; 2.0
Pile type & test direction
total database of 68 tests. 1.8 Closed-ended, tension
O Closed-ended, compression
1.6 O Open-ended, tension
No No Mean COV 1.4 O
O
Open-ended, compression

Method pile tests sites Qc/Qm O


OMXMX MX
1.2

Qc /Qm
K MX MX
MX
1.0WD MX MX
ICP; all piles 68 19 1.03 0.20 K
MX
ICP; open-ended piles 25 13 0.99 0.17 0.8 MX

API RP2A (1993); all 68 19 0.99 0.33 0.6


piles 0.4
0.2
0.0
2.0 1 10 100
Pile type & test direction
1.8 O Closed-ended, tension YSR
K Closed-ended, compression
1.6 O
Open-ended, tension

1.4
Open-ended, compression Figure 7. Distribution of Qc/Qm with respect to YSR: ICP
O
OMX K
O shaft procedure for clays.
1.2 O
Qc /Qm

MX
O MXMX
1.0 WD MX MX
MX

0.8 MX MX
2.0
D
0.6 MX 1.8 O
K
0.4 1.6 O
0.2 1.4 K MX O
O
0.0 O
1.2 O
Qc / Qm

MX
1 10 100 O MX
MX
PT
HU MX
YSR 1.0 MX
WD MX

0.8 MX MX

Figure 5. Distribution of Qc/Qm with respect to YSR: API 0.6 MX

(1993) shaft procedure for clays. 0.4


0.2

2.0 0.0
0 20 40 60 80 100 120 140 160
1.8 O L/D
K
1.6 O

1.4 K O
Figure 8. Distribution of Qc/Qm with respect to pile slen-
MX O
1.2
O
O derness ratio, L/D: ICP shaft procedure for clays.
Qc /Qm

MX
O MX
MX
PT
HU MX
1.0 MX
WD MX

0.8 MX MX

0.6 MX

0.4 7 CONCLUSIONS
0.2
0.0 This paper has summarised the new and modified
0 20 40 60 80 100 120 140 160 cases considered, and the results obtained, in a sub-
L/D
stantial re-examination of the reliability of the API
Figure 6. Distribution of Qc/Qm with respect to pile slen- RP2A and ICP recommendations concerning the
derness ratio, L/D: API (1993) shaft procedure for clays. axial capacity of piles driven in sands and clays. The
study has included performing new site specific ring-
shear, index, oedometer and other laboratory testing
unable to match the exceptionally low capacities inter- at some key sites. Around 60 new pile tests, involv-
preted from medium term tests at Lierstranda, even ing nine widely different geotechnical settings, have
when applying the site-specific interface shear tests been considered.
performed by Ridgway (2004). However, Karlsrud et The re-assessment reinforces the main conclusions
al draw attention to Flaates (1968) finding that shaft drawn by Jardine & Chow (1996): the ICP capacity
capacities grow strongly in the medium to long term, procedures lead to far better reliability parameters
through as yet unidentified ageing processes. than current version of API RP2A for both silica sands
Jardine et al (2005a) suggest that pile driving moni- and clays. The ICP procedures have been extended to
toring combined with checks for long term set up cover a wider range of pile types and loading condi-
effects may be warranted in any cases where the pres- tions. Jardine et al (2005a) also give guidance on
ence of very low plasticity, low YSR clay-silts such as applying the methods to a wider range of geomateri-
those found at Lierstranda that might give rise to als, including carbonate and micaceous sands, very
potential concerns. low plasticity clay-silts and diatomaceous clays.

695

Copyright 2005 Taylor & Francis Group plc, London, UK


ACKNOWLEDGEMENTS Jardine, R.J. and Standing, J.R. 2000. Pile load testing per-
formed for HSE cyclic loading study at Dunkirk, France.
The Authors wish to thank colleagues at University 2 Vols. Offshore Technology Report OTO 2000 007;
College Dublin and Norwegian Geotechnical Institute Health and Safety Executive, London.
Jardine, R.J., Chow, F.C, Overy, R.F. and Standing J.R.
who provided test samples and to colleagues at 2005a. ICP procedures for driven piles in sands and
Imperial College London for their help with testing. clays. Thomas Telford Limited, London.
Jardine, R.J., Standing, J.R. and Chow, F.C. 2005b. Some
observations of the effects of time on the capacity of piles
REFERENCES driven in sand. Submitted to Gotechnique, December
2004.
Aldridge 2004. Personal Communication. Karlsrud, K., Nowacki, F. and Kalsnes, B. 1993. Response in
American Petroleum Institute 1993. RP2A-WSD: Recom- soft clay and silt deposits to static and cyclic loading
mended Practice of Planning, Designing and Construct- based on recent instrumented pile load tests. Proc. SUT
ing Fixed Offshore Platforms Working Stress Design, Int. Conf, Kluwer, Dordrecht, pp. 549584.
20th edition, Washington, pp. 5961. Lehane, B.M. 1992. Experimental investigations of pile
Bogard, D. and Matlock, H. 1998. Static and cyclic load test- behaviour using instrumented field piles. PhD Thesis,
ing of a 30-inch diameter pile over a 2.5-year period. Proc. Imperial College, London. approach. Proc. Conf. on the
Offshore Technology Conf., OTC 8767, pp. 455468. Behaviour of Offshore Structures (BOSS), pp. 2336.
Bond, A.J. 1989. Behaviour of displacement piles in over- Lehane, B.M., Jardine, R.J. and McCabe, B.A. 2004.
consolidated clays. PhD Thesis, Imperial College, London. Response of a pile group in clay to first time one-way
Chow, F.C. 1997. Investigations into displacement pile cyclic tension loading. Advances in Geotechnical
behaviour for offshore foundations. PhD Thesis, Imperial Engineering. Proc. Skempton Memorial Conference.
College, London. Pub. Thomas Telford, London, pp. 700710.
Chow, F.C., Jardine, R.J., Nauroy, J.F. and Brucy, F. 1997. Ridgway, A. 2004. A re-evaluation of driven pile loading
Time-related increases in the shaft capacities of driven tests at Lierstranda and Onsy, Norway. MSc. disserta-
piles in sand. Gotechnique, Vol.47, No.2, pp. 353361. tion, Imperial College, London.
Clarke, J. (ed) 1993. Conf. on Large-Scale Pile Tests in Clay. Saldivar-Moguel, E.E. 2002. Investigation into the behav-
Thomas Telford, London. iour of displacement piles under cyclic and seismic
Clausen, C.J.F. and Aas, P.M. 2001. Capacity of driven piles loads. PhD Thesis, Imperial College, London.
in clays and sands on the basis of pile load tests. Proc. Strick van Linschoten, C.J. 2004. Driven pile capacity in
Conf. ISOPE, Stavanger, Vol. 4, pp. 581586. organic clays, with particular reference to square piles in
Cowley, R.C. 1998. The effect of pile geometry on the design the Belfast Sleech. MSc dissertation, Imperial College,
of piles using the new Imperial College pile design London.
method. MSc dissertation, Imperial College, London. Thompson, G.W.L. and Jardine, R.J. 1998. The applicability
CUR 2001. Bearing capacity of steel pipe piles. Report of the new Imperial College Design method to calcareous
20018 Centre for Civil Engineering Research and sands. Proc. Conf. Offshore Site Investigations and
Codes, Gouda, The Netherlands. Foundation Behaviour. Society for Underwater technol-
Flaate, K. 1968. Baereevnen av friksjononspeler I leire. ogy, London, pp. 383400.
Veglaboratoret, 1968. Tomlinson, M.J. 1996. Presentation at British Geotechnical
Jardine, R.J. 1985. Investigations of pile-soil behaviour, with Society meeting Recent Advances in driven pile
special reference to the foundations of offshore struc- design. Institution of Civil Engineers, London, reported
tures, PhD Thesis, Imperial College, London. in Ground Engineering, Vol.29, No.10, pp. 3133.
Jardine, R.J. and Chow, F.C. 1996. New Design Methods for Williams, R.E., Chow, F.C. and Jardine, R.J. 1997.
Offshore piles, Marine Technology Directorate, London. Unexpected behaviour of large diameter tubular steel
Jardine, R.J., Overy, R.F. and Chow, F.C. 1998. Axial capac- piles. Proc. Int. Conf. on Foundation Failures, Singapore,
ity of offshore piles in dense North Sea sands. ASCE, 1213 May 1997.
JGE, Vol. 124, No.2, pp. 171178. Zuidberg, H.M. and Vergobbi, P. 1996. EURIPIDES, Load
Jardine, R.J. and Saldivar, E. 1999. An alternative interpretation tests on large driven piles in dense silica sands. Proc.
of the West Delta 58A tension pile research results. Proc. Offshore Technology Conf., Houston, OTC7977.
Offshore Technology Conf., Houston, OTC10827.

696

Copyright 2005 Taylor & Francis Group plc, London, UK


A general framework for shaft resistance on displacement piles in sand

D.J. White
Cambridge University Engineering Department

ABSTRACT: The stress history of a soil element adjacent to the shaft of a displacement pile in sand leads to a
general relationship for predicting shaft resistance. The stages in the stress history are linked to the penetration
process at the pile base, and subsequent friction fatigue along the pile shaft. The stress history gives each part
of the relationship physical significance, providing a rational basis for selecting design parameters. Recent field
and laboratory data is used to illustrate each stage of the stress history, with emphasis on the influence of cyclic
loading on friction fatigue and the resulting mean unit shaft resistance.

1 INTRODUCTION This paper aims to


present a general framework for the prediction of
The API design method for shaft resistance has been
shaft resistance based on the governing mech-
largely unchanged for 20 years and uses an earth pres-
anisms identified in recent laboratory and field
sure approach, in which the unit shaft resistance at
investigations.
failure, sf, is assumed to be proportional to the in situ
discuss, within this framework, how modern design
vertical effective stress up to a limiting value.
methods capture the governing behaviour.
Since 1990, a significant body of field data has
shown that an earth pressure approach is inappropriate,
and can be unconservative for long offshore piles. New
design methods, which assume a more realistic distri- 2 SOIL ELEMENT STRESS HISTORY
bution of shaft resistance have been proposed (e.g.
Lehane & Jardine 1994, Jardine & Chow 1996, CUR 2.1 Shaft resistance failure condition
2001, Randolph et al. 1994).
The input parameters to these design methods are In sand, the unit shaft resistance at failure, sf, is gov-
calibrated against databases of pile load tests. The erned by Coulomb friction (Equation 1) and depends
design equations are more complex than the current on the normal effective horizontal stress, rf and the
API approach, but give improved reliability. mobilized angle of interface friction, . To calculate
However, inaccurate predictions can arise in condi- static axial capacity for design, the task is to estimate
tions beyond the database. For example, Thompson & the maximum values of horizontal stress and interface
Jardine (1998) warn that the MTD method (Jardine & friction angle, , in order to establish the unit shaft
Chow 1996) for silica sand leads to a seven-fold over- resistance at failure, sf.
prediction of shaft resistance in carbonate sands. This (1)
discrepancy indicates that, although the MTD method
offers improved reliability in silica sands, not all of the
parameters governing shaft resistance are captured in A reliable design value of  can be selected if the
the method. particle size and pile shaft roughness are known, using
Such warnings are valuable. These new design well-established correlations (Uesegi & Kishida 1986,
methods are vulnerable to misuse since the physical Jardine et al. 1993). There is greater uncertainty in the
significance of each part of the calculation is not prediction of rf.
explicitly linked to the mechanisms that govern shaft The stress history of a soil element adjacent to a
resistance. Also, the teaching of modern pile design displacement pile in sand links rf to the in situ value
methods is problematic since design equations without h0. The stress changes at each stage of the installa-
an obvious physical origin do not sit easily alongside tion and loading process lead to a general expression
conventional course material. for shaft resistance.

697

Copyright 2005 Taylor & Francis Group plc, London, UK


2.2 Stages of stress history
In order to examine the change in horizontal stress
from h0 to rf (and sf), the history of a soil element
which finishes immediately adjacent to the shaft of a
displacement pile is considered Figure 1. The stress
path associated with these events is sketched in
Figure 2. During stages AC, when the soil element is
below the pile, the loading is expressed in terms of
the spherical and deviatoric stress, p and q. Beyond
stage C, when the soil element is adjacent to the pile
shaft, the loading history is expressed as the normal
and shear stresses applied to the pile shaft, r and s.

2.3 Stages AC: in situ conditions pile


shoulder
A soil element that will lie adjacent to the pile shaft
after installation initially lies beneath the pile tip. For
the case of a CPT (or a jacked pile), as the pile tip
approaches, the local mean effective stress rises from
the in situ value and can be approximated as the base
(qb) or cone (qc) resistance at that depth. In the case of a
driven pile the highest stress mobilised during each
hammer blow will be related to qb.
Since qb is typically two orders of magnitude greater
than the in situ mean stress, suppressed axes have been
used on Figure 2 to show the stress changes during Figure 1. Stages in the loading history of a soil element
stage AB. Noting the high strain level associated with adjacent to a displacement pile.
deformation close to the tip of a pile or CPT, it is
assumed that the soil will move up the failure line in
q-p space.
As the soil element passes the pile tip the stress
level reduces and the element exerts an upward shear
stress on the lower part of the pile shaft: the maximum
unit shaft resistance, max. For the case of a CPT, max
is equal to the sleeve resistance, fs (although fs is actu-
ally the average shear stress over 3 diameters behind
the cone tip). CPT sleeve friction lies in the range
0.51.5% qc for typical for silica sands (rising to
23% in some uncemented carbonate sands e.g. North
Rankin, Beringen et al. 1982), indicating a two orders
of magnitude drop in stress as the soil element passes
round the shoulder, which takes the stress state to
point C, having returned across the suppressed axes of
Figure 2.
It is notable that a higher friction ratio is observed in
some carbonate sands, given that these soils are associ-
ated with low pile capacity. A correlation based on qc
and calibrated for silica sand would give a conservative
prediction of fs at North Rankin. Figure 2. Loading history of a soil element adjacent to a
If the difference between dynamic pile driving displacement pile.
resistance and steady CPT penetration is ignored,
qb  qc and max  fs for a closed ended pile. For an unplugged manner, max may be lower than fs due to the
open-ended pile penetrating in a plugged manner, the lower radial displacement.
same assumption can be made since the soil flow pat- The area ratio, Ar, during an increment of penetra-
tern is identical. However, if the penetration is in an tion of an open-ended pile can be defined as the ratio

698

Copyright 2005 Taylor & Francis Group plc, London, UK


of the added volume (i.e. the gross pile volume minus
any soil entering the plug) to the gross pile volume
(Equation 21). Ar  1 corresponds to plugged penetra-
tion or a closed-ended pile. Ar  0 corresponds to a
vanishingly thin-walled pile for which there would be
negligible radial movement.

(2)

This stress path suggests a relationship for max of


the form shown in Equation 3, where a is analogous to
the cone friction ratio, Fr, and b is a positive index.
Alternatively, fs could be substituted for aqc:

(3)

The dependency of max on Ar can be explored using


the analogy of cavity expansion (White et al 2005). Figure 3. Contraction of an interface shear zone leading to
Ignoring volume change in the soil, the proportional friction fatigue, and a reduction in r (White & Bolton 2004).
increase in cavity radius at the pile wall is (1
Ar)
0.5.
The insertion of a closed-ended pile or CPT (Ar  1)
generates greater radial movement at a given radial
coordinate than an open-ended pile. Multiple configurations of this instrument provided
measurements of s between h  4D and 35D.
Negligible reduction in friction was evident over this
2.4 Stages CE: friction fatigue range: s ' f(h).
2.4.1 Mechanism of friction fatigue Similar behaviour was observed in jacked model
As a soil element is passed by the pile shaft, cycles of pile tests conducted at the North Rankin site, offshore
shearing are applied if the pile is driven or jacked. Western Australia. A series of tension tests were con-
Recent research has shown that the horizontal stress ducted on a small model pile of diameter 60 mm and
acting on the shaft of a jacked or driven pile at a given length 2500 mm (L/D  42) that was jacked below the
soil horizon decreases as the distance above the pile borehole base. The results from 25 tests (including
tip, h, increases with continued penetration. This effect three additional tests using a 36 mm diameter model
is termed friction fatigue (Heerema 1980), and is pile) are compared to the CPT profile for the site in
attributed to gradual densification of soil at the pile Figure 4. Although scatter is present, the model pile
surface with continued penetration (White & Bolton friction is comparable to the CPT sleeve friction; the
2002, Randolph 2003). This interface layer is restrained mean residual shaft resistance during pullout was
by the confining stiffness of the far field soil (4G/D). 54 kPa, compared to a mean CPT sleeve friction of
The operative G is high due to the over-consolidation as 55 kPa at the test depths. Despite the North Rankin soil
the pile tip passed, as seen in cone pressuremeter tests. conditions being considered problematic, the simple
Contraction of the interface layer reduces r (Figure 3). result that on a slender model pile (L/D  42), sf  fs
is found.
The good agreement between these model jacked
2.4.2 The influence of cycling on friction fatigue pile tests and fs measurements at North Rankin led to a
Field data and model tests indicate that friction fatigue design value of sf  40 kPa (Renfrey et al. 1988).
is better linked to the number of shearing cycles However, it was later recognized that this value over-
applied to the interface, N, rather than net sheared dis- predicted the full-scale driven pile capacity. A series of
tance, h. DeJong (2001) reports CPT soundings in pullout tests were carried out on 762 mm diameter
medium dense sand using an instrument equipped with hammer-driven well conductors after concerns arose
four friction sleeves and installed in 1000 mm strokes. following the ease of installation of the jacket piles.
These tests showed a far lower shaft resistance than the
jacked model pile tests.
1
Do  pile outer diameter, Di  pile inner diameter CAPWAP analyses from the installation of the
IFR  incremental filling ratio, (plug length)/(tip depth). main foundation piles at North Rankin A provide a

699

Copyright 2005 Taylor & Francis Group plc, London, UK


Shaft (or sleeve) friction (kPa)
0 50 100 150 200 250 300
0
Jacked
model
pile
20
Peak friction
Friction at 1D displacement
Depth below seabed (m)

Residual friction in 1st cycle


40
fs profile (Beringen et al 1982)
fs values (Renfrey et al 1988)
qc profile (Beringen et al 1982)
60

80

Figure 5. Influence of loading cycles (i.e. hammer blows)


100 on shaft resistance at North Rankin.

120
0.0 2.5 5.0 7.5 10.0 12.5 15.0
Cone tip resistance, qc (MPa)

Figure 4. Comparison of jacked model pile shaft resist-


ance and CPT sleeve friction at North Rankin (after Renfrey
et al. 1988).

link between the jacked model pile results and the


conductor load tests. Figure 5 shows the progressive
reduction in shaft resistance during continuous driv-
ing, deduced from CAPWAP analyses. Although
some recovery is evident after a redrive, presumably
due to pore pressure dissipation, the resistance soon
reverted to 15 kPa. Randolph (1988) suggested that
such a recovery could only occur once, since the
redrive would not re-generate the high pore pressure
created by the passage of the pile tip.
This influence of number of cycles, N, on pile shaft
capacity is widely recognized for post-installation
loading. Cyclic stability diagrams are widely used to
predict the degradation of shaft resistance with cycles
of varying magnitude (e.g. Randolph 1988, Lee & Figure 6. Comparison of mean normalized shaft resistance
Poulos 1993). It is logical that cycles applied during and installation cycles in silica and carbonate sands.
installation form part of the same degradation process.
White & Lehane (2004) describe a series of cen-
trifuge tests that measured r on a model pile installed A simple measure of friction fatigue is the ratio of
with and without cycles of shearing, then tested under mean shaft resistance, sf,av, to mean CPT sleeve fric-
cyclic loading. It was found that during continuous tion, fs,av along the embedded pile length. This is a
installation, r was independent of h (as evident in the very crude comparison since the actual number of
two field studies described previously). With cyclic cycles experienced by a given soil element varies with
installation and load testing, the loss of r correlated depth and the shape of the blowcount profile.
better with the number of cycles, N, than distance h, Load test data shown previously for North Rankin
where N included any cycles during installation. (NR) is augmented by a site-specific comparison of
The influence of installation cycles on friction jacked and driven piles at Dunkirk, France (Chow
fatigue is illustrated in Figure 6, using recent field 1997). Also, the shaft resistance mobilized during the
data from driven (high N) and jacked (low N) piles. four EURIPIDES tension load tests (Fugro 1996) has

700

Copyright 2005 Taylor & Francis Group plc, London, UK


been calculated for the distinct shallow (z 29 m,
qc 20 MPa) and deep (z  29 m, qc  40 MPa) sand
strata. In the deep stratum, the value of N includes
only the blows since the pile entered that layer.
A trend of increasing friction fatigue (reducing
sf,av/fs,av) with cycles is evident in Figure 6, with a
stronger effect in the NR carbonate sand, which is more
susceptible to contraction during shear.
Interestingly, the NR conductor tests (which are
widely quoted as an example of surprisingly low pile
capacity) and the EURIPIDES tension tests both mobil-
ized a mean shaft resistance of 25% of the mean fs
over the pile length. Their similarity when viewed in
this way suggests that separate design methodologies
may be unnecessary. Semple (1988) shows that car- Figure 7. Contraction under cyclic loading in a 1 m diameter
bonate and silica sands behave similarly when viewed interface ring shear test (Kelly 2001).
within a critical state framework, although the actual
parameters differ. It may be useful to treat pile behav-
iour the same. A schematic stress path for stages CE is shown on
Figure 2. The butterfly-shaped loops of normal and
2.4.3 Modelling of friction fatigue shear stress, as observed during interface testing, lead
Friction fatigue can be mimicked in constant normal to a progressive reduction in r with cycling. After the
stiffness interface shear box tests, which use an elastic final jacking stroke, or hammer blow, some base load
spring to model the confinement provided by the far may be retained by negative shaft friction on the upper
field soil. During cycling, a net reduction in volume is part of the pile, leaving the soil at stress state E. This
observed, even for dense sand, and the minimum nor- simple stress path overlooks the influence of excess
mal stress within each cycle reduces (Mortara 2001, pore pressures and Poisson strains generated during
DeJong et al. 2003). Comparable patterns of horizon- driving, but serves to highlight the influence of cycling
tal stress degradation during loading of a model pile on r.
are reported by White & Lehane (2004).
The large shear displacements that can be achieved
in the 1 m diameter Sydney University ring shear 2.5 Stage EF (G):application of working
apparatus allow better simulation of pile-soil interface load
behaviour than conventional shear boxes. Figure 7
Stage EF, in which the pile is loaded to failure (or
shows results from 4 tests on dense sand conducted
subjected to a cyclic working load-stage G) can be con-
under constant normal stress (Kelly 2001). After
sidered as an extension of any cycles imposed during
initially dilating, each sample continues to shear at
installation, as shown on Figure 2.
constant volume, until a series of 10 displacement-
An additional effect during the working life of the
controlled cycles are applied. These demonstrate that
pile is set-up. The mechanism in figure 3 leads to ini-
the number of cycles has a greater influence on the
tial conditions for such an increase in r since this con-
contraction of the interface layer rather than the net
traction and unloading will cause the circumferential
displacement. Also, a higher amplitude of cycling
stress to exceed the radial stress (White & Bolton 2004,
leads to greater contraction.
White et al. 2005). Any creep-induced stress equaliza-
An elastic cavity contraction calculation links vol-
tion will increase r.
ume change in the interface layer to changes in r
(Equation 4). For realistic values of cavity stiffness,
4G/D, an interface contraction of a few microns leads 2.6 Prediction of s during stages CF(G)
to a significant loss of r (and sufficient strain to
reduce G, so the response is not truly elastic). This In the analysis of a pile during working conditions, it
high lateral stiffness is evident in the high shaft capac- is widely accepted that pile shaft capacity degrades
ity of tapered piles; only a small taper is necessary to under cyclic loading. To predict the change in shaft
lock-in high friction. However, on a straight pile, resistance between max (stage C, Equation 3) and sf
any contraction of the interface layer leads to a signif- (stage F) requires this mechanism to be extended into
icant loss of friction. the installation phase.
However, to explicitly model the cyclic behaviour
(4) during installation is not realistic within a design
method, since N is unknown prior to pile driving. For

701

Copyright 2005 Taylor & Francis Group plc, London, UK


design, a pragmatic approach is to model the friction relates to the partial displacement of soil around
degradation from max to sf using a reduction factor the tip of an open-ended pile, which leads to a
linked to h, and non-dimensionalised using pile diam- lower radial stress and unit shaft resistance at the
eter, D (e.g. Lehane & Jardine 1994, Jardine & Chow pile shoulder.
1996, CUR 2001, Kolk et al. 2005) (Equation 5), where relates to the degradation of horizontal stress as the
c is a negative index. A minimum value of h must pile tip penetrates further.
be specified in Equation 5 to avoid the prediction of
infinite friction at the pile tip. Research and field data suggests higher degradation
with stiffer soil and smaller piles (i.e. higher 4G/D),
carbonate soils, (which contract more when sheared),
(5) and increasing cycles (which cause contraction in the
shear zone at the pile-soil interface).
Equation 6 is similar to the Jardine & Chow (1996),
If max is assumed to act 1 diameter behind the pile CUR (2001) and Kolk et al. (2005) equations for shaft
tip (which is close to the midpoint of a CPT friction resistance in sand, although these methods were not
sleeve), and ignored below that point, Equation 5 can be derived in the two-step format described by Equations
integrated from h  D to L to give the ratio sf,av/max. 3 and 5, which separates the influence of friction
Curves of this ratio as a function of index c are shown fatigue from stress changes at the pile tip.
on Figure 6 for the case of max being constant with
depth (and h). A reasonable fit to the scattered data is
achieved by c 
(log N)/8. Similar logarithmic 4 CONCLUSIONS
models are used to predict cyclic shaft friction decay
under storm conditions. By considering the stress history of a soil element adja-
The CUR (2001) and Fugro (Kolk et al. 2005) design cent to the shaft of a displacement pile, a general rela-
methods, which were calibrated against a database of tionship for predicting shaft resistance has been
large open-ended driven piles (high N), use an index of described. The stress history is divided into two stages;
c 
0.90 (applied to h/R). In contrast, the Jardine & firstly, the generation of high unit shaft resistance at the
Chow (1996) method, which was initially calibrated pile tip, and secondly the degradation of this resistance
using sf results from jacked piles (low N), uses a through friction fatigue. CPT friction sleeve results are
value of c 
0.38. shown to model the first stage. Field data and labora-
The simple comparison in Figure 7 overlooks the tory simulations link the second stage, friction fatigue,
influence of end condition. The low shaft resistance on to cyclic loading.
the open-ended piles could be due to the lower area This framework provides a rational basis for mod-
ratio, Ar, leading to a lower max. It is difficult to separ- ern pile design methods, within which the underlying
ate the effects of end condition (open- or closed-ended: mechanisms are apparent, and carbonate and silica
low or high Ar) from installation method (driven or sands show comparable behaviour.
jacked: high or low N), since most high quality load test
data is from open-ended driven piles or closed-ended
jacked piles. ACKNOWLEDGEMENTS

3 PREDICTION OF SHAFT RESISTANCE Valuable discussions with Barry Lehane, Mark


Randolph and James Schneider during the preparation
Equations 3 and 5 lead to an expression for the pre- of this paper are acknowledged.
diction of shaft resistance from cone resistance, qc:

REFERENCES
(6)
Beringen F.L., Kolk H.J. & Windle D. 1982. Cone penetra-
tion and laboratory testing in marine calcareous sediments.
This expression includes three empirical indices, each Geotechnical properties, behaviour and performance of
linked to a stage in the stress history during installa- calcareous sediments, ASTM STP777. 179209.
tion (and loading) shown in Figure 2: Chow F.C. 1997. Investigations into the behaviour of dis-
placement piles for offshore foundations. PhD thesis,
relates to the fractional drop in stress from qc to University of London (Imperial College).
max as the soil passes the pile tip. Field data from CUR. 2001. Bearing capacity of steel pipe piles. Centre of Civ.
smooth model piles suggests that a  FR. To expli- Engng. Res. & Codes (CUR) Rep. 20018, Gouda, NL.
citly consider pile roughness, Equation 6 could be DeJong J.T. 2001. Investigation of particulate-continuum
recast with a  rf/qc, and a tan  term added. interface mechanisms and their assessment through a

702

Copyright 2005 Taylor & Francis Group plc, London, UK


multi-friction sleeve penetrometer attachment. PhD Randolph M.F., Dolwin J. & Beck R. 1994. Design of driven
thesis, Georgia Inst. Technology, Atlanta USA. piles in sand. Gotechnique 44(3):427448.
DeJong J.T., Randolph M.F. & White D.J. 2003. Interface load Randolph M.F. 2003. Science and empiricism in pile design.
transfer degradation during cyclic loading: a microscale Gotechnique 53(10):847875.
investigation. Soils & Foundations 43, (4):8193. Renfrey G.E., Waterton C.A. & van Goudoever P. 1988.
Fugro. 1996. EURIPIDES database report, Vols. 15, Fugro Geotechnical data used for the design of North Rankin
Engineers BV, The Netherlands & Geodia SA, France. A platform foundation. Proc. Int. Conf. on Calc.
Lee C.Y. & Poulos H.G. 1993. Cyclic analysis of axially Sediments. 343356.
loaded piles in calcareous soils. Canadian Geotech. J. Ripley I., Keulers A.J.C. & Creed S.G. 1988. Conductor load
(30):8295. tests. Proc. Int. Conf. on Calc. Sediments, Perth. 429438.
Lehane B.M. & Jardine R.J. 1994. Shaft capacity of driven Semple R.M. 1988. The mechanical properties of carbonate
piles in sand: a new design method. Proc. VII Int. Conf. on soils. Proc. Int. Conf. on Calc. Sediments, Perth. 807836.
the Behaviour of Offshore Structures, Boston, 1, 2336 Thompson G.W.L. & Jardine R.J. 1998. The applicability of
Heerema E.P. 1980. Predicting pile driveability: Heather as the new IC pile design method to carbonate sands. Proc.
an illustration of friction fatigue. Ground Engng Conf. Offshore Site Invest. & Fndn Behaviour, Aberdeen.
13(Apr):1537. 383400.
Jardine R.J. & Chow F.C. 1996. New design methods for Uesegi M. & Kishida H. 1986. Influential factors of friction
offshore piles. Marine Tech. Directorate, London between steel and dry sands. Soils and Foundations.
MTD96/103 26(2):3346.
Jardine R.J., Lehane B.M. & Everton S.J. 1993. Friction coef- White D. J. & Bolton M.D. 2002. Observing friction fatigue
ficients for piles in sands and silts. Vol. 28, Proc. Conf. on on a jacked pile. Centrifuge and Constitutive Modelling:
Offshore Site Inv. & Fndn Behaviour, London. 147180. Two extremes. Springman S.M. (ed.) Balkema. 347354.
Kelly R. 2001. Development of a large diameter ring shear White D.J. & Bolton M.D. 2004. Displacement and strain
apparatus and its use for interface testing. PhD thesis. paths during plane strain model pile installation in sand.
Univ. Sydney. Gotechnique. 54, (6):375398.
Kolk H.J., Baaijens A.E. & Senders M. 2005. Design criteria White D.J. & Lehane B.M. 2004. Friction fatigue on dis-
for pipe piles in silica sands. Proc. Int. Symp. on placement piles in sand. Gotechnique. 54, (1): 645658.
Frontiers in Offshore Geotechnics, Perth. White D.J., Schneider J.A. & Lehane B.M. 2005. The influ-
Mortara G. 2001. An elasto-plastic model for sand-structure ence of effective area ratio on the shaft friction of dis-
interface behaviour under monotonic and cyclic loading. placement piles in sand. Proc. Int. Symp. on Frontiers in
PhD thesis. Politecnico di Torino, Italy. Offshore Geotechnics.
Randolph M.F. 1988. The axial capacity of deep foundations
in calcareous soils. Proc. Int. Conf. on Calc. Sediments.
837858.

703

Copyright 2005 Taylor & Francis Group plc, London, UK


Field research into the effects of time on the shaft capacity of
piles driven in sand

R.J. Jardine & J.R. Standing


Imperial College London, UK

F.C. Chow
WorleyParsons Pty Ltd, Australia, formerly Imperial College London

ABSTRACT: The paper describes new field research into the effects of time on the shaft capacity of nine sep-
arate steel pipe piles driven at the Dunkirk dense sand test site. Remarkable increases are shown to develop,
with important differences being observed between the Intact Ageing Characteristic found from fresh, previ-
ously unfailed, piles and those re-tested after prior failures. The results explain why some earlier investigations
into ageing phenomena have produced widely scattered and sometimes confusing results.

1 INTRODUCTION (2000) and Jardine et al (2005b), provide more com-


plete descriptions of the site conditions, installation of
Chow et al (1997), reported field tests on steel piles the R1 to R6 (and C1) piles, the static load tests on all
driven in sand near Dunkirk, northern France that iden- piles and their interpretation.
tified remarkable increases in shaft capacity with time.
These data and other cases from the literature led
Jardine & Chow (1996), to suggest a tentative semi- 2 SITE CONDITIONS
logarithmic ageing relationship for shaft resistance.
Chow et al also considered possible mechanisms for the The test site comprises an area of around 1 hectare,
field data, suggesting that interaction between creep beneath which is found hydraulic silica sand fill over
and an arching action set up around the piles led to post glacial medium sized silica marine sands, with
gains with time of radial effective stresses, and hence occasional sand bands containing organic material;
shaft resistance. groundwater is encountered at around 4 m depth.
This paper summarises further research into the Figure 1 shows six of the CPT traces obtained within
same processes that involved several new (456 mm OD) the GOPAL pile test area, identifying the layering and
steel pipe piles driven in 1998 at the Dunkirk site as showing the predominantly dense in-situ conditions
part of the GOPAL jet grouting research programme (with 10 MPa qc 40 MPa). A wide range of in-situ
(Parker et al 1999). Multiple tests on the six 19 m long and laboratory testing has been conducted in conjunc-
reaction piles (R1 to R6), one 10 m long control pile tion with the piling research; Chow (1997), Connolly
(C1), and two 324 mm OD piles driven to 22 m in (1996), Kuwano (1999), Jardine & Standing (2000),
December 1988 by the CLAROM group (Brucy et al Jardine et al (2001).
1991a, b), enabled the effects of time on fresh (pre-
viously untested piles) to be distinguished from those
applying to multiple re-tests on previously failed piles. 3 PILE LOAD TESTS
Almost all ageing studies to date e.g. Axelsson
(2000), have involved the latter type of test, leading to Piles R1 to R6, along with C1 (and a subsequently jet
potentially confusing results. grouted pile) were driven in August 1998. A series of
Only the key results can be discussed here; Brucy 20 tension tests and one compression test was per-
et al (1991a, b), give more details of the dynamic driving formed on these piles at various intervals over the fol-
analysis of the CLAROM piles and the re-strike tests lowing eight months. These data were combined with
that were performed on them 5 months after their earlier tests performed by Chow (1997), on the two
installation. Chow et al (1997), Jardine & Standing CLAROM piles that were performed around 2070

705

Copyright 2005 Taylor & Francis Group plc, London, UK


reaching a maximum of 0.08 mm/min for the final
loading stage. Most tests took between 10 and 20 hours
to reach failure. The slow testing rates made it possible
in many cases to control loads after reaching peak
capacity so that data could be gathered even when the
response was brittle. However, more rapid tension tests
to failure were included in some cases for operational
reasons.

4 RESULTS FROM FIRST TIME TESTS

The key data from first time static tension tests to fail-
ure on three nominally identical 19 m long piles
(R1, R2, R6) tested at different ages are presented in
Figure 2. The loaddisplacement curves of these aged
Intact piles are virtually identical up to 900 kN (sug-
gesting little effect of time on initial stiffness) but the
235 day capacity is around 2.3 times that available at
9 days. Driving records and non-destructive proof
loading tests demonstrated that the large differences
seen between the piles were due to age and not local
variations in soil properties or installation details.
The trend of first time peak shaft capacity Qs with
time (t) after driving is presented in Figure 3. Here the
data are expressed in terms of the ratio between meas-
ured shaft resistance Qs(t), accounting for pile and plug
submerged self weight, and the short term shaft cap-
acity Qs ICP predicted by the CPT based ICP method
described by Jardine & Chow (1996) and Jardine et al
(2005a). The predictions allowed for minor variations
between installations by adopting the nearest CPT pro-
file to each pile and its final recorded penetration. The
ICP normalization also allowed the average of Chows
first time tension tests on the 22 m long, 324 mm OD,
CLAROM piles to be added (accounting for their dif-
ferent dimensions and nearest CPT profiles) as well as
Figure 1. CPT test profiles for test site. Sounding codes the shaft resistance interpreted from the single first time
refer to nearest test piles (R1, R2, C1 etc.). compression test on C1, the 10 m long control pile.
It has been assumed that the capacity does not
change over the first 24 hours and that the nominal
days after driving. As noted in the introduction, the two early 1-day capacity is the same as the End of Driving
22 m long piles were subjected to re-strike tests 5 resistance assessed by Brucy et al (1991a and b), from
months after installation. However, unlike the shorter the instrumented pile driving. The early 1 day capacity
(11 m long) CLAROM piles reported by Chow et al point plotted is added to the five static measurements
(1997), neither of the 22 m long piles had been sub- to complete the capacity-time plot, showing a sharp
jected to any prior static load test to failure. Some of increase in time over the first 8 months after driving.
the pile tests were also subjected to cyclic loading The steep Intact Ageing Characteristic (IAC) plotted
(Jardine & Standing 2000). on Figure 3 indicates that the ICP calculations match
The static pile tests were load controlled, with the capacity available about 10 days after driving, a
between 15 to 30 increments being applied before result that has been confirmed in a broader study con-
failure developed. Loading progressed according to ducted by the authors (Jardine et al 2005a). An assess-
pre-set criteria designed by the authors. As detailed by ment of ageing effects is required if capacities are to be
Jardine & Standing (2000), the pause periods were estimated at younger or greater ages.
extended until fixed creep criteria had been met. The The IAC plotted on Figure 3 contrasts sharply with
initial creep rate targets were set at 0.01 mm/min and Jardine & Chows (1996), tentative trendline. As shown
then raised in steps in the later stages of the test, below, this discrepancy is related primarily to the

706

Copyright 2005 Taylor & Francis Group plc, London, UK


3500
R1 - 1st test 02/09/1998 ( 9 days)
Similar results were obtained in re-tests on all of the
R2 - 1st test 17/04/1999 (235 days) other piles as may be seen in Figure 5 where traces are
Force applied to pile head - tension positive (kN)

R6 - 1st test 09/11/1998 (81 days)


3000
shown for piles R5 and R6, plotted in terms of the
normalised ICP capacities introduced in Figure 3. It
2500
is apparent that any aged pile loses a substantial part of
2000
its capacity as it undergoes first time failure and relax-
ation to zero load.
1500 The capacity reductions are not obvious from the
first-time loaddisplacement plots. Although the over-
1000 all loading stage plots appear to show a ductile
response, re-tests performed immediately afterwards
500
can show greatly reduced capacities. The combined
data set showed that these immediate losses of capacity
0
0 5 10 15 20 25 30 35 increased as a function of pile age. If the pile is allowed
Pile head displacement (mm)
to rest, it develops a degree of capacity recovery. But
Figure 2. Load displacement behaviour of three first pre-tested piles appear to always fall short of their Intact
time tension tests to failure performed at different ages after equivalents over the age range investigated.
driving. The ageing characteristics of pre-tested piles are
therefore non-unique and depend on (i) when the first
test was performed and (ii) the subsequent history of
testing. Nevertheless, it is interesting that the re-tested
3
capacities tended (with all piles) to scatter around the
Intact Ageing Characteristic (IAC) trendline proposed by Jardine & Chow (1996), as illus-
2.5
R2 CLAROM trated in Figure 5. The gap between the latter trend-line
22 m long
2 and the IAC gives an impression of the potential
C1
growth of brittleness with time.
Qs(t)/QsICP

R6
1.5

R1
1 6 RE-ANALYSIS OF CHOWS DATA BASE
Jardine and Chow
(1996) trendline
0.5 The above findings prompted a re-analysis of Chows
original data base. All re-tests and re-strikes were elim-
0
1 10 100 1000 10000 inated, leading to a greatly reduced dataset. However,
Time after driving (days) as shown in Figure 6, the surviving results (that include
closed ended concrete piles) fall close to the Intact
Figure 3. Tension shaft capacity-time trends for first time Ageing Characteristic interpreted in Figure 3 for the
load tests on steel pipe piles at Dunkirk. Data normalized by recent Dunkirk tests. Note that in this diagram the val-
ICP shaft capacity predictions. ues of Qs (t) have been normalised by the nominal 1
day capacities found from pile driving monitoring as
at Dunkirk.
The three additional sites to Dunkirk relate to the
different behaviours of fresh piles, and those that have Jamuna Bridge site in Bangladesh (Tomlinson 1996)
been failed previously. and two sites in the USA reported by Bullock &
Schmertmann (1995). Soil conditions at Jamuna Bridge
comprise loose to medium dense silty medium fine
5 RE-TESTS ON PREVIOUSLY FAILED PILES micaceous sand and the piles tested were predominantly
concrete piles, 400 and 450 mm square and 20 to 30 m
All of the GOPAL piles were subjected to more than long for the approach structure. One 762 mm diameter
one set of load tests. Figure 4 presents load-deflection tubular open-ended pile was also driven to 78 m depth
traces from re-tests on piles R1 and R3 as examples, and tested statically. The sites in the USA comprised
along with the upper and lower bound first time tests one dense sand site and one loose sand site with con-
(on Piles R2 and R1 respectively) as plotted in Figure 2. crete driven piles,
As before, the piles respond in a similar way on loading Insufficient information existed to compute the
up to around 900 kN but then diverge. The previously ICP Qs values corresponding to each test pile in
failed piles are unable to mobilize the same capacities Figure 6. It appears that the ageing processes lead to
as the Intact installations, and the older retested piles similar results in a spread of sand conditions and pile
showed markedly brittle post-peak characteristics. types.

707

Copyright 2005 Taylor & Francis Group plc, London, UK


3500
R1 - 1st test 02/09/1998 ( 9 days)
R1 - 2nd test 28/10/1998 (57 days)
R1 - 3rd test 26/04/1999 (239 days)
3000 R2 - 1st test 17/04/1999 (235 days)

Force applied to pile head - tension positive (kN)


R3 - 2nd test 20/04/1999 (85 days*)
* test performed same day as first,
but after episide of cyclic loading
2500

2000

1500

1000

500

0
0 5 10 15 20 25 30 35
Pile head displacement (mm)

Figure 4. Load displacement behaviour in tension during re-tests on piles R1 and R3, compared with first-time tests on
R1 and R2.

2.5
Dunkirk
IAC
2 R6
Qs(t)/QsICP

R5(1) & R6(1)


R6(2)
1.5 R5(2)
R6(4)
R6(3) R6(5)
1 R5(3)
R5(1) & R6(1) Jardine and Chow
(1996) trendline

0.5
Line representing
capacity at end of driving

0
1 10 100 1000 10000
Time after driving (days)
Notes:
R5(1) - virgin path for pile R5 (end point estimated to be similar to that for R6)
R5(2) - decrease in capacity from two phases of cycling testing (end point proven by second test)
R5(3) - increase in capacity (end point proven by third test)
R6(1) - virgin path for pile R6 (end point proven by first test)
R6(2) - decrease in capacity from one phase of cycling (end point proven by second test)
R6(3) - further decrease in capacity following another phase of cyclic loading (end point proven by third test)
R6(4) - increase in capacity (end point estimated by drawing line parallel to that proven for R5(5))
R6(5) - decrease in capacity from cyclic failure (end point proven by fourth test)

Figure 5. Normalised tension shaft capacity-time trends for previously failed piles R5 and R6 compared to Intact Ageing
Characteristic of Dunkirk piles and Jardine & Chow (1996) trendline.

708

Copyright 2005 Taylor & Francis Group plc, London, UK


5.0 7. Pre-tested piles show brittle load-displacement
Bullock & Schmertmann (1995)
4.5 Tomlinson (1996) Jamuna mica sand curves and develop further losses in capacity dur-
4.0 ing unloading.
3.5
8. Extreme loading cycles, including pre-testing to
failure and subsequent unloading, degrade shaft
Qs(t)/Qs(t=1)

3.0
capacity and disrupt the growth of capacity with
2.5 Dunkirk
IAC
time.
2.0
Jamuna Bridge 9. The capacity of pre-failed piles recovers with
1.5 tubular pile
time, but at relatively modest rates, giving non-
1.0 monotonic time-capacity traces that plot well
0.5 below the fresh piles IAC.
0.0
0.1 1 10 100 1000 10000
10. The capacity-time traces followed by pre-failed
Time after driving (days) piles are not unique but depend on the sequence
and intensity of prior testing.
Figure 6. Comparison between normalised Dunkirk IAC 11. While high-level cycling is damaging, low-level
results and Chows data set of intact aged piles. load cycling was found to accelerate the benefi-
cial ageing processes in critical experiment run at
Dunkirk.
12. It is essential to separate tests on fresh and pre-
Jardine et al (2005b), discuss the results further
tested piles when studying ageing effects. Elimi-
and report additional experiments. They note that
nating re-strikes and re-tests from Chows dataset
high level cyclic loading degrades capacity but
leads to a subset of first-time tests on fresh
demonstrate that low level cycling can enhance the
piles that conforms with the fresh pile IAC rela-
ageing processes. They also conclude that the com-
tionship found at Dunkirk.
bined field data set is consistent with the explanation
13. Pile capacity calculation procedures that take no
that interaction between creep and an arching mecha-
account of time will be subject to considerable
nism is responsible for the ageing characteristics of
error unless they consider only a tightly specified
piles driven in sand.
age range. In the case of the ICP procedures, the
standard calculation is most likely to match field
capacities in tests conducted around 10 days after
7 SUMMARY AND CONCLUSIONS
installation.
14. The re-tests described in this paper give shaft
A programme of first-time loading and re-test experi-
capacities that scatter sporadically around the ini-
ments has befigen performed on open-ended steel
tially suggested trendline, while the first time
pipe piles driven in a predominantly dense silica marine
IAC lies far above it.
sand at Dunkirk, northern France, leading to the fol-
15. Chow et al (1997), suggested that the time
lowing main conclusions:
dependency of shaft capacity for piles driven in
1. Fresh piles develop substantial increases in shaft sands is due to the radial stresses developed on
capacity in the months following driving, defining the shaft growing through a relaxation (with time)
an Intact Ageing Characteristic (IAC). of a circumferential arching stress field. The test
2. The ageing processes commenced within a few days programme did not include any radial stress
of driving. measurements. But the field test data (including
3. Shaft capacities rose until an apparently stable the observation that low-level cycling can accel-
plateau was attained after around eight months. erate capacity growth) are at least consistent with
4. The long term capacities were more than double this explanation.
those seen in load tests conducted at the typical
testing age (i.e. a few days after driving).
5. It is not certain how scale affects shaft capacity Chow et al (1997), also suggested that the local shaft
variations with time, although both small and large resistances may also grow with time due to changes in
piles show considerable changes within six the degree of restrained dilation that develops as the
months of driving. shaft is loaded to failure. The degree of radial expansion
6. Piles driven in sand develop relatively ductile required for sand grains to unlock from the pile shaft,
load-displacement curves in their first load test to and the stiffness of the restraining soil mass may
failure. However, they are unable to develop the increase with time. Axelsson (2000), reports meas-
same peak capacity when re-tested immediately urements that support the above hypotheses. Physico-
afterwards, indicating that capacity is lost during chemical processes such as corrosion may also play a
the unloading stage. role in some circumstances.

709

Copyright 2005 Taylor & Francis Group plc, London, UK


ACKNOWLEDGEMENTS Chow, F.C. 1997. Investigations into displacement pile
behaviour for offshore foundations. PhD Thesis, Univ.
The above research was funded by the EU (through London (Imperial College).
the GOPAL project) and the UK Health and Safety Chow F.C., Jardine, R.J., Brucy, F. and Nauroy, J.F. 1997.
Time related increases in the shaft capacities of driven
Executive (HSE) and their support is acknowledged piles in sand. Gotechnique, 47, No. 2, pp 353361.
gratefully. The authors thank Mr Eric Parker of Jardine, R.J. and Chow, F.C. 1996. New design methods for
DAppolonia, Genova for his major contribution to offshore piles. MTD Publication 96/103, MTD, London.
the project and the Port Autonome de Dunkerque for Jardine, R.J. and Standing, R.J. 2000. Pile load testing per-
their generous loan of the site. The pile testing was formed for HSE cyclic loading study at Dunkirk, France.
performed in conjunction with Precision Monitoring Two Volumes. Offshore Technology Report OTO 2000 007;
and Control (PMC) of Cleveland (UK); the labora- Health and Safety Executive, London. 60p and 200p.
tory work at Imperial College was conducted by Jardine, R.J., Kuwano, R., Zdravkovic, L. and Thornton, C.
Dr Reiko Kuwano and Mr Tim Connolly; the in-situ 2001. Some fundamental aspects of the pre-failure
behaviour of granular soils. Proc. 2nd Int Symp. On Pre-
soil testing was performed by Building Research failure Behaviour of Geomaterials, IS- Torino, Volume 2.
Establishment (Garston, UK) and Simecsol of Swets & Zeitlinger, Lisse, pp1077113.
Dunkerque, France. Jardine, R.J., Chow, F.C., Overy, R.F. and Standing, J.R.
2005a. ICP design methods for driven piles in sands and
clays. Thomas Telford, London.
Jardine, R.J., Standing, J.R. and Chow, F.C. 2005b. Some
REFERENCES observations of the effects of time on the capacity of piles
driven in sand. Submitted to Gotechnique, December
Axelsson, G. 2000 Long term set up of driven piles in sand. 2004.
PhD Thesis, Royal Institute of Technology, Stockholm, Kuwano, R. 1999. The stiffness and yielding anisotropy of
Sweden. sand. PhD Thesis, Imperial College, University of
Brucy, F., Meunier, J. and Nauroy, J.F. 1991a Behaviour of London.
pile plug in sandy soils during and after driving. Proc Parker, E. J., Jardine, R.J., Standing, J.R. and Xavier, J. 1999.
23rd Offshore Technology Conference, Houston, OTC Jet grouting to improve offshore pile capacity. Proc.
6514, pp 145154. Offshore Technology Conference, Houston, OTC 10828.
Brucy, F., Nauroy, J.F., Le Tirant, P. and Meunier, J. 1991b Tomlinson, M.J. 1996. Presentation at British Geotechnical
Comparison of static and dynamic tests of piles in sand. Society meeting Recent Advances in driven pile design,
Proc Conf. Foundations Profondes, Paris, pp 369378. 2nd October 1996, Institution of Civil Engineers, London,
Bullock, P. & Schmertmann, J.H. 1995, Personal reported in Ground Engng 29(10), pp 3133.
Communications.
Connolly, T. 1996. Hollow cylinder tests on Dunkerque
sand. Internal Report, Imperial College London.

710

Copyright 2005 Taylor & Francis Group plc, London, UK


Design criteria for pipe piles in silica sands

H.J. Kolk & A.E. Baaijens


Fugro Engineers B.V., The Netherlands

M. Senders
University of Western Australia, Perth, previously Fugro Engineers B.V., The Netherlands

ABSTRACT: Previous studies (e.g. Hossain & Briaud 1993, Jardine & Chow 1996) have suggested that CPT-
based design criteria for axial capacity of driven piles in silica sands are more reliable than design criteria based
on soil boring data. These studies were based on a limited number of load tests on driven steel piles at sites
where CPT data were available. To confirm this suggestion, a data base was made consisting of sets of high
quality soil and pile load test data which met the following criteria: i) driven open and closed-ended steel piles,
ii) reliable pile load tests, and iii) well defined soil conditions, including CPT data. Forty-five pile load tests that
meet these criteria were found. In this paper, the database of load test results is compared with current offshore
design criteria (API 1993, MTD 1996) plus a new design method (Fugro 2004).

1 INTRODUCTION for axially loaded, open-ended, driven pipe piles in


sands (Fugro 2004). This study is summarised in this
The traditional method in offshore design practice for paper and two companion papers (Kolk et al. 2005a,
predicting axial capacity of piles driven into silica sands Kolk et al. 2005b).
uses criteria specified in API RP 2A (1993). The pre-
dicted unit friction and end bearing values are pre-
dominantly based on soil classification and estimated
relative density. Various recent comparisons (Hossain & 2 PILE LOAD TEST DATA BASE
Briaud 1993, Jardine & Chow 1996, Fugro 2004) of
pile load test results and axial pile capacities predicted A survey of the public literature and industry files
according to the API RP 2A for piles in sand have was made in order to find pile load tests in sand that
proven that this method is unreliable. Predictions tend meet the following criteria:
to be conservative for relatively short piles in very driven open and closed-ended steel piles
dense sand, while they tend to be unconservative for reliable pile load tests
relatively long piles driven into loose sands. well defined soil conditions, including CPT data.
Various design methods have been developed during
recent years, which suggest that considerably more This survey found 45 pile load tests (24 tension and 21
accurate capacity predictions can be made through compression) that meet these criteria (Table 1). This
direct correlation of pile friction and end bearing with includes the EURIPIDES and Ras Tanajib II pile load
the results of in-situ Cone Penetration Tests (CPTs). test data presented by Kolk et al. (2005b) and Kolk et al.
The main reason why CPT based design criteria have (2005a) respectively. This number compares with 22
not yet replaced those in API RP 2A is probably that pile load tests on steel piles at sites with CPT data
insufficient representative pile load test data were considered by Jardine & Chow (1996) for developing
available to support new criteria. Fortunately, results their design approach.
of various high quality load tests have now become All tests were assigned a Data Quality Factor (DQF)
available. ranging from DQF1 (just acceptable) to DQF5 (very
In view of these recent developments, Fugro per- reliable) based on a (subjective) assessment of the
formed a study on behalf of API which included data reliability of soil data and pile load tests. All sands were
and information needed to improve pile design criteria silica, except those at Jamuna which were micaceous.

711

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 1. Pile load tests in sand (Fugro 2004). For typical offshore piles, ultimate unit friction is
approximately given by (ignoring a small additional
Type Tests Tension Compression contribution attributed to dilation during loading):
Open Euripides 4 (1) 4 (1)
Dunkirk 2 (4) 2 (2) (3)
Hoogzand 2 2
Jamuna 3 (1) 3
Padre Island 2 (4)
Leman Bank 2
Ras Tanajib II 1 1
Subtotal 16 (6) 12 (2)
where a, b and c  empirical coefficients/exponents
specified in Jardine & Chow (1996); f  friction
Closed Hoogzand 1 1 between sand and steel from laboratory shear test;
Ogeechee River 1 4
Akasaka - 1
pa  atmospheric pressure (100 kPa); h  distance
San Francisco - 1 between pile tip and level of pile wall for which unit
Subtotal 2 7 friction is computed; Ro  pile outer radius; Ri  pile
OC Total 18 (6) 19 (2)
inner radius.
The end bearing resistance for unplugged open-
Note: Numbers in parentheses refer to tests performed 200 ended piles is computed according to:
days after installation. All other tests were done sooner.
(5)

3 DESIGN METHODS FOR PILES IN SAND


and for fully plugged pipe piles according to:
3.1 API method
(6)
In the API recommended design method, the ultimate
unit friction, f, and unit end bearing, qu, in granular
soils are determined according to: where Qb,u  pile tip resistance of unplugged pile at
pile head displacement of 0.1D; Qb,p  pile tip resist-
ance of plugged pile at pile head displacement of 0.1D;
(1) D  pile diameter  2Ro; DCPT  CPT cone diameter
(generally  0.036 m); qc  average cone resistance
between 1.5D above pile tip and 1.5D below pile tip.
(2)
In this method, pipe piles are considered as non-
plugging if
where K  coefficient of horizontal earth pressure;
v  effective in-situ overburden pressure;   fric- (7)
tion angle between soil and pile wall; flim  limiting
friction stress; Nq  bearing capacity factor; qlim  where D  pile diameter in m; Dr  Relative density
limiting end bearing stress. of soil below pile tip [%].
Design values for K, , flim, Nq and qlim are given in The MTD criteria specify a unit end bearing for
API RP2A (1993). Selection is based on pile type closed-ended piles which is twice the value computed
(open or closed-ended), soil type and estimated rela- by Equation (6) for plugged open-ended piles.
tive density. Thus the selection of design values is Jardine & Chow (1996) found that frictional cap-
highly subjective, since in situ relative density can acity of piles driven into sand increases with time.
only be evaluated indirectly. Hence, friction in Equation (3) is for a pile loaded 50
days after driving.
3.2 MTD method
Jardine & Chow (1996) of Imperial College, U.K., 3.3 Fugro (2004) method
developed an empirical pile design method which is
The Fugro (2004) method is a modification of the
commonly referred to as the MTD method. The major
MTD method. The modifications are due to the fol-
feature of this method is that ultimate unit friction and
lowing corrections required for the MTD method:
unit end bearing are both directly correlated with
Cone Penetration Test (CPT) cone tip resistance, qc. both pile load test data and laboratory cyclic sand-
In addition, both unit friction and end bearing are steel shear tests (Uesugi et al. 1989), indicate that
highly dependent on pile geometry. both sand grading and steel roughness at the

712

Copyright 2005 Taylor & Francis Group plc, London, UK


interface change such that interface friction angles Table 2. Statistical data of Qc/Qm for tension piles.
become in the order of 290
ultimate unit pile friction at the very pile tip is likely Piles Method Average CoV N
to reduce during compression loading as suggested
OPEN API 93 0.99 0.83 16
by Finite Element analyses (De Nicola & Randolph
MTD 1.22 0.47 16
1993) New 1.07 0.44 16
ultimate unit end bearing criteria in the MTD method API 93* 0.98 0.87 15
are considered too conservative, partly because they MTD* 1.17 0.39 15
apply for a pile head displacement of 0.1D, whereas New* 0.96 0.23 15
offshore practice refers to the value at a pile tip dis- CLOSED API 93 0.33 0.92 2
placement of 0.1D, and party because soil plug MTD 0.53 0.10 2
resistance is ignored. New 0.71 0.02 2
The above modifications were incorporated in modi- ALL API 93 0.92 0.88 18
fied MTD formulae, whereafter empirical coefficients MTD 1.14 0.45 18
New 1.03 0.44 18
and exponents were changed until a statistically good
API 93* 0.90 0.92 17
fit to the pile load test data in Table 1 was obtained. MTD* 1.10 0.44 17
This analysis resulted in the following equations: New* 0.93 0.24 17
Tension:
* Ignoring one Euripides test with doubtful reliability.
(8)
piles in Figure 1. Also shown are trend lines through
Compression: all data points and through the highest quality data
points (DQF4 and DQF5). The selected form of the
(9) trend line is exponential (y  axb) and the optimal
trend is a horizontal line at Qc/Qm  1. It can be
observed that the API method generally underpredicts
pile capacity for relatively short piles and overpre-
(10) dicts capacity for relatively long piles. The MTD
method based data suggest slight underprediction for
relatively short piles and overprediction for relatively
(11) long piles. No clear trend of Qc/Qm varying with L/D
is apparent for the new method.
The Qc/Qm ratios are compared in Figure 2 with an
average normalized cone resistance along the test piles.
This normalized cone resistance, defined as (qc/pa)/
4 PREDICTIONS VERSUS MEASUREMENTS
(v/pa)0.5, is primarily a measure of relative density
and to a lesser extent of in-situ horizontal stress ratio,
4.1 Tension
K0. Typical ranges of normalised cone resistance for
The ultimate tension capacities of the test piles in medium dense, dense and very dense sand are indicated
Table 1 were predicted using the API, MTD and the in Figure 2. Also shown are trend lines through data
new (Fugro 2004) method. The ratio of predicted to points of all tests and through the DQF4 and DQF5
measured capacity, Qc/Qm, for all pile load tests was data points. It can be observed that all three methods
subjected to statistical analysis. The average and coeffi- appear unconservative for sands which are medium
cient of variation (CoV) of this ratio, assuming normal dense and looser, and conservative for dense and very
distributions, are given in Table 2 for various combin- dense sands. However, this inaccuracy is extreme for
ations of pile load tests (N  number of pile load tests). the API method and small for the new method. The
Kolk et al. (2005b) note that there is doubt about the MTD method has an intermediate position relative to
reliability of one of the EURIPIDES pile load tests. these two methods.
Hence, statistical analyses were also made excluding The Qc/Qm ratios were also compared with the time
this test, and are shown as the second set of results for between pile driving and load testing. This comparison
each group of piles in Table 2. It can be observed that is not presented herein due to space limitations. No
the MTD method and new design method are consider- clear trend was apparent in the API based data, due to
ably more reliable than the API method for predicting the large scatter in the data. However, data based on
the tension capacity of open-ended pipe piles. both the MTD method and the new method clearly
The Qc/Qm ratios are compared with embedded indicated an increase in frictional capacity with time.
length in sand/pile diameter (L/D) ratios of the test It seems that these two methods generally overpredict

713

Copyright 2005 Taylor & Francis Group plc, London, UK


API (1993) best fit through all data API (1993)
3 points with DQF 4 or 5 3 50.0% 75.0%92.5%

QAPI/Qmeasured[-]
QAPI/Qmeasured[-]

best fit through all


2 data points with DQF 4 or 5
2

1 best fit through all data points 1 best fit through


all data points

0
0 10 100 1000
0 20 40 60 80 100 120 140
L/D [-] (qc/pa)/(v/pa)0.5[-]

MTD MTD 50.0% 75.0%92.5%


3 3
best fit through all data

QMTD /Qmeasured[-]
points with DQF 4 or 5
QMTD/Qmeasured[-]

best fit through all data points with DQF 4 or 5

2
2
best fit through
all data points
best fit through 1
1 all data points

0
0 10 100 1000
0 20 40 60 80 100 120 140
(qc/pa)/(v/pa)0.5[-]
L/D [-]
2 NEW
NEW 3
QNEW/Qmeasured[-]

50.0% 75.0%92.5%
QNEW/Qmeasured[-]

2 best fit through all data points

1 best fit through all data points with DQF 4 or 5


1
best fit through all data
points with DQF 4 or 5
best fit through all data points 0
10 100 1000
0
0 20 40 60 80 100 120 140 (qc/pa)/(v/pa)0.5[-]
L/D [-] open pipe piles
open pipe piles with DQF 4 or 5
open pipe piles
closed pipe piles
open pipe piles with DQF 4 or 5
closed pipe piles
Figure 2. Predicted/observed tension capacities versus
normalised cone resistance.
Figure 1. Predicted/observed tension capacities versus L/D.

1 using API, MTD and the new method. The average


the capacity of piles loaded within 50 days after driv- and coefficient of variation (CoV) of Qc/Qm, assum-
ing but underpredict tensile capacities of piles tested ing normal distributions, are given in Table 3 for vari-
after a longer set-up period. ous combinations of pile load tests (N  number of
pile load tests).
Again, statistical analyses were also made exclud-
4.2 Compression
ing one EURIPIDES test with doubtful reliability. It
Predictions of total (friction  end bearing) pile can be observed that the MTD and new design methods
capacity, Qc, were made for the pile load tests in Table are considerably more reliable than the API method

714

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 3. Statistical data of Qc/Qm for compression piles. 2
API (1993)
Piles Method Average CoV N best fit through all data points with DQF 4 or 5

OPEN API 93 0.64 0.65 12

QAPI/Qmeasured[-]
MTD 0.95 0.25 12
New 0.99 0.25 12 1
API 93* 0.61 0.69 11
MTD* 0.92 0.26 11
New* 0.94 0.19 11
CLOSED API 93 0.29 0.44 7 best fit through all data points
MTD 0.78 0.30 7
New 1.02 0.18 7 0
0 20 40 60 80 100 120 140
ALL API 93 0.51 0.74 19
MTD 0.89 0.27 19 L/D [-]
New 1.00 0.22 19
2
API 93* 0.49 0.76 18 MTD
MTD* 0.87 0.27 18
New* 0.97 0.19 18

QMTD/Qmeasured[-]
best fit through all data points with DQF 4 or 5
* Ignoring one Euripides test with doubtful reliability.
1
for predicting compression capacity of open-ended best fit through all data points
pipe piles.
Figure 3 presents Qc/Qm ratios for individual pile
load tests versus L/D. Trend lines are again given
through all data points and through the highest qual- 0
ity data points. 0 20 40 60 80 100 120 140
It can be observed that both the API method and L/D [-]
the MTD method tend to underpredict capacity but
become less conservative for relatively long piles. No 2
NEW
clear trend is apparent for the new method.
Figure 3 presents Qc/Qm ratios for individual pile
QNEW/Qmeasured[-]

load tests versus L/D. Trend lines are again given best fit through all data points with DQF 4 or 5
through all data points and through the highest qual-
ity data points. 1
It can be observed that both the API method and
the MTD method tend to underpredict capacity but
become less conservative for relatively long piles. No best fit through all data points
clear trend is apparent for the new method.
Figure 4 present Qc/Qm ratios for individual pile
0
load tests versus the corresponding average normalised 0 20 40 60 80 100 120 140
cone resistances along the test piles. Typical ranges of L/D [-]
the latter for medium dense, dense and very dense sand
are indicated. Also shown are trend lines through data open pipe piles
points of all tests and through the highest quality data open pipe piles with DQF 4 or 5
closed pipe piles
points. It can be observed that the API method appears
unconservative for loose sand and is conservative for
dense to very dense sand. Both the MTD method and Figure 3. Predicted/observed compression capacities
the new method appear appropriate throughout the versus L/D.
whole density range.
Analyses were also made of the individual friction test results. It is also of interest to note that the MTD
and end bearing components of the pile load tests (not method appears unconservative for predicting friction
presented herein). These indicated that the scatter in and conservative for predicting end bearing, whereas
predictions is larger for individual friction and end total capacity is well predicted. This suggests that there
bearing components than for total compression cap- are either compensating errors in the MTD method or
acity. This is due to the inaccuracy of separating outer end bearing from pile load tests is overestimated and
friction and end bearing when interpreting pile load friction from these tests is equally underestimated.

715

Copyright 2005 Taylor & Francis Group plc, London, UK


2 5 SUMMARY
API (1993)
best fit through all data points with DQF 4 or 5 Both the MTD (Jardine & Chow 1996) and the new
QAPI/Qmeasured[-]

(Fugro 2004) studies on tension and compression


piles confirmed that the API method is very inaccu-
1 rate for predicting pile capacity in silica sand.
50.0% 75.0%92.5% The MTD method is significantly more accurate
than the API method. However, the Fugro study showed
that the MTD method is slightly unconservative for
best fit through
all data points friction and conservative for end bearing. The new
(Fugro 2004) design method is more accurate than the
0
10 100 1000 MTD method. Both the MTD and the new method are
(qc/pa)/(v/pa)0.5[-]
conservative for predicting long term capacity.

2
MTD REFERENCES
QMTD/Qmeasured[-]

best fit through all data points with DQF 4 or 5


API 1993. Recommended Practice for Planning, Designing
and Constructing Fixed Offshore Platforms, API
1 Recommended Practice 2A (RP2A), 20th edition.
De Nicola, A. & Randolph, M.F. 1993. Tensile and
Compressive Shaft Capacity of Piles in Sand, Jnl. of
best fit through Geotechn. Eng. Div. ASCE 119(12): 19521973.
all data points Fugro 2004. Axial Pile Capacity Design Method for Offshore
50.0% 75.0%92.5%
Driven Piles in Sand, Report to API.
0 Hossain, M.K. & Briaud, J.L. 1993. Improved Soil
10 100 1000 Characterisation for Piles in Sand in API RP-2A, OTC
(qc/pa)/(v/pa)0.5[-] Paper 7193, Proceedings 25th Offshore Technology
Conference 1: 637654, Houston, Texas.
2 Jardine, R.J. & Chow, F.C. 1996. New Design Methods for
NEW Offshore Piles, MTD publication 96/103, London, U.K.
best fit through all data points Kolk, H.J., Baaijens, A.E., Shafei, K.A. & Dakhil, O.A.
2005a. Axial Load Tests on Pipe Piles in Very Dense Sands
QNEW/Qmeasured[-]

at Ras Tanajib, International Symposium on Frontiers in


Offshore Geotechnics, 1921 September.
1 Kolk, H.J., Baaijens, A.E. & Vergobbi, P. 2005b. Results from
best fit through Axial Load Tests on Pipe Piles in Very Dense Sands. The
all data points EURIPIDES JIP, International Symposium on Frontiers
with DQF 4 or 5 in Offshore Geotechnics, 1921 September.
50.0% 75.0%92.5% Uesugi, M., Kishida, H. & Tsubakihara, Y. 1989. Friction
between Sand and Steel under Repeated Loading, Soils
0 and Foundations. 29(3): 127137.
10 100 1000
(qc/pa)/(v/pa)0.5[-]
open pipe piles
open pipe piles with DQF 4 or 5
closed pipe piles

Figure 4. Predicted/observed compression capacities


versus normalised cone resistance.

The under-prediction of long-term pile capacity by


all three methods indicates an increase of capacity
with time. This increase is less distinct for total com-
pression capacity than for tension capacity. It seems
that the MTD and the new method both underpredict
total compression capacities of piles loaded after 10
days (compared to 50 days for tension capacity).

716

Copyright 2005 Taylor & Francis Group plc, London, UK


Estimating the end bearing resistance of pipe piles in sand using the
final filling ratio

K. Gavin
University College Dublin, Ireland

B.M. Lehane
University of Western Australia, Perth, Australia

ABSTRACT: Recent research has highlighted the effect of the soil core development during pile installation,
on the axial resistance developed during static loading of driven pipe piles in sand. This paper summarises the
experimental findings from recently conducted field tests designed to investigate the factors controlling the
development of base resistance of pipe piles. These provide the framework for a new design method which
incorporates a direct measure of soil core development into the calculation of base resistance. A database of pile
load tests on large scale pipe piles has been assembled, which enables features of full-scale pile behaviour to be
identified as well as facilitating a review of current design methods.

1 INTRODUCTION conventionally instrumented pipe piles. Measurement


of their relative contributions to base resistance has,
The base resistance (Qb) of a pipe pile is a combination however, been achieved using the twin-walled tech-
of the sum of the internal shear stresses mobilised in nique by Paik and Lee (1993), Gavin (1998) and Paik
the soil plug (Qplug) and the resistance mobilised on et al. (2003). The findings from these studies and
the annular base area (Qann): from other recent research on model and field scale
pipe piles is reviewed here first to provide the back-
ground to a proposed new method for estimating base
(1) resistance. This method is shown to compare favourably
with the existing database of base capacity measure-
It is usual to express these as the unit base, plug and ments and to be a significant improvement on other
annular resistances: design methods currently in use.

(2)
2 RECENT INSTRUMENTED PILE RESEARCH

(3) Gavin (1998) describes results from a series of labora-


tory chamber experiments where highly instrumented
model piles were jacked into a loose siliceous normally
consolidated sand. The experiments allowed the meas-
(4)
urement of the soil plug resistance during pile instal-
lation and during static loading (where it has been
where Ab, Aplug and Aann, are the pile base area, plug noted that the majority of piles remain fully plugged
area and annular area respectively. Total and unit base regardless of their plugging behaviour during installa-
resistances referred to in this paper are ultimate val- tion). The test series indicated the significant effect
ues measured at pile head displacements of up to of the soil plug development during installation on
about 1 diameter penetration (unless otherwise stated). the plug capacity that could be mobilized during
It is not possible to separate the contributions static loading. Gavin and Lehane (2003) subsequently
of plug and annular resistance from load tests on proposed the following expressions relating the

717

Copyright 2005 Taylor & Francis Group plc, London, UK


available base resistance of pipe piles to the CPT qc
end resistance (averaged 1.5 pile diameters above and
below the pile tip) and the final incremental filling
ratio ( rate of change of plug height with respect to
the pile penetration) measured at the end of installa-
tion (FFR):

(5a)

(5b)

(5c)

(5d)

Tests on small scale piles have revealed that the


important variables affecting pile plugging (and there-
fore axial capacity) include the driving method (ham-
mer, jacking or vibrating into place), soil density and
overconsolidation ratio. To investigate these effects
(and acknowledging that the majority of piles are
driven to refusal in dense sand layers), a series of field
tests was carried out using both instrumented and un-
instrumented model piles installed in a dense, over- Figure 1. Development of soil core during installation.
consolidated sand deposit. The instrumented 111 mm
diameter pile was jacked into place whilst three single
un-instrumented pipes with diameters of 75, 100 and
114 mm were driven. Separate estimates of the shaft
and base resistance of the driven piles could only be
made during static load testing at the final penetration
depth of the piles. As static tension loading followed
the compression tests, the shaft resistance measured
during the tension loading was assumed to be equal to
80% of the compressive shaft resistance (see Jardine &
Chow 1996). The driven pipe pile base resistance was
therefore estimated by subtracting this value from the
measured total pile resistance. Details of the model
piles and soil conditions are contained in Gavin and
Lehane (2003a,b).
The incremental filling ratios (IFR) recorded during
installation of the jacked pile and the three driven
piles are shown in Figure 1. All piles remained fully
coring (IFR  1) to a pile penetration depth (L) of
0.6 m. The jacked pile subsequently experienced sig-
nificant plugging, with full plugging (IFR  0)
occurring at L  1.1 m. In contrast, all driven piles
penetrated in a partly coring mode up to their max-
imum penetration depth.
The effect of the degree of plugging on peak plug
Figure 2. qann and qplug during instrumented pile installation.
and annular resistances measured during the instru-
mented jacked pile installation are shown in Figure 2.
It is apparent that the annular resistance (qann) is simi- by the Lehane and Gavin (2001) for tests in loose
lar to the cone penetration base resistance (qc) and is sand. The plug resistance increases sharply as the IFR
independent of the IFR value throughout installation; reduces but does not attain a value similar to the qc
this equivalence between qann and qc was also shown value until the pile has penetrated over 5 diameters

718

Copyright 2005 Taylor & Francis Group plc, London, UK


Figure 3. Effect of installation IFR on qplug/qc on model
piles.
Figure 4. Mobilisation of plug and annular resistance dur-
below the level at which the pile had fully plugged ing static loading of a 356 mm pipe pile (Paik et al. 2003).
(with IFR  0). Equation 5a is therefore only valid
after penetration in a fully plugged mode over a num- clear qann again approximates to qc while qplug is very
ber of pile diameters has occurred. As such a high much lower than qc; the qplug/qc ratio of 0.28 is in good
level of additional penetration with IFR  0 is agreement with that predicted by Eqn 6.
unlikely in practice, it is suggested that Equation 5a The evidence in the foregoing, while confirming
should be modified as follows: that sand plugs remain stationary during a static load
test, indicates that its is the degree of sand displacement
(6) at the pile base (as reflected by the FFR value) which
has a controlling influence on the static plug resistance.
The steady state installation plug resistance, qplug,
(measured after a penetration of at least one pile
diameter) of the jacked piles in the loose sand in the 3 DATABASE OF FIELD TESTS
chamber and the jacked field pile in dense sand are
normalised by the CPT qc value at the pile base and A database of 20 instrumented piles with diameters
plotted on Figure 3 against the final filling ratio (D) ranging from 100 to 2000 mm and pile length (L)
(FFR). Normalised qplug values mobilized at failure in to diameter ratios of 1587 was assembled; see Table 1.
static load tests are also shown on Figure 3. Open The piles were mostly driven, (four piles were jacked
symbols are used for installation and static load test into place) in predominantly dense sand. Cone
results are shown by closed symbols. Penetration Test (qc) data were available at all but
All piles (even those which had high filling ratios three sites, where qc was estimated from Standard
during installation) remained fully plugged during all Penetration Test (N) values using the approach sug-
stages of static load testing. A comparison of the instal- gested by Robertson et al. (1988).
lation and load testing data on Figure 3 indicates that The total base resistance (sum of internal and
installation qplug values are typically the same or mar- external load) was measured directly on five piles, the
ginally higher than those recorded in load tests at the jacked pile at Blessington, all three piles at Kochi and
same FFR. A clear trend for qplug/qc to reduce with the the driven pile in Indiana. Measurements of the distri-
FFR is apparent. bution of internal shear stresses (De Nicola &
Paik et al. (2003) also report qplug and qann data for Randolph 1997, Lehane & Gavin 2001) have shown
a 356 mm diameter pile driven into dense sand in that very high shear stresses develop near the toe of
Indiana (USA). The pile exhibited an FFR value 0.77 at the pile, but that these decay rapidly with distance
the final installation depth of 7 m and the plug was above the pile tip (h) and become negligible at h/D 3.
reported to have remained intact during static loading Therefore, in cases where strain gauges were attached
to large displacement (20% of the pile diameter). The to a single walled pile, the base load was taken
measured development of qann and qplug with normal- approximately equal to the load indicated/interpo-
ized pile displacement (s/D) is shown in Figure 4. It is lated from the strain gauges at h  3D. The plug

719

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 1. Database of open ended pile tests.

Location L D qc qb qplug Installation


reference (m) (m) PLR FFR (MPa) (MPa) (MPa) method

Hoogzand 7.0 0.36 0.66 0.66 38 14 9 Driven


Beringen et al.
(1979)
Dunkirk 11 0.32 0.54 0.50 17.7 9.5 7.7 Driven
Brucy et al. 11 0.32 0.52 0.60 17.7 9.5 6.7 Driven
(1989)
Kimitsu 20 1.2 0.77 0.77 39 5.2 3.5 Driven
Ishihara et al.
(1997)
Tokyo 30.6 2 1.08 1 30 4.5 2.2 Driven
Shioi et al.
(1992)
Chiba 40.6 0.8 1 1 30 7.6 5.7 Driven
Kusakabe et al.
(1989)
Hokkaido 40 1.2 0.85 0.85 25.6 7.5 5.6 Driven
Kusakabe et al.
(1989)
Shanghai 79 0.91 0.80 0.80 21.5 8.9 7.7 Driven
Pump et al. 79 0.91 0.85 0.85 21.5 5.6 4.1 Driven
(1998)
Blessington 2 0.11 0.28 0 15 15 15 Jacked
Gavin & 1.4 0.07 0.91 0.83 15 4.8 2.4 Driven
Lehane 1.8 0.10 0.88 0.60 15 7.1 5.4 Driven
(2003) 1.8 0.11 0.85 0.80 15 5.3 3.7 Driven
Kochi 7.1 0.16 0.17 0 9 10 10 Jacked
White et al. 8 0.32 0.45 0.60 5.8 1.7 1.5 Jacked
(2003) 5.7 0.32 0.32 0.55 9 5.5 2.8 Jacked
Indiana 7 0.36 0.83 0.77 22 11 6.3 Driven
(Paik et al.
2003)
Euripides 30.5 0.76 1.0 1.0 52 15 6.9 Driven
Driven
Fugro Ltd. 38.7 0.76 1.0 1.0 52 15 7.2 Driven
(2001) 47.0 0.76 1.0 1.0 70 15 3.4 Driven

resistance is calculated by assuming the annular shown in Table 1, no pile with D  600 mm exhibited
resistance is equal to the qc value at the base. PLR 0.8. Evidence of significant plugging at a site
The degree of soil plugging during installation can where IFR or FFR was not measured was noted at
be crudely represented by the Plug Length Ratio only one location (Hoogzand).
(PLR  Lp/L), which is reported for all piles in the
database. Incremental and final filling ratios are only
available for thirteen of the piles. Although Lehane
4 CURRENT DESIGN METHODS
and Gavin (2004) demonstrate that there is no rela-
tionship between PLR and IFR, significant plugging
4.1 Conventional bearing capacity theory
during driving of large diameter piles in sand is rare,
and it is suggested that for cases where FFR values Such approaches use a modified form of the
are not reported and PLRs exceed about 0.8 (i.e. the Terzaghis bearing capacity theory, and assume the
pile is largely coring throughout installation) the FFR ultimate end bearing resistance (qb) value developed
may be taken approximately equal to the PLR. As by a deep foundation to be directly proportional to the

720

Copyright 2005 Taylor & Francis Group plc, London, UK


4.2 MTD Method (Jardine & Chow 1996)
Jardine & Chow (1996) compiled a database of instru-
mented piles in sand and proposed a design equations
for both plugged (FFR  0) and un-plugged (FFR  0)
piles. The authors suggest that plugging is controlled
by the pile diameter and the relative density of the
soil, with smaller diameter piles in dense soils being
prone to plugging. The following simple expression
to assess the minimum density necessary to form a
rigid plug is given as:

(8)

If plugging is not predicted the contribution of


internal shear stresses is ignored, and the pile base
capacity is given by the load developed on the annu-
lus only:

(9)

(where Di is pipes internal diameter)


Figure 5. Variation of Nq with relative density.
The Jardine & Chow (1996) design guidelines
(which are often referred to as the MTD method) cover
the design of open and closed ended driven piles and
include an expression for calculating the base resist-
vertical effective stress acting at the pile base through ance of closed ended piles at a displacement of 10%
a relationship of the form: of the pile diameter (qb0.1). This ratio was found to
depend on the qc value and the pile diameter, with the
(7) ratio qb0.1/qc decreasing with increasing pile diameter:

where Nq is a bearing capacity factor which depends (10)


on
(the soils friction angle) and the embedment
depth, and 
v is the vertical effective stress at the pile where DCPT is the diameter of the CPT cone.
base. Chow (1997) suggests that the end bearing resist-
Although both Nq and 
v can be assessed with rea- ance of plugged piles is equivalent to the end bearing
sonable accuracy, a major drawback of this approach is of closed-ended piles of equal dimension at very large
the sensitivity of Nq to small changes in
. Sampling displacements, but at typical displacements of 10% of
difficulties for cohesionless soils are such that empirical the pile diameter, the end bearing resistance may be
correlations between
and in-situ tests data are gener- very much lower. Based on their database Jardine &
ally used in design, introducing additional uncertainty. Chow found that for a given pile diameter, the ratio of
Nq values in the API method are assumed to vary qb0.1/qc for a plugged open-ended pile is half that of
with relative density (Dr). Design values are com- a closed ended pile. On this basis equation 10 (which
pared in Figure 5 to values backfigured from the data- is considered applicable to closed ended piles) is
base piles. The range of measured Nq values at a given modified as follows:
Dr value indicate that the base resistance of pipe piles
is not uniquely related to Dr. (11)
Jardine & Chow (1996) noted a tendency for pre-
dictions using the API approach to show significant The MTD design line for the database is shown on
scatter when compared to measured base resistance, Figure 6, where it is seen to provide a reasonable
and to generally under-predict available resistance. lower bound to the database values. However, while
However, as seen on Figure 5, quite a number of noting that the method estimates the base resistance at
actual base resistances are under-predicted by API. a pile head displacement of 10%, significant under
Even when limiting base resistance values recom- prediction of capacity is evident for piles up to 1 m in
mended by API are employed (which depend on Dr), diameter. In addition, recent research has raised doubts
25% of the cases in the database are significantly regarding the diameter effect for closed ended piles,
over-predicted by API. with Gavin & Lehane (1998) demonstrating the

721

Copyright 2005 Taylor & Francis Group plc, London, UK


It appears that the simple expression which
includes FFR in the assessment of plug resistance
provides excellent predictions for a range of pile
diameters, soil densities and pile lengths.

5 CONCLUSIONS

(i) Instrumented pile tests that combined accurate


measurement of plug and annular resistance
with soil core development throughout installa-
tion demonstrate that the degree of plugging
experienced by a pipe pile during installation in
a given soil is the most important variable con-
trolling the base resistance developed during
subsequent static loading.
(ii) A review of two of the currently popular design
methods for evaluating base capacity of pipe
piles has shown that these methods are unreli-
able primarily because they do not take proper
account of the degree of plugging (i.e. soil dis-
Figure 6. Effect of pile diameter on qb/qc. placement) during pile installation on capacity.
(iii) An expression for evaluation of the base resist-
ance of pipe piles, written as a function of the
CPT qc value and final filling ratio (FFR), pro-
vides an excellent match to the available meas-
urements of base capacity.

ACKNOWLEDGEMENTS

The authors wish to thank the technical staff at


University College Dublin (UCD) and Trinity College
Dublin, George Cosgrave, Frank Dillon, David
McCauley and Martin Carney for assistance with the
model pile testing. The work was funded by a UCD
Presidents Research Award.

Figure 7. Effect of IFR on qplug/qc for database.


REFERENCES
tendency for qb to equal qc at large pile displacements
over a range of pile diameters. Re-interpretation of American Petroleum Institute (1993). RP2A-WSD:
the MTD database of closed ended pile tests by White Recommended practice of planning, designing and con-
and Bolton (2003) confirmed this, and further high- structing fixed offshore platforms Working stress
lighted the effect of residual loads, partial embedment design, 20th edition, Washington, 5961.
and the qc averaging technique on the trend for qb0.1/qc Beringen F.L., Windle, D. and Van Hooydonk W.R. (1979),
Results of loading tests on driven piles in sand,
for closed ended piles to reduce with pile diameter. The Proceeding of Conference on Recent Developments in the
safe extrapolation of the MTD approach to open- design and construction of piles, Institution of Civil
ended piles must therefore be questionable. Engineers, London, UK, 213225.
Brucy F., Meunier K. and Nauroy J.F. (1991). Behaviour of pile
4.3 Proposed design method based on FFR plug in sandy soils during and after driving, Proceedings
of the 23rd Offshore Technology Conference, Houston,
The normalized plug resistances indicated by the OTC 6514, 145154.
piles in the database are plotted against FFR in Figure 7, Chow F. (1997), Investigations into the behaviour of dis-
which includes the modified relationship between qc, placement piles for offshore structures, Ph.D Thesis,
IFR and qplug detailed in Equation 6. Univ. of London (Imperial College).

722

Copyright 2005 Taylor & Francis Group plc, London, UK


De Nicola A. and Randolph M.F. (1999). Centrifuge model- Lehane B.M. and Randolph M. (2002). Evaluation of a min-
ling of pipe piles in sand under axial loads, imum base resistance for driven pipe piles in siliceous
Geotechnique, UK, 49(3), 295318. sand, Journal of Geotechnical and Geoenvironmental
Gavin K. (1998). Experimental investigation of open and Engineering., ASCE 128(3) pp 198205.
closed ended piles in sand, Ph.D Thesis, University of Paik K.H. and Lee S.R. (1993). Behaviour of soil plugs in
Dublin (Trinity College). open-ended piles driven into sands Marine Georesources
Gavin K.G. and Lehane B.M. Estimating end bearing capaci- Geotechnology., 11 pp 353373.
ties of driven piles in sand using In-situ tests Proceedings Paik K.H., Salgado R, Lee J. and Kim B. (2003). Behaviour
VII International Conference On Piling and Deep of Open and Closed Ended Piles Driven Into Sands.
Foundations, Vienna, 1998. Journal of Geotechnical and Geoenvironmental
Gavin K.G. and Lehane B.M. (2003). End Bearing Engineering., ASCE, 129 No.4. pp 296306.
Resistance of Small Diameter Pipe Piles in Dense Sand, Pump W., Korista S, Scott J. (1998). Installation and load
Proceedings of the International Conference on testing of deep piles in Shanghai alluvium. Proceedings
Foundation Behaviour, pp 321330. Dundee, September. of the VII International Conference on Deep
Ishihara K., Saito A, Shimmi Y, Mieura Y. and Tominaga M. Foundations, Vienna. Paper 13, pp 1.3.11.3.7.
(1977). Blast furnace foundations in Japan, Proceedings Randolph, M.F. (2003). Science and empiricism in pile
of the 9th Int. Conference on Soil Mechanics and foundation design. Geotechnique 53, No.10, pp 847875.
Foundation Engineering, Tokyo, 157236. Robertson P.K., Campanella R.G. and Whitman A. (1983).
Jardine R.J. and Chow F.C. (1996). New design methods for SPT CPT correlations, Journal of Geotechnical
offshore piles, MTD publication 96/103, Marine Engineering, ASCE , Vol 109, No. 11, 14491459.
Technology Dept., London, UK. Shioi Y., Yoshida O, Meta T. and Homma M. (1992).
Kusakabe O., Matsumoto T, Sanadabata I, Kosuge S. and Estimation of bearing capacity of steel pipe pile by static
Nishimura S. (1989). Report on questionaire: Predictions loading test and stress-wave theory, Application of stress-
of bearing capacity and drivability of piles, Proceedings wave theory to piles, Balkema, Rotterdam. pp 325330.
of the 12th International Conference on Soil Mechanics White, D.J. and Bolton M.D., (2004). A review of field
and Foundation Engineering, Rio de Janeiro, 5, measurements of CPT and pile base resistance in sand.
29572963. ICE Proceedings, Geotechnical Engineering. Accepted
Lehane B.M. and Gavin K.G. (2001). The base resistance of for Publication.
jacked pipe piles in sand, Journal of Geotechnical and White, D.J., Sidhu, H.K, Finlay. T.C.R, Bolton M.D. and
Geoenvironmental Engineering, ASCE, 127(6),473480. Nagayama, T. (2003). Press in piling: The influence of
Lehane B.M. and Gavin K.G. (2004). Discussion- Journal of plugging on driveability Proc. 8th Int. Conf. Of the Deep
Geotechnical and Geoenvironmental Engineering, ASCE. Foundations Institute, New York. pp 299310.

723

Copyright 2005 Taylor & Francis Group plc, London, UK


Evaluation of end-bearing capacity of open-ended piles driven in
sand from CPT data

X. Xu, B.M. Lehane & J.A. Schneider


University of Western Australia (UWA)

ABSTRACT: This paper uses the current UWA database of base capacity measurements to examine the pre-
dictive performance of three CPT based methods used for the assessment of the end-bearing resistance of driven
open-ended piles in siliceous sands. The same database is then employed to validate a new improved method for
base capacity evaluation, which considers the resistance provided by the annulus and plug separately and incorp-
orates the effects of soil displacement and an appropriate qc averaging technique.

1 INTRODUCTION during driving on base (and shaft) capacity, the FFR


is not incorporated explicitly in any of the three
The behaviour of pipe or open-ended piles is expected recently proposed CPT based design methods: Fugro-
to lie between that of full-displacement and non- 04 (Kolk et al. 2005), ICP-05 (Jardine et al. 2005),
displacement piles, as shown by full-scale field and NGI-04 (Clausen et al. 2005) presumably
tests (Szechy 1959, Kishida 1967, Paik et al. 2003), because its measurement is still not standard practice.
laboratory testing chamber studies (ONeill & Raines A new database of pile load tests in sand has been
1991, Foray et al. 1998, Gavin 1998, Paik & Salgado compiled at the University of Western Australia,
2003), and centrifuge model pile tests (De Nicola & UWA (Lehane et al. 2005). This paper presents the
Randolph 1996, Bruno 1999). component of the database containing good quality
In an accompanying paper to this Symposium on base capacity measurements for open-ended piles in
base capacity evaluation of closed-ended (full dis- siliceous sand, and uses it to:
placement) driven piles, Xu & Lehane (2005) demon-
(i) examine the relative reliability of the Fugro-04,
strate that the ratio of unit end-bearing resistance at a
ICP-05, and NGI-04 methods for assessment of
pile tip displacement of 10% of the pile diameter (qb0.1)
the end-bearing capacity of open-ended piles;
to the averaged cone resistance qc in the vicinity of the
(ii) propose a simple means of assessing the end-
tip for such piles is approximately 0.6. The relation-
bearing resistance at a base displacement of 10%
ship between qb0.1 and qc may be expected to be less
of the pile diameter (qb0.1), which involves FFR
consistent for driven open-ended piles in view of the
and an appropriate CPT qc averaging technique.
differences between the mode of penetration of a cone
and that of a coring or partially plugging pile during
installation. However, Salgado et al. (2002), Gavin &
2 NEW DESIGN METHODS
Lehane (2003), and others, have shown that a rela-
tively consistent relationship between qb0.1 for an
Design formulas for the Fugro-04, ICP-05 and NGI-04
open-ended pile and the CPT qc value becomes
methods are provided in Table 1 and additional details
apparent when the effects of sand displacement close
are summarised as follows. These highlight the sig-
to the tip during pile driving are accounted for. This
nificant differences between the respective methods.
installation effect is best described by the incremental
filling ratio (IFR  Lp/z, incremental change of (i) The ultimate end-bearing resistance qbu is defined
plug length Lp relative to change in penetration depth by the ICP-05 and NGI-04 methods as the resist-
z) measured over the final stages of installation and is ance mobilised at a pile head displacement of 10%
referred to as the final filling ratio (FFR). As the FFR of the pile diameter (0.1 D). The Fugro-04 defini-
approaches zero, qb0.1 approaches that of a closed- tion of ultimate end-bearing resistance is the
ended pile with the same outer diameter. However, same as that adopted by API RP2A (1993) i.e.
despite the strong impact of the soil plug response the end bearing resistance mobilised at 0.1 D pile

725

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 1. New design methods for end-bearing resistance (v) The average qc value in the vicinity of the pile
qb0.1 of open-ended pile in siliceous sands. tip,
qc , recommended by Fugro-04 and ICP-05 is
the average qc from a level of 1.5 D above the
Method Design equations pile tip to 1.5 D below the tip1. This averaging
technique is identical to that proposed by the
Fugro-04 qb0.1/pa  8.5  (q
c /pa)
0.5
Ar0.25, pa  100 kPa respective methods for closed-ended piles, despite
ICP-05 If Di 2.0(Dr
0.3), or Di 0.083 qc DCPT/pa, the fact that the influence zone in the vicinity of
then the pile is unplugged, and qb0.1/qc  Ar a typical driven pipe pile (which is usually almost
If not, the pile is plugged, and qb0.1/qc  fully coring during driving) will be less than that
maximum[0.5 0.25 log(D/DCPT), 0.15, Ar] of a closed-ended pile with the same diameter.
NGI-04 qb0.1  minimum[plugged qb0.1, unplugged
qb0.1] The plugged qb0.1 value is calculated as:
qb0.1  qc,tip  FDr; FDr  0.7/(1  3D2r ) with 3 OVERVIEW OF DATABASE
Dr  0.4  ln[qc,tip/22/(paV)0.5)], pa  100 kPa
The unplugged qb0.1 is calculated as: The UWA database of (good quality) base capacity
qb0.1  qb,ann  Ar  qb,plug  (1
Ar) with measurements for open-ended piles in siliceous sand
qb,ann  qc,tip, and qb,plug  12 f L/Di currently comprises data from 13 full-scale static load
f : external skin friction, see Clausen et al. tests at 7 sites for which CPT qc profiles were avail-
2005 able. Table 2 summarises the significant details con-
L: pile embedment depth
cerning each load test. Further features of note are:
Note Ar  1
(Di/D)2; qc  averaged qc near
pile tip; and DCPT  0.036 m At Dunkirk, the first time compression tests on two
piles, CLa and CSa, are included in the database.
Both of these had strain gauges attached to their out-
side walls and pile CSa had a 6.5 mm thick inner
shoe. Final filling ratios (FFRs) and plug length
ratios PLRs (PLR  Lp/z) were about 0.5 and 0.55
tip displacement, referred as qb0.1. In the database
respectively, indicating a relatively high level of
considered here for open-ended piles, differences
plugging during pile installation.
between qbu and qb0.1 are small and therefore the
end-bearing resistance in all methods discussed The EURIPIDES test programme involved a highly
instrumented pile (Zuidberg & Vergobbi 1996),
will be referred as qb0.1 throughout the paper.
driven at one location and then load tested at three
(ii) In the Fugro-04 method, end-bearing resistance is
depths. The pile was then extracted and re-driven to
expressed as a function of the square root of qc and
the deepest penetration achieved in the first instal-
area ratio Ar ( 1
(Di/D)2) raised to the power
lation. The FFRs were in the range of 0.8 to 1.0,
of 0.25, where Di is the internal pipe diameter. The
while PLRs were about unity.
method assumes that the sand plug remains sta-
tionary (i.e. does not fail) during static loading. For tests at the Hoogzand site, the strain gauge
instrumented pile no. I and un-instrumented pile
(iii) The ICP-05 method assumes implicitly that a
no. III (which had a 4 mm thick internal shoe) were
pile plug can fail under static loading and employs
driven open-ended, with only plug length at final
an empirical relationship (based on data obtained
pile penetration being measured. Tension tests
during pile driving) between pile diameter and
were performed after compression tests. For pile
sand relative density (Dr) to assess if the plug will
no. III, the end-bearing resistance was estimated by
fail statically. The ICP-05 formulations are such
assuming that the tension shaft capacity was 75%
they can lead, in certain instances, to the inference
of the shaft capacity in compression.
of a greater base capacity for a pile which is
adjudged (on the basis of the specified relation- In Indiana, a comprehensively instrumented,
double-walled, pile system was used for the first
ship between D and Dr) to have a failing plug than
time at field scale to separate the resistance pro-
a pile which is assessed to have a rigid basal plug.
vided by the annulus (qb,ann) and by the plug
(iv) In the NGI-04 method, the assumptions made
(qb,plug). The FFR and PLR were around 0.8.
regarding the contribution of the inner wall shear
stress are such that the unplugged qb0.1 value is For Ras Tanajib II, the load test performed on a
fully instrumented pile with a pile tip depth of 25 m
generally significantly greater than the plugged
was included in the database. The PLR was 0.81,
qb0.1 value (with an average unplugged to plugged
ratio of 5.7 when applied to the database of base
capacity measurements); this trend is consistent 1
ICP-05 does warn against the use of such an averaging
with the expectation that, for most typical piles, technique when there are any significant weaker layers near
a plug will not fail under static loading. pile tip, but provides no guidance for this case.

726

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 2. UWA database of open-ended piles in sand.

q
c (MPa)
D t L vo qb0.1a qb,plugb qc,tip
Site name m mm m FFR Ar Arb* kPa MPa MPa MPa 1.5Dc Dutchd Reference

Dunkirk CSa 0.324 19.1 11.3 0.45 0.22 0.63 141 7.3 6.0 26.7 24 21.0 Brucy 1991
Dunkirk CLa 0.324 12.7 11.3 0.48 0.15 0.58 141 6.6 5.7 26.7 24 21.8 Brucy 1991
Euripides I30c 0.763 35.6 30.5 0.99 0.18 0.18 320 12.3 7.9 61.5 60.8 55.0 Fugro 2004
Euripides I38c 0.763 35.6 38.7 0.90 0.18 0.26 403 9.9 5.9 50.8 50.8 47.3 Fugro 2004
Euripides I47c 0.763 35.6 47.0 0.89 0.18 0.27 488 15.3 11.7 65.9 66.4 53.1 Fugro 2004
Euripides II47c 0.763 35.6 46.7 0.82 0.18 0.28 477 16.0 12.5 63.3 63.1 53.4 Fugro 2004
Hoogzand Ic 0.356 16 7.0 0.66e 0.17 0.45 100 12.2 10.2 37.7 42.2 36.9 Beringen
1979
Hoogzand IIIc 0.356 20 5.3 0.77e 0.21 0.39 82 11.2 7.1 45.0 45.5 44.1 Beringen
1979
Indiana OEP 0.356 32 7.0 0.8 0.33 0.48 94 7.4 5.7 21.7 21.7 18.0 Paik 2003a
RasTanajib 25a 0.763 38.5 25.0 1.13 0.19 0.19 306 20.7 14.2 92.4 84.9 79.6 Fugro 2004
Shanghai ST1 0.914 20 79.0 0.80e 0.09 0.27 637 5.9 5.2 21.5 21.5 21.5 Pump 1998
Shanghai ST2 0.914 20 79.1 0.85e 0.09 0.22 637 5.1 4.3 21.5 21.5 21.5 Pump 1998
Tokyo TP 2.0 43 30.6 1.08e 0.08 0.08 275 2.0 1.6 30.4 23.9 9.7 Shioi 1992
a
at 0.1 D pile tip displacement, and inferred from axial pile stress at h  2 Di minus external shaft friction from h  0 to
h  2 Di with f  0.01qc; b inferred from qb0.1 by assuming qb,ann  0.6 qc (Dutch); c arithmetic averages of qc over  1.5 D
(employed by ICP-05 and Fugro-04); d qc obtained by Dutch averaging technique, employing effective diameter
D*  D  (A*r )0.5; e no measurement of IFR with FFR assumed equal to PLR.

which was due to a high level of plugging during difficulties2, prompted the authors to infer end-bear-
initial driving from 17 m to 19 m penetration (Fugro ing capacities as the load in the pile wall at a distance
2004) but the FFR was about unity. h  2Di above the pile toe minus the estimated exter-
In Shanghai, two 79 m long pipe piles were instru- nal shaft friction which was assumed equal to 1% of
mented with strain gauges at 9 levels along the pile qc. The base capacities deduced in this way are pro-
shaft and driven through soft clay (30 m) to a vided in Table 2 and are generally in good agreement
dense sand layer. Plug length ratios (PLR) were with those quoted for the same tests in Chow (1997)
around 80% to 85% at final pile penetration, with and Fugro (2004).
no IFR during installation being reported.
In Trans-Tokyo Bay, a 2 m diameter pipe pile (fit- 4 EVALUATION OF FUGRO-04,
ted with a 9 mm thick, 300 mm long external driv- ICP-05 & NGI-04
ing shoe) was driven to a depth of 30.6 m and load
tested to plunging failure. This plunging failure The database summarised in Table 2 was tested against
was attributed to the presence of a relatively soft the Fugro-04, ICP-05, and NGI-04 predictive methods
clay layer close to its tip. The pile was not plugged outlined in Table 1. All parameters, such as averaged
during driving with the final plug level being cone resistance and vertical effective stress, were
2.4 m above the sea bed after driving. No strain assessed independently from the cited references. The
gauge instrumentation data along the pile shaft was ratios of calculated to measured qb0.1 values are plotted
reported and therefore the value qb0.1 given in Table on Figure 1 against the averaged relative densities near
2 is that quoted in Shioi (1992). the pile tips (Dr,tip) for each method. The average and
Measurements of inner and outer skin friction dis- coefficient of variation (COV) of the ratio between cal-
tributions (e.g. ONeill & Raines 1991, Lehane & culated and measured end-bearing capacities are
Gavin 2001), as well as theoretical studies reported included in these figures and may be taken as measures
by Randolph et al. (1991) have shown that the internal of each methods predictive ability. It is evident that:
shear stresses are highest near the tip of the pile and Fugro-04 over predicts the base capacities for pile
decay exponentially with distance above the pile tip; load tests at Tokyo Bay, but provides reasonable
internal skin friction, as a consequence, is very predictions for other piles in the database.
small at a height (h) of two diameters (Di) above the
pile tip. This trend coupled with the general shortage
2
of reliable strain gauge data very close to the tips because of the need to assess the contribution to the
of pipe piles and the associated interpretation strain/load measurement of internal and external friction.

727

Copyright 2005 Taylor & Francis Group plc, London, UK


2.00 30 Dunkirk CLs
Dunkirk CSa
1.75 Tokyo 27 EURIPIDES I30c
EURIPIDES I38c
qb0.1, C [Fugro-04]/qb0.1, M

1.50 qb0.1,C/qb0.1, M=3.6 EURIPIDES I47c


24 EURIPIDES II47c
Indiana OEP
1.25 Ras Tanajib
Fugro-04 21
average=1.29
1.00 cov=0.55
18

h/Di(-)
0.75 Dunkirk
Euripides 15
Hoogzand
0.50
Indiana
Ras Tanajib
12
0.25 Shanghai
Tokyo (a) 9
0.00
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 6
Dr,tip
3
2.00
0
1.75 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4
qb0.1, C [ICP-05] /qb0.1, M

1.50 IFR(-)
1.25
ICP-05 Figure 2. Examples of IFR measurements during pile
average=0.76
1.00 cov=0.29 driving.
0.75 Dunkirk
Euripides
Hoogzand
0.50 Indiana
Ras Tanajib Despite the wide differences between the formula-
0.25 Shanghai tions for qb0.1 employed by these three methods, their
Tokyo (b)
0.00 predictive performances are broadly similar. Given that
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 virtually the same very small database of load tests
Dr,tip was used to calibrate each of the methods, it follows
2.00 that at least two of the three respective formulations
1.75 contains compensating errors. None of the methods
Tokyo
qb0.1,C/qb0.1, M=3.4
allows for the varying effects of soil displacement
qb0.1, C [NGI-04]/qb0.1, M

1.50 shown by each of the database piles (as illustrated on


1.25 Figure 2 for 8 of these piles), implying that their
NGI-04
average=1.00 extrapolation outside of the database to larger diame-
1.00 cov=0.81 ter piles is questionable.
0.75 Dunkirk
Euripides
Hoogzand
0.50 Indiana
Ras Tanajib 5 THE UWA-05 DESIGN APPROACH FOR qb0.1
0.25 Shanghai
Tokyo (c)
0.00 The basis of the new UWA-05 approach presented
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 here is summarised as follows:
Dr,tip
The base capacity of a pipe pile falls between that
Figure 1. Ratios of calculated to measured capacities: of a full-displacement closed-ended pile at FFR  0
(a) Fugro-04, (b) ICP-05; and (c) NGI-04. (when a pile is fully plugged during installation)
and that of a non-displacement or bored pile when
FFR  1.
Lehane & Randolph (2002) show that a sand plug
ICP-05 has the lowest COV and provides a conser- will not fail during static loading (as long as the plug
vative prediction for all pile load tests. Although not is longer than 5 Di) and indicate that the compres-
evident from Figure 1b, this method becomes pro- sion of the soil plug for a pile driven into siliceous
gressively more conservative as the pile diameter sand is a relatively small component of the full tip
increases (with qb0.1/q c reducing to Ar for large displacement. The overall base stress at a displace-
diameter piles). ment of 0.1 D i.e. qb0.1 (which is much lower than the
NGI-04 over-predicts the base capacities for pile true ultimate base stress) is therefore controlled by
load tests at Shanghai and Tokyo Bay and displays the stiffness of the sand below the sand plug; this
a tendency to under-predict in very dense sand. stiffness is shown by Gavin & Lehane (2003) to be

728

Copyright 2005 Taylor & Francis Group plc, London, UK


related to degree of soil displacement in the vicinity 0.7
of the tip as measured by FFR. Dunkirk
Euripides
0.6
Xu & Lehane (2005) show that, for full displace- Hoogzand
ment driven piles, the ratio of qb0.1 to the averaged 0.5
Indiana

cone resistance qc in the vicinity of the tip is Ras Tanajib


Shanghai

qb,plug/qc
approximately 0.6 and that this ratio is not depend- 0.4 Tokyo
ent on the pile diameter.
Recommended values of qb0.1/qc at tip movement 0.3 Equation (2b)
qb,plug/qc=0.6-0.45FFR
of 0.1 D for non-displacement (bored) piles range 0.2
from 0.15 to 0.23 (Bustamante & Gianeselli 1982,
Jamiolkowski & Lancellotta 1988, Ghionna et al. 0.1
1993, Franke 1989). These ratios are also not
dependent on the pile diameter. 0.0
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2
Experimental data and numerical analysis indicate FFR
that the resistance that can develop on the tip annu-
lus at a base movement of 0.1 D varies between about Figure 3. Ratios of qb,plug/ q
c vs. FFR.
0.6 and 1.0 times the CPT qc value (e.g. Bruno 1999,
Salgado et al. 2002, Lehane & Gavin 2001, Paik
et al. 2003, Jardine et al. 2005). A constant of 0.6
for qb,ann/qc is adopted here, which is equivalent to (2b)
the qb0.1/qc ratio for driven closed-ended piles.
The averaged cone resistance qc should be evalu-
ated in the same way as that employed for (2c)
closed-ended piles (Dutch qc averaging technique,
Schmertmann 1978), but, for consistency, needs to Substituting Equation (2) into (1), the ratio of qb0.1/q
c
employ an effective diameter D* related to the is then derived as:
degree of partial plugging during the final stages
of installation. (3a)
FFRs were derived for the database piles by aver-
aging the IFRs (shown on Figure 2) measured over For typical offshore piles with FFR  1, A*r  Ar and:
the final three diameters of pile installation. FFR val-
ues were taken equivalent to PLR values in the (3b)
absence of IFR data. The normalised plug resistances
qb,plug/qc deduced from the qb0.1 values of the database Equation 3a provides predictions for qb0.1 for fully
piles (assuming qb,ann/qc  0.6) are plotted against plugged, partially plugged and fully coring pile instal-
these FFRs on Figure 3. These indicate a trend very lations. The continuous form of this equation is made
similar to that shown by Gavin & Lehane (2005) with possible by employment of the effective area ratio
qb,plug/qc reducing from 0.6 at FFR  0 to 0.15 at term, A*r. The equivalent diameter, D*, employed when
FFR  1. This trend is identical to that inferred from assessing qc is consequently given as D*  D A*r .
consideration of the respective end bearing capacities Although the FFR may be assumed to be unity for a
of full and non-displacement piles. Evidence for typical large diameter offshore pile, its value cannot
imposition of a maximum FFR value of unity is also presently be predicted in advance to assist the design of
provided on Figure 3, which shows that the Ras smaller diameter piles. In such cases and so that the
Tanajiib and Tokyo piles with FFR  1.1 can gener- effects of partial plugging can be roughly incorporated
ate the same plug resistance as the base resistance of in a design proposal for qb0.1, an estimate of the likely
a bored pile. The proposed expression for qb0.1 then FFR may be obtained from the trend line through
follows from: existing measurements shown on Figure 4; this figure
includes FFR and PLR data available for the UWA base
capacity database, as well as PLR measurements from
(1a)
the California (USA) Department of Transportation.
The approximate nature of the proposed trend line
is apparent on Figure 4, which shows how this line
(1b) over-estimates the FFR for two piles at Dunkirk.
Lehane & Gavin (2004) show that the low FFR values
for these piles arises as the piles were driven through
(2a) dense sand and terminate in looser sand.

729

Copyright 2005 Taylor & Francis Group plc, London, UK


1.1 3a provides a good representation of the measured
1 qb0.1 data, with the exception of the Dunkirk site
0.9 where qb0.1 values are over-predicted.
0.8 The UWA-05 predictions may be compared
0.7 FFR measured directly with those of Fugro-04, ICP-05 and NGI-04
0.6 FFR = PLR with reference to Figures 1 and 6. This comparison
FFR

0.5
indicates that, unlike the three other methods, the
0.4 Di(m) 0.2
UWA-05 predictions are reasonable at all test sites,
FFR min 1, with the method having an average of calculated to
1.5m
0.3
Dunkirk measured capacities of 1.08 and a COV of calculated
0.2 CLa & CSa to measured capacities of 0.16.
0.1
0
0.0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
7 CONCLUSIONS
Inner Diameter, Di (m)

Figure 4. Database of FFR (including PLR) measurements.


This paper employs recent research findings to develop
a simple expression relating the end bearing capacity,
qb0.1, of a driven pile in sand, with an averaged CPT
0.6
end resistance (q c ) and an effective area ratio term
Equation (3a) (Ar*), which allows for the effects of partial plugging
0.5 qb0.1/qc=0.15+0.45Ar* during installation. It is demonstrated that this expres-
sion provides better predictions than three other
0.4
recently proposed CPT formulations when tested
against the (relatively small) database of existing base
qb0.1/qc

0.3 Dunkirk
Euripides
capacity measurements.
0.2 Hoogzand
Indiana
Ras Tanajib
0.1 Shanghai REFERENCES
Tokyo

0.0 American Petroleum Institute, RP2A-WSD, 1993. Recom-


0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 mended practice of planning, designing and constructing
Ar* fixed offshore platforms Working stress design. 20th
edition, Washington.
Figure 5. Ratio of qb0.1/ q
c vs. effective area ratio A.*
r. Beringen, F.L., Windle, D. & Van Hooydonk, W.R. 1979.
Results of loading tests on driven piles in sand. in Recent
2.00 development in the design and construction of piles, ICE,
London: 213225.
1.75 Brucy, F., Meunier, J. & Nauroy, J.-F. 1991. Behaviour of pile
plug in sandy soils during and after driving. Proc., 23rd
qb0.1, C [UWA-05]/qb0.1, M

1.50
Annual OTC, Houston:145154.
1.25 Bruno, D. 1999. Dynamic and static load testing of driven
UWA-05
average=1.08 piles in sand. PhD Thesis, University of Western Australia.
1.00 cov=0.16 Bustamante, M. & Gianeselli, L. 1982. Pile bearing capacity
0.75 Dunkirk by means of static penetrometer CPT. Proc., 2nd European
Euripides Symposium on Penetration Testing, Amsterdam: 493499.
Hoogzand
0.50
Indiana Chow, F.C. 1997. Investigations into the behaviour of dis-
0.25
Ras Tanajib placement piles for offshore foundations. PhD Thesis,
Shanghai
Tokyo
Imperial College.
0.00 Clausen, C.J.F., Aas, P.M. & Karlsrud, K. 2005. Bearing
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 capacity of driven piles in sand, the NGI approach. Proc.
Dr,tip ISFOG, Perth.
De Nicola, A. & Randolph, M.F. 1999. Centrifuge modeling
Figure 6. Ratio of calculated over measured capacities for of pipe piles in sand under axial loads. Gotechnique,
UWA-05. 49(3): 295318.
Foray, P., Balachowski, L. & Colliat, J.-L. 1998. Bearing
capacity of model piles driven into dense overconsolidated
6 PREDICTIONS FOR qb0.1 BY UWA-05 sands. Canadian Geotechnical Journal 35(2): 374385.
Franke, E. 1989.Prediction of the bearing behaviour of piles,
The predictive performance of Equation 3a is assessed especially large bored piles. in Proc. XII ICSMFE, RIO
on Figures 5 and 6. Figure 5 indicates that Equation de Janeiro.

730

Copyright 2005 Taylor & Francis Group plc, London, UK


Fugro Engineers B.V. (Fugro) 2004. Axial pile capacity Lehane, B.M., Schneider, J.A. & Xu, X. 2005. Evaluation of
design method for offshore driven piles in sand, P-1003, design methods for displacement piles in sand. UWA
Issue 3, to API, August 2004. report GEO: 05341.1.
Gavin, K. & Lehane, B.M. 2005. Estimating the end bearing ONeill, M.W. & Raines, R.D. 1991. Load transfer for pipe
resistance of pipe piles in sand using the final filling piles in highly pressured dense sand. J. Geotech. Engrg.,
ratio. In Proc. ISFOG, Perth. ASCE, 117(8): 12081226.
Gavin, K. 1998. Experimental investigation of open and Paik, K. & Salgado, R. 2003. Determination of bearing
closed ended piles in sand. PhD Thesis University of capacity of open-ended piles in sand. J. Geotech. &
Dublin (Trinity College). Geoenv. Engrg. ASCE, 129(1): 4657.
Gavin, K.G. & Lehane B.M. 2003. Shaft friction for open- Paik, K., Salgado, R., Lee, J. & Kim, B. 2003a. Behavior of
ended piles in sand. Canadian Geotechnical J., 40(1): open- and closed-ended piles driven into sands.
3645. J. Geotech. & Geoenv. Engrg. ASCE, 129(4): 296306.
Ghionna, V.N., Jamiolkowski, M., Lancellotta, R. & Pump, W., Korista, S. & Scott, J. 1998. Installation & load
Pedroni, S. 1993. Base capacity of bored piles in sands tests of deep piles in Shanghai alluvium, in Proceedings,
from in situ tests. in Deep Foundation on Bored and VII International Conference on Piling & Deep
Auger Piles, ed. V. Impe, Balkema, Rotterdam: 6775. Foundations 1: 3136.
Jamiolkowski, M. & Lancellotta, R. 1988. Relevance of Randolph, M.F., Leong, E.C. & Houlsby, G.T. (1991). One-
in-situ test results for evaluation of allowable base resist- dimensional analysis of soil plugs in pipe piles.
ance of bored piles in sands, in Deep Foundation on Bored Gotechnique 41(4): 587598.
and Auger Piles,, ed. V. Impe, Balkema, Rotterdam: Salgado, R., Lee, J. & Kim, K. 2002. Load tests on pipe piles
107120. for development of CPT-based design method. Final
Jardine, R.J., Chow, F.C., Overy, R. & Standing, J. 2005. ICP Report: FHWA/IN/JTRP-2002/4.
design methods for driven piles in sands and clays, Schmertmann, J.H.1978.Guidelines for cone test, perform-
Thomas Telford. ance, and design. U.S. Fed. Highway Administration,
Kishida, H. 1967. Ultimate bearing capacity of pipe piles in FHWATS-78209.
sand. Proceedings of the 3rd Asian Regional Conference Shioi, Y., Yoshida, O., Meta, T. & Homma, M. 1992.
on Soil Mechanics and Foundation Engineering: 196199. Estimation of bearing capacity of steel pipe pile by static
Kolk, H.J., Baaijens, A.E. & Sender, M. 2005. Design cri- loading test and stress-wave theory (Trans-Tokyo Bay
teria for pipe piles in silica sands. In Proc. ISFOG Perth. Highway). in Application of stress-wave theory to piles,
Lehane, B.M. & Gavin, K. 2004. (Discussion). Determination Balkema, Rotterdam: 325330.
of bearing capacity of open-ended piles in sand. Szechy, C.H. 1959. Tests with tubular piles. AcTa Technica
J. Geotech. & Geoenv. Engrg. ASCE, 130 (6): 656658. of the Hungarian Academy of Science 24: 181219.
Lehane, B.M. & Gavin, K.G. 2001. Base resistance of Xu, X. & Lehane, B.M. 2005. Evaluation of end-bearing
jacked pipe piles in sand. J. Geotech. & Geoenv. Engrg., capacity of closed-ended piles in sand from CPT data. In
ASCE, 127(6): 473480. Proc. ISFOG, Perth.
Lehane, B.M. & Randolph, M.F. 2002. Evaluation of a min- Zuidberg, H.M. & Vergobbi, P. 1996.EURIPIDES, Load
imum base resistance for driven pipe piles in siliceous sand. tests on large driven piles in dense sands, Proc. 28th OTC:
J. Geotech. & Geoenv. Engrg. ASCE, 128(3): 198205. 193206.

731

Copyright 2005 Taylor & Francis Group plc, London, UK


Evaluation of end-bearing capacity of closed-ended pile in sand from
cone penetration data

X. Xu & B.M. Lehane


University of Western Australia (UWA)

ABSTRACT: The end-bearing capacity of closed-ended displacement piles in sand is often linked to either an
averaged CPT end resistance (q ) over a defined influence zone in the vicinity of the tip or to the actual cone
c
resistance at the pile tip level (qc,tip). This paper uses a database compiled at UWA to examine the sensitivity and
suitability of commonly employed averaging techniques for qc. The component of the UWA database contain-
ing high quality base capacity measurements indicates that the ratio of the end bearing resistance at a pile base
displacement of 0.1 D (qb0.1) to the q c value derived using the Dutch averaging technique, is with few excep-
tions, equal to 0.6  0.1 for driven piles; a design qb0.1/q c ratio of 0.50 is suggested for closed-ended piles.
Unlike some other published approaches relating qb0.1 to qc, this qb0.1/q c ratio shows no systematic dependence
on sand relative density, stress level and pile diameter.

1 INTRODUCTION (iii) propose a simple means of assessing the base


resistance at a base displacement of 10% of the
The estimation of end-bearing resistance of a dis- pile diameter.
placement pile utilising cone end resistance (qc) is
one of the earliest direct applications of CPT data. A
number of recent publications have re-examined the
2 NEW DESIGN METHODS
relationship between qb and qc; these have included
Fugro (2004), Jardine et al. (2005), and Aas et al.
Design formulas for three methods are provided in
(2004) who present the Fugro method (Fugro-04), the
Table 1 and some pertinent details are summarised as
Imperial College method (ICP-05), and the Norwegian
follows:
Geotechnical Institute method (NGI-99) respectively.
The three methods were based on individually assem- (i) The ultimate end-bearing resistance (qbu) is
bled and assessed databases, which do not include all defined by the ICP-05 and NGI-99 methods as the
the available load tests presented in this paper because resistance mobilised at a pile head displacement
of different data quality control standards and emer- of 10% of the diameter (D). Such a definition
ging new case histories. In addition, while Fugro-04
and ICP-05 employ a database of base capacity meas- Table 1. Design methods for end-bearing resistance of
urements, the NGI-99 approach did not consider the closed-ended pile in sand.
shaft and base resistance components separately.
A new database of pile load tests in sand has been Methods Design equations
compiled at the University of Western Australia
(UWA). This paper presents the component of the Fugro-04 qbu  qb0.1 at pile head displacement of 0.1D
qb0.1/pa  8.5  (q/p )0.5, p  100 kPa
database containing high quality base capacity meas-
c a a
urements for closed-ended piles in siliceous sand, and qc  qc averaged 1.5 D above and below pile tip
ICP-05 qbu  qb at pile head displacement of 0.1 D
uses it to: qbu  qc [1-0.5log(D/DCPT)]; DCPT  0.036 m

qc  qc averaged 1.5 D above and below pile tip
(i) examine the relative reliability of the Fugro-04,
ICP-05, and NGI-99 design methods for assess- a lower bound qbu  0.3 qc applies
ment of the end-bearing capacity of closed- NGI-99 qbu  qb at pile head displacement of 0.1D
qbu  qc,tip  FDr
ended piles; FDr  0.8/(1 + Dr2)
(ii) investigate various averaging techniques Dr  0.4  ln[qc,tip/22/(pa v)0.5)], pa  100 kPa
employed for qc in the region of the pile tip;

733

Copyright 2005 Taylor & Francis Group plc, London, UK


does not consider a piles elastic shortening, averaging technique for resistance data meas-
which can be substantial for long piles. The ured with a 36 mm diameter cone.
Fugro-04 definition of ultimate base resistance (iv) The NGI-99 qbu value is obtained by multiplying
is the same as that adopted by API RP2A (1993) the cone end resistance at pile tip level by a rela-
i.e. the end bearing resistance mobilised at a tive density correction factor FDr, which leads to
base displacement of 0.1 D referred to as qb0.1; qbu/qc ratios reducing with increasing Dr.
this definition is employed here to ensure a con-
sistent comparison of mobilised end bearing for
piles of varying axial compressibility.
3 OVERVIEW OF DATABASE
(ii) The Fugro-04 assumes that qbu  qb0.1 is pro-
portional to the square root of qc.
The UWA database of base capacity measurements
(iii) The ICP-05 method proposes that qbu/qc varies
for closed-ended piles in siliceous sand currently
with the pile diameter, reducing by a factor of 3
comprises data from 29 full-scale static load tests at
as D increases from 0.1 m to 0.9 m. The existence
13 sites for which CPT qc profiles were available.
of this scale effect is contentious e.g. White &
Table 2 summarises the significant details concerning
Bolton (2005) argue that an apparent diameter
each load test. Further features of note are:
dependence of this ratio arises because of the
greater zone of influence around a large diam- Clay layers were present at Kallo and Salt Lake
eter pile and the consequent need for a suitable Valley, where piles were driven through a soft clay

Table 2. UWA database of closed-ended piles in sand.

qc (Mpa)
D L qb0.1 qc,tip v,tip
Site name Test no. (m) (m) (Mpa) (Mpa) 1.5D Dutch (kPa) Reference
a
Akasaka AK 1C 0.2 11 18.2 27.5 29.9 25.7 147.4 BCP (1971)
Akasaka AKa 6B 0.2 4 4.3 8.6 8.4 6.2 54.4 BCP (1971)
Akasaka AKa,c 6C 0.2 11 18.1 27.5 29.9 25.7 147.4 BCP (1971)
Baghdad BGa,b P1 0.285 11 5.1 6.5 6.0 4.7 151.9 Altaee (1992)
Baghdad BGa,b P2 0.285 15 7.1 6.2 7.1 6.1 189.3 Altaee (1992)
Drammen Da,b A 0.28 8 1.1 3.0 2.9 2.6 89.5 Gregersen
(1973)
Drammen Da,b D/A 0.28 16 1.8 5.0 5.2 4.4 177.2 Gregersen
(1973)
Dunkirk DKa DK1/L1C 0.102 7.4 11.1 15.4 15.2 11.8 102.0 Chow (1997)
Dunkirk DKa DK2/L1C 0.102 5.96 10.7 13.7 13.1 9.3 88.0 Chow (1997)
Hoogzand Ga,b,c II-C 0.356 6.75 13.6 29.6 28.6 26.3 97.5 Beringen
(1979)
Hsin ta HTa TP4 0.609 34.25 3.1 7.7 9.0 6.4 310.5 Yen (1989)
Hunters point HPa,b,c S 0.273 9.15 5.0 8.6 8.3 6.6 99.6 Briaud (1989)
Indiana INEW CEP 0.356 6.87 11.0 18.2 18.1 16.4 89.6 Paik (2003)
Kallo KAa CPT250 0.25 8.85 12.6 20.4 18.8 13.3 100.4 De Beer (1979)
Kallo KAa I 0.908 9.69 9.0 27.9 27.3 14.7 108.9 De Beer (1979)
Kallo KAa II 0.539 9.71 10.7 28.1 29.3 18.8 109.1 De Beer (1979)
Kallo KAa III 0.615 9.82 9.7 29.2 30.1 16.3 110.2 De Beer (1979)
Kallo KAa IV 0.815 9.8 9.2 28.9 29.0 15.1 110 De Beer (1979)
Kallo KAa V 0.406 9.33 10.7 25.9 25.0 16.6 105.3 De Beer (1979)
Kallo KAa VII 0.609 9.37 8.5 26.5 24.9 16.0 105.7 De Beer (1979)
Labenne LBa,b LB1/L1C 0.102 5.95 4.4 4.5 4.5 4.3 79 Lehane (1992)
Labenne LBa LB2/L1C 0.102 1.83 4.6 6.3 6.3 5.5 34.0 Lehane (1992)
Ogeechee OGb H-11 0.457 3 3.6 7.1 6.3 3.6 42.7 Vesic (1970)
Ogeechee OGb,c H-12 0.457 6.1 10.3 10.9 12.2 8.8 74.2 Vesic (1970)
Ogeechee OGb,c H-13 0.457 8.9 12.7 15.1 14.8 10.9 103.6 Vesic (1970)
Ogeechee OGb,c H-14 0.457 12 12.8 6.7 13.8 10.5 135.6 Vesic (1970)
Ogeechee OGb,c H-15 0.457 15 15.43 15.1 15.2 13.1 166.2 Vesic (1970)
Sermide SMb Sermide 0.508 35.85 8.9 16.4 16.3 14.5 315 Appendino
(1981)
Salt lake SLNEW 1700 south 0.324 23.2 7.8 24.6 20.7 12.3 207.0 Rollins (1999)
a
included in ICs database; b included in NGIs database; c included in Fugros database.

734

Copyright 2005 Taylor & Francis Group plc, London, UK


layer into dense sand. Interbedded clay layers were capacities may be taken as measures of each methods
also detected in the sand site at Hsin Ta. predictive ability. The calculations were performed
All piles were driven with the exception of 7 jacked following the procedures given in Table 1 and are
piles: 1C and 6B from Akasaka, DK1 and DK2 in summarised in Figure 1. It is evident that:
Dunkirk, LB1 and LB2 in Labenne, and CPT250
in Kallo; the average diameter of these piles is less The Fugro-04 method tends to over-predict, espe-
cially in loose sand sites (Drammen and Hsin Ta).
than the total average D in the database.
This appears to be because the method indirectly
Pile embedments range from 1.8 m to 36 m,
implies that qc has to be greater than 7.2 MPa to
while D varies from 102 to 908 mm. Measured
lead to a qb0.1/q
c ratio of less than unity. The over-
qb0.1 capacities vary from 1.1 MPa to 18 MPa.
estimation at Kallo and Salt Lake Valley could, as
Piles were instrumented with strain gauges or a tip
discussed later, be attributed to the use of an inappro-
load cell, allowing for a reasonable determination
priate qc averaging technique.
of qb0.1 with allowance for residual loads at most
sites. At Salt Lake Valley, qb0.1 was estimated using 4.5
Davissons procedure and scaled according to the calculationmethod:Fugro-04 AK
4.0 average= 1.43
measured total load at a base displacement of 0.1 D. cov=0.53
BG
3.5 D
Residual loads had to be, by necessity, assumed to DK

qb0.1,C/qb0.1,M
be zero at Ogeechee while those at Baghdad were 3.0 G
2.5 HT
estimated based on a procedure proposed by HP
Fellenius (1989). 2.0 I
Only two sites (Indiana and Salt Lake Valley) have 1.5 KA
LB
not been included in any previous databases, indi- 1.0 OG
cating a dearth of high quality data available in the 0.5
SM
public domain. On inspection of Table 2, it is also SL
0.0
noteworthy that only 2 out of the 29 load tests have 0 0.2 0.4 0.6 0.8 1
been included in all three databases employed for
Fugro-04, ICP-05, and NGI-99. It follows that there (a) Pile diameter (m)
have been variable subjective judgments made 1.8
concerning the reliability of various published pile calculation method: ICP-05 AK
1.6 average= 0.95
load tests. Of the 29 tests in the UWA database, only BG
1.4 cov= 0.30 D
7, 21, and 13 load tests on closed-ended piles are DK
1.2
qbu,C/qb0.1,M

included in the Fugro-04, ICP-05, and NGI-99 G


1.0 HT
databases respectively. Furthermore, while sand sites HP
are naturally variable, only one representative 0.8 I
KA
CPT profile is reported by the authors for each of 0.6 LB
the various case histories at 8 of the 13 test sites. 0.4 OG
Average qc profiles were employed in this study at SM
0.2 SL
sites with multiple reported CPT soundings. 0.0
It was found, in general, that there was good agree- 0 0.2 0.4 0.6 0.8 1
ment between the ultimate end-bearing capacities (b) Pile diameter (m)
inferred in this study and those deduced in assembly
of the Fugro-04, ICP-05, and NGI-99 database. 2.5
calculation method: NGI-99 AK
Corrections to allow for different definitions of ultim- average=1.0 BG
ate capacity (see Table 1) were not significant 2.0 cov=0.44 D
DK
because the average length of the piles in the database
qbu,C/qb0.1,M

G
was less than 11 m. 1.5 HT
HP
I
1.0 KA
4 EVALUATION OF FUGRO-04, LB
ICP-05, & NGI-99 0.5 OG
SM
SL
The UWA database was tested against the Fugro-04, 0.0
ICP-05, and NGI-99 predictive methods. All param- 0 0.2 0.4 0.6 0.8 1
eters, such as averaged cone resistance and vertical (c) Pile diameter (m)
effective stresses, were assessed independently, The
average and coefficient of variation (COV) of the Figure 1. Ratios of calculated to measured capacities:
ratio between calculated and measured end-bearing (a) Fugro-04, (b) ICP-05; and (c) NGI-99.

735

Copyright 2005 Taylor & Francis Group plc, London, UK


The ICP-05 method results in an average calcu- 1.2 AK
lated to measured ratio close to unity. This is per- BG
1.0 D
haps not surprising as most of the data points (21
DK
out 29) have been included in developing the 0.8 G

qb0.1/qc
method itself. HT

In general, the NGI-99 method under-estimates 0.6 HP


I
base capacity, except in loose sand sites (Drammen 0.4 KA
& Hsin Ta) and non-homogeneous stratigraphies LB
(Kallo & Salt Lake Valley). This may be partly OG
0.2
SM
because the approach used to establish and cali- SL
brate the NGI-99 method did not try to separate the 0.0
0 0.2 0.4 0.6 0.8 1
end-bearing and shaft capacities.
Pile diameter (m)
As shown in Figure 1, it would appear that, all
three methods do not perform well in loose sand; this Figure 2. Normalised qb0.1D by qc averaged over 1.5 D.
is likely to be because of the limited number of case
histories in loose sand.
Furthermore, the somewhat arbitrary selection of following four qc averaging techniques currently in
an average qc (over 1.5 D) has the potential to lead popular use:
to the erroneous trends. The effect of the choice of the
averaging technique is now discussed. (a) Dutch method (Van Mierlo & Koppejan 1952,
Schmertmann 1978), where the cone resistances
are averaged over an influence zone extending
from 8 D above, and between 0.7 D and 4 D below
5 SENSITIVITY OF qc AVERAGING
the pile toe following a minimum path rule.
TECHNIQUES
(b) LCPC (Bustamante & Gianeselli 1982), where an
equivalent arithmetic average of qc values 1.5 D
The qb0.1/qc trend with pile diameter established for
above and below pile tip is obtained and a filter
the UWA database is shown in Figure 2, which employs
rule eliminates extreme qc values.
values of cone resistance averaged 1.5 D above and
(c) an arithmetic average over an defined influence
below pile tip. The database employed is the same as
zone e.g. 1.5 D, 2 D, and 8 D.
that used for ICP-05, except case histories for which
(d) a geometric average over 8D above and 4 D below
only SPT data were available have been excluded and
the pile tip.
data from Ogeechee, Indiana, Sermide and Salt Lake
Valley have been added. The ICP-05 design curve is The qc profiles for all case histories in the database
evidently a reasonable fit to the qb0.1/qc data. However, were digitised and a computer code written to exam-
due to the limited data points for large diameter piles, ine the implications of these techniques.
the interpretations of the tests results from Kallo are It was found that:
critical to the inference of the trend for qb0.1/qc to
(i) there is no significant difference between qc val-
reduce with pile diameter. Piles were driven through
ues averaged by applying the LCPC method or by
soft clay with a small penetration into underlying
simply arithmetic averaging values over 1.5 D.
dense sand at both the Kallo and Salt lake Valley sites.
Their averages are also very close to the cone
In such cases, Eslami & Fellenius (1997), Randolph
resistances at the pile tip qc,tip, with averaged
(2003), White & Bolton (2005), amongst others, sug-
ratios of qc,1.5 D/qc,tip and qc,LCPC/qc,tip equal to
gest that the design cone resistance should be
1.01 and 1.02 respectively. Figure 3(a) illustrates
weighted to reflect the overlying weaker material and
the comparison for the LCPC technique.
the partial embedment into dense sand.
(ii) Figures 3(b) and 3(c), which employ methods
White & Bolton (2005) present a re-interpretation
(a) and (d), indicate distinct differences between
of the partial embedment of pile tip into a dense layer
average values and qc,tip at homogenous and non-
in site Kallo. The qc averaging technique employed,
homogeneous sites. The averaged ratios of
which was selected based on proposals of Meyerhof
qc,Dutch/qc,tip and qc,geometric average/qc,tip are equal to
& Valsangkar (1977), necessitates generation of an
0.83 and 0.93, if the Salt Lake Valley and Kallo
idealised linear qc variation with depth in the vicinity
sites are excluded, indicating that qc values are
of the tip, identification of an interface between two
not sensitive to the averaging techniques at
distinctly different layers and judgment on the level at
homogenous sites.
which the influence of the dense layer is first sensed
in the CPT. This procedure is therefore rather subject- The Kallo case history provides an opportunity to
ive and hence consideration is given here to the investigate the suitability of averaging techniques as

736

Copyright 2005 Taylor & Francis Group plc, London, UK


40 AK
qc (MPa)
BG 0 10 20 30 40 50
D
30 7
DK
qc, LCPC (MPa)

G 8
HT range of load test depths
20 HP 9
I
KA 10
LB
10 OG 11
SM

Depth (m)
SL 12
0
(a) 0 10 20 30 40 13

40 14 36 mm Cone
AK 250 mm Cone
BG 15 qc, Dutch
D qc,1.5 D
30 DK 16 Pile load tests
qc, Dutch (MPa)

G
HT 17
20 HP
I 18
KA
LB 19 (a)
10 OG
SM
20
SL
0 qc (MPa)
(b) 0 10 20 30 40
0 3 6 9 12 15
40 15
AK 36 mm Cone
averaged over 8D above 16
qc, geometric average (MPa)

BG Jacked Pile, B=350 mm


and 4D below pile tip level D qc, Dutch
30 17
DK

clay
qc, 1.5 D
G 18
HT
20 HP 19
I 20
KA
Depth (m)

10 LB 21
OG
SM 22

sand
SL 23
0
(c) 0 10 20 30 40 24
qc,tip (MPa) 25
26
Figure 3. Comparisons of averaged qc by (a) LCPC;
27
(b) Dutch; (c) geometric average over
4D/8D.
clay

28 (b)
29
it includes the end resistance measured during continu- 30
ous jacked installation of a 250 mm diameter cone/pile.
The end resistance measured with this cone and that Figure 4. Measured and predicted qc profiles at sites:
of a standard 36 mm diameter cone are compared on (a) Kallo, Belgium; (b) Perth, Australia.
Figure 4(a) with the end resistance of the 250 mm
diameter cone predicted using the Dutch averaging
technique applied to the qc data. Evidently, relatively load of this pile was recorded continuously during
good agreement is obtained, apart from an over- pile jacking as the pile moved from a soft clay layer
prediction between 10.5 and 11.5 m. Also shown are into a 6 m thick layer sand and then back into a firm
the qc values (Dutch) for the pile load tests. clay. Predictions obtained using all other averaging
A further example of the suitability of the Dutch techniques considered were inferior.
method for an instrumented 350 mm square concrete It is therefore presumed here that, as the Dutch
jacked pile in Perth is presented in Figure 4(b) for a case method provides the most accurate estimates of the
history reported by Lehane et al. (2003). The base steady state penetration resistance of piles, this

737

Copyright 2005 Taylor & Francis Group plc, London, UK


steady state resistance may be more uniquely related 1.4
to qb0.1 than any other measure of CPT end resistance. Jacked AK
1.2 BG
Ogeechee River D
1.0 DK
G

qb0.1/qc
6 RE-EVALUATION OF A DESIGN 0.8 HT
HP
APPROACH 0.6 I
KA
The database of qb0.1 values normalised using the 0.4 LB
Dutch qc averaging procedure are presented in Figure 5. OG
0.2 SM
It is evident that: SL
0.0
(i) qb0.1/ qc ratios are higher for the smaller diam- 0 0.2 0.4 0.6 0.8 1
eter piles, most of which were jacked. The mean Pile diameter (m)
qb0.1/
qc value for the 7 jacked piles in the data-
base is 0.9 and significantly higher than the Figure 5. Normalised qb0.1 by averaged qc (Dutch).
mean ratio for the driven piles. A number of
publications (e.g. Randolph 2003, Gavin &
Lehane 2003) indicate that the higher base (iv) In light of the poor extent of the existing data-
capacity of jacked piles is due to their tendency to base, a qb0.1/
qc ratio of 0.5 is recommended for
have larger residual base stresses; these stresses design purposes. This ratio is slightly larger than
lead to stiffer base response and hence a higher the value of 0.4 recommended by Randolph
stress mobilised at a displacement of 0.1 D. (2003), which is an independent review of a
(ii) The qb0.1/ qc values for the five 457 mm diam- slightly smaller database to that considered here.
eter driven piles tested at the Ogeechee River site
are greater than unity, even though the quoted end
bearing capacities did not include the additional
(unknown) residual base stresses. ACKNOWLEDGEMENTS
(iii) Interpreted qb0.1 values are affected significantly
by the assessment of the residual base stress. To The authors wish to acknowledge the assistance and
illustrate (an extreme) example of this effect, willingness to share data and opinions of various per-
qb0.1 values without residual load correction and sonnel involved in the development of the Fugro-04,
with correction (using the Altaee et al 1992 ICP-05, and NGI-99 methods.
interpretation, which employs the procedure
recommended by Fellenius 1989) are plotted on
Figure 5 for the two piles tested in Baghdad. REFERENCES

Aas, P.M., Clausen, C.J.F. & Lacasse, S. 2004. Bearing


capacity of driven piles in sand based on pile load tests.
7 CONCLUSION Presentation, 7 May 2004, Texas A&M, IOBT.
Altaee, A., Fellenius, B.H. & Evgin, E. 1992. Axial load trans-
(i) The database provides little evidence of a fer for piles in sand 1: Tests on an Instrumented Precast
dependence of qb0.1/ qc on stress level, diameter Pile. Canadian Geotechnical Journal 29(1): 1120.
American Petroleum Institute, RP2A-WSD, 1993. Recom-
and relative density, provided that an appropri-
mended practice of planning, designing and constructing
ate averaging technique for qc in the vicinity of fixed offshore platforms Working stress design, 20th
the pile tip is employed. Xu & Lehane (2005) edition, Washington.
deduced the same conclusion from a compre- Appendino, M. 1981. Interpretation of axial load tests on
hensive series of centrifuge tests involving three long piles. Proc. of the International Conference on Soil
pile diameters in a variety of sand stratigraphies Mechanics and Foundation Engineering 2: 593598.
tested over a range of stress levels. BCP-Committee 1971. Field tests on piles in sand. Soils and
(ii) The Dutch averaging technique to deduce qc has Foundations 11(2): 2949.
been shown to capture the effects of strongly Beringen, F.L., Windle, D. & Van Hooydonk, W.R. 1979.
Results of loading tests on driven piles in sand. Proc.
layered startigraphy and is recommended for
Conf. on Recent Development in the Design and
such cases. Construction of Piles, ICE, London: 213225.
(iii) Excluding the Ogeechee River site and the cor- Briaud, J.-L. & Tucker, L. M. 1989. Axially loaded 5 pile
rected base capacities at Baghdad, the database group and single pile in sand. Proc. 12th International
of driven piles indicates qb0.1/
qc  0.59 with a Conference on Soil Mechanics and Foundation
COV of 0.15. Engineering: 11211124.

738

Copyright 2005 Taylor & Francis Group plc, London, UK


Bustamante, M. & Gianeselli, L. 1982. Pile bearing capacity Lehane, B.M., Pennington, D. & Clark, S. 2003. Jacked end-
predictions by means of static penetrometer CPT. Proc. bearing piles in the soft alluvial sediments of Perth.
2nd European Symposium on penetration testing 2: Australian Geomechanics 38(3): 123133.
493500. Meyerhof, G.G. & Valsangkar, A.J. 1977. Bearing capacity
Chow, F.C. 1997. PhD Thesis. Investigations into the behav- of piles in layered soils. Proc. 9th Int. Conf. Soil
iour of displacement piles for offshore foundations, Mechanics & Foundation Engineering (1): 645650.
Imperial College. Paik, K., Salgado, R., Lee, J. & Kim, B. 2003. Behavior of
De Beer, E., Lousberg, D., De Jonghe, A., Carpentier, R. & open- and closed-ended piles driven into sands. Geotech-
Wallays, M. 1979. Analysis of the results of loading tests nical and Geoenvironmental Engineering 129(4):
performed on displacement piles of different types and 296306.
sizes penetrating at a relatively small depth into a very Randolph, M.F. 2003. Science and empiricism in pile foun-
dense layer. Proc. Conf. on Recent Development in the dation design. Geotechnique 53(10): 847875.
Design and Construction of Piles, ICE, London: 199211. Rollins, K.M., Miller, N.P. & Hemenway, D. 1999.
Eslami, A. & Fellenius, B.H. 1997. Pile capacity by direct Evaluation of pile capacity prediction methods based on
CPT and CPTu methods applied to 102 case histories. cone penetration testing using results from I-15 load
Canadian Geotechnical Journal 34(6): 88690. tests. Transportation Research Record 1675: 4050.
Fellenius, B.H. 1989. Prediction of pile capacity. Proc. Schmertmann, J.H. 1978. Guidelines for cone test, perform-
Symposium on Predicted and Observed Behaviour of ance, and design. U.S. Federal Highway Administration,
Piles, ASCE Special Geotechnical Publication 23: FHWATS-78209
293302. Van Mierlo, W.C. and Koppejan, A.W. 1952. Lengte en
Fugro Engineers, B.V. (Fugro) 2004, Axial pile design draagvermogen van heipalen, Bouw, January.
method for offshore driven piles in sand. Fugro Report Vesic, A.S. 1970. Tests on instrumented piles, Ogeechee
No. P1003, Issue 3 to API, 5 August 2004. River site. Journal of the Soil Mechanics and Foundations
Gavin, K. & Lehane, B.M. 2003. End bearing of small pipe Division ASCE SM2: 561584.
piles in dense sand. BGA International Conference on White, D.J. & Bolton, M.D. 2005. Comparing CPT and pile
Foundations: 321330. base resistance in sand, ICE, Geotechnical Engineering
Gregersen, O.S., Aas, G. & DiBiagio, E. 1973. Load tests on 158, 314.
friction piles in loose sand. Proc. 8th Int. Conf. Soil Xu, X. & Lehane, B.M. 2005. Centrifuge study of displace-
Mechanics & Foundation Engineering Moscow (2): ment pile behaviour in sand. in prep.
109117. Yen, T.-L., Chin, C.-T. & Wang, R.F. 1989. Interpretation of
Jardine, R.J., Chow, F.C., Overy, R. & Standing, J. 2005. ICP instrumented driven steel pipe piles. Foundation
design methods for driven piles in sands and clays, Engineering: Current Principles and Practices ASCE:
Thomas Telford. 12931308.
Lehane, B.M. 1992. PhD Thesis. Experimental investiga-
tions of pile behaviour using instrumented field piles.
Imperial College.

739

Copyright 2005 Taylor & Francis Group plc, London, UK


The influence of effective area ratio on shaft friction of
displacement piles in sand

D.J. White
University of Cambridge, UK

J.A. Schneider & B.M. Lehane


University of Western Australia, Perth, Australia

ABSTRACT: As a cone penetration test (CPT) induces similar strain paths to closed ended pile installation,
its results are directly applicable to the evaluation of radial stress on closed ended piles. For CPT data to be used
to evaluate the shaft capacity of open-ended piles, an understanding of the relative magnitude of radial stress on
open and closed-ended piles is necessary. This relative magnitude is explored using cavity expansion theory to
simulate the stress field as soil flows around a pile tip. The end condition of the pile affects the flow field. These
analyses allow the difference in shaft friction on open and closed-ended piles in sand to be linked to the pile area
ratio and plugging behaviour, with friction angle having a small influence. A function of the effective area ratio,
which combines the area ratio of the pile and the incremental filling ratio, is proposed for estimating the differ-
ence in radial stress along open-ended piles as compared to closed-ended piles.

1 INTRODUCTION suggested based on the mode of installation during an


increment of penetration, as defined by the effective
The ultimate local shaft friction, f, along a displace- area ratio, Ar, which is the ratio of the added volume
ment pile is a function of the radial stress, rf, and (i.e. the gross pile volume minus any soil entering the
interface friction angle, f, at failure as shown by plug) to the gross pile volume1:
Lehane et al. (1993):

(1) (2)

The form of Equation 1 is similar to that of the current


Large-diameter thin-walled caissons, which are
American Petroleum Institute (API) (2000) design
increasingly being considered in new offshore fron-
method for sand, in which rf is specified as the in
tiers, displace a minimal volume of soil compared to
situ vertical effective stress, vo, multiplied by an
their gross area (Ar 0), leading to only a small
earth pressure coefficient, Kf. Values of Kf  0.8 and
stress increase above K0 for monotonic installation.
1.0 are recommended for open and closed-ended piles
Conversely, an open-ended pile penetrating in a
respectively, and f is linked to density and particle
plugged manner causes the same soil displacement as
size (API 2000).
a closed-ended pile (Ar  1). For IFR  0, Ar  4t/D0
Hence, closed-ended piles are predicted to mobil-
which is 0.1 for typical offshore piles, reducing to
ize 25% more shaft resistance than open-ended piles.
0.01 for thin-walled caissons.
The effective radial stress on the shaft of a displace-
This paper examines the influence of pile end con-
ment pile is greater than K0vo in most sands, reflect-
dition on shaft friction using cavity expansion as an
ing an increase in radial stress during pile installation.
analogy for the penetration process. Area ratios ranging
The higher value for a closed-ended pile is logical,
from zero to unity are considered, leading to a tenta-
since a greater volume of soil must be displaced, and
tive design recommendation.
is observed in experimental studies (Gavin & Lehane
2003).
This simple classification as closed- and open- 1
Do  pile outer diameter. Di  pile inner diameter
ended piles is a limitation. Instead, a classification is IFR  incremental filling ratio, (plug length)/(tip depth).

741

Copyright 2005 Taylor & Francis Group plc, London, UK


2 BACKGROUND the deformation of a soil element below the pile tip is
influenced by the end condition.
Although the API (2000) method links shaft friction The R* modification in Equation 3b can be recast
to the in situ effective stress, it is now more common as a multiplication factor of (R/R*)c  A
c/2 r on
to link shaft friction to cone penetration test (CPT) tip Equation 3a. This form shows more clearly how R*
resistance, qc. This is logical since the soil adjacent to and the variable c lead to a difference in predicted
the pile shaft experienced a stress related to qc (as the shaft resistance due to end condition.
pile tip passed that point) more recently than vo.
Lehane & Jardine (1994) link the radial stress on a
closed-ended pile after installation and consolidation, 3 MODELLING OF OPEN AND
rc, to qc and the normalized distance behind the pile CLOSED-ENDED PILES
tip, h/R:
Cavity expansion (CE) analysis is widely used, with
success, to model penetration problems such as the
(3a) cone penetration test as well as pile capacity (e.g.
Salgado et al. 1997, Yu & Houlsby 1991).
In this paper, the radial stress created by the expan-
where h is the distance above the tip of the pile, R is sion of a cavity to accommodate piles of varying effect-
the radius of the pile, and h/R takes a minimum value ive area ratio is examined as a means of predicting the
of 8. The constants a and c take values of 0.024 influence of effective area ratio on pile shaft resistance.
and
0.33 respectively. The (h/R)c term accounts for For open-ended piles, the streamlines of soil flow
the phenomenon of friction fatigue that can be differ from the closed-ended case, depending on the
attributed to cyclic installation effects (Lehane 1992, effective area ratio, Ar. This is shown conceptually in
Poulos 2000, Randolph 2003, White & Bolton 2004, Figure 1, with the corresponding radial stress increas-
and White & Lehane 2004). ing with Ar, based on observations by White & Bolton
Equation 3a can be divided into two parts such that (2004) and Gavin & Lehane (2003).
aqc represents a maximum radial stress close to the The lower radial displacement required for open-
pile tip, and (h/R)c quantifies the attenuation of this ended or partially-plugged penetration leads to a
stress along the pile shaft due to friction fatigue. lower radial stress, r (and hence shaft friction)
To account for differences in radial stress on open- behind the pile tip. In this paper, cavity expansion
ended as compared to closed-ended piles, a modifica- analysis is used to link the ratio of radial stress on
tion factor is applied in one of two ways. Either, a open and closed-ended piles to Ar.
reduction factor can be applied to the maximum radial
stress on the pile shaft (API 2000, NGI 2001, Randolph 3.1 Cavity expansion analysis
2004, and Tomlinson 2001), or the rate of radial stress
reduction along the shaft can be changed (Jardine & The proportional reduction in r as a function of Ar is
Chow 1996, and Fugro 2004/Kolk et al 2005). examined using the cavity expansion solution for a
The former method represents a modification of the soil following a Coulomb failure criterion.
variable a, while the latter method can be a modifica- Close to the pile (i.e. the cavity wall), the soil is at
tion to c, or to R. Jardine & Chow (1996) recommend failure, and the stress distribution can be found by
substituting R* for R, where R*/R  A0.5 r assuming combining the yield criterion (Equation 4) with the
unplugged penetration (IFR  1, Equation 3b). R* radial equilibrium equation (Equation 5). The result-
for an open-ended pile is the radius of an equivalent ing expression describes the decay in radial stress
closed-ended pile of the same solid area. within the plastic zone in terms of the Rankine passive
earth pressure coefficient Kp  (1  sin )/(1
sin )
(Equation 6). In this expression m  1 for a cylindrical
(3b) cavity and 2 for a spherical cavity.

(4)
There is a stronger logical basis for applying a
modification to the variable a than to c or R to
account for end condition. It is the penetration mode
(5)
close to the tip, leading to the maximal stress aqc, that
is influenced by the end condition, not the friction
fatigue behaviour along the shaft, quantified by
(h/R)c. A soil element adjacent to the shaft is no (6)
longer influenced by the pile end condition. However,

742

Copyright 2005 Taylor & Francis Group plc, London, UK


Figure 1. Schematic streamlines of soil flow and profiles of radial stress (r  radial displacement of soil element at pile wall).

The cavity pressure corresponding to pile or CPT


penetration is typically 20100 times the in situ verti-
cal stress, reflected in values of bearing capacity fac-
tor Nq. For this stress ratio, the normalized plastic
radius (beyond which the soil is no longer at failure)
is in excess of 20 for practical values of friction angle 
(using the derivation given by Yu & Houlsby 1991).
Therefore, in the region of interest for this study
(r/R 20), the radial stress decay is given by
Equation 6, and depends only on the friction angle.

3.2 Cavity expansion analogy for end condition


Profiles of the radial decay in radial stress normalized
by the pressure at the cavity (or pile) wall for spher-
ical and cylindrical cavity expansion are shown in
Figure 2. The radial distance from the centreline of a
solid pile, r, is normalized by R.
The profiles show only a weak dependence on fric-
tion angle. Although cavity expansion pressures in
sand are very sensitive to the constitutive parameters, Figure 2. Cavity expansion analysis of radial stress on
Figure 2 shows that the radial decay in stress away open and closed-ended piles.
from the cavity is not.
The radial stress on an open-ended pile (Ar 1) of open-ended pile installation as when a closed-ended
radius R is estimated from the stress decay around a pile displacing the same volume of soil is installed.
solid (closed-ended) pile of equivalent radius Req as The open-ended pile wall lies at a radius, r  R,
follows. It is assumed that soil in the far field experi- where the enclosed area is R2. At the equivalent solid
ences the same displacements and stresses during pile wall the enclosed area is R2eq. Therefore, following

743

Copyright 2005 Taylor & Francis Group plc, London, UK


the earlier definition of effective area ratio as the
added volume ( R2eq) divided by the gross pile volume
( R2), the wall of an open-ended pile lies at a radius
of R/Req  A
0.5 r , giving the second x-axis of Figure 2.
This method of estimating the stress on an open-
ended pile from the radial profile of stress around an
equivalent closed-ended pile was used previously by
Whittle (1992), Carter et al (1980), and Randolph
(2003) for undrained strain path and cavity expansion
analyses.
A small deviation from this assumption arises dur-
ing drained penetration due to any volume change in
the soil between the equivalent closed-ended pile and
the wall of the open-ended pile.
If it is assumed that the same proportional drop in
stress occurs immediately behind the pile tip, from qb
(or qc) to r, independent of the pile end condition,
then the above analogy can be used to evaluate
r,open/r,closed close to the pile tip. Figure 3. End condition index, b, vs. friction angle.
This stress ratio can be linked to Ar and friction
angle by substituting the cavity expansion analogy for
end condition (which assumes that the wall of an
open-ended pile lies at normalized radius R/Req away shown that the mean of these values is appropriate for
from an equivalent solid pile) into Equation 6. Hence: a constant- analysis (Collins et al 1992). In addition,
the friction angle will vary with stress level in the
manner shown by Bolton (1986), but will cause min-
imal change in b (Figure 3).
(7)
The corresponding radial stress ratios,

r,open/r,closed for typical (unplugged) open-ended piles
with D0/t  20 and 40, corresponding to Ar  0.2 and
where the end condition index, b, is defined as: 0.1 are highlighted on Figure 2. These are in the range
4060 percent, based on the cylindrical cavity expan-
sion curve for   35.
(8) It should be noted that for very thin-walled suction
caissons (Ar 0.01) the assumption of a plastic
radius beyond the pile wall may not apply, so
Based on the proposed hypothesis, this expression Equation 7 will no longer hold and the in situ stress
provides a theoretical solution for a reduction factor will influence the stress at the pile wall.
to extend Equation 3a to open-ended piles. Noting
that the net deformation is cylindrical, a value of
m  1 is often adopted for the analysis of penetration 4 RADIAL STRESS VARIATION NEAR PILE
problems. Cylindrical rather than spherical cavity
expansion was used by Salgado et al (1997), who The cylindrical cavity expansion stress field given by
showed good agreement between analytical cavity Equations 46 can be extended to include the reduc-
expansion solutions and CPT data. tion in stress behind the tip. This extension provides
A cylindrical analogy is also supported by the cali- insight into the stress field around a pile shaft, which
bration chamber results of Houlsby & Hitchman may influence the widely observed phenomenon of
(1988) who showed that cone resistance, qc, cor- increasing shaft capacity with time (set-up).
relates better with h than either mean stress, p
Field and model tests indicate that the radial stress
(which would be expected for a spherical cavity which acts a short distance behind the tip of a closed-
expansion process), or 
v. ended pile or CPT is typically between 1 and 2.5 per-
The end condition index, b, is relatively insensitive to cent of qc (DeJong 2001; White & Lehane 2004). This
friction angle (Figure 3), and a single value for typical range, multiplied by a sand-steel friction coefficient of
friction angles and cylindrical expansion is b  0.35. 0.35, agrees with typical CPT friction ratios in sand.
The actual operational friction angle at a given Ignoring any elastic response, the stress field in
location will vary from the peak to critical state value this region can be modelled in an approximate fash-
as deformation progresses. Numerical analyses have ion by adding a cavity contraction stage to the stress

744

Copyright 2005 Taylor & Francis Group plc, London, UK


can be incorporated into a CPT-based design approach
(Equation 3a) as follows:

(9a)

Substituting Equation 7 into Equation 9a:

(9b)

where Ar varies with IFR (Equation 2). Equation 8


links the end condition index, b, to friction angle  and
for typical values b  0.35, if a cylindrical analogy is
used. For spherical expansion this value is doubled.
Since the IFR is not known prior to installation, it
Figure 4. Simple analysis of radial variation in stress near is generally conservative for design to assume that
pile after unloading behind tip. IFR  1. For large diameter field tests of relatively
thick walled piles (D0/t  20), average IFR values are
typically 1.1 to 1.15 (Ar  0.1), although values as
low as 0.75 were measured near the end of driving for
distribution shown in Figure 2. This contraction fol- the EURIPIDES piles (Ar  0.4). This partial plug-
lows the yield condition and equilibrium equations ging may lead to higher levels of radial stress on
shown previously, but with failure occurring in the instrumented test piles as compared to typical piles
near field with 
r . The resulting radial variation used for offshore structures. Data in Gavin & Lehane
in stresses for a cavity pressure reduction to 1% of the (2005) show that for driven piles in medium dense to
initial value is shown in Figure 4. very dense sands, IFR generally increases with pile
It is notable that in the near field (r/R 4), the cir- diameter. For pile diameters greater than 0.5 m, IFR is
cumferential stress exceeds the radial stress, due to generally between 0.8 and 1.1, while for smaller piles
the unloading process. These conditions would tend (D 0.5 m) IFR is generally between 0.5 and 0.9, for
to cause any subsequent creep-induced equalization which Ar may be as high as 0.65. As most piles in
of stresses to create an increase in 
r at the pile wall, databases used to calibrate pile design methods in
and hence lead to set-up of shaft resistance (White & sand are 0.5 m diameter or less, partial plugging is
Bolton 2004). likely to have a significant effect on the interpretation
Figure 4 shows a two- to ten- fold reduction in 
r of those databases.
for values of 3  r/R  6 compared with the 100 Figure 5 compares the influence of area ratio on
times reduction at the pile shaft. This very simple r,open/r,closed shown by this study with values
analysis of the stress change is comparable to in-soil implied by recent pile design methods.
stress measurements during the passing of a nearby The R* modification method recommended by
pile tip presented by Gavin & Lehane (2003) and Jardine & Chow (1996) and Fugro (2004)/Kolk et al
Leung et al (1996). Whilst the on-pile stress drops by (2005) implies a reduction factor with the exponent b
two-orders of magnitude (as evidenced from CPT set equal to
c/2 and an IFR of unity assumed.
friction ratio data), a smaller reduction in stress The API (2000) and Jardine & Chow (1996) meth-
occurs remote from the pile. ods indicate the highest values of 
r,open/r,closed. In
the case of the Jardine & Chow (1996) method, this
difference can be attributed to the extension of this
5 DESIGN IMPLICATIONS method from the closed-ended approach of Lehane &
Jardine (1994). The R* extension was calibrated using
The preceding analysis has linked the reduced shaft the CLAROM open-ended pile tests at Dunkirk, France
friction on an open-ended pile (compared to a closed- (Brucy et al 1991). Partial plugging occurred during
ended pile) to the lower radial stress generated due driving of these relatively small-diameter pipe piles,
to the smaller volume of soil displaced during installa- leading to an effective area ratio of Ar between 0.55
tion. This approach is in agreement with recommen- and 0.65. (following Equation 1), which is excessive
dations by API (2000), NGI (2001), Randolph (2004), for offshore piles. If the Jardine & Chow (1996)
and Tomlinson (2001). The effect of end condition recommendation shown on Figure 5 is shifted to this

745

Copyright 2005 Taylor & Francis Group plc, London, UK


the influence of end condition is required in order to
design open-ended piles (Ar 1). This paper presents
cavity expansion analyses that simulate the ratio of
shaft friction behind the tip of open- and closed-
ended piles. These analyses provide a theoretical
basis for linking pile end condition to shaft resistance.
In general, these analyses agree with current design
methods. However, discrepancies have been shown
between methods that are calibrated against overall
pile capacity. This indicates that compensating errors
may exist when accounting for the attenuation of
radial stress due to (a) end condition and (b) friction
fatigue, among other factors such as time effects, and
the estimation of base resistance.

ACKNOWLEDGEMENTS

We gratefully acknowledge an IPRS scholarship and


a UWA university postgraduate travel award for the
second author to visit the University of Cambridge.
Figure 5. Ratio of radial stress on open and closed-ended
piles as a function of area ratio.
REFERENCES

value of Ar, good agreement is found with the analy- API 2000. Recommended Practice for Planning, Designing
sis in this paper. and Constructing Fixed Offshore Platforms-Working Stress
It should be noted, however, that Jardine & Chow Design, RP2A-WSD 21st ed., American Petroleum Inst.
Bolton, M.D. 1986. The strength and dilatancy of sands.
(1996) do not explicitly define 
r,open/r,closed, but Gotechnique, 36(1):6578.
couple this calculation with friction fatigue further Brucy, F., Meunier, J. & Nauroy, J.F. 1991. Behaviour of a
along the pile shaft (as described in Section 2). Since pile plug in sandy soils during and after driving. Offshore
their method was calibrated against load test results, Technology Conference, Houston, OTC6514:145154
any unconservatism implied by Figure 5 must be Carter, J.P., Randolph, M.F. & Wroth, C.P. 1980. Some
counteracted by a conservative assessment of friction aspects of the performance of open- and closed-ended
fatigue further along the shaft, by the underestimation piles. Num. Methods in Offshore Piling, ICE:165170.
of base resistance (Kolk et al 2005), or other factors. Collins, I.F., Pender, M.J. & Wang, Y. 1992. Cavity expansion
A more conservative assessment of the radial stress in sands under drained loading conditions. Int. J. Numerical &
Analytical Methods in Geomechanics 16: 323.
differences between open and closed-ended piles is to DeJong, J.T. 2001. Investigation of particulate-continuum
reduce r,closed in direct proportion to the area ratio interface mechanisms and their assessment through a
(b  1). A similar approach was presented by Gavin & multi-friction sleeve penetrometer attachment. PhD Thesis,
Lehane (2003), in which a linear reduction to a stress Georgia Institute of Technology, Atlanta, GA, USA:361 pp.
ratio of 0.1 is proposed. This method closely matches Fugro Engineers B.V. (Fugro) 2004. Axial pile design
the spherical cavity expansion analysis in this paper method for offshore driven piles in sand. Fugro Report
(Figure 5). No. P1003, Issue 3 to API, 5 August 2004:122 pp.
The cylindrical cavity expansion analysis in this Gavin, K.G. & Lehane, B.M. 2003. The shaft capacity of
paper agrees closely with the recommendations by pipe piles in sand. Canadian Geotech. J., 40 (1):3645.
Gavin, K.G. & Lehane, B.M. 2005. Estimating the end bear-
Fugro (2004)/Kolk et al (2005), and Tomlinson ing resistance of pipe piles in sand using the Final Filling
(2001), and provides a theoretical basis to justify how Ratio Proc. Int. Symp. on Frontiers in Offshore Geotechnics
these design methods account for pile end condition. Houlsby, G.T. & Hitchman, R. 1988. Calibration chamber tests
of a cone penetrometer in sand, Gotechnique, 28(1):
3944.
Jardine, R.J. & Chow, F.C. 1996. New design method for off-
6 CONCLUSIONS shore piles. MTD Publication 96/103, Marine Technology
Directorate, London:48 pp.
Design methods for piles in sand are increasingly based Kolk, H.J., Baaijens, A.E. & Senders, M. 2005. Design cri-
on results from the CPT, which is a closed ended pen- teria for pipe piles in silica sands. Proc. Int. Symp. on
etrometer (Ar  1). Therefore, an understanding of Frontiers in Offshore Geotechnics.

746

Copyright 2005 Taylor & Francis Group plc, London, UK


Lehane, B.M. 1992. experimental investigations of pile behav- Salgado, R., Mitchell, J.K. & Jamiolkowski, M. 1997.
iour using instrumented field piles, PhD Thesis, University Cavity expansion and penetration resistance in sand. ASCE
of London (Imperial College), London, U.K.:615 pp. J. Geotech. & Geoenvironmental Engng, 123(4):344354.
Lehane, B.M., Jardine, R.J., Bond, A.J. & Frank, R. 1993. Tomlinson, M.J. 2001. Foundation Design and Construction,
Mechanisms of shaft friction in sand from instrumented 7th Edition, Prentice Hall, Sydney:569 pp.
pile tests, J. of Geotech. Engng, 119 (1), ASCE:1935. White, D.J. & Bolton, M.D. 2004. Displacement and strain
Lehane, B.M. & Jardine, R.J. 1994. Shaft capacity of driven paths during plane strain model pile installation in sand.
piles in sand: a new design approach. Proceedings, Gotechnique, 54(6):375398.
Behavior of Offshore Structures (BOSS) 2336. White, D.J. & Lehane, B.M. 2004. Friction fatigue on dis-
NGI 2001. Bearing capacity of driven piles in sand. Report placement piles in sand. Gotechnique, 54(10):645658.
525211-2, Norwegian Geotechnical Institute (NGI), Oslo. White, D.J. 2005. A general framework for shaft resistance
Poulos, H.G. 2000. Some aspects of pile skin friction in cal- on displacement piles in sand. Proc. Int. Symp. in
careous sediments. Engineering for Calcareous Sediments, Frontiers in Offshore Geotechnics. Perth.
Balkema, Rotterdam:457471. Whittle, A.J. 1992. Assessment of an effective stress analy-
Randolph, M.F. 2003. Science and empiricism in pile foun- sis for predicting the performance of driven piles in clays.
dation design. Gotechnique, 53 (10):847875. Proceedings, Offshore Site Investigation and Foundation
Randolph, M.F. 2004. Comments on Fugro Report to API on Behaviour, London, Vol. 28:607643.
Pile Capacity in Sand. submitted to Fugro Engineers Yu, H.S. & Houlsby, G.T. 1991. Finite cavity expansion in dila-
B.V., Leidschendam, The Netherlands:2 pp. tant soils: loading analysis. Gotechnique, 41(2):173183.

747

Copyright 2005 Taylor & Francis Group plc, London, UK


A centrifuge study of the monotonic and cyclic resistance of
piles and pile groups in sand

C. Gaudin & B.M. Lehane


Centre for Offshore Foundation Systems/School of Civil & Resource Engineering, University of Western Australia

P.F. Wallis
Arup (Brisbane)

ABSTRACT: This paper describes a series of centrifuge experiments in which model displacement piles and
pile groups were subjected to monotonic and cyclic tension loading. The experiments employed a new tech-
nique that allowed pile group installation and load testing to be conducted without the need to halt the cen-
trifuge. Notwithstanding scale effects associated with the development of shaft friction on miniature piles, the
experiments provide strong indications that the cyclic resistance of pile groups is significantly higher than that
of single piles at comparable levels of tension cycling.

1 INTRODUCTION of five piles. Reference cyclic tests on single piles as


well as static tests on single piles and pile groups
Pile groups are often subjected to cyclic tension loading were also performed to assist interpretation of the
by wave or wind loading. Methodologies to assist the effects of cycling.
design of such groups are, however, not well developed This particular project was instigated by the needs
and practitioners usually take the approach of adopt- of designer engineers at Arup in Singapore, who
ing a larger factor of safety than would be used for a required guidance on the assessment of cyclic effects
statically loaded pile group. The difficulty is com- for pile groups planned for a newly proposed Ferris
pounded by the fact that there are very few reported Wheel the Singapore Flyer (similar to the Millennium
case histories of cyclically loaded pile groups. Wheel constructed in London). Each pile group for
Jardine & Standing (2000) addressed the effects of this proposal would be founded in medium dense and
high and low level cyclic tensile loading on single steel dense sand and was anticipated to contain 5 piles; the
tubular piles driven in sand. Their data show that, while pile groups were required to sustain low level cycling
high level cyclic loading leads to a progressive degrad- from wind loading over the structures design life.
ation in available shaft capacity, low level cycling
actually enhanced the shaft capacity. The effects of
2 TESTS APPARATUS AND PROCEDURE
cycling on shaft friction have been also examined by
White & Lehane (2004) in centrifuge tests on single
2.1 Experimental setup
displacement piles in dense sand. Shaft capacity was
shown to degrade with the number of cycles; this The experiments were undertaken in the geotechnical
degradation was considered to be due to contraction of drum centrifuge at The University of Western Australia
the sand at the pile-sand interface during cycling and (UWA). A complete description of the drum cen-
the consequent reduction of the horizontal stresses act- trifuge facility at UWA is given by Stewart et al. (1998).
ing on the shaft. Five model 125 mm long aluminium piles with a
While some insights into the effects of cycling on diameter of 10 mm and wall thickness of 1 mm were
single piles have been obtained, there is a dearth of employed in each pile group tested; single (reference)
information available concerning the performance of piles had the same configuration. The piles in the
cyclically load pile groups in sand. This paper addresses groups were arranged as shown on Figure 2 with a
this need by presenting the results from a systematic relatively rigid pile cap. The spacing ratio, defined as
model scale study of the relative effects of the magni- the distance between the centre of the centre pile to
tudes of tension load cycles and sand density on groups the centre of a corner pile (s) normalised by the pile

749

Copyright 2005 Taylor & Francis Group plc, London, UK


diameter (D) was either 3 or 4. The surface of the The five load cells are mounted on the cap of the
piles was sandblasted in order to achieve a centre line pile group in order to achieve the required spacing
average roughness RCLA about 7 m. The normalised ratio (Figure 2). After the five piles have been jacked
roughness Rn, defined as the ratio between RCLA and monotonically into the sand sample, one after another
the mean particle size of the sand D50 was about 0.035. (the centre one first then corners 1 to 4; see Figure 2),
Specific plug-inload cells were developed to facili- the cap is mounted on the actuator and driven towards
tate their connection to the piles after installation. the pile group in order to achieve the connection and
These load cells constitute a key feature of the experi- then to apply the load. This system permits modeling
ments described here as they permit installation and of the full sequence from installation to loading of a
load testing of the pile group at the desired centrifugal pile group without stopping the centrifuge.
acceleration without the need to halt the centrifuge.
The plug-in load cells are shown in Figure 1. They 2.2 Soil preparation and experimental program
are divided in two parts (i) a lower part containing the
load cell, which is screwed to the head of each pile and The sand used to prepare the centrifuge sample is a
(ii) an upper part which is connected to the cap and fine sand, which is usually used for centrifuge testing
the actuator. Features of the plug-in load cell include at UWA. Classification data are presented in Table 1.
a mechanical connection to catch and engage piles Two saturated samples were prepared for this study.
after their installation and an electrical connection to For sand preparation, a specialized sand placement
link the load cell to the data acquisition system. actuator was fitted to the tool table in place of the
The mechanical connection is assisted by circular multi-axis loading actuator. The sand was poured into
collars mounted on the lower part of the plug-in load the channel via a hose at 20 g, keeping the height of
cell. These are tightened by a rubber band and are fall constant by driving back the actuator as the height
located in a specific slot on the upper part after the of sand in the channel increased. Water was then added
latter has moved inside the lower part. The connection to the sample and subsequently drained out to induce a
allows the pile to be subjected to either compression suction and allow the centrifuge to be halted. The sand
or tension loading. surface was levelled after which the channel was spun
The electrical connection is achieved using silver back to the target acceleration level. Water then was
plates placed at the bottom of the lower part and spring allowed to infiltrate from the base of the channel to a
connectors placed inside the upper part. The four height approximately 10 mm above sample surface.
silver plates, quarter circle shaped, are connected to The complete height of each sand bed was 168 mm.
the four strain gauges of the load cell. Four spring Although the preparation process was similar for these
connectors are placed in a square arrangement in the samples, the relative density achieved was different.
upper part and are connected to the actuator and then This discrepancy between target relative densities is
to the data acquisition system. When the mechanical likely to be due to the difficulties encountered while
connection is completed, the spring connectors are in attempting to control the height of fall accurately
contact with the silver plates, enabling transmission
of the electrical signal of the load cell to the data
acquisition system.

Figure 2. Plug-in load cell mounted on the cap and pile


group arrangement.

Table 1. Classification data of the UWA sand.

D60 D50 D10


emax emin Gs (mm) (mm) (mm) Cu Cc

0.76 0.49 2.65 0.22 0.20 0.12 1.77 1.08


Figure 1. A pile and a plug-in load cell.

750

Copyright 2005 Taylor & Francis Group plc, London, UK


when the sand is placed. Core samples were taken Installation curves of single piles in both samples (not
after completion of all tests in each sample to assess presented here) have shown (i) almost identical axial
the relative density, giving 75% for the sample 1 and stress profiles for installation in sample 1 (ii) some-
45% for the sample 2. As the relative density of the what less regular profiles in sample 2 with deviations of
two samples were different, sample 1 was spun at 80 g up to 20% between installation curves and (iii) a simi-
while sample 2 was spun at 160 g in order to achieve lar installation resistance in the two samples, although
similar installation resistance and similar pull-out their acceleration levels differed by a factor of 2.
capacity for both samples. The mean final installation load ranged from 3 kN
The experimental programme is summarised in to 3.5 kN in both samples. The base resistance may be
Table 2. Typical (and representative) results are pre- estimated by assuming that the average shaft shear
sented here while all results are included in Gaudin & stress measured in tension tests was approximately the
Lehane (2004). same as the compression average shear stress developed
during installation. This calculation indicates that
about 90% of the installation resistance is derived
2.3 Pile installation
from the pile base. Assuming that the base resistance of
Typical piles axial stresses (defined as the head load the pile is roughly equal to the tip resistance of a cone
divided by the tip area) recorded during installation penetrometer, the relative density of both samples was
are presented in Figure 3a for a pile group in sample 2. assessed using the correlation between the tip resist-
ance and vertical effective stress given by Lunne &
Table 2. Testing program. Christofferson (1983). Results are plotted in Figure 3b
and show that (i) relative densities are consistent with
Sample Test Pile s/D Loading Amplitude1 those obtained from the core samples (ii) the relative
(%)
density of the sample 1 is nearly constant with depth
but that of sample 2 increases slightly with depth. The
1 S1-S1-M Single Monotonic installation axial stresses of the individual group piles
1 S1-S1-C Single Cyclic 1248 presented in Figure 3a show that corner piles require a
1 S1-G3-M Group 3 Monotonic
1 S1-G3-C Group 3 Cyclic 2956
higher installation load than the centre pile. The final
2 S2-S1-M Single Monotonic axial stress for corner pile 1 is typically 10% higher
2 S2-S2-M Single Monotonic than that of the centre pile. Final axial stresses of corner
2 S2-S1-C Single Cyclic 1336 piles 2, 3 and 4 are again 10% higher than for corner
2 S2-S2-C Single Cyclic 13 pile 1, with no significant increase between the three
2 S2-G3-M Group 3 Monotonic corner piles. These trends illustrate the relative influ-
2 S2-G3-C Group 3 Cyclic 6 ence of the disturbance created by penetration of piles
2 S2-G4-M Group 4 Monotonic in sand which is primarily due to increases in the lat-
2 S2-G4-C1 Group 4 Cyclic 936 eral stresses within the soil mass rather than increases
2 S2-G4-C2 Group 4 Cyclic 972
in in-situ density (Gaudin et al., 2005).
1
Expressed as a percentage of the monotonic peak average
shaft shear stress of the monotonic single pile test.
3 MONOTONIC LOADING
Pile axial stress (MPa) Relative density (%)
0 20 40 60 0 20 40 60 80 100 3.1 Single pile
0 0
Center pile The average shear stress developed along the pile dur-
20
20 Corner pile 1
ing pull-out is defined as the tensile load divided by
Corner pile 2
Corner pile 3 40
the shaft area of the pile embedded in the soil. The
40 Corner pile 4 average peak shear stress (p) will subsequently be
Depth (mm)

used for comparative purposes.


Depth (mm)

60
60 Typical monotonic pull-out curves for single piles
80
are presented in Figure 4 for both samples (denoted
80
100
s1 and s2). All exhibit a similar peak average shear
stress (p) about 80 kPa, illustrating the good repeat-
100 120 Sample 2 ability of the tests within the two samples. The similarity
Sample 1 of p values for both samples, despite their different
120 140
(a) (b)
relative densities, coupled with the similarity of the
pile installation axial stresses (qc) for the two sam-
Figure 3. Pile axial stresses during installation (a) and ples confirm the validity of approaches which relate
inferred relative density (b). p to the CPT qc value.

751

Copyright 2005 Taylor & Francis Group plc, London, UK


100 400
90 350
Average shear stress (kPa)

80

Tensile Load (N)


300
70
250
60
200
50
150
40 S2-S1-M Center pile
30 S2-S2-M
100 Corner pile 1
20 S1-S1-M 50 Corner pile 3
10 0
0 1 2 3 4 5 6
0
0 1 2 3 4 5 6 Displacement (mm)
Displacement (mm)
Figure 5. Typical load-displacement curves for pile group
Figure 4. Average shear stress-displacement curves for tests performed in sample 1.
single piles in samples 1 and 2.
100

Average shear stress (kPa)


90
This p value of 80 kPa is higher than the range of 80
40 to 50 kPa observed in centrifuge experiments on 70
60
equivalent buried piles (i.e. same diameter, sand dens- 50
ity, g-level, pile length and pile surface roughness, 40
Lehane et al., 2005). Monotonic jacked pile installation 30 Center pile
Average corner piles
in sand can therefore be expected to lead to almost 20
Single pile
twice the shaft resistance of a buried pile with the 10
0
same roughness. As will be seen later, however, the 0 1 2 3 4 5
cycling imposed during actual pile installations in Displacement (mm)
the field is likely to reduce their available shaft capacity.
The peak average shear stress is reached after Figure 6. Comparison between performance of group piles
0.6 mm, i.e. 6% of the diameter which contrasts with and single pile (sample 1).
a displacement of about 1% normally seen for field
scale piles. This difference is due to the localised (and The comparison between a single pile and a pile
non-scalable) dilation of the shear band at the soil- group is presented in Figure 6 where p is plotted
pile interface (Lehane et al., 2005). against head displacements for the single pile, the
centre pile and the corner. It can be seen that p for the
3.2 Pile group corner piles is similar to that of the single pile although
the pile axial stress recorded during the installation
The pile loads recorded during the tension tests on
was higher for corner piles than for the single pile. In
group piles in sample 1 are plotted against vertical
the same way, p for the centre pile (which was the first
displacement in Figure 5. Several trends, which are in
group pile to be installed) is lower than that of the sin-
keeping with those seen in all static tension tests on
gle pile although their respective installation axial
groups, are noted:
stresses were similar.
The loads recorded in the two opposite corner piles The net result is that the tension capacity of the pile
are identical (pile nos 1 and 3). This trend was true group is about 94% of the capacity of five individual
of all group tests and therefore, to simplify presen- piles. This contrasts with data indicated in a case history
tation, only the average load of the two corner piles described by Briaud et al. (1988) involving a com-
is plotted in subsequent figures. pression pile group comprising 5No. 273 mm diameter
The peak tension load of the centre pile is about piles in medium dense sand for which the shaft cap-
30% lower than the tension capacity of the corner acity efficiency was assessed to be 1.83. Such a con-
piles. A smaller difference between centre and cor- trast is believed to be due to (i) the higher relative
ner pile capacity of about 15% was observed in capacity of the single centrifuge piles because of the
sample 2 at both s/D  3 and 4. The reduced absence of installation cycles (White & Lehane,
capacity of the centre pile may be partly due to the 2004) and (ii) potentially higher densification effects
specific installation sequence. This point is still in the sand due to driving of the field scale piles.
however under investigation. The stiffnesses of the single piles are significantly
Interestingly (and perhaps surprisingly) no signifi- higher than those of the group piles (Figure 6). The
cant difference between the stiffness of the centre lower stiffness of group piles, arising from load inter-
pile and corner piles was observed. action, is a well known characteristic; displacements

752

Copyright 2005 Taylor & Francis Group plc, London, UK


to mobile peak uplift resistance are typically about 100
2.5 times those required for the full development of p

Average shear stress (kPa)


90
on single piles. 80
A comparison between the two spacing ratios (not 70
60
presented here) reveals no significant differences
50
between the performance of pile groups with spacing
40
ratio of 3 and 4. p values for the centre pile and cor- 30
S2-S2-C
S2-S1-M
ner piles are almost identical at both s/D values. 20 S2-S1-C
S1-S1-C
10
S1-S1-M
0
4 CYCLIC LOADING 0 1 2 3 4 5 6
Displacement (mm)
4.1 Single pile
Figure 7. Cyclic responses of single piles compared to
The response of single piles at various levels of cyclic monotonic responses.
tension loading are presented in Figure 7 for samples
1 and 2.
This figure shows that the behaviour of a single 4.0 S2-S1-C
pile under tensile cyclic loading depends on (i) the 3.5

Displacement (mm)
density of the sample and (ii) the amplitude of the 3.0
cyclic loading. Low amplitude cyclic loading does 2.5
not lead to a significant degradation in the perform- 2.0 S2-S2-C S1-S1-C
ance of a pile in dense sand (S1-S1-C). After 160
1.5
cycles and about 1.7 mm of vertical displacement (i.e.
1.0
17% of the diameter of the pile), the post cycling value
of p mobilised in a monotonic pull out (S1-S1-C) is 0.5

about 89% of that developed under monotonic loading 0.0


0 50 100 150
with no previous cycling (S1-S1-M). No. of cycles
The accumulation of the vertical displacements of
single piles during tension cycling in both samples is Figure 8. Accumulation of vertical displacements of the
presented in Figure 8. The evolution of the displace- pile under cyclic loading.
ment of the single pile in sample 1 (with Dr  75%)
is approximately linear up to 110 cycles. The pile
100
head displacements generated during each cycle sub-
90
sequently increase slightly and it might be inferred
Average shear stress (kPa)

80
that if further cycling was imposed on this pile, a rapid
70
acceleration in the rate of accumulation of pile head
60
displacement may be expected after about 250 cycles.
50
Unlike sample 1, low amplitude cyclic loading leads
to a significant degradation in the performance of the 40

pile in the medium dense sand of sample 2 after rela- 30 Center pile mono (S1-G3-M)
20 Average corner pile mono (S1-G3-M)
tively few cycles. After 60 cycles and an accumulation Average corner pile cyclic (S1-G3-C)
of 4 mm displacement (i.e. 40% of the diameter of the 10 Center pile cyclic (S1-G3-C)
pile), the post-cycling, p value attained (S2-S1-C) 0
0 1 2 3 4 5
was only 54% of that developed under monotonic Displacement (mm)
loading with no previous cycling (S2-S1-M).
The high amplitude of cyclic loading (S2-S2-C) Figure 9. Cyclic response of the pile group in sample 1
generates almost immediate failure with a very fast compared to the monotonic response.
accumulation of vertical displacements which reach
2 mm (i.e. 20% of the diameter of the pile) after only
4.2 Pile group
2 cycles.
The performance of a single pile is evidently sig- The response of pile groups under cyclic loading is
nificantly affected by cyclic loading with the effects presented in Figure 9 for sample 1. Results from sam-
being more pronounced in looser sand at the same ple 2 indicate similar trends and are not presented
level of cycling (expressed as a fraction of the static here. The average p for the pile group, the centre pile
pile capacity). The post cycling capacity is related to and the corner piles are plotted against the corres-
the degradation (i.e. accumulated displacement) gen- ponding vertical displacements, and compared to
erated by cycling. the response observed in static/monotonic loading.

753

Copyright 2005 Taylor & Francis Group plc, London, UK


3.0 1
S2-G4-C2
Centre group pile (square symbols)
2.5 Single piles (circular symbols)
Displacement (mm)

Dr=45% (closed symbols)


0.8 Dr=75% (open symbols)
2.0

1.5
S1-G3-C 0.6

cy /max static
1.0
2-ways cyclic
loading
0.5
2 1-way cyclic
S2-G3-C S2-G4-C1 0.4
0.0 70 loading
0 100 200 300 400 500 NF
NF
No. of cycles
0.2
NF
150
Figure 10. Accumulation of vertical displacement of pile 55
groups under cyclic loading
0
0 0.2 0.4 0.6 0.8 1
mean/max static
Table 3. Summary of results (all stresses expressed in kPa).

Sample 1 Sample 2 Figure 11. Cyclic interaction diagram.

SP s/D  3 SP s/D  3 s/D  4


(S2-S1-C). No effect of the spacing ratio (s/D) was
p 80 74 80 74 74 observed.
The performance of pile groups is evidently less
Low amplitude cycles affected by cyclic loading than single piles. This find-
cymean 24 32 28 20 17 ing is apparent from the summary of results presented
cymax 38 42 18 34 28 in Table 3 and the cyclic interaction diagram plotted on
pcy 71 73 43 75 74 Figure 11. The latter diagram uses the monotonic (pre-
No. cyc. 140 500 60 280 280
cymean/p 0.3 0.43 0.35 0.27 0.23
cycling) static capacity to normalise the average and
cymax/p 0.48 0.57 0.35 0.46 0.38 amplitude of shaft shear stresses (mean and cy) applied
pc/p 0.89 0.99 0.54 1.01 1.00 during cycling. The number of cycles to failure (arbi-
trarily selected as when the rate of accumulation of
High amplitude cycles displacement accelerates) is shown on this figure; the
cymean 40 27 symbol NF means that no failure was observed.
cymax 72 48
pcy 62
Figure 11 illustrates (a) the lower cyclic resistance
No. cyc. 2 100 of single pile compared to group piles and (b) the
cymean/p 0.5 0.37 higher potential for cyclic failure in looser sand.
cymax/p 0.9 0.65
pc/p 0.84
5 CONCLUSIONS
Note: Italics indicate that no further cycles were applied as
the pile had accumulated excessive displacement. Centrifuge tests have been performed to investigate the
behaviour of single piles and pile groups jacked into
dense and medium dense sand, and subjected to
Figure 10 presents the accumulation of vertical dis-
monotonic and cyclic tension loading. The experi-
placements of the piles under cyclic loading for every
ments employed a new system which allowed the full
pile group tested.
sequence of installation and loading of a pile group to
Compared to the single pile, the accumulation of
be performed without the need to halt the centrifuge.
vertical displacements during cycling is much lower
The study has revealed the following trends for mono-
for the pile group at the same level of cycling. After
tonically jacked piles in medium dense and dense sand:
300 low amplitude cycles, the vertical displacement
of the pile groups in sample 1 and 2 are less than 1 mm 1 Monotonically jacked single piles have a higher
(S1-G3-C, S2-G3-C, S2-G4-C1). High amplitude capacity than buried piles and their ultimate shaft
cycling (S2-G4-C2) leads to greater accumulation of shear stress is proportional to the end resistance
vertical displacement and a 17% reduction in the mobilised during installation.
post-cycling p value compared to an un-cycled pile. 2 The capacity efficiency (capacity) for pile groups
This reduction is, however, small compared to that with spacing ratios (s/D) of 3 and 4 and comprising
observed for a single pile at the same level of cycling 5 piles is slightly less than one. This value arose

754

Copyright 2005 Taylor & Francis Group plc, London, UK


because the centre pile exhibited a capacity which REFERENCES
was between 15% and 30% less than that of a sin-
gle pile and the corner piles in the group. Briaud J.L., Tucker L.M. & Ng. E. 1989. Axially loaded 5
3 The pile groups required a pile cap displacement pile group and single pile in sand. Proc. Int. Conf. on Soil
of about 2.5 times that of a single pile to mobilise Mechanics and Foundation Engineering, Rio de Janeiro:
11211124.
their peak static tension capacity i.e. the stiffness Gaudin C. & Lehane B.M. 2004. Static and cyclic testing of
efficiency (stiffness) was 0.4. centrifuge single piles and pile groups in sand. GEO
4 At the same level of cycling (with respect to their report 04336, The University of Western Australia: 34 pp.
static capacities), single piles have a lower cyclic Gaudin C., Schnaid F. & Garnier J. 2005. Sand characterisation
resistance compared to group piles. by combined centrifuge and laboratory tests. International
5 Pile head displacements accumulate during ten- Journal of Physical Modelling in Geotechnics. (In Press).
sion cycling and cyclic failure occurs shortly after Jardine R.J. & Standing J.R. 2000. Pile load testing per-
the rate of accumulation of displacement per cycle formed for HSE cyclic loading study at Dunkirk, France.
starts to accelerate. Offshore Technology Report OTO 2000-007: 60 pp.
Lehane B.M., Gaudin C. & Schneider J.A. 2005. Shaft
6 Cyclic degradation occurs more quickly in looser capacity of rough model piles buried in dense sand.
sand. Gotechnique (In Press).
7 Under both monotonic and cyclic loading, pile Lunne T. & Christofferson H.P. 1983. Cone penetrometer
groups with a spacing ratio (s/D) of 3 and 4 exhibit interpretation for offshore sands. Offshore Technology
almost identical responses. Conference, Houston, Texas. OTC4464: 181192.
Stewart D.P., Boyle R.S. & Randolph M.F. 1998. Experience
with a new drum centrifuge. Proceedings of the
ACKNOWLEDGEMENTS International Conference Centrifuge 98, Tokyo, Vol. 1:
3540.
White D.J. & Lehane B.M. 2004. Friction fatigue on dis-
The authors gratefully acknowledge the support pro- placements piles in sand. Gotechnique 54(10): 645658.
vided by COFS and the Arup DTF fund towards the
research presented in this paper.

755

Copyright 2005 Taylor & Francis Group plc, London, UK


Correlations for shaft capacity of offshore piles in sand

J.A. Schneider & B.M. Lehane


University of Western Australia, Perth, Australia

ABSTRACT: The shaft capacity of displacement piles in sand is influenced significantly by many factors
including the (i) level of soil displacement imposed during installation, (ii) nature and method of pile installa-
tion (jacking/driving) (iii) dilation at the sand-pile interface, (iv) interface friction angle, (v) direction of load-
ing (compression/tension), and (vi) elapsed time between installation and load testing. Research on these
factors as well as features of the Imperial College and Fugro design methods are examined in the light of
recently published data from high quality instrumented load tests. This examination leads to the proposal of a
new formulation relating shaft friction with the CPT end resistance.

1 INTRODUCTION to a variety of formulations for rc with the following


general form (Lehane & Jardine 1994, Jardine et al.
While typical design compression loads for offshore 2005b, Fugro 2004, Randolph 2004, White 2005):
piles are now between 30 to 60 MN (Zuidberg &
Vergobi 1996), there are just a few published case his-
tories in sandy soils that report load tests on piles with (2)
a maximum applied load in excess of 15 MN. The
design of most offshore piles therefore relies on
extrapolation from the bounds of the current database where v0 is the vertical effective stress, pa is a refer-
of load tests and, as such, needs to be based on a ence stress  100 kPa and R is the outer radius of the
sound theoretical understanding of the mechanisms pile (with diameter, D, used for some methods).
controlling pile capacity. Equation (2) is used both by the Imperial College
The ultimate local friction that can develop on the (IC) design method (ICP-05, Jardine et al. 2005b)
shaft of a pile (f) in sand has been confirmed by and the Fugro design method (Fugro 2004). These
Lehane et al. (1993) to be a function of the radial methods are evaluated here in the light of recently
stress after installation and subsequent stress equal- published data from full scale load tests on instru-
ization (rc), the change in radial stress during the mented piles to assist in the development of a refined
loading stress path (rd), and the interface friction version of these approaches. This discussion consid-
angle (f): ers shaft capacity in non-calcareous sands. All sands
discussed in this paper are primarily siliceous, except
(1) for piles driven in micaceous sands at Jamuna Bridge,
Bangladesh.
Lehane & Jardine (1994) show that the value of rd
is relatively small for pile diameters greater than 2 ICP-05 AND FUGRO-04 METHODS FOR
about 300 mm in siliceous sands and hence the radial SHAFT CAPACITY
effective stress at peak friction (rf) may be con-
sidered equivalent to rc for offshore piles. The ICP-05 design method is based on results from
It has long been established that &f correlates well load tests on jacked closed-ended instrumented piles
with the CPT end resistance qc (e.g. Bustamante & (Lehane et al. 1993, Chow 1997) and was calibrated
Gianeselli 1982). This observation used in conjunc- for open-ended piles primarily using tests on driven
tion with an observation made using the Imperial piles at Dunkirk (Brucy et al. 1991). Neglecting
College (closed-ended) instrumented pile that rc increases in radial stress due to interface dilation,
varies with the distance behind the pile tip (h) has led equations (1) and (2) were combined to deduce

757

Copyright 2005 Taylor & Francis Group plc, London, UK


the following expression for local ultimate shaft fric- three of these sites (EURIPIDES, Jamuna Bridge, and
tion (f): Ras Tanajib II) was 763 mm and pile lengths (L) var-
ied from 30 to 78 m. The fourth site (Pigeon Creek,
Indiana) included single tests on 7 m long, open and
closed-ended 356 mm diameter piles. It is noteworthy
that the diameter to wall thickness (t) ratios varied
(3) from 20 for the piles with D  760 mm to 11 for the
356 mm diameter pile; these D/t ratios are consider-
where f/fc is the ratio of capacity to that in compres- ably less than that used for typical offshore piles.
sion (with f/fc  1 in compression and 0.72 for ten- These new test results are combined in this paper
sion loading on a pipe pile) and R* is the equivalent with data from IC instrumented pile tests conducted
pile radius (Req) defined as: in Dunkirk and Labenne to form a database of instru-
mented pile load tests, which is summarised in Table 1.
(4) Notable features of this database include:

where R is the outside radius of a pile and Ri (which (i) Most piles were load tested statically between 2
is zero for a closed-ended pile) is the internal radius and 10 days after installation, although a number
of a pipe pile. The lower limit placed on h/R* in of longer term tests (with equalization periods of
Equation (3) arose because the deepest instrument between 100 and 2000 days) are also included.
position on the IC instrumented (closed-ended) pile (ii) The relative densities (Dr) of the sands at the
was located at 8R above the pile tip. Equation (3) sites vary from medium dense to very dense. At
allows for greater shaft friction to be developed near EURIPIDES and Pigeon Creek, piles were
the tip of a pipe pile as the cut-off is at h  8R*. driven through medium dense sand into very
The Fugro-04 method for shaft friction is based dense sand, resulting in the development of a
primarily on the results of load tests on instrumented large proportion of shaft friction close to the pile
large diameter open-ended piles driven predomin- tip (over 80% in the lower third of the piles). In
antly in a coring manner in dense sands. Separate contrast, the relatively constant Dr of the mica-
equations are provided for tension and compression: ceous sands at Jamuna Bridge, and the reduction
in Dr with depth at Dunkirk, appear to have led to
a more uniform distribution of shaft friction
(5a) along the pile lengths.
(iii) Unlike typical offshore piles for which the incre-
mental filling ratio (IFR) is very close to unity,
the IFRs in the database are generally less than
one. As the degree of soil displacement at the pile
tip (of which IFR is a measure) has an important
influence on the radial stresses and hence shaft
(5b) friction that can develop on a displacement pile, it
may be inferred that shaft frictions for the pipe
piles in the database may be greater than those of
equivalent full scale offshore piles.
(iv) The comparatively low IFRs observed at Jamuna
(5c) bridge may be partly because of the relatively
where Equation 5a applies to piles in compression high compressibility of this (micaceous) sand.
with h/R*  4, Equation 5b applies to piles in com- (v) Average CPT friction ratios (Fr) were typically
pression with h/R* less than four, and Equation 5c relatively constant at a given site but varied
applies to piles in tension. The constant volume inter- between sites/sand deposits e.g. Fr  1.6% at
face friction angle is assumed as 29, which was typ- EURIPIDES and only 0.3% close to one pile at
ical of the interface friction angles for the case Ras Tanajib II. Although this variability in Fr is
histories used to calibrate the model. certainly partly attributable to instrument factors
(e.g. variable wear on the friction sleeve), it may
be indicative of natural differences between the
3 INSTRUMENTED LOAD TESTS response of the sands to penetrometer installation.
(vi) Some discrete zones with a high Fr value (greater
Data from four sand sites involving tests on instru- than 3) were observed at the EURIPIDES and
mented open-ended piles driven in sand have recently Pigeon Creek sites. Strain gauge data in these
become available. The pile diameter (D) employed at zones showed pile shaft friction values that were

758

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 1. Summary of characteristics of pile load tests discussed in this paper.

Age End CPT F3r


Site Pile (days) Condition L (m) D (m) D1r IFR2,3 (%) Symbol References

Dunkirk CL 160 O 11.3 0.32 0.750.15 0.650.35 0.90.45  Brucy et al. 1991; Chow 1997
Dunkirk CS 190 O 11.3 0.32 0.750.15 0.550.25 0.90.45  Brucy et al. 1991; Chow 1997
2000 
Dunkirk IC 1 C 7.4 0.10 1.0/0.75 0.90.45  Chow, 1997
EURIPIDES I-30.5 7 O 30.5 0.76 0.45/0.75 1.00.05 1.10.3  Kolk et al. (2005a); Fugro (2004)
EURIPIDES I-38 2 O 38.7 0.76 0.45/1.0 0.90.05 1.60.35  Kolk et al. (2005a); Fugro (2004)
EURIPIDES I-47 11 O 47 0.76 0.45/0.95 0.90.1 1.50.45 Kolk et al. (2005a); Fugro (2004)
EURIPIDES II-47 6 O 46.7 0.76 0.45/0.95 0.90.15 1.60.4
Kolk et al. (2005a); Fugro (2004)


540
Jamuna Bridge5
759

PS1 81 O 44 0.76 0.40.2 0.550.25 0.70.45 Fugro (1995)


Jamuna Bridge5 PS3 101 O 44 0.76 0.40.2 0.60.3 0.70.45  Fugro (1995)
Jamuna Bridge5 PS1D 4 O 80 0.76 0.450.2 0.70.45 Fugro (1995)
Labenne IC 1 C 6 0.10 0.65/0.45 Lehane et al. (1993)
Pigeon Creek OEP 41 O 7.0 0.36 0.55/1.0 0.800.05 0.550.05  Paik et al. 2003
Pigeon Creek CEP 41 C 6.9 0.36 0.55/1.0 0.550.05 Paik et al. 2003
Ras Tanajib II I-17 10 O 17 0.76 1.2/*4 0.850.2 0.30.1  Kolk et al. (2005b); Fugro (2004)
Ras Tanajib II I-25 100 O 25 0.76 */1.24 1.00.1 0.850.25  Kolk et al. (2005b); Fugro (2004)
1
Apparent relative density based on Lunne & Christoffersen (1983); no corrections for OCR or soil compressibility applied. Relatively uniform layers expressed as mean and
standard deviation. Distinct layering is indicated as two values separated by a slash.
2
IFR  incremental filling ratio, (plug length)/(tip depth). Do  pile outer diameter. Di  pile inner diameter.
3
Averaged over zone where approximately 80 percent of shaft friction capacity was mobilized.
4
A cemented sand, silt, and clay layer (*) existed from approximately 13 to 21 m. This layer was not included in analyses.
5
Sands at Jamuna Bridge contained significant amounts of mica.

Copyright 2005 Taylor & Francis Group plc, London, UK


not typical of sand and are not included in the distances greater than 10 diameters above the tip.
data presented here. Measurements obtained in These tendencies may have been influenced by
the zone of cemented sands, silts, and clays at plugging behaviour and instrument performance
Ras Tanajib II are also not presented here. during key experimental studies associated with
the development of each method.
4 COMPARISON OF DESIGN METHODS TO Shaft friction in tension is lower than that in com-
RECENT CASE HISTORIES pression, and the relative predictive ability of the
ICP-05 and Fugro-04 methods is comparable for
The measured variations of f /qc (within an average of tension and compression. The database average sug-
10 days of installation) with the normalized distance gests that tension shaft friction is 75 percent of that
above the pile tip (h/D) for the pile tests summarized in in compression, but this fraction is likely to vary
Table 1 are presented for closed-ended piles in com- with pile geometry and site conditions.
pression, pipe piles in compression and pipe piles in The expressions adopted for the ICP-05 and
tension on Figures 1a, 1b, and 1c respectively. The Fugro-04 methods evidently lead to quite different
ICP-05 and Fugro-04 predictions for these cases are
also included on these figures. The predictions adopt 100
representative database values of D/t equal to 20. In
addition, an average interface friction angle of 29 is
assumed for the ICP-05 method and the dilation com-
ponent of f (see equation 1) has been removed. The
exact location of the trend lines for the ICP and Fugro
methods will differ slightly for individual piles, but ICP-05
h/D

general trends illustrated in Figure 1 are representative. 10


These figures indicate that: Fugro-04
There is a fair degree of scatter in the observed
variations of f /qc with h/D for all three cases con-
sidered. Ageing, natural variability (e.g. in qc), Data are
variations in f and measurement. unreliable
inaccuracies contribute to observed scatter.
Nevertheless, the data support (to a first order 1
approximation) the form of equation (2). 0.000 0.005 0.010 0.015 0.020 0.025
The Furgo-04 method tends to over-predict shaft f/qc
friction near the tip of piles, while the ICP-05
method tends to over-predict shaft friction at Figure 1b. Comparison of design methods to field data for
open ended piles in compression (see Table 1 for legend).

100 100

ICP-05 ICP-05
h/D
h/D

10 10
Fugro-04 Fugro-04

Data may be
unreliable

1 1
0.000 0.005 0.010 0.015 0.020 0.025 0.000 0.005 0.010 0.015 0.020 0.025
f/qc f/qc

Figure 1a. Comparison of design methods to field data for Figure 1c. Comparison of design methods to field data for
closed ended piles in compression (see Table 1 for legend). open ended piles in tension (see Table 1 for legend).

760

Copyright 2005 Taylor & Francis Group plc, London, UK


variations of f /qc with h/D. Nevertheless, both offshore pile). These figures highlight important dif-
expressions lead to a similar mean and coefficient ferences between the respective formulations:
of variation for load test results in this database (i) Both methods predict lower rc values for pipe
(Fugro 2004), illustrating the presence of compen- piles compared to closed-ended piles, except
sating errors. close to the tip. This tendency is opposite to that
In summary, it is apparent that the ICP-05 and shown by experimental data (Chow 1997, and de
Fugro-04 formulations do not provide a very satisfac- Nicola & Randolph 1999), suggesting a cut off
tory fit to the database of observations. However, as as a function of h/R is more appropriate than h/R*.
illustrated by Jardine et al. (2005b) and Fugro (2004), (ii) The influence of the D/t ratio is more signific-
the methods are considerably more reliable than the ant for Fugro-04 and the radial stress ratio of
existing API recommendations (API 2000). Direct cor- rc-open/rc-closed is substantially higher for
relations between f and qc have been shown by Briaud ICP-05 than that adopted by Fugro-04.
& Tucker (1988), and others, to be superior to the earth (iii) Fugro-04 predictions for rc are well in excess of
pressure type approach adopted in API (2000). The those predicted by ICP-05 close to the tip, but are
ICP-05 and Fugro-04 methods are an advance on these typically 50% of the ICP-05 values at h/D  10
direct correlations as they incorporate the added for open ended piles.
dependence of f /qc on the relative depth of the pile tip (iv) Differences between rc predictions for the two
(h/D), the direction of loading (compression vs. ten- methods are less significant for closed-ended
sion) and the pile end condition (open and closed- piles.
ended). The ICP-05 approach also incorporates the The Fugro-04 and ICP-05 methods therefore differ
important effects of dilation on the shaft capacity of substantially in their respective assumptions regard-
small diameter piles and the variation of the interface ing the influence of the D/t ratio and the reduction
friction angle with the particle size and shaft roughness. rate of f with h/D. These two effects are combined in
an expression such as equation (2), which also does
5 FORMULATIONS FOR SHAFT FRICTION not distinguish between the effects on rc of stress
relief (or unloading) immediately above the tip and
The ICP-05 and Fugro-04 expressions for rc in com- friction fatigue due to cycling further back from the
pression are compared on Figures 2(a) and 2(b) for tip (White & Lehane 2004). The separate effects of
closed-ended piles and pipe piles with D/t ratios of 20 the pile end condition, stress relief near the tip and
(typical of database pile) and 50 (typical of full scale friction fatigue may be expressed in an alternative
version of Equation (2) as:
Fugro-04, D/t=20
Closed-ended (6a)
Fugro-04, D/t=50
D/t=20
ICM-96, D/t=50
D/t=50
ICM-96 D/t=20
(6b)
20 20
18 18 where [a(v0 /pa)b ]  a1  a2, c  c1  c2 and h  0.
16 16 The first term in equation (6a) captures the stress
14 14 relief in the vicinity of the pile tip (noting that rc
would have a similar magnitude to qc at h  0) and is
12 12
proportional to the sleeve friction measured in a CPT.
h/D

h/D

10 10 The [a2(h/R)
c2] term reflects the degradation in lat-
8 8 eral stress due to friction fatigue or cycling, which
6 6 varies with the number and nature of installation
4 4 cycles1. Assuming that c1 is independent of the pile and
2 2 sand type, it follows that Fugro-04 incorporates a much
0 0 higher value of c2 compared to ICP-05 (compare
0 0.5 1 0 1 2 3 equation 3 and 5). This may well be because of the
harder driving conditions relative to the piling ham-
'rc (open) 'rc (Fugro-04)
(a) (b) mer employed (and hence large number of installation
'rc (closed) 'rc (ICM-96)

Figure 2. Comparison of ratio of radial stress for ICP-05 1


This is an approximation given that the friction fatigue
and Fugro-04 methods as a function of (a) pile geometry and term is not likely to vary systematically with h/D
c2 for all
(b) design method. cases.

761

Copyright 2005 Taylor & Francis Group plc, London, UK


20 100
18
16
14
12

h/D
h/D

10 10
8
6
4
2
0 1
0 0.25 0.5 0.75 1 1.25 0.000 0.005 0.010 0.015 0.020 0.025
IFR f/qc

Figure 3. IFR at the end of driving for Dunkirk and Figure 4. Influence of time on shaft friction at Dunkirk and
EURIPIDES pile installation (see Table 1 for legend). EURIPIDES, lines indicate increase in capacity with time at
a particular site for selected points (see Table 1 for legend).

cycles imposed) for many of the pipe piles used to


calibrate Fugro-04. The reduced friction that can be where
developed on a pipe pile compared to a closed-ended
pile is accounted for by the (R*/R)c multiplier in (7b)
equation 6b, which has a value of (R*/R)0.38 and
(R*/R)0.9 for the ICP-05 and Fugro-04 methods In deducing this formulation, it is noted that:
respectively. The higher multipliers adopted in ICP-05 (i) Equation (7b) is a modification to Equation 4 to
and Fugro-04 appears to have arisen because of an account for the effects of partial plugging.
increased level of displacement (partial plugging) (ii) White et al. (2005) propose a d value of 0.7  0.1
during installation of pipe piles. This is apparent on based on considerations of cylindrical expansion.
Figure 3 which plots IFR values measured over the (iii) The value of [(D*/D)d qc a1 (h/D)
c1] should be
last 20 diameters of penetration (where the majority limited to a minimum value of Kov with
of friction was developed; see Figure 1). The IFRs are [a1  (h/D)
c1] being related to length of the
evidently less than unity and higher than would be equalization period and the CPT Fr.
typical of a large diameter offshore pile. (iv) The (h/D*)
c1 term is used in preference to
Equations (2), (3) and (6) do not include a correc- (h/D)
c1 in equation 6b to account for the possi-
tion factor for the strong positive effects of ageing for bility of a more rapid decrease in radial stress
shaft friction in sand (e.g. as recently discussed by behind the tip of a coring or partially plugging
Axelsson (1998) and Jardine et al. (2005a)). The effect pile as compared to a closed ended pile.
of time on the database f values is highlighted on (v) Evidence from multi friction sleeve cones
Figure 4, which indicates the increase in f /qc meas- (DeJong 2001) suggests that the value of c1 is
ured between 200 days and 2000 days at Dunkirk likely insignificant and between about 0.05 and
and between 6 days and 540 days at EURIPIDES. It is 0.15. In general, the value of c2 should increase
apparent that the relative increases with time differ with the number of hammer blows, although c2
between sites and do not show a systematic depend- will also be influenced by characteristics of instal-
ence on h/D. f evidently increases by between 50 and lation cycles and normal stiffness conditions,
100% per log cycle of time (over the equalisation peri- among other factors.
ods considered). It may be surmised that such a rate of (vi) Although a minimum value of rc exists (which
capacity gain cannot continue indefinitely. will be related primarily to the sands friction
On the basis of the foregoing, the following angle), imposition of this limit would have virtu-
equation is proposed as a suitable formulation for ally no effect on the shaft capacity evaluated
evaluating the equalized radial effective stress on a using equation (7) for typical offshore pile sizes.
displacement pile:
Taking the above considerations into account and the
(7a) need to reduce the number of empirical parameters

762

Copyright 2005 Taylor & Francis Group plc, London, UK


because of the shortage of experimental data, the fol- Brucy, F., Meunier, J., & Nauroy, J.-F. 1991. Behaviour of
lowing simplified form of Equation (7) was com- pile plug in sandy soils during and after driving. OTC
pared in Lehane et al. (2005a) to a database of open 6514, Proc., 23rd Annual OTC, Houston:145154.
and closed ended piles: Chow, F.C. 1997. Investigations into the behaviour of dis-
placement piles for offshore structures, PhD Thesis,
University of London (Imperial College), U.K.:766 pp.
(8) DeJong, J.T. 2001. Investigation of particulate-continuum
interface mechanisms and their assessment through a
A (f/fc) ratio of 1.0 and 0.75 was assumed for multi-friction sleeve penetrometer attachment. PhD Thesis,
compression and tension loading respectively, and Georgia Institute of Technology, Atlanta, GA, USA, 361 pp.
rd was evaluated using the recommendations of de Nicola, A. & Randolph, M.F. 1999. Centrifuge modeling
ICP-05. Comparison to a database of pile load tests of pipe piles in sand under axial loads. Gotechnique,
showed that Equation (8) led to mean ratios of calcu- 49(3):295318.
lated to measured capacities (Qc/Qm) of 0.91 and 1.02 Fugro Engineers B.V. (Fugro) 1995. Interpretation report on
for open and closed ended piles, respectively. The reduced scale pile load tests. Jamuna Bridge,
Bangladesh, Report No. K-2380/207.
corresponding coefficients of variation for Qc/Qm Fugro Engineers B.V. (Fugro) 2004. Axial pile design
were 0.23 and 0.27, which are slightly lower than method for offshore driven piles in sand. Fugro Report
Fugro-04 and ICP-05. However, while similar trends No. P1003, Issue 3 to API, 5 August 2004:122 pp.
are shown for Equation 8 and the ICP-05 method Jardine, R.J., Standing, J.R., & Chow, F.C. 2005a. Field
when compared to a pile database, it is noteworthy research into the effects of time on the shaft capacity of
that Equation (8) predicts significantly lower friction piles driven in sand, Proc., ISFOG, Perth:5 pp.
capacities when extrapolating to larger diameter piles Jardine, R.J., Chow, F.C., Overy, R., & Standing, J. 2005b.
(which typically install in a coring manner). ICP design methods for driven piles in sands and clays,
Thomas Telford, London.
Kolk, H.J., Vergobbi, P. & Baaijens, A. 2005a. Results from
6 CONCLUSION axial load tests on pipe piles in very dense sands: the
EURIPIDES JIP, Proc., ISFOG, Perth:6 pp.
Review of ICP-05 and Fugro-04 design methods for Kolk, H.J., Shaffei, K.A. & Baaijens, A. 2005b. Axial load
tests on pipe piles in very dense sands at Ras Tanajib.
piles in sand in the light of recently published instru- Proc., ISFOG, Perth:6 pp.
mented pile test data has revealed compensating Lehane, B.M., Jardine, R.J., Bond, A.J., & Frank, R. 1993.
errors in their formulations. This has arisen because Mechanisms of shaft friction in sand from instrumented
of the large number of variables affecting the shaft pile tests, J. of Geotech. Eng., 119(1):1935.
capacity of displacement piles in sand and the great Lehane, B.M. & Jardine, R.J. 1994. Shaft capacity of driven
shortage of reliable case history data (particularly for piles in sand: a new design approach. Proceedings,
typical large scale offshore pipe piles). It is suggested Behavior of Offshore Structures (BOSS):2336.
that the methods lumped approaches to deal with Lehane, B.M., Schneider, J.A., & Xu, X. 2005. Evaluation of
the separate effects of friction fatigue and the pile end design methods for displacement piles in sand. UWA
report GEO: 05341.1.
condition as well as the treatment of partial plugging Lehane, B.M., Schneider, J.A., & Xu, X. 2005a. The UWA-
are a limitation. A new formulation to address these 05 method for prediction of axial capacity of driven piles
deficiencies is presented, which compares favourably in sand from CPT data. Proc., ISFOG, Perth:6 pp.
to a pile load test database. Lunne, T., & Christoffersen, H.P. 1983. Interpretation of
cone penetrometer data for offshore sands. OTC4464,
Proceedings, 15th Annual OTC, Houston:181192.
REFERENCES Paik, K., Salgado, R., Lee, J., & Kim, B. 2003. Behavior of
open- and closed-ended piles driven into sand. J. of
American Petroleum Institute (API) 2000. RP2A-WSP: Geotech. and Geoenv. Eng., 129(4):296306.
Recommended practice of planning, designing, and con- Randolph, M.F. 2004. Comments on Fugro Report to API on
structing fixed offshore platforms working stress Pile Capacity in Sand. submitted to Fugro Engineers
design, 21st ed, Washington DC:226 pp. B.V., Leidschendam, The Netherlands:2 pp.
Axelsson, G. 1998. Long-term increase in shaft capacity of White, D.J. 2005. A general framework for shaft resistance
driven piles in sand. Proc., 4th Int. Con. on Case on displacement piles in sand. Proc., ISFOG, Perth:6 pp.
Histories in Geotechnical Eng., St. Louis, Missouri, White, D.J., & Lehane, B.M. 2004. Friction fatigue on dis-
paper 1.25. placement piles in sand. Gotechnique. 54(10):645658.
Bustamante, M., & Gianeselli, L. 1982. Pile bearing cap- White, D.J., Schneider, J.A., & Lehane, B.M. 2005. The
acity by means of static penetrometer CPT. Proc., influence of effective area ratio on shaft friction of dis-
2nd European Symposium on Penetration Testing, placement piles in sand, in Proc., ISFOG, Perth.
Amsterdam:493499. Zuidberg, H.M., & Vergobbi, P. 1996. EURIPIDES, Load
Briaud, J.L., & Tucker, L.M. 1988. Measured and predicted tests on large driven piles in dense silica sands. OTC
axial load response of 98 piles. Journal of Geotechnical 7977, Proceedings, 28th OTC, Houston:193206.
Engineering, ASCE, 114(9):9841001.

763

Copyright 2005 Taylor & Francis Group plc, London, UK


Axial load tests on pipe piles in very dense sands at Ras Tanajib

H.J. Kolk & A.E. Baaijens


Fugro Engineers B.V., The Netherlands

K.A. Shafei & O.A. Dakhil


Saudi Arabian Oil Company, Saudi Arabia

ABSTRACT: An extensive load testing programme was conducted on a 0.76 m outer diameter pipe pile in Ras
Tanajib, Saudi Arabia. The objective was to obtain reliable soil and axial resistance data for offshore type piles
in very dense sands typical for the Arabian Gulf, offshore Saudi Arabia. The paper presents details of the pile load
test equipment, instrumentation and test programme. Soil data and test results are presented. The latter include
distributions of soil resistance along the test pile and load-displacement data.

1 INTRODUCTION sampling and SPTs. Soil conditions at these three loca-


tions are similar and are summarised in Table 1.
An extensive load testing programme was conducted CPT cone resistances in the Unit I sands reached val-
in 1998 on a 0.76 m outer diameter pipe pile in Ras ues up to 70 MPa. Cone resistances up to 120 MPa were
Tanajib, Saudi Arabia. A highly instrumented pile measured in Unit III sands. These high resistances
was driven to 17 m depth and static compression and required a combination of drilling and CPT testing
tension tests were performed. After pulling the pile, a through a casing. The tests were performed with a
casing was driven to 17 m bgl, the soil plug removed
and the instrumented test pile was driven through this
casing to 25 m depth. Subsequently, another series of Table 1. Soil stratigraphy.
static compression and tension tests was performed.
Depth below
The soils consist of dense to very dense silica sands, ground level [m]
except between 13 m and 21 m depth where clay and
claystone layers occur. The tests were performed to Unit From To Description
complement a pile test programme performed in 1983
(Al-Shafei et al. 1994). These earlier tests were per- Ia 0.0 3.5 FINE to MEDIUM
formed about 13 m south of this pile test location. The SAND, silty, dense
objective of the tests was to obtain reliable data on Ib 3.5 9.0 FINE to MEDIUM
offshore type piles in very dense sands typical for the SAND, silty, very dense
Arabian Gulf offshore Saudi Arabia. The project was Ic 9.0 10.8 FINE to MEDIUM
SAND, silty, dense
performed by Fugro Engineers B.V. (The Netherlands), IIa 10.8 13.5/14.5* FINE SAND, silty to
assisted by FugroSuhaimi Ltd. (Saudi Arabia) on SILT, sandy, weakly
behalf of Saudi Arabian Oil Company (Saudi Aramco). cemented
IIb 13.5/14.5* 17.5/18.1* CLAY/SILT, very stiff
to CLAYSTONE, weak
2 SOIL CONDITIONS IIc 17.5/18.1* 19.0/20.0 FINE SAND, silty, very
dense/weakly cemented
The water table at the site was at about 7 m bgl. Soil IId 19.0/20.0 21.1 CLAY/SILT, sandy,
conditions at the test location were defined by 2 Cone very stiff to interlayered
weak CLAYSTONE,
Penetration Tests (CPT) and 3 boreholes (BH) per-
SILTSTONE,
formed in 1996 for the Ras Tanajib II programme. SANDSTONE
The first two boreholes were made by drilling and III 21.1 30.0 FINE SAND, silty,
sampling the vertical section where the two CPTs very dense
were made previously. Thus, disturbed samples were
obtained. The third borehole included undisturbed * Depths for CPT-96/1 and CPT 96/2, respectively.

765

Copyright 2005 Taylor & Francis Group plc, London, UK


Cone resistance [MPa]
0 20 40 60 80 100 120 140
0
Dr = 100%
5

10
Depth [m]

15 Cemented sands,
Test Depth 1: 17 m
silts and clays
20

Dr = 40% Test Depth 2: 25 m


25

Dr = 60%
30 Dr = 80%
CPT 96/1 CPT 96/2 average
35

Figure 1. Cone resistance data and relative density at test location.

10 cm2 piezo-cone, except for the deepest part of the


CPTs with a cone resistance in excess of 80 MPa, where Force Ring
a 5 cm2 cone was used (Fig. 1). Direct shear tests
yielded an average effective (soil) friction angle of Ground Level

about 33 degrees in the Unit I sands and approximately


Add on I
35 degrees in the Unit III sands. All sands were silica 6.6 m Add on II
(carbonate content 10%). Water Level 8.15 m
-7.0m
The soil descriptions and test results from unit II
samples vary significantly. However, the cone resist-
ance profile is fairly consistent throughout the profile
(Fig. 1). Plasticity indices in the cohesive layers varied
between 15 and 45%, with the lower values associated Add on I
Instrumentation Section 6.6 m
with claystone and the higher values with clay. The only 12.5 m
laboratory strength tests on these cohesive materials
consisted of two Unconfined Compression Tests, which
gave UCS values of 1.5 MPa (at 15 m bgl) and 1.4 MPa
(at 20 m bgl). 17 m Bottom of Casing
17 m
The CPT equipment used in the 1983 investigation
at Ras Tanajib I was not capable of penetrating very Instrumentation Section
12.5 m
dense sands. Hence, a discontinuous CPT profile with
significantly lower cone resistance values was achieved
at that time. More importantly, the 1983 soil investiga-
tion programme, performed at a distance of 12 m from
Ras Tanajib II, did not reveal the clay and claystone 25 m
layers found in 1996. This difference is primarily
attributed to the lack of continuous soil profile data in Figure 2. Pile load test setup.
1983 as compared to the continuous CPTs in 1996.
in this section was to measure the soil-pile interaction
3 TEST PILE AND INSTRUMENTATION in the soils of interest, i.e. from 4.5 m to 17 m bgl in
the first test and from 17 m to 25 m bgl in the second
The test pile had an outer diameter of 0.76 m. The test (Fig. 2). The add-on section above this instru-
lower 12.5 m of the pile was highly instrumented and mented section had no instrumentation (except near
had a wall thickness of 38.5 mm. The instrumentation the pile head) and had a wall thickness of 42 mm.

766

Copyright 2005 Taylor & Francis Group plc, London, UK


The increase in tangential strain gauges was made
since the EURIPIDES programme had proven the need
for both axial and tangential strain gauges for an accur-
ate assessment of axial load.
The toe load cells were more robust since those used
in the EURIPIDES programme were damaged after a
significant amount of heavy pile driving.
The most significant change was that the instru-
mentation and cables were mounted in steel protection
channels. The EURIPIDES channels consisted essen-
tially of epoxy material on which a thin metal sheet
was glued for protecting the epoxy against abrasion.
Unfortunately, the channels near the toe of the EURIPI-
DES test pile were torn off due to high stresses at the
pile toe during pile driving in very dense sand. Thus,
instrumentation near the pile toe was lost with increased
driving resistance. Each channel of the Ras Tanajib II
test pile consisted of two welded side strips and a
rounded steel cover plate, which was welded to these
Figure 3. Cross-section of instrumented pile section.
strips after the moulding epoxy resin (potting) was
added (Fig. 3). These channels were significantly
stronger and proved to survive the entire driving and
The instrumented section was provided with four
testing programme. Unfortunately, the channels also
instrumentation cable channels 90 degrees apart from
affect the strain gauge measurements. In particular
each other on the outside of the pile (Fig. 3).
the external pressures on the protection channels have
Instrumentation was mounted at various levels along
a significant influence. The effects of the steel chan-
the instrumented section with spacing of 0.3 pile diam-
nels was accounted for in the data reduction process
eters near the pile toe increasing to 5.3 pile diameters
through a combination of pile calibration testing and
at the top of the instrumented section. The amount of
ABAQUS Finite Element analyses of a model of the
instrumentation near the pile toe was enhanced in order
test pile subject to various combinations of soil stresses.
to determine wall end bearing and pile wall inner and
The instrumented pile section was subjected to
outer friction.
repeat axial loading after the instruments were mounted
The key instrumentation included:
and hooked up to a data acquisition system. This served
9 instrumentation levels with 4 sets of axial and two purposes:
tangential strain gauges
removal of residual stresses in the test pile;
5 levels with 2 total pressure cells facing the inside
calibration of both pile and instruments.
or outside of the pile
3 levels with 2 pore pressure cells and 1 level with This approach was also successfully used for the
1 pore pressure cell facing the inside or outside of EURIPIDES pile testing programme.
the pile
2 toe load cells
3 levels with 2 thermocouples 4 LOADING SYSTEM
1 level with 2 accelerometers.
The pile load testing rig developed for the EURIPI-
The majority of the instrumentation was of the same
DES programme was used for the Ras Tanajib tests
proven design as used in the 1995 EURIPIDES pile
since it had proven to be a reliable and efficient sys-
load test programme. Details on design and perform-
tem (Zuidberg & Vergobbi 1996). This special test rig
ance of these (and other) instruments are given by
was designed and fabricated in order to facilitate chan-
Zuidberg & Vergobbi (1996).
ging over from compression to tension testing (and vice
The main differences with the EURIPIDES instru-
versa) as well as changing over from the testing mode
mented section are:
to the pile installation/removal mode.
(a) Shorter instrumented section (12.5 m versus 27 m) This space frame concept with pin connections
(b) Both axial and tangential strain gauges at all lev- served two purposes. Firstly, it simplified erection and
els (versus limited tangential strain gauges) dismantling of the rig. Secondly, it avoided out of plane
(c) Improved toe load cells bending and subsequent deformation. This made the
(d) Steel protected instrumentation channels, all on rig an order of magnitude stiffer than the conventional
outside (versus 2 inside/2 outside). beam loading system.

767

Copyright 2005 Taylor & Francis Group plc, London, UK


The reaction system of this test rig consisted of six Table 2. Overview of tests.
size 0.76 m outer diameter, 15 m penetration pipe piles,
equally spaced at six metres distance from the test pile. Penetration Set-up period
Load was applied from the rig to a force ring on the depth [m] Date [days]
add-on pile by three double acting hydraulic jacks. Each
17 May 1721, 1998 8
jack had an effective stroke of 1 metre and a capacity 25 September 1213, 1998 102
of 10 MN in compression and 5 MN in tension.
Details on design and performance of this test rig
are provided by Zuidberg and Vergobbi (1996). in outer wall roughness may be due to higher radial
stresses on the outer pile wall.
Check of the instrumentation read-outs after retrieval
5 PILE INSTALLATION AND REMOVAL confirmed that negligible zero shifts occurred. This
confirmed the success of the pre-installation pile load-
The pile was driven to 17 m bgl for the first test, using ing and calibration effort.
an IHC S-90 hydraulic hammer with an energy setting The design of the steel protection channels and of the
of 85 kJ. Strain and acceleration in the pile head was instruments proved highly successful since virtually
measured during driving together with measurement all instrumentation was in good order after pile driving.
of soil level inside the pile. The latter indicated continu-
ous plug development during driving (coring behav-
iour). The plug ratio was 99% at the end of driving. 6 PILE LOAD TESTS
After completion of the pile load tests at 17 m, the
pile was completely pulled out of the soil using the load Pile load tests at each penetration consisted of alter-
test rig. The soil plug was fully recovered and removed nating series of compression (C) and tension (T) tests.
by jetting. Subsequently, a 42 inch OD casing was A constant rate of pile head displacement of 1 mm/min
driven to 17 m penetration and followed by soil plug was applied in these tests. The compression tests were
removal inside the casing to 16.8 m depth. Thereafter generally performed to a pile head displacement in
the test pile (provided with a second add-on) was driven the order of 200 mm (0.25 D) and the tension tests
to 25 m penetration. The average delivered hammer to a pile had displacement of about 80 mm (0.1 D).
energy was in the order of 80 to 85 kJ. The final plug An overview of the tests is given in Table 2.
ratio was 81%, which was the result of continuous plug- The soil investigation and the soil plug data revealed
ging during initial driving to 19 m penetration accord- that the upper 14 m of the test pile was embedded in
ing to the plug measurements during driving. After dense to very dense sand and the lower part was in
reaching final penetration, the annulus between test layers of predominantly clay, claystone and silt in the
pile and casing below 7 m bgl was subsequently filled 17 m test. The pile toe in this test was in clay. The test
with loose sand in order to reduce the risk for pile started 8 days after driving. The excess pore pressures
buckling during pile load testing (Fig. 2). in the clay layers were directly after driving in the
When the test pile was brought to the soil surface order of 1 MPa outside the pile and 4 MPa on the
from 17 m penetration it remained fully plugged. inside. The majority of these pressures dissipated dur-
Samples from this plug were taken prior to and during ing the first 2 days after driving. They were in the
plug removal by jetting. Visual classification of these order of 0.25 MPa at the start of the compression test,
data confirmed the general soil profile given in Table which was about 2 to 3 times the hydrostatic pressure.
1, except that Unit IIa seemed more cemented than The compression test was followed by a tension test
suggested by the CPTs and previous borings. and another compression and tension test (C-T-C-T).
Subsequently, the pile instrumentation was checked The lower 4 m of the test pile (from 21 to 25 m bgl)
and the roughness of the pile wall (both inside and out- in the 25 m test was embedded in very dense sand.
side) was measured. The upper 4 m (from 17 to 21 m bgl) was embedded in
Comparison of the roughness measurements before layers of predominantly clay, claystone and sand. The
installation and after pile extraction from 17 m revealed test started with a compression test 102 days after pile
that: installation. The total test series consisted of 3 com-
pression tests and two tension tests (C-T-C-T-C).
the outer wall roughness had reduced from about
29 m to 13 m;
the inner wall roughness had reduced from about 7 TEST RESULTS
28 m to 15 m.
7.1 Resistance data 17 m test
The reduction in roughness is attributed to abra-
sion of the pile wall by sand during pile installation Figure 4 shows the pile head load-displacement curve
and removal. The apparently slightly larger reduction for the 17 m test. The maximum load in the first

768

Copyright 2005 Taylor & Francis Group plc, London, UK


25

20
Compression Test 1 Compression Test 2
15
Pile Head Load [MN]

10 After Tension 1: Push Pile


Back
5

0
0 50 100 150 200 250 300 350 400
-5 Tension Test 1
Tension Test 2
-10

-15
Pile Head Displacement [mm]

Figure 4. Force at pile head versus pile head displacement for 17 m test.

Load [MN]
-20 -15 -10 -5 0 5 10 15 20 25 30
0

2
11.1% OD
4 10% OD

6 25% OD
Depth [m]

8 15% OD
10

12
Cemented sands,
14 silts and clays
16

18

20

Figure 5. Distribution of axial load in pile for selected pile head displacement at 17 m.

compression and the first tension tests at 17 m is 20.5 outer friction at levels in excess of 3 m (4 D) above
and 12.9 MN respectively. Figure 5 shows the axial load the pile toe since it is unlikely that there is significant
distributions along the pile for a selected number of friction between the soil plug and the inner pile wall.
pile head displacements. These curves illustrate that However, inner wall friction nearer to the pile toe is
relatively limited resistance is provided by the sand significant, particularly during compression loading.
layers to 11 m bgl. Hence, the average unit friction of the inner and the
Figure 6 shows the distributions of total unit friction outer wall of the pile near the pile toe may be in
from the first compression and the first tension tests between the presented total unit friction values and
at 17 m and 25 m bgl. The distributions for the 17 m half these values.
test apply for a pile head displacement of 205 mm and It is of interest to note that the unit friction values in
86 mm respectively. The total unit skin friction was compression in the sands to 11 m bgl are in the order
determined by taking the difference in axial load at of 200 kPa which is significantly higher than values
two subsequent strain gauge levels and dividing this recommended in API (1993). It is also of interest to
by the outer pile wall area. The resulting unit friction note that the unit friction values in tension are about
value is plotted between these two strain gauge levels. half these values, which again differs significantly from
This value is probably in the order of the average API recommendations.

769

Copyright 2005 Taylor & Francis Group plc, London, UK


Average Total Unit Friction per Pile Segment [kPa]
-4000 -3000 -2000 -1000 0 1000 2000 3000 4000
0
-200 -100 0 100 200
0

5
Tension 17 m
4

10
8

Tension 17 m Compression 17m


Depth [m]

12
15

Compression 17 m

20

Tension 25 m
Compression 25 m
25

30

Figure 6. Distribution of total unit skin friction at 17 and 25 m for max PHL.

30

25 Compression Test 2

20 Compression Test 1

15 After Tension 2: Push Pile


Pile Head Load [MN]

Back
10 After Tension 1: Push Pile Back

0
0 50 100 150 200 250 300 350
-5
Tension Test 2
-10 Tension Test 1

-15
Pile Head Displacement [mm]

Figure 7. Force at pile head versus pile head displacement for 25 m test.

7.2 Resistance data 25 m test


compression tests in very dense sands where a signifi-
Figure 7 shows the pile head load-displacement curve cantly larger rate of increase was observed.
for the 25 m test. The maximum load in the first com- Figure 8 shows the axial load distributions along the
pression and the first tension tests is 25.1 and 10.4 MN pile for a selected number of pile head displacements.
respectively. It is of interest to note that the compression These curves confirm that the sand inside the casing
capacity does not increase significantly with increasing between 7 and 17 m bgl offers negligible frictional
pile head displacement. This is unlike EURIPIDES resistance. These curves also illustrate that the majority

770

Copyright 2005 Taylor & Francis Group plc, London, UK


Axial Load [MN]
-15 -10 -5 0 5 10 15 20 25 30
10

12
10% OD
10.6% OD 14

16
15% OD
18
Depth [m]

Cemented sands,
silts and clays 25% OD
20

22

24

26

28

30

Figure 8. Distribution of axial load in pile for selected pile head displacement at 25 m.

of the resistance is obtained from the very dense sand Four instrumentation channels are essential in view of
unit below 21 m bgl. redundancy and/or correction for bending moments.
Figure 6 shows the distributions of total unit friction Stress conditions at a pile toe are highly complex
from the first compression and the first tension test. and cannot be assessed from strain gauge data only.
The distributions for the 25 m test apply for a pile head Internal and external radial pressure data facilitate
displacement of 203 mm and 81 mm respectively. For interpretation of test results although interpret-
reasons discussed with the 17 m test above, actual unit ation of all data remains complex.
wall resistance values below 22 m bgl may be down to The unit friction values are significantly higher than
half the presented total friction values. those recommended in API. The distribution as well
In view of the above, total unit friction values plotted as the comparison between tension and compres-
at 21.2 and 22.7 m bgl are indicative of outer friction sion also differs significantly.
in sand at this level. These values are in the order of Derivation of unit end bearing values from the pile
900 to 1300 kPa for compression and 0.3 to 0.4 times load test measurements is not straight forward in
these values for tension. These values differ signifi- view of the complex stress conditions at the pile tip.
cantly from values used in conventional offshore design
practice (API 1993).
ACKNOWLEDGEMENTS

8 CONCLUSIONS The management of Saudi Aramco are thanked for


funding this research project, and permitting publica-
The following conclusions can be drawn from the Ras tion of this paper.
Tanajib II pile load test programme:
The soil investigation for Ras Tanajib II revealed a REFERENCES
significant amount of clay, claystone and silt layers,
unlike the earlier investigation at Ras Tanajib I. Al-Shafei, K.A., Cox, W.R. and Helfrich, S.C. 1994. Pile
Pre-stressing and calibration of instrumented pile load tests in dense sand: Analysis of static test results,
sections prior to load testing are essential for reli- OTC Paper 7381, Proceedings 26th Offshore Technology
able data reduction. Conference, Houston, Texas, pp. 83102.
The surface of the pile became smoother, probably API 1993. Recommended Practice for Planning, Designing
as a result of abrasion by sand during pile driving. and Construction Fixed Offshore Platforms, API
Recommended Practice 2A (RP2A), 11th edition, 1979,
A wealth of high quality soil and pile instrument- 20th edition, 1993.
ation data was obtained. Zuidberg, H.M. and Vergobbi, P. 1996. EURIPIDES, Load
Axial strain gauges need to be complemented by Tests on Large Driven Piles in Dense Silica Sands, 28th
tangential strain gauges for an accurate assessment Offshore Technology Conference, Houston, Texas, paper
of axial load in a pile. No 7977, Vol. 1, pp. 193206.

771

Copyright 2005 Taylor & Francis Group plc, London, UK


Piles

Copyright 2005 Taylor & Francis Group plc, London, UK


Bearing capacity of driven piles in clay, the NGI approach

K. Karlsrud, C.J.F. Clausen & P.M. Aas


Norwegian Geotechnical Institute, Oslo, Norway

ABSTRACT: A comparison between calculated and measured capacities of driven piles in clay shows that the
API RP2A calculation method from 1993 over-predicts the capacity of piles in normally consolidated clay of
low plasticity. The authors propose a revised calculation method called NGI-99. This method includes correc-
tions related to the undrained shear strength, time between pile driving and testing, and pile tip condition dur-
ing driving. Details of the method are presented together with comparisons between calculated and measured
pile capacities. The proposed method gives a good agreement between measured and calculated capacities for
most of the well-documented large scale pile tests included in the study.

1 INTRODUCTION back-figured. Other tests were carried out as pile seg-


ment tests from the bottom of a drilled and cased hole.
Various semi-empirical methods have been proposed Key soil parameters for the interpretation are total
over the last 30 years to calculate the axial capacity of unit weight, porewater pressure, clay undrained shear
driven piles in clay. Fourteen such methods are referred strength su and clay plasticity Ip. The reference
to in Clausen & Aas (2001). The design of offshore undrained shear strength used by NGI for Database 1
piles has traditionally been carried out as recom- are from direct simple shear (DSS) tests consolidated
mended by the American Petroleum Institute, API to the in situ vertical stress. This strength is considered
RP2A, API (1993). The present API recommenda- to represent the best estimate of the actual in situ
tions are mainly based upon the results presented by strength on a vertical shear plane. For many of the pile
Dennis & Olson (1983). tests where NGI has been involved, DSS testing had
During the last 20 years, the Norwegian Geotech- been carried out. For other sites, the DSS strengths
nical Institute (NGI) has carried out, or participated were estimated based upon other available test results
in a number of field tests on driven piles in clays. The (see Section 4).
results from these tests, together with results published Another important matter is the length of time the
by others, have allowed a comparison between the pile pile is allowed to rest between driving and test-loading.
capacities calculated by the API (1993) method and The skin friction values quoted in Table 1 correspond
those actually measured. to the estimated value after dissipation of 90% of the
The authors present the results of such compari- excess porewater pressures due to pile driving.
sons, and propose a new calculation method, called Section 5 presents the time correction used with the
NGI-99, that gives improved agreement between pre- NGI-99 method for cases where a correction of the
dicted and observed capacities. An accompanying measured capacity was required.
paper by Clausen et al (2005) deals with the capacity The results in Table 1 may be used to calculate -,
of driven piles in sand. - and -values, defined by:

(1)
2 DATABASE 1
(2)
Two different data bases with results from pile load
tests in clay were established by NGI. They are referred
to as Database 1 and Database 2. (3)
Database 1 (Table 1) includes results from a number
of instrumented pile tests where the measured axial where skin is the measured or calculated pile skin
forces in the piles at different depths allow the local friction, sDSS
u is the DSS undrained shear strength, and
skin friction between two instrumented levels to be vo is the in situ vertical effective stress.

775

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 1. Database no. 1, measured local pile skin friction.

1 2 3 4 5 6 7 8 9 10 11 12

Pile Pile Open1) Depth vo sDSS


u s meas2)
Site name no. Closed (m) Ip (%) (kPa) (kPa) (kPa)   8/7   9/8  9/7

Haga B-piles 1 2 2.7 15 51 40 18 0.784 0.450 0.353


2 2 3.9 15 72 42 40 0.583 0.952 0.556
3 2 4.5 25 82 48 49 0.585 1.021 0.598
Onsy A1 4 2 10 38 62 17 17.7 0.274 1.041 0.285
A2 5 2 17.5 47 100 24 21.8 0.240 0.908 0.218
A3 6 2 25 41 138 35 29.2 0.254 0.834 0.212
A4 7 2 32.5 35 176 45 34.8 0.256 0.773 0.198
CI 8 2 7.5 36 51 15 10.5 0.294 0.700 0.206
CI 9 2 12.5 43 75 19 13.8 0.253 0.726 0.184
CI 10 2 17.5 48 100 24 21.6 0.240 0.900 0.216
CI 11 2 22.5 43 125 31 22.7 0.248 0.732 0.182
CI 12 2 27.5 39 150 38 25.3 0.253 0.666 0.169
CI 13 2 32.5 35 176 45 36 0.256 0.800 0.205
B1 14 1 10 38 62 17 17.2 0.274 1.012 0.277
Lierstranda A7 15 2 10 24.5 78 23 11.4 0.295 0.496 0.146
A8 16 2 17.5 16 126 29 11.3 0.230 0.390 0.090
A9 17 2 25 12.5 181 38 12.9 0.210 0.339 0.071
A10 18 2 32.5 12 237 50 11 0.211 0.220 0.046
B2 19 1 10 23.5 78 23 13.5 0.295 0.587 0.173
Pentre NGI A5 20 2 20 13 174 50 22.3 0.287 0.446 0.128
A6 21 2 27.5 17 243 69 60.1 0.284 0.871 0.247
Pentre LDPT 22 1 20 13 179 50 17 0.279 0.340 0.095
23 1 30 13 263 75 51 0.285 0.680 0.194
24 1 40 18 350 100 79 0.286 0.790 0.226
25 1 50 18 445 125 78 0.281 0.624 0.175
Tilbrook NGI A 26 2 7.5 20 112 450 175 4.018 0.389 1.563
C 27 2 7.5 20 112 450 200 4.018 0.444 1.786
C 28 2 13.5 24 192 430 240 2.240 0.558 1.250
B 29 2 20 35 280 500 230 1.786 0.460 0.821
B 30 2 23.75 34 330 700 350 2.121 0.500 1.061
Tilbrook Compr. 31 1 7.5 20 112 450 100 4.018 0.222 0.893
LDPT Tension 32 1 12.5 25 172 425 200 2.471 0.471 1.163
-- 33 12 17.5 33 240 460 210 1.917 0.457 0.875
-- 34 12 22.5 35 308 670 300 2.175 0.448 0.974
West Delta 35 1 15 55 56 17 15.5 0.304 0.912 0.277
36 1 25 35 102 22 24 0.216 1.091 0.235
37 1 35 32 149 31 29 0.208 0.935 0.195
38 1 45 35 184 39 31.5 0.212 0.808 0.171
39 1 55 52 219 49 46.5 0.224 0.949 0.212
40 1 65 65 252 63 65 0.250 1.032 0.258
Bothkennar IC-pile 41 2 3.5 35 30 17 18 0.567 1.059 0.600
42 2 5 52 40.2 20 21 0.498 1.050 0.522
Cowden IC-pile 43 2 3.5 19 54 180 65 3.333 0.361 1.204
44 2 5 17 72.5 93 75 1.283 0.806 1.034
Canons Park IC-pile 45 2 3.5 54 46 78 65 1.696 0.833 1.413
46 2 5 40 60 120 99 2.000 0.825 1.650
Houston Unv 47 2 6.1 42 85 86 60 1.012 0.698 0.706
Drammen Pier # 16 48 1 25 17 243 56 28 0.230 0.500 0.115
Lake Oromieh New pile 49 1 33 20 215 55.7 43 0.259 0.772 0.200
1)
Pile tip condition during driving : 1  Open 2  Closed; 2)
Measured skin friction.

776

Copyright 2005 Taylor & Francis Group plc, London, UK


0.4
Database 1
49 data points

0.3

= skin measured/'vo
0.2

NC = 0.08.(Ip-10%)0.3
Min 0.05 NC 0.25
0.1

Data points from


LDPT at Pentre

0
0 10 20 30 40 50 60 70
Clay plasticity, Ip (%)

Figure 1. Measured pile skin friction in terms of -values, Figure 2. Measured pile skin friction in terms of -values
Database 1. for tests in soft clay, Database 1.

lowest values. It is also noted that the low


The -values back-figured from Database 1 are -values back-calculated for the NC clays are not
compared to the API recommendations on Figure 1. reflected in the nearly constant -values for low plas-
The filled-in points on this plot represent values from ticity clays.
the large diameter pile tests (LDPT) at Pentre and
Tilbrook, Clarke (1992).
For a strength ratio  of 0.20.3, most of the test 3 DATABASE 2
results fall well below the API line. This matter is
further investigated on Figure 2 where the back- This database contains results from 135 pile load tests
calculated -values are plotted against plasticity Ip. carried out at 52 different locations. The complete
All data points on this figure with -values smaller data base is included in NGI (2000). All except four
than 0.3 are from clay deposits that are close to nor- tests are in the public domain. Database 2 includes the
mally consolidated (NC). The fully drawn -line detailed soil layering and key soil parameters at each
shown on Figure 2, is taken to be representative for of the pile test locations.
NC clay deposits, and can for 10% Ip 55% be To calibrate the NGI-99 method, 36 of the best
expressed as: documented and most relevant pile tests from
Database 2 were considered. All of these piles are
(4) tubular steel piles. The 36 tests were selected using
the following criteria:
The dotted Min line on Figure 2 is taken to repre-
sent the lower bound value measured by all the pile The highest quality rating on soil and pile data.
tests studied by NGI. The lower bound skin friction A pile outer diameter of more than 0.2 m.
only depends upon vo and Ip, and not upon the A pile tip depth of more than 10 m.
undrained shear strength. This minimum value is A time between driving and testing of more than
given by: 14 days.

(5) The undrained shear strengths and pile capacities


reported in the source documents were modified by
the procedures described in Sections 4 and 5. To cali-
(6) brate the NGI-99 calculation method, the following
anchor point tests were selected, i.e. the model
An inspection of the   sDSS
u / vo values in Table 1 parameters were adjusted in order to obtain a good
shows that   0.22 is a representative average of the agreement with the measured results.

777

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 2. Database 2 anchor points. the measured pile capacity. This is due to (1)
increased horizontal effective stresses between pile
Pile test location Reference and soil as the excess porewater pressure set up dur-
ing pile driving dissipates, and (2) an ageing con-
Long Beach, USA Doyle & Pelletier (1985)
tinuing even after full dissipation of the excess water
Pentre LDPT, UK Clarke (1992)
Tilbrook LDPT,UK Clarke (1992) pressures. In addition, the pile capacity may increase
Onsy, Norway Karlsrud et al (1992) as a result of previous test loading of the pile, see
West Delta, USA Chan & Birrell (1998) Karlsrud & Haugen (1985) and Karlsrud et al (1992).
Drammen E18, Norway Tvedt & Fredriksen (2003) The reference pile capacity used by the NGI-99
Sandpoint, USA Fellenius et al (2004) method is the capacity measured 100 days after pile
driving. Since none of the piles studied were tested
after exactly 100 days, a correction for the effects of
4 UNDRAINED SHEAR STRENGTH time is needed. NGI-99 assumes that the strength
CORRECTIONS increase after dissipation of the excess porewater
pressures can be expressed as:
The undrained shear strength measured on a clay
sample in the laboratory depends upon a number of (8)
factors, including :
Sampling method where t is the time between driving and test loading,
Sample consolidation Q(t) is the capacity after t days and Q(100) the refer-
Direction of shear loading ence capacity after 100 days. This approach assumes
Type of test used that at both time t and time 100 days, full dissipation
Strain rate has taken place. The 10 value above is a dimension-
less capacity increase for a ten-fold time increase.
As an example, unconsolidated undrained (UU) The value of 10 could in theory be determined
triaxial tests are more influenced by sample disturb- from pile tests carried out on identical piles at differ-
ance than consolidated tests, but are often tested at a ent times. In practice, the value of 10 needs to be
higher strain rate. The API RP2A recommends that estimated from load tests carried out on the same pile
the strengths to be used for pile design are obtained at different times after driving. Since the measured
from UU tests. The NGI-99 method uses the same sUU u change in pile capacity is likely to be influenced both
reference strength. If other test types were used, one by the time since driving, and by the previous test
needs to convert to sUUu . In the NGI-99 method, the loading, it is not a straightforward matter to isolate
different tests are assumed to give the following NC the time effect.
strength ratios . Based upon a few tests in Database 2, supplemented
For each of these test types, it is assumed that the by the results given in Flaate (1968), the authors
strength of an overconsolidated sample (OC) can be decided to use the following 10 value for time cor-
expressed as: rection of the measured skin friction values:

(7) (9)

where OCR is the overconsolidation ratio, pc/vo.


With known test type, su, vo and Ip, the above can be (10)
used to obtain sUU
u and OCR.
If equation (7) gives an OCR value that is smaller where Ip and OCR are average values along the pile
than 1.0, this means that either the undrained shear shaft.
strength has been underestimated, or the in situ verti-
cal stress overestimated. For such cases the NGI-99
method assumes that the effective stress is correct, 6 THE NGI-99 CALCULATION METHOD
and the undrained shear strength is calculated for
OCR  1.0. Section 2 showed that the API (1993) recommenda-
tions, referred to as API-93, do not predict the low
skin friction values measured in NC clays of low plas-
5 EFFECT OF TIME BETWEEN PILE ticity. This effect was first documented by Karlsrud
DRIVING AND TESTING et al (1992). The important effect of clay plasticity
upon pile capacity for piles in NC clays has later been
Practical experience shows that the time between pile confirmed by full-scale pile tests in Drammen, Norway,
driving and pile testing has an important effect upon Tvedt & Fredriksen (2003), and in Sandpoint, Idaho,

778

Copyright 2005 Taylor & Francis Group plc, London, UK


Ip > 55% Table 3. NC strength ratios for different test types.
1
Test type NC strength ratio  NC
40%
Unconfined compression UCT  0.22
0.8 30% UU triaxial compression UU  0.25
CIU/CAU triaxial CMP  0.32
20% compression
= skin/SuRef

0.6 Direct simple shear DSS  0.22


15% Field vane FVT  0.18  0.24  Ip/100
FVT 0.3
12%
0.4
A C Table 4. Effect of pile tip condition.
B
0.2 Group Tests  Tip C/M avr.
Ip < 10% A : API-93
B : NGI-99, open-ended 1 26 0.251 Open 1.06
C : NGI-99, close-ended 2 47 0.251 Closed 1.00
0
0.1 1 10 3 28 110 Open 0.78
4 27 110 Closed 0.65
Strength ratio = suRef/'vo (Log scale)

Figure 3. Comparison between NGI-99 and API-93 -values. The factor F tip is taken as 1.0 for a pile driven
open-ended. Based upon results given by Clausen &
Fellenius et al (2004). It is therefore proposed to Aas (2001), and the results in Table 4 below, the fol-
revise the API-93 -values as indicated on Figure 3. lowing F tip expression is proposed for a pile driven
 ! 0.25: For NC clays with  0.25, the skin closed-ended:
friction is given by:
(18)
(11)
(19)
(12)
0.25 !  ! 1.0: For clays with 0.25  1.0,
where the reference strength suRef shall be taken as sUU
u
the -value is determined by a linear interpolation
or 0.25/0.22  1.14 times sDSS
u . Dividing (11) by (12)
between   NC and   1.0 as indicated on Figure
leads to : 3, allowing for the log scale of this plot.
The authors recommend to check that the calcu-
(13) lated skin friction is not smaller than Min  vo, where
Min is given by equations (5) and (6).
The tip resistance acting against a closed or plugged
The value to be used for NC, determined by calcu-
pile is taken as 9 times the undrained reference shear
lations with the two databases, is given by equation
strength. For piles subjected to long-term tensile loads,
(4). The NC value for UU tests is 0.25 (Table 3),
a crack could form at the pile tip, before the clay
which leads to :
strength is fully mobilised in reversed end bearing.
(14)
7 COMPARISON BETWEEN CALCULATED
(15) AND MEASURED CAPACITY

 " 1.0: For the overconsolidated clays with Figure 4 compares calculated NGI-99 skin friction to
  1.0, it is proposed to calculate the skin friction the measured values from Database 1. For the major-
from: ity of the data points, NGI-99 gives a good prediction.
It should be noted that there is a fairly large inherent
scatter in the pile and/or soil test data.
(16)
Ratios of calculated to measured pile capacities for
the 36 Database 2 tests are plotted on Figures 5 and 6.
(17) Seven of the eight anchor point values fall within

779

Copyright 2005 Taylor & Francis Group plc, London, UK


Calculated NGI-99 pile
capacity/Measured capacity
0 0.5 1 1.5 2
0
Database 2
36 data points

Tip depth below ground surface (m)


C/Mavr = 1.03
20 CoV = 0.26
Onsy
Tilbrook

40 Drammen
Sandpoint

Pentre
60
Anchor point tests West Delta
for NGI-99 calibration

Figure 4. Comparison between calculated and measured Long Beach


80
skin friction, Database 1 and NGI-99 calculation method.
Figure 6. Comparison between calculated and measured
pile capacities, Database 2 and NGI-99 calculation method.
Calculated NGI-99 pile capacity/Measured capacity

2
Database 2 Local skin friction & undrained shear strength (kPa)
36 data points 0 50 100 150
C/Mavr = 1.03 0
1.5 CoV = 0.26 Pre-drilled Ip (%)
to 15 m
Depth below ground surface (m)

10
12

14
1 20

Measured 18
30
NGI-99
0.5
40
SuUU 23
Anchor point tests
for NGI-99 calibration
0 50
0 10 20 30 40 50 60 70 16
API-93
Clay plasticity, Ip (%)
60
Figure 5. Comparison between calculated and measured Figure 7. Comparison between measured and calculated
pile capacities, Data Base 2 and NGI-99 calculation method. local skin friction, test pile Pentre.

20%, which only reflects that these points were For the soft clays with   0.25 to 1.0 the effect
used for the calibration of the NGI-99 method. of closing the pile tip is thus to increase the average
All of the 128 pile tests on steel piles in Database 2 skin friction by 6%. For the stiff clays with
were analysed in order to see if an effect of open/closed   110 a 20% increase is found.
pile tip during driving could be demonstrated. The Even if a calculation method gives the correct
NGI-99 calculation method was used, but with the Ftip answer for the total pile capacity, it may not give the
value in equation (16) taken as 1.0 for all piles. Each correct distribution of skin friction with depth. In
pile test was placed into one of four groups (Table 4), such case, the method could be non-conservative if
and the average ratio between the calculated and used for layered soil profiles. Figures 7 and 8 compare
measured capacity (C/M) was found for each group. the measured and the predicted local skin friction

780

Copyright 2005 Taylor & Francis Group plc, London, UK


Local skin friction & undrained shear strength (kPa) Results from pile load tests indicate that piles
driven closed-ended in stiff clays have higher skin
0 250 500 750 friction than open-ended piles. The NGI-99 method
0
includes a factor that reflects this observation.
SuUU NGI-99 gives a good agreement between measured
Depth below ground surface (m)

5 and calculated capacities for most of the published


Measured
and well documented large scale pile tests. However,
tension pile
there is considerable scatter and uncertainty when
10
it comes to the precise effect of plasticity upon the
skin friction in soft clays. Further pile load tests in
15 soft clays of medium and low plasticity are highly
desirable.
For design purposes the authors recommend to use
20 Measured several calculation methods, including the NGI-99
compr. pile and API-93 methods, complemented by e.g. Karlsrud
25 & Nadim (1990) and Jardine & Chow (1996). If for a
NGI-99 open given case, these methods lead to considerable differ-
NGI-99 closed
API-93 ences in the calculated capacity, the pile tests in the
30 data base that have the closest similarity to the case
studied should be used for guidance.
Figure 8. Comparison between measured and calculated
local skin friction, two test piles at Tilbrook.

ACKNOWLEDGEMENTS

values by the NGI-99 method for the LDPTs at Pentre The work presented herein was supported by Norsk
and Tilbrook, Clarke (1992). Hydro, Statoil and NGI. The authors gratefully
For the nearly NC silty clay deposit at Pentre, acknowledge their generous support.
Figure 7, it is seen that NGI-99 captures the variation
of the measured skin friction with depth quite well.
For the overconsolidated clays at Tilbrook, Figure 8,
there is only a small difference between the API-93 REFERENCES
and the NGI-99 methods. Both methods result in
American Petroleum Institute 1993. Recommended Practice
a good agreement with the measured values. The for Planning, Designing and Constructing Fixed Offshore
low skin friction measured in the top 12 m of the Platforms Working Stress Design. API RP 2A-WSD,
compression pile remains to be explained, Nowacki 20th Edition, Washington, 1 July 1993.
et al (1992). Chan J.H.C. & N.D. Birrell 1998. Project Overview and
Organization Tension Pile Study. OTC Paper no. 8762,
Houston, May 1998.
Clarke J. 1992 (Editor) Large-Scale Pile Tests in Clay. Proc.
8 CONCLUSIONS Large-Scale Fully Instrumented Pile Tests in Clay, ICE,
London, June 1992, ISBN 0-7277-1917 1, T. Telford Ltd.
A comparison between pile axial capacities calcu- Clausen C.J.F. & P.M. Aas 2001. Capacity of Driven Piles
in Clays and Sands on the Basis of Pile Load Tests.
lated by the API-93 method and the capacities actu-
Proceedings of the 11th (2001) International Offshore
ally measured shows that the calculated values can be and Polar Engineering Conference, ISOPE, Volume II,
34 times higher than the measured ones. This large Stavanger June 2001. p. 581586.
difference was only found for piles in normally con- Clausen C.J.F., P.M. Aas & K. Karlsrud 2005. Bearing
solidated clays of low plasticity. The authors therefore Capacity of Driven Piles in Sand, the NGI Approach. Proc.
propose a modification of the -factors, used by the of the ISFOG Conference, Perth, WA, September 2005.
API-93 method, that leads to a better agreement Dennis N.D. & R.E. Olson 1983. Axial Capacity of Steel
between calculated and measured capacities. Pipe Piles in Clay. Proc. Geotechnical Practice in
The proposed NGI-99 method uses the same refer- Offshore Engineering, Austin, Texas, April 1983.
Doyle E.H. & J.H. Pelletier 1985. Behaviour of a Large
ence undrained shear strength as API RP2A, i.e. UU
Scale Pile in Silty Clay. Proc., 11th ICSMFE, San
triaxial values, and provides conversion factors in Francisco 1985, Vol. 3, p.1595.
case other strength types are used. Fellenius B.H., D.E. Harris & D.G. Anderson 2004. Static
The calculated capacity corresponds to a time of Loading Test on a 45 m Long Pipe Pile in Sandpoint,
100 days after pile driving. The NGI-99 method Idaho. Canadian Geotechnical Journal, Vol. 41, No. 4,
includes time corrections on the measured pile capacity. August 2004, pp. 613628.

781

Copyright 2005 Taylor & Francis Group plc, London, UK


Flaate K. 1968. Bearing Capacity of Friction Piles in Clay. Axial Loading Based on Recent Instrumented Pile Load
NGF Stipend 19671968, Veglaboratoriet, Oslo, 1968. Tests. Society of Underwater Testing, London.
Jardine R.J. & F.C. Chow 1996. New Design Methods for Norwegian Geotechnical Institute 2000. Bearing Capacity
Offshore Piles. Marine Technology Directorate Ltd., of Driven Piles, Piles in Clay. Internal report no. 525211-1,
Publication MTD 96/103, London 1996. 23 March 2000.
Karlsrud K. & T. Haugen 1985. Axial Static Capacity of Nowacki F., K. Karlsrud & P. Sparrevik 1992. Comparison
Steel Model Piles in Over-Consolidated Clay. Proc. 11th of Recent Tests on OC Clay and Implications for Design.
ICSMFE, San Francisco 1985. Proc. Large-Scale Pile Tests in Clay, ICE, London 1992,
Karlsrud K. & F. Nadim 1990. Axial Capacity of Offshore edited by J. Clarke, Thomas Telford Ltd.
Piles in Clay. OTC paper 6245, Houston May 1990. Tvedt G. & F. Fredriksen 2003. E18 Ny motorvegbru i
Karlsrud K., B. Kalsnes & F. Nowacki 1992. Response of Drammen. Prvebelastning av peler. Proceedings from
Piles in Soft Clay and Silt Deposits to Static and Cyclic the conference on Rock Blasting and Geotechnics, Oslo.

782

Copyright 2005 Taylor & Francis Group plc, London, UK


Lateral pile design of the Ursa tension leg platform

Earl H. Doyle
Consultant, Sugar Land, Texas, USA

E.T. Richard Dean


Soil Models Ltd, London, UK

Jason A. Newlin
Shell International E&P Inc, Houston, Texas, USA

ABSTRACT: The piles for Ursa Tension Leg Platform (TLP) were installed in 1998. This TLP is one of the
most heavily loaded in the world, due to great water depth and large weight of water displaced by the hull. Large
diameter piles were used, without a foundation template. Calculations indicated that design lateral loads would
give pile deflections greater than the database on which API design criteria are based. Centrifuge tests were con-
ducted to develop criteria for large lateral deflections and group effects. Results are used to develop a design
basis and are compared with other design assumptions including existing Matlock/API criteria.

1 INTRODUCTION exceeded the test data on which the API lateral cri-
teria was based. A program of centrifuge tests was con-
The Ursa field is located 130 miles southeast of New ducted at Cambridge University, England to develop
Orleans. It encompasses Mississippi Canyon (MC) lateral design criteria including group factors for static
blocks 808, 809, 810, 852, 853 and 854. The field is and cyclic loading.
produced from a single Tension Leg Platform (TLP)
located in MC block 809 in 3,916 ft (1194 m) water
depth. Production began in March 1999. 2 CENTRIFUGE TEST PROGRAM
The Ursa TLP was designed to simultaneously with-
stand hurricane force waves and winds. The total design The use of geotechnical centrifuges in offshore geot-
weight of water displaced by the hull of was about echnics and foundation design is discussed by Rowe
97,500 tons. Global analyses for the design conditions (1983), Murff (1996) and others. Centrifuge model-
resulted in pile load conditions that were significantly ing and scaling techniques are discussed by Schofield
larger than previous TLPs. A foundation design was (1980), Craig et al. (1988), Taylor (1994) and others.
adopted with four groups of four piles each. Each pile Many of the larger geotechnical centrifuges of the
was designed with the tendon receptacle as an inte- world are listed by Ng (2004).
gral part of the pile, so that the pile consists of a drive The centrifuge test program for the Ursa TLP was
head, tendon receptacle, main pile body and driving reported by Doyle et al. (2004). The tests simulated
shoe. Each is 96( (2438.4 mm) in diameter and the lateral response of a 100-inch (2540 mm) diam-
417 ft (127.1 m) long, weighing approximately 380 t eter pile embedded in clay to 200 ft (61 m). For the pile
(344.7 tonne). groups, a spacing-to-diameter ratio (s/D ratio) of 3.08
At the location originally planned for the Ursa was used. Beam-column analyses on the prototype
platform, soil conditions within the zone where lat- pile using the Matlock (1970) criteria and expected
eral loading is important consist of normally consoli- maximum design loads were used to develop the model
dated clay. The interpreted soil shear strength was pile sizes and test parameters. The final prototype pile
0.029 ksf (1.4 kPa) at the mudline and increased lin- design used at Ursa, while close to the pile and soil
early at a rate of 1.1 to 1.2 kPa/m to a penetration of studied in the centrifuge test program, did not exactly
182 feet (55.5 m). match the pile and soil used in the test program.
During the preliminary design phase it was realized However, the results were judged to be applicable to
that the calculated lateral displacement of the piles the Ursa TLP.

783

Copyright 2005 Taylor & Francis Group plc, London, UK


Venting tube

Connector to links
of loading system

Load line

Pile, od 100" 4.9'


M1
M2
Mudline
M3

M4
Clay
Figure 1. Elevation view of centrifuge model test arrange- Layer A M5

Pile Section 1. Length 117.8 feet


ment, showing two piles contained in a steel tub of clay
consisting of three layers, A, B, and C with different con- M6
solidation histories.
M7

Four centrifuge tests were performed. Figure 1 M8


shows an elevation sectional view of one of the tests. M9
This test involved a two-pile group loaded laterally,
with pile heads connected by an articulated link loaded Clay M10
by a loading block. Figure 2 shows a typical model Layer B
pile. Instrumentation included up to 15 strain-gauge M11
bridges M1 to M15 measuring bending moments in
M12
the pile. There were load and displacement transduc-
ers at the pile caps.
The clay used in the model tests was speswhite
kaolin. Its properties are described by Clegg (1981), M13
Airey (1984) and Al-Tabbaa (1984, 1987).
All four tests were performed at 100 g centrifuge
gravities. Pile loading was displacement-controlled.
Test 1 was a static lateral loading test on a single pile,
taken to failure. Test 2 was primarily a static loading Pile Section 2. Length 56.8 feet
test on two piles loaded in line. The piles were spaced
at 3.08 diameters from centerline to centerline. Test 3 M14
involved cyclic lateral loading of a single pile. Test 4
involved cyclic lateral loading of a two-pile group Clay
loaded in line with the piling spacing again set at Layer C
3.08 diameters.
Results were analyzed using a polynomial fitting
technique and integrations and differentiations to
obtain lateral pile deflections along the pile lengths, M15
pile inclinations, lateral soil resistances, and shear
forces in the pile, as well as the measured bending Plug
moments. The analysis method was described by
Doyle et al. (2004). Firm
base

Figure 2. Instrumented model pile, with dimensions


3 CENTRIFUGE TEST RESULTS shown scaled to full prototype size. Total pile length at
prototype scale was 200 feet (61 metres). Prototype diam-
The centrifuge testing program produced a huge vol- eter was 100 inches (2540 mm). Some features shown were
ume of data. In this paper, only one aspect of the results required for testing purposes, but were not part of the Ursa
is discussed. It was possible to infer the ultimate lateral piles, such as the vent tube, plug, and base.

784

Copyright 2005 Taylor & Francis Group plc, London, UK


Pult factor Pult factor
0 2 4 6 8 10 0 2 4 6 8 10
0 0
Range of
centrifuge
2 2 data

4 4

6 6
Z/D

Z/D
8 8
Range of
centrifuge
10 data 10

12 Matlock cyclic (API, 2000) 12


Matlock static, adjusted, Ursa design
Matlock static (API, 2000) sensitivity 3.33
14 14

Figure 3. Centrifuge test results for lateral resistance factor Figure 4. Centrifuge test results, comparison with Matlock
Pult plotted against normalized depth z/D below mudline, and static criterion adjusted for sensitivity, and curve used for
comparison with calculations using Matlock (API, 2000). the Ursa pile design.
Effect of soil sensitivity ignored in Matlock calculations.

First, in reviewing the basis of the Matlock (1970)


soil resistances along the length of the pile, and to and API (2000) criteria, it was noted that pile deflec-
express this as a factor Pult times the soils shear tions in the relevant API database did not exceed about
strength. Adapting notation from Section 6.8 of API 25% of the pile diameter. In contrast, deflections cal-
(2000): culated with the Matlock criteria for the Ursa piles
exceeded 25% of the pile diameter, and deflections
significantly larger than 25% were observed in the cen-
(1) trifuge tests. It was suggested that the Matlock criteria
may not be applicable to the Ursa TLP pile design.
where pu is the ultimate lateral soil resistance (units of Another possible reason for the low load factors
force per unit pile length), c is the undrained shear could have been the severe remolding observed near
strength of the soil, and D is the pile outer diameter. the soil surface. Matlock (personal communication)
The undrained shear strength was measured during suggested that piles that undergo large deflection might
the centrifuge tests using a small-scale in-flight have a smaller ultimate pile factor. He suggested it
in-situ vane, and the results were correlated with mois- could be as low as 1/(Soil Sensitivity) times the factor
ture contents taken before and after testing. calculated using his original criteria.
Figure 3 shows selected results for Pult obtained Figure 4 shows the centrifuge data again, together
from the four centrifuge tests, plotted against the nor- with a curve for Matlocks recommendation regarding
malized depth z/D, where z is depth below mudline. sensitivity, and an Ursa design curve discussed later.
The results include some from cyclic tests in which A sensitivity of 3.33 was used, which would be typ-
CCTV visual observations during the tests suggested ical of many Gulf of Mexico clays. This gives a factor
that severe remolding was occurring in the clay, at least of 1/3.33  0.3, which was applied to the static
near the mudline. The horizontal extent of the bars indi- Matlock criteria for the purpose of these calculations.
cates the range of values of Pult inferred from the tests. The results of this comparison suggest that, for
Also shown are predictions of the Matlock static depths of 4 to 6 diameters below mudline, Matlocks
and cyclic methods, as described by API (2000). recommendations on sensitivity are in broad agree-
The plot shows that the inferred ultimate load fac- ment with the test data, but would be excessively con-
tors were mostly significantly below the criteria devel- servative below this.
oped by Matlock. Two potential causes of this were Based on these considerations, a design curve was
considered, as follows. selected for the Ursa design, as shown in Figure 4.

785

Copyright 2005 Taylor & Francis Group plc, London, UK


The design curve takes account of the low Pult values applied 11 feet (3.35 m) below the pile top. Local
from the mudline to z/D  8. It is fitted in such a way scour of one diameter (8 ft, 2.44 m) was assumed, to
that, below z/D  10, the Pult factors are the same as account for possible deleterious effects during instal-
the Matlock factors for cyclic loading. lation of a swinging pile prior to insertion. Sixty-ksi
(414 MPa) yield-strength steel was used throughout
4 URSA LATERAL PILE DESIGN the pile main body section.
Figure 5 shows results of the beam-column calcu-
The Ursa pile design is based on geotechnical condi- lations for the three cases. The lateral deflections are
tions reported by Doyle (1999). These conditions very small below about 20 pile diameters below mud-
were slightly different from the properties of the clay line. The bending moments are also small below this
in the centrifuge tests. However, the test conditions depth. The maximum moments for cases 2 and 3 were
were judged to be close to the final Ursa conditions. about the same, but the maximum for case 3 occurred
Lateral resistance factors inferred from the centrifuge lower down the pile.
tests were judged to be applicable. The pile size is The test results for z/D  8 were consistent with the
shown in Table 1. Matlock cyclic p-y curve. That curve implies that per-
Beam column analyses were conducted using the manent soil strain softening occurs at lateral deflections
above pile size. For this discussion only three of the greater than 3 yc. For the case here, taking a constant e50
lateral analysis cases are presented and these illustrate of 0.02, 3yc is 14.4( (366 mm), strain softening
the worst-case conditions. The cases were: should not occur for lateral deflections less than this.
In Figure 5a, a lateral deflection of 366 mm plots at
1. Matlocks cyclic criteria using a group pile factor about z/D  8 and 10 for Cases 2 and 3, respectively.
2. Centrifuge case: Ultimate pile factor based on the This explains why the test results below z/D  8 indi-
centrifuge tests (Fig. 4) and cated Pult factors similar to Matlock for cyclic loading.
3. Sensitivity case: Ultimate factor based on a soil For Figure 5c, the plotted pile stress is the sum of the
sensitivity factor of 1/3.33  0.3 (Fig. 4). stresses due to axial load and moment. Axial stress was
For the group pile analyses, it was assumed that the calculated by dividing axial load by cross-sectional
group factor for a 3.08 s/D ratio was 0.68 for the trail- area A. Bending stress was calculated by dividing bend-
ing pile and 1.0 for the leading pile. In order to ing moment by elastic section modulus Z. The steps in
account for the trailing pile group factor, the inter- the curves correspond to step changes of wall thickness,
preted shear strengths were multiplied by the group see Table 1. Ideally, changes of wall thickness would be
factor. The resulting shear strengths were used to cal- located well away from the place of maximum bending
culate the load-deflection (p-y) curves using the pro- moment, partly because section changes can induce
cedures recommended in API (2000). The ultimate stress concentrations. Figure 5c illustrates how the soil
load factors that were used are shown in Figures 3 & 4. resistance can affect this (effects of stress concentra-
The design load case was a vertical load of 5,930 tors were not incorporated in the plotted curves).
kips (26.4 MN) applied at an angle of 6.5 plus an The results show that the least conservative case is
installation tolerance of an additional 1. The result- Case 1, assuming standard Matlock cyclic criteria and
ing axial and lateral loads were 5,880 kips (26.2 MN) accounting for group effects. The most conservative
and 774 kips (3.44 MN) respectively. Loads were is Case 3, when the Matlock static criteria are modi-
fied by the soil sensitivity. Case 2, which was used in
the final Ursa design, is between the other two cases.
Table 1. Ursa pile sections and sizes.

Length, Length, Diameter, Wall thickness,


feet metres mm mm 5 DISCUSSION

7 (Top) 2.14 2057.4 76.2 In designing the Ursa piles, it was judged that the cen-
10 (Rec) 3.05 2082.8 101.6 trifuge test results produced criteria that most closely
6.25 1.91 Transition 107/76.2 fitted the design premise. Without these results, the
6.75 2.06 2438.4 76.2 design would have been based on the Matlock cyclic
10 3.05 2438.4 63.5 criteria, with ultimate resistance factor modified by
10 3.05 2438.4 57.1 some appropriate number (in this case 0.68 for the
90 27.43 2438.4 50.8
trailing pile) to account for group effects. This would
10 3.05 2438.4 44.4
10 3.05 2438.4 38.1 have produced an un-conservative pile design.
252 76.81 2438.4 34.9 The maximum calculated deflection at the mudline
5 (Shoe) 1.52 2438.4 50.8 for this Matlock-cyclic case was 25% of the pile
diameter and therefore it could have been assumed
Rec  Receptacle. that the criteria used was sufficient for the design

786

Copyright 2005 Taylor & Francis Group plc, London, UK


(a) -10 Lateral deflection, mm loading. Another way to view the design process would
- 400 0 400 800 1200 1600
be to use the Matlock criteria (adjusted for group
effects) and design a pile stiff enough to limit deflec-
0
tions to 25% of the pile diameter. The centrifuge tests
showed that the criteria used required a pile design
10 with thicker walls to resist the loading.
Depth below mudline, metres

The Sensitivity case appears to be a conservative


Case 1
20 assumption if no other data are available, at least for
Case 2 depths greater than z/D  6. It would have resulted in
30
wall thicknesses sufficient to resist the loads imposed
Case 3 by the design conditions. This case is not unreason-
able based on visual observations during and after the
40 centrifuge cyclic tests. Figure 6 shows a plan view of

50 (a)

60

(b) -10 Bending moment in pile, MN.m


- 20 0 20 40 60 80
0

10
Depth below mudline, metres

Case 2
20
Case 1
30

40 Direction
of pile
movement
50 Pile
Case 3

60 (b)
Directions of pile displacements during test

(c) -10 Max axial+bending stress in pile, MPa


approx 3D approx 3D
- 100 0 100 200 300 400
0

10 Case 2
Depth below mudline, metres

20
Case 1

30

40

Pile, diameter D approximate edge of


50
disturbed zone
Case 3
60 Figure 6. Post-test observations of the soil surface for the
third test, large displacement cyclic lateral loading of a sin-
Figure 5. Results of beam-column analyses for three analy- gle pile (a) photograph showing loading apparatus, pile, and
sis cases (a) lateral deflections, (b) peak bending moments, soil (white) with remolded area and disturbed zone, and
(c) max stresses in pile due to axial load and bending moments. (b) plan sketch of observations.

787

Copyright 2005 Taylor & Francis Group plc, London, UK


a depression observed around the pile cap after the thickness of about 3/8( (about 10 mm) for the 96( o.d.
end of the third test, involving cyclic lateral loading piles. The cost of this increase was minor in compari-
of a single pile. Around the pile is a zone of settlements son to the major benefits it provided of risk reduction
with remolded soil. Inspection of soil below the sur- and increased safety over the lifetime of the Ursa TLP.
face suggested a conical region of disturbed soil It is suggested that the criteria developed in the cen-
possibly bounded by shear zones. trifuge test program could be used under similar geot-
The remolding of soil in a zone around the pile echnical, geometric and loading conditions. However,
may also be help to explain some results of the group care should be taken to ensure that the deflections are
pile cyclic loading tests, for which the spacing-to- of the same order of magnitude as in the tests, in terms
diameter ratio s/D was only 3.08. In effect, zones of of ratio to the pile diameter. If deflections are signifi-
remolding around closely-spaced piles may overlap, cantly larger, the effects of remolding may be greater,
thereby creating a zone large enough to negate group and may extend to a greater normalized depth z/D.
effects if the s/D ratio is small enough. The centrifuge test results were consistent with the
proposal that, if no other data are available, piles with
large deflections and group effects may be conserva-
6 FURTHER STUDY tively designed by reducing Matlocks factors by the
soil sensitivity or equivalently by using the original
One centrifuge test series is not sufficient for use in factors and using the remolded strength instead of the
all TLP pile designs. The series described in this paper undisturbed strength.
does demonstrate that additional work needs to be
conducted to better understand laterally loaded piles
that have large deflections and group affects. ACKNOWLEDGEMENTS
As discussed by Doyle et al. (2004), Matlocks
static criteria gave remarkably good agreement with The Authors thank the operator of the Ursa field,
centrifuge test results on the statically loaded single Shell Exploration and Production Company, and its
pile and the statically loaded lead pile in a group. This partners, BP, ExxonMobil and ConocoPhillips for
suggests that large static lateral deflections, of the permission to publish the data. We thank the staff of
order of those seen in the centrifuge tests, do not Cambridge Universitys Schofield Center for their
necessarily reduce Matlocks ultimate lateral resistance efforts, which made the centrifuge tests possible.
factor for static loading. However, large cyclic deflec- We thank Professors Malcolm D. Bolton and
tions and group affects will change the Matlock cri- Andrew N. Schofield (Cambridge), Jitendra S. Sharma
teria and should be further investigated by more (University of Saskatchewan) and Arun J. Valsangkar
centrifuge testing. (University of New Brunswick).
Large strain finite element analyses, coupled with We especially thank the co-designer of the Ursa
centrifuge tests, simple conceptual models, and devel- piles, Mr. J. H-C. Chang, who is in our fondest mem-
opments in constitutive modeling for large strain ory as a most dedicated and thorough engineer, col-
cyclic loading, may assist with the further develop- league and friend. Mr. Chang also performed the
ment of understanding of processes that occur in the beam column analyses reported in this paper.
soil, and the relevant parameters, particularly near the
mudline where the greatest remolding occurs under
cyclic loading.
REFERENCES

Airey, D.W., 1984. Clays in simple shear apparatus. Ph.D


7 CONCLUSIONS thesis, Cambridge University
Al-Tabbaa, A., 1984. Anisotropy of clay. M.Phil thesis,
A centrifuge test program was used to develop ulti- Cambridge University
mate lateral resistance factors for the Ursa TLP pile Al-Tabbaa, A., 1987., Permeability and stress-strain response
design. Analysis of the test results presented in this of speswhite kaolin. Ph.D thesis, Cambridge University
paper indicate that the use of Matlocks cyclic criteria API, 2000. Recommended Practice for Planning, Designing
modified by a group factor would have produced an and Constructing Fixed Offshore Platforms Working
un-conservative pile design. Stress Design, RP2A-WSD, American Petroleum Institute
Figure 5b implies a 20% increase in maximum Clegg, D., 1981. Model piles in stiff clay. Ph.D thesis,
Cambridge University
bending moment for Cases 2/3 compared to Case 1. Craig, W.H., James, R.G. & Schofield, A.N. (eds), 1988
A 20% increase in bending moment requires a 20% Centrifuges in Soil Mechanics, Balkema
increase in section modulus (for the same grade of Doyle, E.H. 1999. Pile Installation Performance for Four
steel) to maintain a similar level of stress utilization. TLPs in the Gulf of Mexico, Paper OTC 10826, Offshore
In the present case, this represents an increase in wall Technology Conference, Houston, May

788

Copyright 2005 Taylor & Francis Group plc, London, UK


Doyle, E.H., Dean, E.T.R, Sharma, J.S., Bolton, M.D., Ng, C., 2004. Geotechnical Centrifuges in the World.
Valsangkar, A.J. & Newlin J.A., 2004. Centrifuge http://www.ust.hk/webgcf/GCFCenter.html, Hong Kong
Model Tests on Anchor Piles for Tension Leg Platforms, University of Science and Technology,
Paper OTC 16845, Offshore Technology Conf, Houston, Rowe, P.W., 1983. Use of large centrifugal models for off-
May. shore and nearshore works, Proc Int Symp Geotechnical
Matlock, H., 1970. Correlations for design of laterally loaded Aspects of Coastal and Offshore Structures, Bangkok,
piles in soft clay, Paper OTC 1024, Offshore Technology eds.Yudbhir and A.S.Balasubrananiam, Balkema, 2133
Conf, Houston, May Schofield, A.N., 1980. Cambridge geotechnical centrifuge
Murff, J.D., 1996. The geotechnical centrifuge in offshore operations, Geotechnique, 30(3), 227268
engineering. OTRC Honors Lecturer, Offshore Technology Taylor, R.N. (ed), 1994. Geotechnical Centrifuge Technology,
Conference, Houston, Texas Blackie Academic & Scientific

789

Copyright 2005 Taylor & Francis Group plc, London, UK


Case study on soil plugging of open-ended steel pipe piles in Tokyo Bay

T. Matsumoto & P. Kitiyodom


Kanazawa University, Kanazawa, Japan

ABSTRACT: In this work, a case study on soil plugging of two large diameter open-ended steel pipe piles,
which were constructed in Tokyo Bay, is presented. Analysis of the load-displacement relationships of these
piles are carried out using a hybrid numerical program KWAVE. Good agreements between the analysis results
and the measurement values are found. Then a parametric study is carried out to investigate possible methods
to increase the bearing capacity of the pipe piles due to the increase in the soil plugging effect.

1 INTRODUCTION Moreover, the parametric study is carried out to investi-


gate possible methods to increase the bearing capacity
Open-ended steel pipe piles are widely used for founda- of the pipe piles due to the increase in the soil plug-
tions both on land and offshore. During driving process ging effect.
of these piles into the soil, a soil column known as the
soil plug is formed inside the piles. As penetration con-
tinues, the frictional resistance between the inner pile 2 OUTLINE OF TEST PILING IN TOKYO BAY
shaft and the soil plug may develop, and may prevent
further soil intrusion. Depending on the relative move- Trans-Tokyo Bay Highway (TTBH) connects Kawasaki
ment between the pile and the soil plug, the pile is con- City and Kisarazu City over the Tokyo Bay. TTBH
sidered to be perfectly plugged, imperfectly plugged consists of the tunnel section (9.4 km) and the bridge
or unplugged. The question then arises how the effects section (4.7 km). The bridge piers are supported by
of the soil plug on the pile performance during pile driven open-ended steel pipe piles.
driving and during static loading in compression. The test piling on two open-ended pipe piles was
An empirical design formula has been developed in performed in 1989 through 1990 at a point of the bridge
Japan for open-ended steel pipe piles having diameters route about 3 km offshore from Kisarazu City. Pile
less than 1.0 m. However, foundation engineers some- specifications are summarised in Table 1. Two test
times encounter troubles in the use of the design for- piles, designated as T1 and T2, were driven by a steam
mula, since open-ended steel pipe piles having outer hammer. The soil that intruded into the pile was excav-
diameters larger than 1.0 m are constructed in some ated during interruption of pile driving so that pile T2
cases. had no soil inside it (had no internal shaft resistance).
With an aim to clarify the effects of the soil plug on The compression load tests of the piles were per-
the bearing capacity of driven open-ended pipe pile, the formed after 2 months from the end of pile driving.
test piling was conducted on two large diameter open- The tension load test was conducted on pile T1 after 1
ended pipe piles in Tokyo bay (Momiyama et al. 1992). year from the end of the compression load test. Figure 1
In order to investigate the influence of plugging shows the SPT N-values at the site.
effects on the load-displacement relationship, in this
paper, an analytical study on soil plugging of these
open-ended pipe piles was conducted using an exten- Table 1. Pile specifications.
sion version of a hybrid numerical program KWAVE
developed by Matsumoto & Takei (1991). The program Property Pile T1 Pile T2
has been extended so that it can include the silo model
of the soil inside the pile with consideration of soil Outer diameter (mm) 1600 1600
deformability and can model cross steel brace which Wall thickness (mm) 24 24
Length (m) 50.95 63.15
is attached inside the pile. Embedment length (m) 26.65 32.75
It was found that there are good agreements between Youngs modulus (kPa) 2.06  108 2.06  108
the analysis results and the measurement values.

791

Copyright 2005 Taylor & Francis Group plc, London, UK


SPT N-value Axial force (MN)
0 40 80 120 160 200 240 280 320 360 0 4 8 12 16 20
5 0
0 -20

-5 -15
-10 -10
-10
-15 -5
G.L. G.L.
0 -20
Depth form G.L. (m)

-20

T.P. (m)
-25 5
T.P. (m)

-30 10
-30
-35 15

-40 20

-45 25 -40
-50 30
Pile T1
-55 35
-50 Pile T2
-60 40

-65 45
50 Figure 3. Distributions of the axial forces of piles T1 and T2.

Figure 1. SPT N-values at the test site in Tokyo Bay.


indicated again that the internal shaft resistance of
pile T1 was negligible.
Pile head load (MN)
3 ANALYSIS OF TEST PILING
0 2 4 6 8 10 12 14 16 18 20
0
3.1 Analytical method
pile toe
5 Although analysis of driving response of an open-
ended pipe pile may be carried out by a finite element
analysis (Liyanapathirana et al. 1998, 2000, 2001),
Settlement (mm)

10
a simplified method was used in this study.
pile head In order to consider the influence of the soil plug
15 on the behaviour of an open-ended pipe pile during
driving as well as static loading, the pipe pile is mod-
20 elled as shown in Figure 4 according to Randolph &
Simons (1986). The soil plug is modelled as a series
of masses and springs, namely pile within a pile
25 Pile T1
(Heerema & de Jong 1980), with frictional forces
Pile T2 between the soil nodes and pile nodes, so that the
30 interaction between the soil plug and the pile can be
introduced in the analysis. The pile/soil modelling has
Figure 2. Load-settlement relations of piles T1 and T2. been incorporated in a computer program KWAVE by
Matsumoto & Takei (1991).
The static soil spring at the pile base node and the
The load-settlement relations of piles T1 and T2 soil plug base node are estimated by means of Equations
are compared in Figure 2. It is seen from Figure 2 that (1) and (2) following Muki (1961) (after Poulos &
the load-settlement curves of piles T1 and T2 are Davis 1974), and the static soil spring at the pile shaft
almost identical until the pile head load of 12 MN was nodes are estimated by mean of Equation (3) follow-
reached, indicating that the mobilisation of the inter- ing Randolph & Wroth (1978):
nal shaft resistance was negligible in pile T1.
The distributions of axial forces along the piles are (1)
compared in Figure 3. The rate of decrease in the axial
force down the pile shaft is similar for piles T1 and T2,
which mean that the mobilisation of the shaft resistance
including the internal and external shaft resistances (2)
was almost identical in piles T1 and T2. This result

792

Copyright 2005 Taylor & Francis Group plc, London, UK


Pipe pile External where G0 is the shear modulus of the soil at the low
soil strain, Rf is the hyperbolic curve fitting constant,  is
the shear stress, and f is the shear stress at failure.
With the tangent shear modulus in Equation (5),
Soil plug
the soil spring at the pile base node, the soil spring at
the soil plug base node and the soil spring at the outer
pile shaft nodes (Equations (1) to (3)) are estimated
by means of Equations (6) to (8):
Soil plug

(6)

(7)

(8)
Figure 4. Modelling of pipe pile, external soil and soil plug.

where G is the shear modulus of the soil,  is the


Poissons ratio of the soil, ro is the outer radius of the in which   oroRf/f, Pb is the mobilised base load,
pile and ri is the inner radius of the pile. Pf is the ultimate base load,  is the nonlinear influ-
ence factor, and o is the shear stress at the outer pile-
soil interface.
(3) As the sides of the soil plug are constrained by the
inner shaft of the pile, one-dimensional Youngs modu-
lus, Ec, of the soil plug is obtained using Equation (9).
where rm  2.5 L(1
) and L is the pile length.
Note that the internal shaft spring kin is employed
for calculation of mobilisation of the internal shaft (9)
friction. Although the value of kin could be related to
shear deformation of the soil plug, i.e. relative dis-
placement between the centre and the edge of the soil The validity of the method was insured through the
plug, the value of kin was set to be a large value within back analyses of the model pile driving test and model
which the calculation becomes stable. static load test results (Kitiyodom et al. 2004).
The maximum internal friction, in(max), is
expressed as 3.2 Analytical condition
In the test site, only the profile of SPT N-values was
(4) available (see Figure 1). Hence, the shear wave veloc-
ity of the soil was estimated by means of an empirical
where  is the coefficient of friction between the soil relation as shown in Equation (10) following Imai
plug and the pile, r0 the initial effective radial stress (1977).
in the soil plug before loading, v the increment of
the effective vertical stress in the soil plug after load- (10)
ing and K0 is the coefficient of earth pressure at rest.
To consider nonlinear behaviour of the soil in the From Equation (10), the shear modulus of the soil
analyses, a hyperbolic relationship between shear stress at the low strain, G0, can be calculated as follows:
and shear strain is employed as suggested by Chow
(1986). The tangent shear modulus of the soil, Gt, is
(11)
given by
where the assumed soil density of 1.8 ton/m3 was used.
(5) The other parameters used in the analyses are sum-
marised in Table 2. Both analyses assuming the soil

793

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 2. Parameters used in the analysis. Load applied to the pile head (MN)
0 4 8 12 16 20
Property Value
0
Friction coefficient,  0.3
Coefficient of earth pressure at rest, K0 0.5 5
Poissons ratio of the soil plug,  (drained) 0.3
(undrained) 0.49

Settlement (mm)
Rf (outer pile shaft) 0.9 10
(pile base) 0.99 Pile T2
15

Measured w0
20
Measured wt

25 Calculated w0
Calculated wt
Qout
Qint 30

Figure 6. Calculated and measured load-settlement curves


of pile T2.

Axial force (MN)


0 4 8 12 16 20
Qtoe Qplug Qtoe 0 10
5 5
Figure 5. Components of resistance at pile toe. Calculated
10 Measured 0
15
Distance from pile head (m)

-5
plug to be fully drained or fully undrained were car-

Elevation from T.P. (m)


20
ried out. Note that in undrained condition there is no G. L. -10
increase in the effective stress of the soil plug, so the 25 -15
value of v in Equation (4) was set as 0. 30 -20
3.3 Analytical results 35 -25
As shown in Figure 5, the soil plug resistance, Qplug, 40 -30
can be calculated easily from the force equilibrium 45 -35
condition at the pile toe using Equation (12).
50 -40
(12) 55 -45
Pile T2
60 -50
where P is the applied force at the pile head, Qout is the
outer pile shaft resistance and Qtoe is the resistance
against the annulus of the pipe pile bottom. Figure 7. Calculated and measured axial forces along
However, in practice it is difficult to distinguish pile T2.
between the outer and the inner pile shaft resistance
from the results of pile load test. Fortunately, as Figure 6 shows the calculated and measured load-
described before, in Momiyama et al. (1992) the soil settlement relations of pile T2. Both the settlements at
that intruded into pile T2 was excavated during inter- the pile head, w0, and the pile toe, wt, are shown in the
ruption of pile driving so that pile T2 had no soil inside figure. The calculated and measured distributions of
it (had no internal shaft resistance). axial forces along pile T2 are shown in Figure 7. There
In this study, back analysis of pile T2 was carried are good agreements between the calculated results
out first. Using the parameters obtained from the and the measured values. In this analysis, the shear
back analysis of pile T2, the analyses of pile T1 were modulus of the soil at the low strain calculated from
carried out. The analysis results were compared with Equation (11) was reduced by a factor of 0.6 to obtain
the site measurements. a good matching. The distribution of the outer shaft

794

Copyright 2005 Taylor & Francis Group plc, London, UK


Load applied to the pile head (MN) Axial force (MN)
0 10 20 30 40 50 60 70 80 90 100 0 4 8 12 16 20
0 0 5

100 5 Calculated 0
10 Measured -5

Distance from pile head (m)


200

Elevation from T.P. (m)


15 -10
Settlement (mm)

300 Pile T1
(drain) 20 -15
G. L.
400 25 -20
500 30 -25
Measured w0
600 Measured wt 35 -30
Calculated w0 40 -35
700
Calculated wt Pile T1
45 -40
800 (undrain)
50 -45
Figure 8. Calculated and measured load-settlement curves
of pile T1 (drained condition). Figure 10. Calculated and measured axial forces along pile
T1 (undrained condition).

Load applied to the pile head (MN)


Load (MN)
0 5 10 15 20
0 0 5 10 15 20
0
Total
20
Qout
20 Pile T1
Settlement (mm)

Pile T1 (undrain) Qplug


40
Settlement (mm)

(undrain)
Qtoe
40
60
Measured w0
60
Measured wt
80
Calculated w0
Calculated wt 80
100

Figure 9. Calculated and measured load-settlement curves


of pile T1 (undrained condition). 100

Figure 11. Calculated load-settlement relations of pile T1


(undrained condition).
resistance was estimated from the measured axial
forces at the pile head load of 16 MN.
Figures 8 and 9 show the calculated and measured The calculated and measured distributions of axial
load-settlement relations of pile T1 assuming the soil forces along pile T1 are compared in Figure 10. The soil
plug to be fully drained and fully undrained, respect- plug is assumed to be fully undrained in the calcula-
ively. It can be seen that the calculated results, which tion. There are good agreements between the calculated
assumed the soil plug to be fully drained, overesti- results and the measured values.
mate the measured ultimate capacity of the pile. The Figure 11 shows the calculated load-settlement rela-
calculated results, which assumed the soil plug to be tions of pile T1 assuming fully undrained condition in
fully undrained, match very well with the measured the soil plug. The total pile head load, the total outer
values. So it may be stated that the soil plug inside the shaft resistance, Qout, the total inner shaft resistance,
pile was in fully undrained condition during even the Qplug, and the total toe resistance, Qtoe, are plotted
static compressive pile load test. against the pile head settlement. It is seen from the

795

Copyright 2005 Taylor & Francis Group plc, London, UK


Load applied to the pile head (MN)
0 10 20 30
0
Measured
10
Calculated
20 Ldrain = 10m

Settlement (mm)
30
Ldrain = 5m
40
Ldrain = 1m
50

60

70
Pile T1
80

Figure 13. Calculated load-settlement curves varying


Ldrain (without cross steel rib).
Ldrain
Load applied to the pile head (MN)
0 10 20 30
0
Measured
Figure 12. Length of the fully drained section. 10 Calculated
20 Ldrain = 10m
Settlement (mm)

calculated results that the total inner shaft resistance 30


is relatively small compared to the total resistance of 40 Ldrain = 1m
the pile. Note that in the analyses of pile T1, the initial Ldrain = 5m
50
effective radial stress in the soil plug, r0, was assumed
as r0  K0v0 where v0 is the effective overburden 60
pressure, and that the bearing capacity of the pile toe
70 Pile T1
and the soil plug base was assumed as 45000 kPa.
In Matsumoto et al. (2005), the influence of the gen- 80
eration of pore pressures in the soil plug during pile
driving and of the consolidation of the soil plug after Figure 14. Calculated load-settlement curves varying
Ldrain (with cross steel rib).
pile driving on the behaviour of an open-ended steel
pipe pile is discussed. They suggest that the set-up phe-
nomena due to the consolidation of the soil plug after Figure 13 shows the calculated load-settlement rela-
pile driving could not be expected. tions of pile T1 varying the length of the fully drained
section as 1, 5 and 10 m. It can be seen from the figure
that the bearing capacity of the pile increases as the
length of the fully drained section increases. Rapid
3.4 Parametric study
increase in the bearing capacity can be seen for the
In the previous section, it was shown that the pile-soil length of the fully drained section greater than 5 m
modelling which was employed in the present study can (3 times the pile diameter).
simulate the behaviour of the pile, the soil and the soil Figure 14 shows the calculated load-settlement rela-
plug during static loading, if adequate soil parameters tions of pile T1 having the cross steel brace attached
are selected. In this section, a parametric study was inside the pile, varying the length of the fully drained
carried out to investigate possible methods to increase section as 1, 5 and 10 m. It can be seen from the figure
the bearing capacity of the pipe piles due to the increase that the bearing capacity of the pile increases as the
in the soil plugging effect. length of the fully drained section increases. Compared
Two methods are considered in this paper. One of with the corresponding cases in Figure 13, the increase
them is to increase the length of the fully drained sec- rate of the bearing capacity of the pile having the cross
tion in the soil plug from the soil plug base, Ldrain, (see steel brace is higher than that of the pile without the
Figure 12). The other is to increase the inner shaft cross steel brace.
interface area per unit pile length by attaching the Figure 15 shows the calculated distribution of the
cross steel brace inside the pipe pile. inner shaft resistances at the pile head load of 30 MN.

796

Copyright 2005 Taylor & Francis Group plc, London, UK


Inner shear resistance (kPa) Heerema, E.P. & de Jong, A. 1980. An advanced wave equa-
0 400 800 1200 tion computer program which simulates dynamic pile
0 5 plugging through a coupled mass-spring system. Proc.
5 Pile T1 in TTB(1) 0 Int. Conf. on Numerical Methods in Offshore Piling,
London: 3742.
Distance from pile head,x (m)

10 -5

Elevation from T.P. (m)


Imai, T. 1977. P- and S-wave velocities of the ground in
15 -10 Japan. Proc. 9th ICSMFE, Tokyo: 257260.
20 -15 Kitiyodom, P., Matsumoto, T., Hayashi, M., Kawabata, N.,
G.L. Hashimoto, O., Ohtsuki, M. & Noji, M. 2004. Experiment
25 -20
on soil plugging of driven open-ended steel pipe piles
30 Calculated -25 in sand and its analysis. Proc. 7th Int. Conf. on the
35 -30 Application of Stress-Wave Theory to Piles, Selangor:
at P = 30 MN 447458.
40 -35
Liyanapathirana, D.S., Deeks, A.J. & Randolph M.F. 1998.
45 -40 Numerical analysis of soil plug behaviour inside open-
50 -45 ended piles during driving. International Journal for
Numerical and Analytical Methods in Geomechanics, 22:
Figure 15. Calculated distribution of the inner shaft resist- 303322.
ances (with cross steel rib). Liyanapathirana, D.S., Deeks, A.J. & Randolph M.F. 2000.
Numerical Modelling of large deformations associated
with driving of open-ended piles. International Journal
It is seen that the inner shaft resistance increases for Numerical and Analytical Methods in Geomechanics,
exponentially from the top of the soil plug to the pile 24: 10791101.
base, showing a so-called silo effect. Liyanapathirana, D.S., Deeks, A.J. & Randolph M.F. 2001.
Numerical modelling of the driving response of thin-
walled open-ended piles. International Journal for
Numerical and Analytical Methods in Geomechanics, 25:
4 CONCLUDING REMARKS 933953
Matsumoto, T. & Takei, M. 1991. Effects of soil plug on
A case study on soil plugging of two large diameter behaviour of driven pipe piles. Soils and Foundations
open-ended steel pipe piles, which were constructed 31(2): 1434.
in Tokyo Bay, was presented. Analyses of the load- Matsumoto, T., Wakisaka, T. & Numata, A. 2005. One
displacement relationships of these piles were carried dimensional wave propagation analysis of an open-ended
out using a hybrid numerical program KWAVE. pipe pile with consideration of the excess pore pressure
It was found that the pile/soil modelling which was in soil plug. Proc. International Symposium on Frontiers
in Offshore Geotechnics, Perth (to be presented).
employed in the present study can simulate the behav-
Momiyama, Y., Honma, M., Katayama, T. & Maruyama, T.
iour of the pile, the soil and the soil plug during static 1992. Vertical loading test on large-diameter steel-pipe
loading, if adequate soil parameters are selected. piles for Trans-Tokyo Bay Highway, TSUCHI-TO-KISO
From the parametric study, it was found that two JSSMFE 40(2): 4752 (in Japanese).
methods are effective to increase the bearing capacity Poulos, H.G. & Davis, E.H. 1974. Elastic solution for soil
of the pile due to the increase in the soil plugging effect. and rock mechanics. New York: Wiley.
One of them is to increase the length of the fully drained Randolph, M.F. & Simons, H.A. 1986. An improved soil
section in the soil plug. The other is to attach the cross model for one-dimensional pile driving analysis. Proc. of
steel brace inside the pipe pile. By the latter method, the 3rd Int. Conf. on Numerical Methods in Offshore
Piling, 117.
equivalent radius of separated inner section becomes
Randolph, M.F. & Wroth, C.P. 1978. Analysis of deform-
smaller and the silo effect is effectively promoted, ation of vertically loaded piles. Journal of Geotechnical
compared to the hollow pile. Engineering ASCE 104(12): 14651488.

REFERENCES

Chow, Y.K. 1986. Analysis of vertically loaded pile groups.


International Journal for Numerical and Analytical
Methods in Geomechanics, 10: 5972.

797

Copyright 2005 Taylor & Francis Group plc, London, UK


Foundation capacity of piled offshore platforms

N. Morgan, I.M.S. Finnie & G. Stewart


Lloyds Register EMEA, Aberdeen, United Kingdom

J.L. Price
Formerly of Lloyds Register

ABSTRACT: Foundation system capacity analyses can be used to demonstrate the suitability of offshore jacket
structure foundations when the capacity of individual piles or pile-groups is considered marginal. This extends
the more commonly adopted approach of considering only the piles or pile-groups as the single foundation elem-
ents, for which the prescribed design capacities are to be achieved. One simplified analysis method is to assume
that the jacket behaves as a rigid-body supported by the piles, which are treated as a whole foundation system,
rather than individual components. This paper presents an example parametric study to assess the effect of plat-
form self-weight on foundation system capacity, and provides an explanation of the evident trends with respect
to the progressive failure of the entire foundation. It demonstrates that the system capacity may appreciably exceed
first-pilefailure in certain particular circumstances and, conversely, an increase or shift in platform weight may
have negligible effect in other circumstances.

1 INTRODUCTION additional topsides weight or risers are to be added


to an existing platform, or
Axial pile design for offshore platforms is tradition- the environmental criteria have been revised and
ally based on achieving target factors of safety that considered to be more onerous than originally pre-
take into account functional loading, gravity loading, dicted, or
extreme storm loading, which are compared to the the consequence of loss of a pile due to fatigue
predicted capacity of individual piles. With due consid- damage is to be quantified as part of life-extension
eration for structural issues, it is often justifiable to studies.
allow for load-sharing between piles in a group and
thus to consider the average loads and capacities
within a pile-group to determine factors of safety 2 RIGID-BODY FOUNDATION ANALYSIS
(Toolan & Horsnell 1980). Taking this one step fur-
ther, the system capacity of the entire platform foun- The pushover response of the entire platform struc-
dation can also be determined through consideration ture and foundation system, taking into consideration
of load-sharing between pile-groups. This is reasonable yielding and buckling of individual components
as long as the structure is able to allow load-sharing necessitates the use of full non-linear analyses (e.g. as
between piles or pile-groups, with acceptable redun- described by DNV-Sintef Bomel 1999).
dancy. In order to establish platform suitability Lloyds However, before embarking on such analysis, it
Register routinely resorts to the use of system-capacity often proves pragmatic to establish the system cap-
analyses in certain scenarios, such as the following acity of the foundation only, assuming that the jacket
examples; behaves as a rigid-body. The required input for this
rigid-body foundation analysis (RBFA) should be
when ambiguity or debatable uncertainty exists in readily available from a standard linear-elastic jacket
the pile design, or analysis, which is usually mandatory for every plat-
the observed behaviour during the pile installation form design. In essence, RBFA can be seen as either a
is found to be noticeably different from that origi- precursor to, or replacement for more sophisticated
nally predicted, or non-linear jacket pushover analyses.

799

Copyright 2005 Taylor & Francis Group plc, London, UK


2.1 Analysis method each isolated pile or pile-group would be shown to sat-
isfy code requirements. The condition where the fac-
The RBFA program developed by Lloyds Register is
tored reactions equal the factored pile or pile-group
based on the fundamental assumption that the pile-head
capacities is usually termed first-pile failure, although
displacements conform to a rotation and vertical trans-
it should be appreciated that this is for factored loads
lation of the jacket and the resultant axial pile reactions
and capacity. System failure is considered to occur
correspond to pre-defined responses, which follows
where uncontrolled displacement of the platform itself
similar approaches as those of Toolan & Horsnell
occurs, irrespective of any prior individual pile failure.
(1980) and Stewart & van der Graaf (1990). In its sim-
plest form, the method assumes that lateral pile failure
is not governing and that the pile reactions are vertical. 3.1 Example
The total foundation response for a number of piles
In order to best-illustrate the differences between sys-
n is solved for the limiting equilibrium of vertical
tem capacity and first-pile failure, and also the effect
loads and two component moments, whilst satisfying
of the magnitude and position of the platform self-
the corresponding n
3 planar equations to ensure that
weight, the following specific example is considered
the assumed rigid geometry is maintained.
throughout this paper.
The vertical translation and rotation of the jacket is
A jacket with a square footprint is supported on 4
iteratively varied until the magnitude of the resultant
identical piles at each corner. The plan position of the
vertical pile reactions equals the platform self-weight
piles and their factored load-displacement responses
and the magnitude of the environmental overturning
are presented in Figures 1 and 2, respectively.
moment (EOTM) is maximized for a considered load-
The platform self-weight is (initially) considered to
ing direction.
be central, and equal to 49MN, which may be assumed
to include the appropriate partial load factors.
2.2 Application The pile response, in Figure 2, is near-linear up to
RBFA is ideally undertaken within a partial factor the peak, followed by a slight reduction to a residual at
framework. For instance, adopting API RP2A LRFD larger displacements, with tension and compression
(1993) philosophy, it would be expected that the ratio response being similar, perhaps corresponding to a
of maximum EOTM to the design 100 year storm event pile developing most of its capacity in shaft sticktion.
EOTM would exceed the minimum recommended Following a linear-elastic structural analysis the input
load-factor (i.e. 1.35 for extreme conditions plus allow- would include an applied EOTM profile, which would
ing for dynamic amplification), if the recommended include wind, waves & current and any other relevant
partial dead and live load factors and also the maximum factors.
recommended pile resistance factor (e.g. 0.8) had been
applied. 3.2 Capacity envelopes
This EOTM ratio could conceivably be (a) deter-
mined for the maximum attainable overturning moment, Figure 3 presents the example EOTM envelopes and
corresponding to irrecoverable plunging failure of the design applied moment envelopes (which may be
piles and uncontrolled rotation of the platform (i.e. assumed to include any applicable factors) for loading
foundation-governed pushover or system failure), or of the base-case example in all plan-directions.
(b) for the overturning moment that causes first-pile
or pile-group failure, corresponding to conventional 3.2.1 First-pile failure
design (i.e. in-place) analysis, or (c) the overturning In Figure 3 the diamond-shaped envelope corresponds
moment that causes excessive pile-head displacements, to the EOTMs that induce first-pile failure. (These cor-
from the perspective of serviceability or structural respond to plunging failure of the compression piles
vulnerability. Project-specific performance standards that most directly resist the overturning moment).
could provide definition of minimum required EOTM These can be compared to the design EOTM, corres-
ratio. ponding to the innermost squarish envelope. It can be
seen from Figure 3 that the first-pile failure envelope
lies within the target EOTM, for most directions, which
3 SYSTEM CAPACITY AND FIRST PILE
would normally be considered unsatisfactory in terms
FAILURE
of conventional design philosophy. However, this is not
necessarily the case if system capacity is considered.
Pile sizing is usually based on linear jacket analysis,
where platform self-weight and environmental load
components are individually considered to establish the 3.2.2 System failure
governing pile-head loads and then individually fac- In Figure 3 the outer square envelope corresponds to
tored according to the applicable code. On that basis the maximum EOTM that could be sustained by the

800

Copyright 2005 Taylor & Francis Group plc, London, UK


B1 10 B2 33.75 22.5 400
45
Various COGs 8
considered
6 56.25

4 200
Base Case 67.5

EOTM about y axis MNm


COG
Y Coordinate m

2
Broadside 78.75
0
0
-10 -8 -6 -4 -2 0 2 4 6 8 10 0
-2 -400 -200 200 400
0 0
-4

-6
-200
-8
A1
-10 A2
X Coordinate m
-400
Diagonal 45 All Moments and pile
coordinates are based upon the EOTM about x axis MNm
centre of coordinates (0,0) Design Moment First Pile Failure Pushover Failure

Figure 1. Jacket layout at mudline. Figure 3. EOTM envelopes for 49MN platform weight.

30
Figure 3 would lie near the envelope with shaded circles
20 in Figure 4a. An interesting feature of Figure 4c is the
appearance of facets or cliffs normal to the diagonals,
10 which are indicative of an EOTM insensitivity to self-
Axial Load MN

weight when the weight reduces below the value at


0 which the facet first emerges. Sections through the
-0.15 -0.05 0.05 0.15 0.25 3-dimensional view of the failure surface (Fig. 4c) are
-10 presented on Figure 4b.
Normal Pile Response
For diagonal load-cases there is a critical self-weight
-20 0.8xLoad, below which the system capacity is not influenced by
1xDisplacement self-weight. This occurs because the self-weight is
0.8x Load and adequately carried by the outrigger piles (i.e. A2 & B1
-30 Displacement
Axial Displacement m for 45, see Fig. 1) up to system failure, which is then
limited by the capacity of the piles in-line with the
Figure 2. Pile load-displacement response. applied EOTM. This also happens to be the underlying
reason why the system capacity of tripod structures is
entire foundation system. This exceeds that corres- coincident with first-pile failure.
ponding to first-pile failure and corresponds to the For broadside load-cases, it turns out that an increase
situation where all effective pile resistance has been in self-weight reduces the EOTM capacity, for reasons
mobilised. If system capacity is considered it can be explained later.
seen that the target or applied EOTM envelope now To illustrate the differences observed between
falls inside the system capacity envelope and there- broadside and diagonal loading, the individual pile
fore the foundation could be considered satisfactory. responses are considered for the base-case centred
That is provided that the jacket structure could ably platform weight (of 49MN).
redistribute load, an issue which could then be inves- Broadside loading
tigated using a full non-linear structural analysis. It was seen on Figure 3 that first-pile and system fail-
ure coincide for broadside loading. Figure 5 presents
the variation of pile reactions with applied overturn-
3.3 Effect of self-weight magnitude
ing moment for this case.
In general, an increase in self-weight results in a In this case the symmetric orthogonal load causes
reduced maximum sustainable EOTM, as depicted on the piles on row 2 to fail in compression simultaneously.
Figures 4a to 4c, which are also for the considered plat- There is a near-linear proportionately between the
form example. The system capacity envelope shown in pile reactions and overturning moments. Initially, at

801

Copyright 2005 Taylor & Francis Group plc, London, UK


11.25 800 25 First pile and system failure

22.5 20
33.75 600 15

Pile Reaction MN
45 10
EOTM about y axis MNm

56.25 400 5
67.5 0
78.75 200 -5 0 50 100 150 200 250 300 350 400
-10
90
0 -15
-800 -600 -400 -200 0 200 400 600 800 -20
-200 -25
0 Degrees, C.o.g = (0,0) Total EOTM MNm
-400 A2 B2 B1
OBM Weight MN A1 0.8xComp Cap 0.8xTens Cap
11 22
33 38.5 -600
44 49.5
55 60.5 Figure 5. Pile response for broadside loading.
66 71.5 -800
77 82.5
EOTM about x axis MNm First pile failure and B2 yields
25 shedding load to A2 & B1
(a) Effect of self-weight
20
Pile Reaction MN
15
10
90
Factored On Bottom Weight MN

Wave Direction 5
80 0 0
70 11.25
22.5 -5 0 100 200 300 400 500 600
60 33.75 -10
50 45 -15
-20
40
-25
30
45 Degrees, C.o.g = (0,0) EOTM MNm
20 A2 B2 B1
10 A1 0.8xComp Cap 0.8xTens Cap
0
0 200 400 600 800 1000 Figure 6. Pile response for diagonal loading.
Total EOTM MNm
zero overturning moment, all the piles exert the same
(b) Cross section through certain wave directions reaction as the centred self-weight is equally distrib-
uted. Eventually, the row 2 piles reach their compres-
sion capacity. More overturning moment could only be
resisted if the row 1 piles were to reduce their reactions
and go into tension. However, this cannot be achieved
without losing the ability to sustain the self-weight.
Diagonal loading
Figure 6 shows the pile response for diagonal (45)
loading, also with a centred self-weight.
B2 is the most heavily-loaded pile and is the first to
reach its peak capacity. However, the EOTM can con-
tinue to increase considerably through a process of Pile
A1 going into tension, and compressive loads shed-
ding to A2 and B1. A system failure mechanism only
develops when these piles also undergo plunging fail-
ure, and both vertical and moment equilibrium can not
(c) 3D view of failure surface be simultaneously maintained. For this case, the max-
imum EOTM for diagonal loading exceeds that for
Figures 4ac. Effect of self-weight and wave direction on broadside loading, even though first-pile failure occurs
EOTM at system failure. at a smaller EOTM.

802

Copyright 2005 Taylor & Francis Group plc, London, UK


First pile failure First pile failure
25 25
20 20
System failure 15

Pile Reaction MN
15
10
Pile Reaction MN

10 System
5
failure
5 0
0 -5 0 100 200 300 400 500 600
0 100 200 300 400 500 600 -10
-5
-15 Pile reactions vary
-10
-20 with eccentric self-weight
-15 Pile reactions vary with
-25
eccentric self-weight
-20
45 Degrees, C.o.g = (-4,0) EOTM MNm
-25
A2 B2 B1
45 Degrees, C.o.g = ( +4,0) EOTM MNm A1 0.8xComp Cap 0.8xTens Cap
A2 B2 B1
A1 0.8 x Comp Cap 0.8 x Tens Cap
Figure 8. Beneficial self-weight eccentricity.
Figure 7. Detrimental self-weight eccentricity.
1000
900
3.4 Self-weight eccentricity
EOTM at system failure MNm
800
Self-weight eccentricity can reduce or increase the System Failure
700 System Failure
on Fig. 8
EOTM corresponding to first-pile failure and system 600
on Fig. 6
failure, depending on whether or not the self-weight System Failure
500
overturning moment counteracts the EOTM, and on Fig. 7
whether it is pile compression or tension failure that 400

dominates the overall response. 300


200
3.4.1 Detrimental effect of self-weight eccentricity 100
Figure 6 showed that for the centred self-weight base- 0
case Pile B2 was most heavily loaded for the diagonal -6 -4 -2 0 2 4 6
(45) load direction, but that there was a large increase X C.o.G m
in EOTM between first-pile failure and system fail- Y C.o.G = -5 m Y C.o.G = 5 m Y C.o.G = 0 m
ure. Figure 7 shows a result for a diagonal load with a
self-weight eccentricity of 4 m towards Pile Row 2. Figure 9. Influence of self-weight eccentricity.
The eccentricity increases the still-water compres-
sion load on Pile B2, which has a direct impact on the 3.4.3 Summary of eccentricity effect
EOTM corresponding to first-pile failure. Additionally, Figure 9 provides a summary illustration of the effect
since Pile A2 is also initially carrying more self-weight of self-weight eccentricity for the diagonal (45) load
load then it fails at a lower EOTM, with a consequent case. It can be seen that self-weight eccentricity ceases
reduction of about 40% in the system capacity from to have a significant effect when the eccentricity is
that seen for the centred self-weight condition (Fig. 6). such that the platform is approaching failure in the
still-water (or operating) condition, for example where
3.4.2 Beneficial effect of self-weight eccentricity the y CoG is at 5 m (Fig. 1). Thus design for the still-
If the CoG was located 4 m towards Row 1, then the water condition can actually have a dominant influence
still-water compression load on Pile B2 would initially on the system capacity under environmental loading.
be much less, consequently with a greater proportion Although it should be appreciated that these results
of its capacity being available to resist EOTM. This is are specifically related to the considered example, the
illustrated by the pile responses shown in Figure 8. total axial pile capacities have been assumed to be
Although Pile B2 is still the first-pile to fail, it does about twice the self-weight, assuming a centred CoG,
so at an EOTM that is 50% greater than for the centred which represents a common situation.
self-weight condition. Interestingly however, the
EOTM at system failure is only slightly higher. This is
because Pile B1 takes more self-weight load and is 4 SUPPLEMENTARY ISSUES
less able to assist in resisting increasing EOTM after
first-pile failure, and redistribution of load when RBFA is intentionally a simplified representation
approaching system failure is limited. of a complex situation. As such there are many

803

Copyright 2005 Taylor & Francis Group plc, London, UK


supplementary issues that can have a non-trivial In its simplest form the RBFA considers axial pile
effect in some circumstances. Some of these effects response only. A more sophisticated modification can
act to increase the apparent capacity, whilst others allow local moment fixity to be considered.
decrease it. These are briefly discussed. A P-delta effect arises due to the shift in self-weight
CoG as the platform displaces, which creates an add-
itional de-stabilising overturning moment in the same
4.1 Factored pile pre-peak stiffness
plane as the EOTM. This can be iteratively addressed
When modelling steel behaviour in pushover analyses by adjusting the self-weight CoG. However, this effect
it is common to factor both load and displacement to is likely to be small at the displacement levels corres-
retain the material stiffness. Although this could also ponding to pile failure.
be done for the pile response, it is usually less fiddly Undertaking the parametric study such as that
to factor the resistance only, perhaps with the justifi- described here using a non-linear jacket modelling-
cation that the soil rigidity index (or equivalent) is package is possible but would prove excessively time-
retained. In practice, the exact magnitude of the pre- consuming. However, when jackets have been modelled
peak pile stiffness usually has an insignificant effect for project-specific circumstances, it generally turns-
on the system behaviour. This is because the combined out that RBFA results are within about 5% of the full
load-sharing response is affected by the relative dif- non-linear platform analysis result when foundation
ferences in pile stiffnesses and also since the individ- failure governs or provides the trigger for structural
ual responses are near-linear until first-pile failure, as collapse.
shown on Figures 5 to 8.
4.5 Pile fatigue failure
4.2 Brittleness and ductility
Where platform piles have marginal fatigue lives the
For the given example, which included only a modest RBFA technique can also be used to investigate the
strain-softening effect, it is likely that the maximum effect of pile separation on the overall platform foun-
EOTM could be sustained with development of per- dation capacity. If a pile becomes detached through
manent foundation displacement and rotation. Platform fatigue failure, RBFA can be performed with a pile
stability would be regained after the applied EOTM missing and this will have the effect of shrinking the
reduces. Subsequent reloading should not significantly system failure envelope for all or most of the environ-
reduce the maximum EOTM unless, for example, the mental directions.
displacements were sufficient to lessen set-up effects.
On the other hand, irrecoverable damage to the foun-
dations can occur where the pile response is brittle, 5 CONCLUSIONS
showing significant post-peak softening, such as for
drilled and grouped piles in soft rocks. In this case This paper has outlined the method of RBFA and the
system collapse may be near-coincident with first-pile benefits it may offer as either a pre-cursor or replace-
failure, and the EOTM may be progressively reduced ment to more sophisticated non-linear pushover analy-
with each pile failure. ses. The following primary conclusions are reached:

4.3 Dynamic response The method requires simple input which is readily
available after conventional linear-elastic structural
The foregoing analyses were based on static response. analysis.
Dynamic response of rigid-body foundation systems The foundation failure envelope is platform specific,
has been studied by Stewart (1992) who noted that depending on the geometry and pile capacities.
ductile systems could withstand a significant over- For broadside load cases where foundation capacity
load above the static EOTM. This diminishes with is marginal using conventional linear-elastic analy-
increasing brittleness. Quantification of this would sis, there is likely to be little benefit in undertaking
require a non-linear dynamic analysis. more sophisticated analyses given the coincidence
of the first-pile and system failure envelopes.
4.4 Structural aspects The greatest difference between first-pilefailure and
system failure is likely to be for diagonal load-cases.
Although RBFA may overestimate the system cap- This is because more of the available pile capacity
acity of the platform, it provides a reasonable indica- is utilized at system failure for this situation.
tion of the load levels when severe structural yield may It turns out that the magnitude and position of self-
occur. In the limit, platform failure could occur through weight has most effect for broadside load-cases, with
pile bending, rather than continued plunging failure of a critical self-weight for diagonal cases below which
the piles. it has little effect.

804

Copyright 2005 Taylor & Francis Group plc, London, UK


REFERENCES Stewart, G., and van der Graaf, J.W., 1990. A Methodology
for Platform Collapse Analysis Based on Linear Super-
API RP2A LRFD, 1993. Recommended Practice for Planning position, OTC 6311.
Designing and Constructing Fixed Offshore Platforms Stewart, G., 1992. Non-linear Structural Dynamics by the
Load and Resistance Factor Design, First Edition, Pseudo-Force Influence Method Part II: Application to
American Petroleum Institute. Offshore Platform Collapse, ISOPE Conference.
DNV-Sintef-Bomel 1999. Best Practice Guidelines for use Toolan, F.E., and Horsnell M.R., 1992. Institution of Civil
of non-linear analysis methods in documentation of ultim- Engineers. Numerical methods in offshore piling.
ate limit states for jacket type offshore structures. Ultiguide.

805

Copyright 2005 Taylor & Francis Group plc, London, UK


Soil characterization for consistent reliability in the Load and
Resistance Factor Design of pile foundations

K.C. Foye & R. Salgado


Purdue University, West Lafayette, Indiana, USA

ABSTRACT: Load and Resistance Factor Design (LRFD) is gaining acceptance as the method to replace the
traditional Working Stress Design (WSD) approach to pile foundation design. The key improvements of LRFD
over WSD are the ability to provide a more consistent level of reliability and the possibility of accounting for
load and resistance uncertainties separately. In order for LRFD to fulfill its promise for designs with more consis-
tent reliability, the methods used to execute a design must be consistent with the methods assumed in the devel-
opment of the LRFD factors. In this paper, a methodology for the estimation of soil parameters for use in design
equations is proposed that should allow for more statistical consistency in design inputs than is possible in tra-
ditional methods. The effectiveness of this simple soil characterization tool is evaluated using reliability analysis.

1 INTRODUCTION design parameters must be accounted for at the time


the reliability analysis is done, and that soil parame-
Geotechnical engineers are beginning to use Load ters for use in a design equation, to calculate a con-
and Resistance Factor Design (LRFD) as a design servatively assessed mean (CAM) value of resistance,
method for foundations. The advantage of LRFD over must be determined in a reproducible way that is con-
traditional Working Stress Design (WSD) is the abil- sistent with the resistance factor. This is a crucial issue
ity to account for uncertainties in load and resistance among several that must be addressed before reliability-
separately. The basic load and resistance factor design based design methods, such as LRFD, reach their
inequality is full potential in geotechnical design (Becker 1996,
Kulhawy & Phoon 2002).
(1) Various statistical procedures have emerged to
address the task of conservatively assessing a design
soil profile. In this paper, we outline a simple method
where RF is the resistance factor, Rn is the nominal to find a CAM. Also, we show how this procedure is
(unfactored) resistance, (LF)i is the load factor multi- useful in improving the repeatability of a design soil
plying load Qi for a particular load combination. Foye profile and how it can lead to designs with more con-
et al. (2005b) demonstrate a method to find values of sistent reliabilities. We use the Cone Penetration Test
RF for foundation design using reliability analysis. (CPT) to illustrate the use of this procedure.
Reliability analysis and the resulting RF values are tools
to assess the impact of various uncertainties on design.
Regardless of the design method used, designers 2 A METHOD TO CONSISTENTLY
interpret a number of in-situ or laboratory tests to DETERMINE CHARACTERISTIC
assemble a design soil profile for foundation design. RESISTANCE
Usually, information from a variety of sources is used
to make judgments about the most likely applicable The following steps are recommended as a method to
values for design. To be safe, engineers conservatively consistently determine CAM soil profiles. We are tar-
assess these parameters to account for scatter in the geting the interpretation of CPT results specifically,
results, testing irregularities, and concerns about the but other tests can also be considered using this method.
accuracy of the measurements.
When LRFD factors are developed using the frame-
Step 1. Group relevant data together
work of reliability analysis, load factors, resistance
factors, and the characteristic resistance are inter-linked. In the first step, several CPT soundings, as well as
This means that the methods used to determine the any soil sampling data, are compared to decide what

807

Copyright 2005 Taylor & Francis Group plc, London, UK


soil types are present at the site and where they are where X is the standard deviation of some random
represented in the various soundings. For each CPT variable X and X is the mean of the same variable.
performed, the individual layers of soil are identified. For the case of our CPT profile, notice that each CPT
Considering sand, a layer is defined in this context as measurement provides a data pair (qc,i, zi) where z is the
a volume of soil with approximately the same relative depth. We can detrend and normalize each data pair
density. Soundings are compared and the data are com- (qc,i, zi):
bined where measurements are being taken within the
same soil layer. Often, when CPT soundings are accom-
panied by geophysical data, geotechnical engineers (4)
can readily assess cases where CPT measurements are
made in the same soil layer. Geophysical tests are valu-
able tools to estimate soil layering and uniformity at where wi are instances of random variable W, which
an offshore site. represents the scatter of data points about the mean
Note that this procedure is not substantially different trendline. Next, the mean of these w values (w) is
from current practice. Geotechnical engineers routinely estimated
make judgments about values determined in in-situ test-
ing. Estimates of design parameters from these testing
logs are supported by information from all the avail- (5)
able tests, not just the sounding under consideration.
By combining data from the same soil layer, we are
building support for CAM values in the same way. where n is the number of data points considered. In the-
ory, if function f(z) defines precisely the mean of the
Step 2. Determine trends in data with depth relationship between qc and z, and a large number of
data pairs (qc,i, zi) are available, w  0. The COV rep-
Many soil properties vary with depth, even within the resenting the uncertainty of f(z) is found by applying
same soil layer. For instance, CPT qc measurements the sample standard deviation formula to the w values:
follow a power function with depth for the same value
of relative density as seen in the results from Salgado
(2002):
(6)
(2)
Note that since we have divided by f(z) in (4), the
standard deviation (Equation 6) of W is the COV rep-
where pA is a reference stress of 100 kPa, DR is rela- resenting the uncertainty of our trend line f(z) follow-
tive density, h is the lateral effective stress, and c1, ing the definition in Equation (3).
c2, and c3 are coefficients related to intrinsic sand The use of Equations (4) to (6) is one technique to
properties. The equation can also be rewritten with analytically determine the scatter in measurements that
depth in place of h. have a trend with depth. The N method, discussed by
In many cases, small portions of the CPT profile Foye et al. (2005a) can expedite assessments of the
will roughly conform to a linear trend. Thus, linear COV of these trended data. The first step is to observe
regression is a simple and powerful tool to quickly the bounds and mean trend of the data. An example
determine mean trends with depth. Data for each soil mean trend and data bounds are illustrated in Figure 1
layer can be analyzed separately to build a complete for qc. For a particular depth, the value of the mean and
profile of trend lines. In the case of the CPT, the result the range (difference between minimum and maximum
of these linear regressions will be a profile of mean qc bound values) can be computed. The standard devia-
as a function f(z) of depth z. tion is then found using
Step 3. Assess the scatter of data about the trend
(7)
Once we find the trend of qc values with depth, the next
step is to quantify the scatter of actual qc measurements
about this trend. We use the coefficient of variation where  is the standard deviation. Values of N for
(COV) to quantify this scatter. The coefficient of vari- different values of n are tabulated in Table 1. Table 1
ation is defined as is derived from work by Tippett (1925). It is applic-
able to sets of normally distributed data for which the
(3) number of data points is limited, the range of data
is known, and the average standard deviation of the

808

Copyright 2005 Taylor & Francis Group plc, London, UK


8 12 16 20 24 8 12 16 20 24
3 3 3 3
upper bound on data CAM trendline,
upper sand layer
4 4 4 4
mean trend
depth (m)

5 5

depth (m)
5 5

6 6 6 6

CAM trendline,
7 7 7 lower sand layer 7
lower bound on data
8 8 8 8
8 12 16 20 24
8 12 16 20 24
qc(MPa)
qc(MPa)
Figure 1. Example CPT qc profile with mean trend and
range lines drawn. Mean trend fit using Equation (2). Figure 2. Visual approximation of CAM function for a
CPT profile The trend lines are drawn so that 80% of the
data points occur to the right of the line.
Table 1. Values of sample range for   1
(N) on the normal distribution (after Tippett
1925). are many possible criteria to select CAM values (for
instance, the 95% exceedance criterion of ACI 1999).
n N n N For the purposes of this paper, we choose an 80%
exceedance criterion. Using this criterion, and assum-
2 1.128379 17 3.587886 ing the scatter of property values about the mean to be
3 1.692569 18 3.640066 normally distributed, each CAM value can be deter-
4 2.058751 19 3.688965 mined as a value 0.84 standard deviations less than
5 2.325929 20 3.734952 the mean:
6 2.534413 50 4.498153
7 2.704357 100 5.0152
8 2.847201 200 5.492108 (8)
9 2.970027 300 5.755566
10 3.077506 400 5.936396
11 3.172874 500 6.073445 By applying (8) to the entire CPT profile, we can
12 3.258457 600 6.183457 quickly find our CAM profile.
13 3.335982 700 6.275154 When a large number of data points is available,
14 3.406765 800 6.353645 the procedure can also be approximated visually.
15 3.471828 900 6.422179 Figure 2 illustrates an example where the CAM line
16 3.531984 1000 6.482942 for an approximately linear qc profile can be drawn
visually such that 80% of the data points lie above the
CAM line.
population based on this data sample is sought. The
COV can then be computed using Equation (3). It is
possible that for some geotechnical quantities, the COV 3 ASSESSING THE USEFULLNESS OF
varies with the mean value or with depth. In these cir- THE CAM PROCEDURE
cumstances, it is conservative to select the greatest
computed COV to represent the scatter in the data. Two potential advantages of the CAM method out-
lined above are considered: repeatability and con-
sistent reliability. To examine the repeatability of the
Step 4. Decrease the mean trend to find the
method, we consider the case of six CPT soundings
CAM trend
over an onshore building site on a profile of uniform
The last step is to find the CAM trend by decreasing sand layers. Each of these six soundings is encountering
values of qc in our mean trend by some amount. There the same sand layers at about the same depths. Thus,

809

Copyright 2005 Taylor & Francis Group plc, London, UK


all of these soundings are relevant measures of the same where LL is the live load and DL is the dead load act-
soil and can be combined for statistical treatment. In ing on the pile base. Table 3 contains the Probability
this example, the soundings are combined in random Density Functions (PDFs) used to describe the uncer-
order to simulate a condition where we do not know the tainty of each of the variables in Equation (10).
outcome of the CPT in advance. To compute the reliability of individual design cases,
Figure 3 parts (a)(f) show the combined CPT qc the method by Low & Tang (1997) is used to compute
profiles, their mean trendlines, and CAM trendlines. the reliability index . The reliability index is a meas-
Starting with part (a), one additional CPT qc profile is ure of the relative reliability of a design. Higher values
added to the plot until all 6 soundings are included in of indicate more reliable designs (lower probabil-
part (f). ities of failure) against a particular failure mode. In
To investigate the influence of additional soundings this case, we are considering the end-bearing of a pile.
on a hypothetical final design, the shaft and end bearing A value equal to 3.0 has often been advocated as
capacity of a driven 18( (457 mm) diameter pipe pile corresponding to an adequate level of safety in geo-
is considered. In this example, the pipe is to be driven to technical design (Paikowsky 2004). We can compute
a depth of 28 m to bear in a relatively dense sand layer. values of cb, qc, Ab, LL, and DL corresponding to this
Computed values of shaft and base resistance level of safety. Two sets of values are output from this
using the different CAM qc profiles appears in Table 2. reliability analysis, limit state (LS) values and nom-
For the purposes of this example, the unit shaft resist- inal design values. Limit state values satisfy Equation
ance is computed as a constant value 0.002 multiplied (10) and indicate the most likely values of these vari-
with values of qc from the CAM qc profile (Lee et al. ables that will lead to failure. Nominal design values
2003). Base resistance is computed using an average are the variable values that are used during design,
of CAM qc values near the pile base. Also appearing and are set as inputs to the reliability analysis. To sat-
in Table 2 is the incremental change in computed isfy our target reliability index of 3.0, the nominal
resistance as CPT soundings are added to the CAM design values are adjusted until the reliability analysis
profile. There is a slight decrease in computed resist- indicates  3.0. To allow the use of (1) in design,
ance observed. This decrease can be attributed to the values of (LFi)* and RF* can be computed as
increased scatter in data resulting from combining
CPT logs. However, this decrease is relatively small
compared to usual tolerances in geotechnical design. (11)
This example highlights how relatively insensitive
the CAM method is to the number of CPT soundings
included in the analysis. This is an important result
since a single profile is to be assigned to the sand for (12)
design purposes, regardless of the number of sound-
ings. This result does not suggest that fewer borings are
advisable. However, it does suggest that when we are where subscript LS denotes limit state values and sub-
able to identify individual, relatively uniform soil layers, script n denotes the nominal design values.
perhaps through experience or geophysical surveys, we Since we desire values of resistance factor that are
can arrive at consistent and repeatable design values consistent with code-prescribed load factors, we find
using any number of CPT soundings in these layers. an adjusted resistance factor RF
To assess the impact of the CAM method on the
reliability of a design, we need to perform reliability
analyses. The following discussion draws upon con- (13)
cepts presented in Foye et al. (2005b). For this example,
we consider a hypothetical CPT design method for
pile base resistance. In this method, nominal pile base where (LF)DL and (LF)LL are the code-prescribed load
resistance Rn is computed using factors. Note that this equation is conservative for all
live-to-dead load ratios. Another purpose for finding
adjusted resistance factors is so that we can compare
(9) them apart from the load factors. In this example, we
take (LF)DL  1.2 and (LF)LL  1.6 (ASCE-7 1996).
where cb is a pile base resistance factor and Ab is the Figure 4 plots values of RF versus COVqc, deter-
pile base area. Considering a live-load plus dead-load mined when a mean profile is used to find qc (step 2)
case, a limit state to be avoided is or when CAM values are used. We are conservatively
assuming the COV of qc to equal that found for the
scatter in the CPT profile according to step 3. Notice
(10) in the figure that the value of RF required to satisfy

810

Copyright 2005 Taylor & Francis Group plc, London, UK


0 20 40 60
0 10 20 30 40 50
0 0
0 0
data
data mean line
mean line CAM line
CAM line
10 10
10 10

Depth (m)
Depth (m)

20 20 20 20

30 30 30 30

0 10 20 30 40 50 0 20 40 60
qc (MPa) qc (MPa)
(a) CPT sounding 5 (d) CPT soundings 5, 2, 1, and 3
0 20 40 60
0 20 40 60
0 0
0 0
data
data mean line
mean line CAM line
CAM line
10 10 10 10
Depth (m)
Depth (m)

20 20 20 20

30 30 30 30

0 20 40 60 0 20 40 60
qc (MPa) qc (MPa)
(b) CPT soundings 5 and 2 (e) CPT soundings 5, 2, 1, 3, and 4

0 20 40 60
0 20 40 60
0 0 0 0

data data
mean line mean line
CAM line CAM line

10 10 10 10
Depth (m)
Depth (m)

20 20 20 20

30 30 30 30

0 20 40 60 0 20 40 60
qc (MPa) qc (MPa)
(c) CPT soundings 5, 2, and 1 (f) CPT sounding 5, 2, 1, 3, 4 and 6

Figure 3. Six CPT soundings through uniform sand layers with mean and CAM trendlines.

811

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 2. Shaft and base resistance computed using the combined CPT profiles in Figure 3.

Nominal shaft resistance Nominal base resistance

Incremental Incremental
Number of CPTs Capacity (MN) change Capacity (MN) change

1 0.376 4.035
2 0.358
4.68% 4.127 2.30%
3 0.369 3.14% 4.167 0.97%
4 0.337
8.69% 4.212 1.06%
5 0.342 1.29% 4.154
1.37%
6 0.300
12.17% 4.162 0.19%

Table 3. Probability Density Functions (PDFs) used 0.06 0.08 0.1 0.12 0.14 0.16
to describe example design method uncertainties.

Variable PDF type COV X/X 1n 0.48 0.48

DL normal 0.15 1.05


0.44 0.44
LL lognormal 0.25 1.15 Adjusted RF
cb lognormal 0.3 1
qc normal variable2 variable2 0.4 0.4
Ab lognormal 0.02 1
1
this is a correction used to compute the statistical 0.36 0.36
mean (X) of a variable from its nominal design computed using CAM
value (Xn). Details of this correction appear in Foye computed using mean
et al. (2005a). 0.32 0.32
2
for the purposes of this example, we are varying
qc COV to represent different hypothetical soil
profiles. Values of qc COV are assumed to equal the 0.06 0.08 0.1 0.12 0.14 0.16
qc COV
value of COV found using Step 3. Correction X/Xn
is found by dividing the mean found using Step 2 by
the CAM value found using Step 4. Figure 4. Adjusted Resistance Factor (RF) computed when
CAM or mean CPT profiles are used as input to design.

our  3.0 safety criterion is relatively stable when


CAM values are used in design compared with the design values for two reasons: first, and most import-
case when mean values are used. It is not practical to antly, it provides a statistically consistent method to
have values of RF that change depending on slight analyze data from a particular soil layer, replacing
variations in site conditions. Thus, it is more practical arbitrary selection with a consistent procedure; sec-
to use a CAM procedure to estimate design input ond, the CAM procedure tends to stabilize the reli-
parameters. Additionally, our research shows that an ability of design checks completed using particular
80% exceedance CAM provides the most stable RF values. This method does not replace the engi-
results for foundation design. CAM values based on neers responsibility to determine which data are rele-
exceedance criteria greater than 80% become difficult vant to the design problem, but rather supplements
to evaluate consistently using limited data and can indi- the tools available to analyze them. The method also
cate physically unrealistic soil parameters. Values based agrees with engineers intuition to try to incorporate
on exceedance criteria less than 80% may not suffi- as much available information as possible into design
ciently account for the variabilities typically associ- decisions. Finally, the method outlined in this paper is
ated with soils. a simple technique that parallels materials testing
acceptance criteria already in use in other disciplines.
As geotechnical engineers move towards the use
4 CONCLUSION of rationally-developed LRFD methodologies, tech-
niques to arrive at statistically consistent input values
In summary, the conservatively assessed mean will become valuable tools to guide factor develop-
(CAM) procedure is a valuable tool in selecting ment and allow a gradual transition in design practice.

812

Copyright 2005 Taylor & Francis Group plc, London, UK


REFERENCES Foundation Design Codes and Soil Investigation in view
of International Harmonization and Performance Based
ACI. 1999. Building Code Requirements for Structural Con- Design, Hayama, Japan. 3148.
crete (318-99) and Commentary (318R-99). American Lee, J., R. Salgado, and K. Paik. 2003. Estimation of Load
Concrete Institute, Detroit. Capacity of Pipe Piles in Sand Based on Cone Penetration
Becker, D. E. 1996. Eighteenth Canadian Geotechnical Col- Test Results. Journal of Geotechnical and Geoenviron-
loquium: Limit States Design for Foundations. Part I. An mental Engineering, ASCE. 129(6). 391403.
overview of the foundation design process. Canadian Low, B. K., and W. H. Tang. 1997. Efficient Reliability
Geotechnical Journal. 33(6). 956983. Evaluation Using Spreadsheet. Journal of Engineering
Foye, K. C., R. Salgado, and B. Scott. 2005a. Assessment Mechanics. 123(7). 749752.
of Variable Uncertainties for Reliability-Based Design Paikowsky, Samuel G. 2004. Load and Resistance Factor
of Foundations. Journal of Geotechnical and Geoenviron- Design for Deep Foundations. NCHRP Report 507.
mental Engineering. ASCE. accepted for publication. Washington, D.C.: Transportation Research Board.
Foye, K. C., R. Salgado, and B. Scott. 2005b. Resistance Salgado, Rodrigo. 2002. CONPOINT Beta Version: Users
Factors for Use in Shallow Foundation LRFD. Journal of Manual. Purdue University. West Lafayette, Indiana.
Geotechnical and Geoenvironmental Engineering. ASCE. http://bridge.ecn.purdue.edu/rodrigo/research/
accepted for publication. theory01.html.
Kulhawy, F. H, and K. K. Phoon. 2002. Observations on Tippett, L. H. C. 1925. On the Extreme Individuals and the
Geotechnical Reliability-Based Design Development in Range of Samples Taken from a Normal Population.
North America. Proceedings. International Workshop on Biometrika. 17(3/4). New York: Macmillan. 364387.

813

Copyright 2005 Taylor & Francis Group plc, London, UK


Buckling considerations in pile design

S. Bhattacharya, T.M. Carrington & T.R. Aldridge


Fugro Limited, United Kingdom

ABSTRACT: Buckling instability is one of the more destructive forms of pile failure. Buckling of piles
can be classified into two groups; (a) Global buckling, where a part or full length deforms longitudinally
as in Eulers buckling of unsupported struts; (b) Local buckling where the cross-section of the pile deforms
and the damage is localised. Global buckling is currently considered in design where piles are partially
exposed or driven in extremely soft soil or during installation under driving stresses. Recent studies have
shown that fully embedded end-bearing piles passing through saturated loose to medium dense sand can
buckle if the surrounding soil liquefies in an earthquake. There have been a number of cases where offshore
piles have collapsed during driving due to progressive closure of the internal dimensions the initiat-
ing mechanism being local buckling. This paper summarizes the different cases where buckling should be
considered in pile design. Mechanisms of collapse of offshore piles by local buckling are discussed in a
companion paper.

1 INTRODUCTION 1.2 Limit State of Collapse and Limit State of


Serviceability
1.1 Buckling as a mode of failure
The failure of piled foundations can be classified into
Buckling instability is one of the more destructive two groups:
forms of pile failure. It is sudden and is the cause
(a) Structural failure of the pile whereby the load
of failure of many, if not most structures. The import-
carrying capacity of the foundation drops, see for
ance of buckling instability in structural design
example Figure 1. The figure shows plastic hinges
cannot be underestimated. McRobie (2002) in his
formed in the piles during the 1964 Niigata earth-
introductory lecture on buckling to undergradu-
quake. The fundamental failure mechanisms that
ates states; If you ever intend to design a structure,
do not even think of skipping these (buckling) lec-
tures. This form of failure mechanism dominates
the design of slender members carrying substantial
axial loads. Piles are slender members normally used
to transfer the axial load of the superstructure to
the deep bearing strata. Bond (1989) collated embed-
ded lengths and diameters of piles used in practice.
The study shows that the length to diameter ratio
of piles ranges between 25 and 100. These can be
considered as slender columns, in the absence of soil
support.
Buckling of piles is currently considered in pile
design under the following headings:

1 Partially exposed piles, as in jetties or offshore


platforms where part of the pile is in water or air. Figure 1. Structural failure of piles by forming plastic
2 Piles in very soft soil (clay). hinges Hamada (1992). A piled foundation that collapsed
3 During pile installation by driving. during the 1964 Niigata earthquake.

815

Copyright 2005 Taylor & Francis Group plc, London, UK


can cause plastic hinge formation in a pile are shear reviewed with respect to the current understanding.
failure, bending failure and buckling failure. The A simplified design graph is recommended to avoid
above three forms of failure are often known as global buckling of piles in liquefiable soils.
LIMIT STATE OF COLLAPSE. It must be men-
tioned that each of these types of failure can cause
a complete collapse of the foundations. 2 REVIEW OF CODES OF PRACTICE FOR
(b) Failure by excessive settlement. Often the settle- PILE DESIGN
ment of piled foundations exceeds the acceptable
limits of the structure, which is essentially SER- This section of the paper reviews the design guide-
VICEABILITY LIMIT STATE. In this type of fail- lines against buckling of piles in some of the most
ure, the piles may not fail structurally. used codes of practice.
This paper deals with the buckling aspect of Limit
State of Collapse.
2.1 Eurocode 7 and Part 5 of Eurocode 8
Eurocode 7 (1997) suggests that:
1.3 Structural design of piles Slender piles passing through water or thick deposits
Structurally most piles are designed against bending of very weak soil need to be checked against buckling.
failure due to lateral loads. The semi-empirical P-y This check is not normally necessary when piles are
concept is normally used to design the piles. However, completely embedded in the ground unless the char-
this approach cannot be applied if buckling under axial acteristic undrained shear strength is less than 15 kPa.
loading is a possibility for the member under consid- For design of piles in seismic areas, Eurocode 8
eration. These considerations would lead to the fact advises designers to design against bending due to iner-
that, if part of the pile loses lateral support during its tia and kinematic forces arising from the deformation
design period, the pile should be treated as unsup- of the surrounding soil. It says:
ported column. The structural design of the pile in the Piles shall be designed to remain elastic. When this
unsupported zone should be designed as a column is not feasible, the sections of the potential plastic
carrying lateral loads. hinging must be designed according to the rules of
A recent investigation, Bhattacharya et al. (2004), Part 13 of Eurocode 8.
has revealed that fully embedded end bearing piles Eurocode 8 (Part 5) also says:
passing through loose to medium dense sand can Potential plastic hinging shall be assumed for:
buckle under the axial load alone if the surrounding soil a region of 2d from the pile cap
liquefies in an earthquake. Buckling of fully embedded a region of 2d from any interface between two
piles in extremely soft clay is known, but should also be layers with markedly different shear stiffness (ratio of
considered in loose to medium dense sand in liquefi- shear moduli 6)
able areas. This approach should be applied equally to where d denotes the pile diameter. Such region shall
earthquake or wave induced liquefiable soils. be ductile, using proper confining reinforcements.

1.4 Purpose of this paper 2.2 American Petroleum Institute (API)


This paper aims to list the cases where buckling Clause 3.3.1.b of API (2000) recommends the
should be considered in design. Buckling of piles has following:
been subdivided into two groups: Column buckling tendencies should be considered
for piling below the mudline. Overall column buck-
(a) Local buckling, where the transverse section of
ling is normally not a problem in pile design, because
the pile deforms. In practice, this is often observed
even soft soils help to inhibit overall column buckling.
at the pile tip.
However, when laterally loaded pilings are subject to
(b) Global buckling, like Eulers buckling of an unsup-
significant axial loads, the load deflection (P-)
ported strut, where the longitudinal section of the
effect should be considered in stress computations. An
pile deforms.
effective method of analysis is to model the pile as a
Checking against local buckling is crucial for thin beam column on an elastic foundation.
walled sections and is an important consideration dur- Clause 6.10.2 of API (2000) states:
ing the installation of piles, particularly when driving General column buckling of the portion below the
into extremely hard soil or rock. A companion paper, mudline need not be considered unless the pile is
Aldridge et al. (2005) in this symposium deals with pile believed to be laterally unsupported because of
tip damage. Therefore, this paper does not address local extremely low soil shear strengths, large computed
buckling. The codes of practice for pile design are lateral deflections, or for some other reason.

816

Copyright 2005 Taylor & Francis Group plc, London, UK


API (2000) considers stresses in a pile during driv- 3 WHERE BUCKLING IS IMPORTANT
ing. The code advises designers to have a minimum
pile wall thickness to avoid local buckling. The rec- 3.1 Pile as a beam-column
ommendations are:
A pile can be best described as a beam-column i.e. a
For piles that are to be installed by driving where
column section carrying lateral loads. A general equa-
sustained hard driving is anticipated, the minimum
tion can be described following Heelis et al. (2004).
piling wall thickness used should not be less than

(1)

(2)
where
t  wall thickness in mm, where
D  diameter, in mm. EI  Flexural rigidity of the pile;
P0  External axial compressive force applied at
2.3 Japanese Road Association code the top of the pile i.e. x  0
(JRA 1996) f(x) is the friction per unit length
k(x) is the modulus of subgrade reaction.
The guidelines for designing piles in liquefiable soils The above equation suggests that if part of the soil
are shown in Figure 2. The code advises practicing surrounding the pile loses its effective stress, then
engineers to design piles against bending failure due f(x)  0 and k(x) will be near zero, and the equation
to lateral loads arising out of inertia or slope move- reduces to Eulers buckling equation. The theoretical
ment (lateral spreading). The code discourages the buckling load can be estimated by equation 3.
additions of effects due to inertia and lateral spread-
ing. To check against the bending failure due to lateral
spreading, the code recommends that the non-liquefied (3)
crust above the liquefied soil exerts passive pressure
(qNL in Fig. 2) and the liquefied soil offers 30% of the
total overburden pressure (qL in Fig. 2). where Leff  Effective length of the pile in the unsup-
Eurocode 8 (1998), JRA (1996) focus on bending ported zone. This depends of the boundary condition
strength and omit considerations of the bending stiff- of the pile below and above the support loss zone, see
ness necessary to avoid buckling in the event of soil Bhattacharya et al. (2004).
liquefaction. API (2000) code does consider column
buckling, but only for soils having low shear strength,
i.e. soft clay. The following sections point out that buck- 3.2 Role of lateral load in buckling
ling needs to be considered even for fully-embedded Rankine (1866) recognized that the failure load of
piles passing through loose to medium dense sand structural columns predicted by equation 3 is more
where there may liquefy for any reason. than the actual failure load (PF) i.e. equation 3 is
unconservative. This is because buckling is very sen-
sitive to imperfections and lateral loads. The collapse
also involves an interaction between elastic and plas-
tic modes of failure. Lateral loads and geometrical
imperfections both lead to the creation of bending
moments in addition to axial loads. Bending moments
have to be accompanied by stress resultants that dimin-
ish the cross-sectional area available for carrying the
axial load, so the failure load PF is less than the plastic
squash load (PP) given by A. y (A  area of the pile
section, y is the yield stress of the material). Equally,
the growth of zones of plastic bending reduces the
effective elastic modulus of the section, thereby
reducing the critical load for buckling, so that PF Pcr.
Furthermore these processes feed on each other, as
explained in Horne & Merchant (1965). As the elastic
critical load is approached, all bending effects are
magnified. If lateral loads in the absence of axial load
Figure 2. Japanese Roadways Association (JRA) code. would create a maximum lateral displacement 0 in the

817

Copyright 2005 Taylor & Francis Group plc, London, UK


critical mode-shape of buckling, then the displacement 2 Initial imperfection or lack of straightness. Figure 5
 under the same lateral loads but with a co-existing shows a pile attached to a towing bollard in an off-
axial load P is given by: shore pile installation. This creates an initial eccen-
tric moment.
3 Loss of lateral support due to liquefaction or scour.
(4) Recent investigation has shown that fully embed-
ded end bearing piles passing through saturated,
loose to medium dense sand can buckle under the
axial load alone if the surrounding soil liquefies in
The same magnification factor applies to any initial an earthquake. The stress in the pile section will
out-of-line straightness of the pile in the mode shape initially be within the elastic range, and the buckling
of potential buckling. Correspondingly, all curvatures length will be the entire length of the pile in liquefied
are similarly magnified and so are the bending strains soil. Figure 6 shows a failure of a fully embedded
induced in the column by its lateral loads or eccen- pile by buckling in a centrifuge test.
tricities. The progression towards plastic bending fail- 4 Partially exposed pile. This is often encountered in
ure is accelerated as axial loads approach the elastic jetties or offshore platforms.
critical load (Pcr). Not only do axial loads induce extra 5 Piles in extremely soft clay. Buckling of slender
bending moments (P- effects), but the full plastic steel piles in soft, quick clay in Trondheim (Norway)
bending resistance cannot be mobilized due to the has been reported by Brantzaeg & Elvegaten (1957).
fact that part of the pile section is required to carry
the axial loads. Equation 4 indicates that for a column
carrying an axial load of half its Euler load, that lat-
eral displacements and therefore bending moments
would be 1/(1
0.5) or 100% bigger than those cal-
culated ignoring axial load effects. This is important
if significant lateral loads must also be carried.

3.3 List of cases where buckling is important


The cases where buckling needs special attention are
listed below:
1 During installation by driving. The stability of
slender piles during driving has been dealt with by
Burgess (1976). This is also a design consideration
in offshore installations, see Figures 3 and 4. Figure
3 shows a typical offshore installation and Figure 4
shows the pile stick up. Once the pile is in the sleeve,
it is important to check the buckling potential under
the action of the lateral forces due to the wave Figure 4. Pile stick up.
loading and the hammer weight.

Figure 3. A typical offshore pile installation. Figure 5. Attachments at the bottom of the pile.

818

Copyright 2005 Taylor & Francis Group plc, London, UK


6 Buckling of pile due to dredging operation in a (column carrying lateral loads) element with bi-axial
marine harbour. It has been reported by Sovinc bending. If the section of the pile is a long column,
(1981) that a piled marine harbour was seriously analysis would become extremely complex and an
damaged during a dredging operation due to soil explicit closed-form solution does not exist. The solu-
movement. tion of such a problem demands an understanding of
7 Local buckling at the sleeve during driving. the way in which the various structural actions inter-
8 Local tips buckling due to faulty shoe design. act with each other i.e. how the axial load influences
the amplification of lateral deflection produced by the
lateral loads. In the simplest cases i.e. when the sec-
4 SIMPLIFIED APPROACH TO AVOID tion is a short column, the superposition principle
GLOBAL BUCKLING OF PILES can be applied i.e. direct summation of the load
effects. In other cases, careful consideration of the
As mentioned earlier, the part of the pile in liquefiable complicated interactions needs to be made.
soil should be treated as an unsupported column. A pile Designing such a type of member needs a three-
not only has axial stress but also may have bending dimensional interaction diagram where the axes are:
stresses in two axes due to the lateral loads. The pile Axial (P), major-axis moment (Mx) and minor-axis
represents a most general form of a beam-column moment (My). The analysis becomes far more com-
plicated in presence of dynamic loads. The above
complicated non-linear process can be avoided by
designing the section of the pile as a short column
i.e. for concrete section length to least lateral dimen-
sion less than 15 (British Code 8110) or a slenderness
ratio (effective length to minimum radius of gyration)
less than 50.
Figure 8 shows the study of 14 reported case histor-
ies of pile foundation during earthquakes, after
Bhattacharya (2003) and Bhattacharya et al. (2004).
The case histories were from four different earth-
quakes. Six of the piled foundations survived while
others suffered severe damage. Essentially, it is
assumed that the pile is unsupported in the liquefiable
zone. For each of the case histories, the Leff of the pile
in the liquefiable region is plotted against the min-
imum radius of gyration (rmin) of the pile. rmin is intro-
duced to represent piles of any shape (square, tubular,
circular) and is given by I/A where I is the second
Figure 6. Buckling of a fully embedded pile in a centrifuge moment of area; and A is the cross sectional area of
test, after Bhattacharya (2003). the pile section. For a solid circular section, rmin is

Figure 7. Failure of Adriatic harbour during dredging operation, Sovinc (1981).

819

Copyright 2005 Taylor & Francis Group plc, London, UK


0.8 kept as short column i.e. for concrete section
length to least lateral dimension less than 15 or a slen-
0.7
derness ratio (effective length to minimum radius of
0.6 Good gyration) less than 50.
0.5 performance The main assumptions in developing the design
chart are:
(rmin) m

0.4 Poor
performance 1 The piles are either solid concrete section having E
0.3
(Youngs Modulus) of 22.5 GPa or steel tubular
0.2 section having E of 210 GPa.
0.1 2 The piles are not in a single row and at least in
2  2-matrix form this ensures that the pile heads
0 are restrained against rotation but free to translate.
0 10 20 30 40 50
3 The thickness of the steel pile is based on
Effective length (Leff) m
equation (1).
Figure 8. Study of 15 case histories, Bhattacharya et al.
(2004).
5 CONCLUSIONS

Minimum dia of pile from buckling consideration Buckling of pile can be classified into two groups:
2.25 global buckling and local buckling. In global buck-
2 Concrete pile ling, the pile deforms longitudinally leading to lateral
Diameter of pile (m)

1.75
1.5
Steel tubular pile instability of the entire structure. On the other hand in
1.25
local buckling, the cross section of the pile deforms
1 leading to a localized damage. In either case, the load
0.75 carrying capacity of the pile reduces drastically and
0.5 may lead to complete collapse of the foundation. Eight
0.25 cases have been listed where buckling is a design con-
0 sideration.
4 6 8 10 12 14 16 18 20
Global buckling should be considered for fully
Thickness of liquefiable layer (m)
embedded end-bearing piles passing through loose to
Figure 9. Minimum diameter to avoid buckling of piles,
medium dense where they may liquefy for any reason.
Bhattacharya and Tokimatsu (2004). This can be avoided by reducing the slenderness ratio
of the pile in the likely-to-be-unsupported zone. A
simplified approach to avoid buckling under such situ-
ations has been described.
0.25 times the diameter of the pile and for a hollow
circular section rmin is 0.35 times the outside diameter
of the pile. Leff is dependent on the thickness of the
liquefiable zone, depth of embedment and the fixity REFERENCES
at the pile head. In the figure, a line representing a
slenderness ratio of 50 could differentiate the good Aldridge, T.R, Carrington, T.M and Kee, N.R. 2005.
Propagation of pile tip damage during installation,
performance piles from the poor performance. It is
Proceedings of ISFOG 2005, Australia.
worthwhile to note that slenderness and buckling dif- API 2000. Recommended practice for Planning, Designing
ferentiated between the good and poor performance and Constructing Fixed Offshore Platforms Working
irrespective of whether the ground surface was sloped Stress Design. American Petroleum Institute.
or not. Thus the study shows that piles should be Bhattacharya, S. 2003. Pile instability during earthquake
designed as short columns, i.e. large diameter piles liquefaction, PhD Thesis; University of Cambridge (U.K).
are better. Bhattacharya, S., Madabhushi, S.P.G and Bolton, M.D.
Figure 9 shows a typical graph showing the min- 2004. An alternative mechanism of pile failure in liquefi-
imum diameter of pile necessary to avoid global buck- able deposits during earthquakes, Geotechnique 54,
No 3, pp 203213.
ling depending on the thickness of the liquefiable
Bhattacharya and Tokimatsu 2004. Essential criteria for design
soils following Bhattacharya & Tokimatsu (2004). If of piled foundations in seismically liquefiable areas,
the diameter of a pile is chosen based on Figure 9, Proceedings of the 39th Japan National Conference on
then non-linear P- analysis can be avoided and the Geotechnical Engineering, Niigata, 7th to 9th July 2004.
lateral load amplification effects, explained in section Bond, A.J. 1989. Behavior of displacement piles in over-
3.2 are minimal. Essentially, the section of the pile is consolidated clays, PhD Thesis, Imperial College (U.K).

820

Copyright 2005 Taylor & Francis Group plc, London, UK


Brandtzaeg, A. and Elvegaten, E.H. 1957. Buckling tests of Heelis, M.E., Pavlovic, M.N. and West, R.P. 2004. The ana-
slender steel piles in soft, quick clay, Proceedings of the lytical prediction of the buckling loads of fully and par-
4th International Conference on Soil Mechanics and tially embedded piles, Geotechnique 54, No 6, pp 363373.
Foundation Engineering (ICSMFE), London, 12th to McRobie, A. 2002. Buckling and stability, Undergraduate
24th August 1957. Volume II, pp 1923. lecture notes; University of Cambridge (U.K).
Eurocode 7 1997. Geotechnical design, Brussels, European Rankine, W.J.M. 1866. Useful rules and tables, London.
Committee for Standardization. Sovinc, I. 1981. Buckling of piles with initial curvature,
Eurocode 8 (Part 5) 1998. Design provisions for earthquake Proc of the International Conference on soil mechanics
resistance of structures-foundations, retaining structures and foundation engineering, Volume 8, pp 851855.
and geotechnical aspects, European Committee for stand- JRA 1996. Japanese Road Association, Specification for
ardization, Brussels. Highway Bridges, Part 5, Seismic Design.
Hamada 1992. Large ground deformations and their effects Rankine, W.J.M. 1866. Useful rules and tables, London.
on lifelines: 1964 Niigata earthquake, Technical report Horne, M.R and Merchant, W. 1965. The Stability of Frames,
NCEER-92-0001, Volume 1. Pergamon.

821

Copyright 2005 Taylor & Francis Group plc, London, UK


Propagation of pile tip damage during installation

T.R. Aldridge, T.M. Carrington & N.R. Kee


Fugro Limited, United Kingdom

ABSTRACT: This paper reviews the mechanisms that can result in the localized pile tip damage of offshore steel
tubular piles during installation, and details the conditions under which this may lead to progressive pile col-
lapse. The potential for local tip damage is assessed for all stages of installation and related to soil conditions.
It is believed that such considerations are not included in current offshore codes of practice, and it is hoped that
their use should help to prevent this type of local buckling and progressive collapse occurring in future.

1 INTRODUCTION buckling of the tip, particularly for long large


diameter piles. The control of the pile during low-
There have been a number of cases where large diam- ering must be such as to prevent this.
eter offshore steel pipe piles have collapsed during driv- (c) Pile in sleeve: tip damage may result from the forces
ing, and measurements of the internal dimensions of between the pile and the sleeve, resulting from
the pile have indicated progressive closure of the pile out-of-verticality of the pile, the pile and hammer
with penetration. The pile collapse has either led to pre- weight and the wave or current loading on the pile
mature refusal, or to problems with insert pile instal- whilst in the sleeve. Dynamic effects from wave
lation. Fugro has been involved as a consultant in a loading should be incorporated into the analysis.
number of such cases, and has used the observations (d) Obstructions within the soil: the pile tip may be
made in these cases as the basis for developing simple deformed by hitting obstructions within the soil,
analytical methods which it is considered help to such as boulders. The force given by any obstruc-
explain these occurrences, and it is hoped will help tion can be estimated from basic soil mechanics
designers to avoid such problems in future. principles, but dynamic effects may increase the
peak forces and should also be considered.
(e) Hard or cemented layers within the soil: may
2 POTENTIAL FOR PILE TIP DAMAGE
result in pile tip deformation, particularly where
DURING INSTALLATION
the pile is battered or the surface of the hard stra-
tum is not horizontal, so that the pile initially
There are many phases during pile installation when
impacts on one side only.
there is the potential for piles to be damaged, particu-
(f) Pile tip configuration: an external chamfer at the
larly at the pile tip. These include general handling on
pile tip will tend to force the soil beneath the pile
deck; lifting to the vertical; lowering into stabbing
wall to the outside of the pile during driving. This
cones; lowering through pile sleeves; impact on any
results in increased external stresses on the pile tip
obstacles on or beneath the seabed; and finally the
compared with a straight-ended pile, and in hard
forces and stresses applied by the soil and hammer
enough soils could cause the pile tip to fold inwards.
during the piling operation. It is also often the case that
a pile may be left in the sleeve during environmental
loading, either with or without the hammer in place.
3 FORCE TO CAUSE LOCAL DAMAGE
This range of cases is as follows:
(a) Lifting and handling on deck: whilst the lifting When the first of these types of problem occurred, we
process will always be checked for global bend- were not aware of any published solution to the local
ing, the potential for tip damage to be caused by buckling of the end of a steel tube. We therefore
axial and lateral loads applied locally to the pile tip assessed the point load that would cause such buck-
should also be checked. ling using upper bound theory for an assumed plastic
(b) Lowering into the pile sleeve stabbing cone: even hinge mechanism. In this approach, a mechanism of
very slow speeds of impact may result in lateral plastic hinges and deformation under an applied load

823

Copyright 2005 Taylor & Francis Group plc, London, UK


For a force applied at a 1:4 angle to the pile wall, this
resulted in Equation 2:

(2)

It is notable that for the large diameter steel pipe


 piles used offshore, the local lateral and axial buck-
y ling is not dependent on D/t, as is often assumed, but
 simply on the pile wall thickness and steel yield stress.
ny Equations 1 and 2 above may therefore be used to
check the tendency of the pile tip to deform locally
during the six critical stages of an installation given in
Section 2. The dynamics of the loading can be critical
in many of these cases, and soil resistances should be
uprated to allow for the dynamic soil resistance dur-
ing pile driving (e.g. Roussel, 1979). The dynamics of
the pile itself must also be incorporated into the ana-
lyses, particularly for the pile handling on deck, low-
ering into the stabbing cone and also whilst within the
Figure 1. Plastic hinge mechanism assumed when deriving sleeve. The effect of the water within the pile should
upper bound estimate of lateral load to cause local tip buckling. also be considered.

may be assumed, and will always give an upper bound 4 PROPAGATION OF TIP DEFORMATION
to the load actually required to cause deformation.
The assumed mechanism in this case is presented in 4.1 General
Figure 1. If the rotations on the plastic hinges are Whilst an initial deformation of the pile tip may cause
taken as  and  and the ratio of the width to length of some increased resistance to driving, and possibly some
the area defined by the plastic hinges is assumed to be pile deviation, the greatest threat to the installation is
n, then it is possible to derive the upper bound value that the deformation will develop further, resulting in
of the lateral load at the pile tip that could cause such a completely collapsed pile. Model tests in our labora-
a buckle to occur. tories and also observations offshore indicate that the
This approach, with n adjusted to give the lowest deformed area of the pile firstly develops as shown in
value of lateral load, gives the lateral load Flateral to Figure 2.
cause plastic hinge formation at the tip of a uniform Figure 3, measured on a collapsed pile offshore,
large diameter thin-wall tube as: shows the development of the collapsed geometry
(dashed line) compared with the original circular shape
(1) as the pile is driven deeper through stiff soils, showing
that the pile progressively collapsed until it formed a
Since this result was simply an upper bound value, we peanut shape at depth.
used laboratory tests on steel tubes to confirm the To avoid such propagation occurring, it is important
magnitude of the force required. We then used more to define the combination of pile and soil conditions
sophisticated 3D elasto-plastic finite element analysis which would tend to make a local buckle propagate
of a tube subject to a lateral point load at the tip, which such that the pile would completely close.
again confirmed the validity of Equation 1. We have It will be noted that Equations 1 and 2 do not incorp-
subsequently found a more recent derivation (Health orate the effect of any internal soil pressure resisting
and Safety Executive, 2001) which gives a solution the collapse. This is because in this mechanism the piles
within 15% of Equation 1 for a thin-wall tube. It may do not collapse at a given elevation, but simply cut
be noted that the area of application of the load has down through the soil, with the pile walls steadily driv-
been assumed to be small in the above analyses, but ing inwards and together. This is clearly indicated by the
that very similar equations may be derived for a uni- fact that in most observed cases of this kind of collapse,
form lateral pressure loading applied over a given area the elevation of the top of the soil plug remains close to
of the pile close to the tip. mudline. Whilst it requires some lateral soil resistance
Further laboratory tests and 3D finite element to prevent the pile simply springing back to its original
analyses were then performed to determine the near- shape, the soil plug is therefore not being compressed
axial force that would cause a local buckle at the tip. laterally by any significant amount at any elevation.

824

Copyright 2005 Taylor & Francis Group plc, London, UK


soil and pile behaviour must therefore be compared to
define whether the soil is both stiffer and stronger
than the pile.

4.2 Stiffness
We could not find a published solution for lateral load-
ing on the end of a pipe, but Roark and Young (1975)
give the inwards deflection lateral under a lateral load w
per unit length for a ring of diameter D and wall thick-
ness t supported at each side as:

(3)

where Epile is Youngs Modulus and pile the Poissons


Ratio for the pile.
Finite element analyses we have performed of the
end section of a large diameter pile indicate that the
Figure 2. Progressive development of lateral deformed lateral stiffness is similar to that derived for a ring of
area during driving after initial tip damage. approximately 0.5D length. Using this equivalent
length and also taking a Poissons Ratio of 0.3 for the
pile steel indicates that the pile wall lateral deflection
can be related to the force applied by the soil by:

(4)

For the soil, the lateral stiffness of a loaded area at


depth may be taken from Poulos and Davies (1974)
which gives a lateral displacement lateral of a rigid cir-
cular area of radius r at depth in a soil of Poissons Ratio
soil and Youngs Modulus Esoil as:

(5)

Figure 3. Cross-sections showing progressive pile collapse If it is assumed that the soil pressure is applied over
during driving following an initial deformation of the pile tip. an 0.5D diameter, then Equations 5 and 6 can be used to
compare the stiffness of the pile and soil response. This
comparison indicates that the stiffness of the pile will
be low enough to allow propagation if:

There are considered to be two criteria which


define whether a local buckle will propagate: (6)
1 there must be enough pressure from the soil onto the
initially small buckled area to cause the pile to con-
For a steel pile and soil Poissons Ratio values of 0.3
tinue to yield; this is a function of the soil and pile
(drained, cohesionless) or 0.5 (undrained, cohesive)
strength;
Equation 7 indicates that the soil stiffness values
2 the soil elastic response must be stiffer than that of
required to cause propagation for a 2.0 m diameter pile
the pile. Otherwise, any inward pile deformations
of a range of wall thickness are as given in Table 1.
will spring back and recover elastically.
It is clear from Equation 7 and Table 1 that the
When calculating the propensity of an initial small wall thickness at the pile tip is absolutely critical in
buckle at the pile tip to propagate at a given site, the preventing the propagation of any initial pile tip

825

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 1. Minimum soil stiffness expected to cause propa- Equation 8 suggests that for typical offshore piles
gation for 2.0 m o.d. pile, based on stiffness considerations. of 50 mm wall thickness and 345 MPa yield stress, an
initial deformation of 0.2 m length would not develop
Pile wall thickness (mm) 40 50 75 100 further unless the soil yield pressure (or qc value) was
of the order of 65 MPa or higher. Such qc values are
Soil stiffness (clay) (MPa) 6.3 12 42 98
Soil stiffness (sand) (MPa) 7.6 15 50 120 common in very dense overconsolidated sands but are
not encountered in un-cemented clays. From strength
considerations, propagation is possible for piles driv-
Table 2. Minimum soil parameters expected to cause prop- ing in very dense sands or into cemented soil layers, but
agation for 2.0 m o.d. pile, from stiffness considerations. would not be expected whilst driving through clays.
Once a deformation of length 0.5 m had developed,
Pile wall thickness (mm) 40 50 75 100 then Equation 8 indicates that it would still require
extremely hard clays to cause further propagation, but
su (clay) (kPa) 16 30 105 245 most sands or cemented strata would do so.
qc (sand) (MPa) 2.0 7.5 12.5 30.0

4.4 Conclusion on propagation


deformation. Adopting typical values for soil stiffness The propagation of a small initial deformation requires
of 400su in clay (Butler, 1974) and 4qc in sands (Lunne that the soil be both strong enough and stiff enough to
& Christoffersen, 1983) transforms Table 1 to give typ- deform the pile further. For typical offshore pile dimen-
ical soil parameter values that might cause damage sions, both criteria would generally be met, resulting in
propagation for a 2.0 m diameter pile of various wall the potential for progressive deformation, when driving
thicknesses, as given in Table 2. through very dense cohesionless strata or cemented
It may be seen that, from stiffness considerations, to layers, but not when driving through clay strata.
prevent propagation of initial tip damage when driving
through hard clays (su  200 kPa) and dense sands
(qc  30 MPa), may require that the pile tip either has 5 COMPARISON WITH RECOMMENDED
a wall thickness of the order of 100 mm, or that it is PRACTICE
made of a very high yield strength steel.
It is of interest to review the standard practice for off-
4.3 Strength shore jacket piles, as given by the American Petroleum
Institute in the 21st Edition of their Recommended
The yield pressure in the soil at depth may be taken
Code of Practice for the Planning, Design and Con-
from standard bearing capacity formulations. In clay,
structing of Fixed Offshore Platforms, API RP-2A
the ultimate lateral pressure qu may be based on a
(2000). Section 6.10 addresses the recommended
static resistance of the order of 9 to 12 times the
wall thickness of piles installed by driving, although
undrained shear strength, but this can be uprated by of
Sections 6.10.2 to 6.10.5 only consider local buckling
the order of 50% to allow for dynamic effects. This
caused by overall pile axial and bending stresses, but not
results in a yield pressure of 15 to 20 su which is then
local buckling of the tip. Section 6.10.6 entitled Min-
very similar to values obtained from CPT testing, i.e.:
imum Wall Thickness states that D/t should be small
enough to preclude local buckling, although again the
(7)
equations referred to relate to local column buckling
under axial and lateral loading, not tip buckling. This
In sands, it is also considered that the most reliable section also states that if hard driving is expected, the
estimate of the yield pressure would be based on the wall thickness should be greater than 6.35  D/l00
ultimate resistance given by the CPT cone resistance, qc. (mm), giving, for example, a minimum wall thickness
These values of yield in the soil may then be com- of only 28 mm for a 2.134 m diameter pile.
pared with the force required to continue the progres- The equations used to develop Table 2 indicate that
sive yield of the pile, by developing the upper bound a large diameter pile of 28 mm wall thickness could
plasticity approach to consider uniform pressure on a develop a complete collapse from an initial 0.2 m
progressive increase in the loaded pile area, as shown sized area of tip deformation if the soil yield pressure
in Figure 2. The resulting equation shows that for an exceeded about 20 MPa. This could therefore easily
initial deformation of length y to continue propaga- occur in sands, but would not be expected in clays, since
ting, the pressure from the soil must be at least: an undrained shear strength in excess of 1000 kPa
would be required. Application of the wall thickness
criteria given in the current API code for hard driv-
(8) ing would therefore not tend to result in pile collapse

826

Copyright 2005 Taylor & Francis Group plc, London, UK


after small amounts of initial tip damage in regions 7 CONCLUSIONS
where the soils are predominantly clays, but could do
so in areas where overconsolidated sands or cemented This paper presents equations which can be used to
strata are encountered. assess whether pile tip damage may occur for thin-
Section 6.10.8 concerns the use of a Driving Shoe walled large diameter pipe piles during offshore
which is stated to be used to assist piles to penetrate handling and installation, and whether such damage
through hard layers or to reduce driving resistances. may result in progressive pile collapse when it is
There is currently no guidance given on the thickness attempted to drive the piles to their target penetration.
or length of the driving shoe, although the previous The equations appear to be consistent with observa-
(20th) Edition of API gave a recommendation for the tions of damage during actual offshore installations
consideration of the use of a pile shoe at least one which the authors have reviewed in the past. It is
diameter in length with a wall thickness at least 1.5 believed that such considerations are not specifically
times the value in Section 6.9.6. For a 2.0 m diameter included in the existing codes of practice. It is there-
pile, the previous edition would have indicated a wall fore considered that inclusion of such recommenda-
thickness for the shoe of 42 mm. It is therefore appar- tions for avoiding lateral pile tip damage and subsequent
ent that the only recommendation relating to tip damage propagation should be incorporated into codes of
under hard driving results in lower wall thicknesses practice, in such a way that they are related directly to
than are normally used offshore, and also in lower wall pile geometry, the pile steel parameters and easily
thicknesses than have collapsed during hard driving measured soil parameters.
offshore.
In conclusion, the API code does not appear to pro-
vide recommendations on wall thickness in the form
that would be needed to ensure that the pile dimensions REFERENCES
were such as to prevent the occurrence of progressive
pile collapse following initial local deformation of American Petroleum Institute. 2000. Recommended Practice
the pile tip, related to given soil conditions. for Planning, Designing and Constructing Fixed Offshore
PlatformsWorking Stress Design. API Recommended
Practice 2A-WSD (RP 2A-WSD), 21st Edition.
6 OFFSHORE EXPERIENCE American Petroleum Institute. 1993. Recommended Practice
for Planning, Designing and Constructing Fixed Offshore
The authors have used the approaches outlined in this Platforms - Working Stress Design. API Recommended
paper when analyzing two major offshore cases of Practice 2A-WSD (RP 2A-WSD), 20th Edition.
progressive pile collapse. In both cases, the piles were Butler, F.G. 1974. General Report on State-of-the-Art.
Review Session III: Heavily overconsolidated clays.
drilled out following installation, clearly showing that Conf. on Settlement of Structures, Cambridge.
pile collapse had occurred. Equations 1 and 2 were Health and Safety Executive. 2001. A study of pile fatigue
used in each case to estimate the magnitude of lateral during driving and in-service and of pile tip integrity.
and axial loads on the pile tips that would have initi- Offshore Technology Report 2001/018.
ated tip damage, and these were compared with loads Lunne, T. & Christoffersen, H.P. 1983. Interpretation of cone
estimated for each stage of the installations, as out- penetrometer data for offshore sands. Proceedings, 15th
lined in Section 2 of this paper. Whilst in each case Annual Offshore Technology Conference. Houston, 1983,
there must remain some uncertainty as to the actual Paper OTC 4464.
mechanism or mechanisms that initiated the pile tip Poulos, H.G. & Davis, E.H. 1974. Elastic solutions for soil
and rock mechanics. London: John Wiley and Sons Inc.
damage, application of Equations 6 and 8 clearly Roark, R.J. & Young, W.C. 1975. Formulas for stress and
showed that once such initial tip damage had occurred, strain. London: McGraw-Hill Book Company.
the combinations of pile geometry and soil parameters Roussel, H.J. 1979. Pile driving analysis of large diameter
at each site were such that progressive pile collapse high capacity offshore pipe piles. PhD Thesis. Tulane
would have been expected to result, as observed. University, March 1979.

827

Copyright 2005 Taylor & Francis Group plc, London, UK


One dimensional wave propagation analysis of an open-ended pipe
pile with consideration of the excess pore pressure in soil plug

T. Matsumoto, T. Wakisaka & A. Numata


Kanazawa University, Kanazawa, Japan

ABSTRACT: In this study, one-dimensional wave propagation analysis program KWaveFD is extended in
order to incorporate the soil plug inside the open-ended pipe pile. When an open-ended pipe pile is driven into
a saturated soil, high pore pressures are generated in the soil plug. The effect of generation of pore pressures in
the soil plug is also considered in the extended KWaveFD. Furthermore, one-dimensional consolidation behaviour
of the soil plug after pile driving is analysed by finite difference method, using the distribution of pore pressures
at the end of the pile driving as initial condition. The behaviour of the pile under static loading can also be cal-
culated using the increased inner-shaft resistance due to the consolidation of the soil plug. In this paper, results
of numerical calculation of the above-mentioned behaviour of a driven open-ended pipe pile are demonstrated.

1 INTRODUCTION the end of pile driving is analysed with consideration


of the friction between the soil plug and the inner pile
Open-ended pipe piles are widely used for foundations shaft using finite difference scheme, with the distribu-
of offshore structures. During pile driving of an open- tion of pore pressures at the end of pile driving as ini-
ended pipe pile into the ground, a soil column called tial condition.
as the soil plug is created inside the pile. A number of It is noted that this paper is concerned with pile driv-
research works on the behaviour of soil plug have been ing for undrained penetration, although all driven piles
done (e.g. Yamahara 1964a, b, Heerema & de Jong will not be installed in an undrained manner. Therefore,
1980, Randolph 1987, Paikowsky et al. 1989, the results of this paper should be limited to soil con-
Paikowsky & Whitman 1990, Randolph et al. 1991, ditions with a low coefficient of consolidation, such
1992, Leong & Randolph 1991, Matsumoto & Takei as clay.
1991, Liyanapathirana et al. 2001, Kitiyodom et al.
2004, Matsumoto et al. 2004).
The behaviour of the soil plug, however, is not fully 2 METHODS OF ANALYSIS
clarified yet. Some observations have shown that high
pore pressures are generated in the soil plug when an 2.1 Finite difference approximation of wave
open-ended pipe pile is driven into a saturated ground equation
(Matsumoto et al. 1995, Matsumoto 1998). In addition In KWaveFD, the differential equation (Eq. (1)) for the
to generation of pore pressures in the soil plug during one-dimensional wave propagation in the pile was
driving, consolidation of the soil plug after the pile solved by means of a finite difference scheme
driving should be considered because of the fact that (Wakisaka et al. 2004). In Eq. (1), the influence of skin
the inner shaft resistance may increase as the effective friction,  (or end resistance, qb), is explicitly consider-
stress increases during the consolidation. ed, unlike the characteristic solutions of the wave
In this study, one-dimensional wave propagation equation.
analysis program KWaveFD (Wakisaka et al. 2004)
was extended to calculate the wave propagation in the
soil plug as well as the pile by means of finite difference (1)
method. A particular feature of the extended KWaveFD
is that generation of pore pressures in the soil plug is
calculated assuming fully undrained condition in the where t is the time, x the coordinate along the pile axis,
soil plug during driving. Moreover, the one-dimen- cp the wave velocity,  the skin friction, and wP, A, U
sional consolidation behaviour of the soil plug after and are the displacement, the cross-sectional area, the

829

Copyright 2005 Taylor & Francis Group plc, London, UK


circumferential length and the density of the pile. The The soil spring, kb, the damping constant, cb, and the
origin of x coordinate is placed at the pile head and additional soil mass, Mb, of the pile toe and the soil plug
the positive direction is taken in the direction toward the base are estimated from Eqs. (4) and (5), respectively,
pile toe. The shaft resistance, , acting upward against in which rout is the outer pile radius, rin the inner pile
the pile is taken as positive. radius, and G, Vs, s and  are the shear modulus, the
In the extended KWaveFD, the shaft resistance, ,
is the sum of the outer shaft resistance, out, and the
inner shaft resistance, in. The outer and the inner cir- Pile node
cumferential lengths of the pile, Uout and Uin, are taken
into account. dashpot cv
Finite difference approximation for Eq. (1) is (viscous)
expressed by Eq. (2) to deal with change in section slider max
properties as shown in Figure 1.
Soil adjacent to pile

dashpot cr
spring kout (radiation)

(2)

Soil far from pile (fixed)

where t is the time increment, x the element length Figure 2. Outer shaft resistance model. (after Randolph &
and E the Youngs modulus. Subscripts i and j Simons, 1986).
denote node number and time step, respectively.
The wave propagation in the soil plug is also gov-
Pile node
erned by Eq. (1), although cp is replaced by the bar
wave velocity in the soil plug, csp,  by in, by the Dashpot cb2
density of the soil plug, sp, U by Uin, and A is replaced Slider qb
by the cross-sectional area of the soil plug, Asp. Note
that in acting downward against the soil plug is taken
as positive. Finite difference expression for the wave
propagation in the soil plug is basically the same as Lumped
Eq. (2). The wave propagations in the pile and the soil Spring kb
mass Mb
plug are simultaneously solved to obtain the behav- Dashpot cb1
iours of the pile and the soil plug.
The outer shaft resistance, out, is modelled accord-
ing to Randolph & Simons (1986)(Fig. 2). The base
resistance of the pile toe and the soil plug base is mod-
elled according to Deeks & Randolph (1995) (Fig. 3). Figure 3. Pile base resistance model. (after Deeks &
The inner shat resistance is modelled as shown in Randolph, 1995).
Figure 4. The spring having stiffness, kin, was intro-
duced to calculate the mobilisation of the inner shaft
Pile node
resistance,  in.
The soil spring, kout, and the radiation damping, cr, of
the outer shaft resistance are estimated from Eq. (3). dashpot cv
(viscous)
slider max

spring kin

Soil plug

Figure 1. Notation used. Figure 4. Inner shaft resistance model.

830

Copyright 2005 Taylor & Francis Group plc, London, UK


shear wave velocity, the density and Poissons ratio of 2.3 Consolidation process of the soil plug after
the soil. pile driving
The viscous damping of the shaft resistance, which
The consolidation process of the soil plug is calcu-
occurs after the pile-soil slippage, is modelled by Eq. (6)
lated using the method proposed by Randolph et al.
according to Gibson & Coyle (1968), Heerema
(1991). The consolidation of the soil plug is governed
(1979) and Litkouhi & Poskitt (1980).
by the following equations:

(3)
(8)

(9)
(4)

(10)

(5) (11)

where t is the time, z the depth from the top of the soil
plug,  the effective vertical stress, u the excess pore
pressure, k the permeability, Esc the one-dimensional
(6) modulus, w the unit weight of fluid, c the coefficient
of consolidation, w the soil displacement,  the effect-
ive unit weight of the soil, in the inner shaft resistance,
where 0 is the reference relative velocity  1 m/s. inmax the maximum inner shaft resistance, and is the
coefficient determining  inmax.
2.2 Calculation of pore pressure in the soil plug The value of inmax is updated at each calculation
during pile driving step, according to the calculated value of  at the pre-
vious calculation step.
It may be reasonable to assume fully undrained con- The equations (8) to (11) are approximated in finite
dition for the soil plug during driving. In the present difference forms as follows:
analysis method, the soil plug is modelled as two-phase
material consisting of elastic soil skeleton and com-
pressible pore fluid. It is also assumed that the soil plug
deforms one-dimensionally without radial strain. From
the above assumptions and the principle of effective (12)
stress, the following relation can be derived:

(13)
(7)

in which  is the total vertical stress,  the effective


vertical stress, u the excess pore pressure, Esc the con-
strained Youngs modulus (one-dimensional modulus)
of the soil skeleton, Kf the bulk modulus of fluid, n the (14)
c
porosity and E eq the equivalent one-dimensional
modulus.
c
The equivalent one-dimensional modulus, Eeq , of
the soil plug is used in the wave propagation analysis
and the vertical strains, , are calculated in each step (15)
of calculation. Using the calculated strains, the pore
pressures in the soil plug are calculated by means of
Eq. (7). (16)

831

Copyright 2005 Taylor & Francis Group plc, London, UK


0.0 Proposed method
0.2 Theory Tv = 0.1
Tv = 0.2
0.4 Tv = 0.4
0.6 Tv = 0.6
Tv = 0.8
0.8
Permeability
z/H

1.0
k = 8.6  10-6 m/day
1.2 One-dimensional modulus
1.4 E c = 8.7  103 kPa
Unit weight of fluid
1.6 3
w = 9.8 kN/m
1.8 Coefficient of consolidation
2.0 c = 7.64  103 m2/day
0.0 0.2 0.4 0.6 0.8 1.0 Plugvlength L = 10 m
u/u0 Drainage length H = 5 m

Figure 5. Comparison of consolidation ratio curves obtained


from the proposed method and the theory.

where t is the time step for each calculation, z the


length of the soil plug element, subscript i the node
number, subscript j the time step number, i,0 is the
effective vertical stress of the soil plug at the start of
consolidation (or at the end of pile driving).
Note that the consolidation process is calculated Figure 6. Configuration of pile and ground.
with the initial distribution of the excess pore pressures
at time t  0, which was calculated in the pile driving Table 1. Specifications of the pile.
analysis, and that the displacement of the bottom of
the soil plug is fixed during consolidation. Length (m) 11
In order to confirm the calculation method men- Outer diameter (mm) 800
Inner diameter (mm) 780
tioned above, Terzaghis one-dimensional consolida- Cross-sectional area (m2) 0.0248
tion problem was solved and the calculated results were Youngs modulus (kPa) 2.4  108
compared with the theoretical values in Figure 5. Density (ton/m3) 7.8
The uniform distribution of initial pore pressure, u0, Wave velocity (m/s) 5547
was assumed and the drainage was allowed at the top Mass (ton) 2.13
and bottom of the soil plug, and the soil properties are
indicated in Figure 5. The results are indicated in the
normalised form. It can be seen from the figure that Table 2. Soil parameters used in the calculation.
the proposed method simulates very well the one-
dimensional consolidation problem. Parameter Layer 1 Layer 2

Soil density, t 1.9 ton/m 3


2.0 ton/m3
3 NUMERICAL EXAMPLE Undrained shear strength, cu 0.3v0 198 kPa
Youngs modulus, E 500cu 99 MPa
In order to discuss the increase in the bearing cap- Poissons ratio of soil 0.3 0.3
acity of a driven open-ended pipe pile with elapsed time skeleton, 
Bulk modulus of fluid, Kf 2000 MPa 2000 MPa
after the end of pile driving, an open-ended steel pipe Degree of saturation, Sr 100% 100%
pile driven into a saturated clay ground is calculated
for an example, using the analytical method proposed
in this paper.
clay. It is assumed that the height of the soil plug is
equal to the embedment length of 10 m.
3.1 Pile and ground description
The assumed soil properties and the soil resistance
Pile and ground assumed in this example calculation parameters are listed in Table 2. The initial effective
are shown in Figure 6 and Table 1. The ground is a nor- overburden pressure, v0  , increases proportionally
mally consolidated saturated soft clay (Layer 1) having with depth. The undrained shear strength, cu, of the soft
a thickness of 10 m underlain by a stiff clay (Layer 2). clay was estimated as cu  0.3v0. And, Youngs mod-
The open-ended steel pipe pile is driven to the stiff ulus of the soil was estimated empirically as 500cu.

832

Copyright 2005 Taylor & Francis Group plc, London, UK


Hence, both cu and E also increases proportionally
with depth. The shear modulus, G, is obtained using 3
the assumed value of Poissons ratio,   0.3. The
void ratio, e, was assumed by

Force (MN)
2
(17)

where  ref is the reference effective vertical stress 1


(set as 44.1 kPa), e0 is the void ratio at the reference
state (assumed as 0.786), and Cc is the compression
index (assumed as 0.45). 0
0 5 10 15 20
The coefficient of permeability, kv, in the vertical
direction was estimated by the following equation Time (ms)
(Sekiguchi et al, 1981) based on experimental evi-
dence (for example, Taylor, 194): Figure 7. Impact force on the pile head.

(18)
30
where kv 0 is the value of kv at the reference state 25 Pile head
(assumed as kv0  8.6  10
4 m/day), and k is a Displacement (mm)
20
constant (assumed as k  0.434Cc).
c 15
The equivalent one-dimensional modulus, Eeq , is
calculated from Eq. (7) and the corresponding equiva- 10 Soil plug head
lent Poissons ratio for undrained deformation,  eq, is
obtained. 5
The maximum shaft resistance down the inner and 0
the outer pile shafts during pile driving was assumed
-5
to be equal to the undrained shear strength, cu. 0 100 200 300 400 500 600
The shear modulus and the undrained shear strength Time (ms)
of Layer 2 were assumed as shown in Table 2. The
end-bearing capacity, qb, of Layer 2 was estimated as Figure 8. Displacements of the pile head and the soil plug.
qb  6cu.
The values of  and in Eq.(6) were set as   1.0
and  0.2.
Excess pore pressure (kPa)
3.2 Calculation results 0 200 400
0
The impact force shown in Figure 7 was applied on t = 200 ms
the pile top in the pile driving analysis. In order to 1
t = 400 ms
account for the influence of the residual stresses at the
end of pile driving, the analysis of pile driving for 3 2 t = 600 ms
consecutive blows was carried out, applying the impact 3
Depth from G.L. (m)

force in Figure 7 at a time interval of 200 ms.


The calculated displacements of the pile head and 4
the top of the soil plug are shown in Figure 8. The per-
manent set per blow, S, in each blow is 7.2 mm. The 5
permanent set per blow of the soil plug does not
6
increase so much with the number of blows and that
the total settlement of the soil plug is only 1 mm at the 7
end of the 3rd blow.
The distributions of the pore pressures in the soil 8
plug at the end of each blow are shown in Figure 9. It
is seen that the distributions at the end of the 2nd and 9
the 3rd blows are similar. Hence, the pore pressure 10
distribution at the end of the 3rd blow at t  600 ms
was adopted for the initial pore pressure distribution Figure 9. Distributions of the pore pressures in the soil
in the calculation of consolidation after the pile driving. plug at end of each blow.

833

Copyright 2005 Taylor & Francis Group plc, London, UK


Friction (kPa) Effective vertical stress (kPa)
-8 -6 -4 -2 0 2 4 6 0 20 40 60 80 100
0 0
t = 600 ms t=0
1
t = 1 day
Outer shaft
2 Inner shaft 2 t = 5 day
resistance
resistance

Depth from G.L. (m)


t = 30 day
3
Depth from G.L. (m)

4 t = 100 day
4

5
6
6

7
8
8

9 10
10
Figure 12. Change in distribution of effective vertical
stresses in the soil plug with elapsed time after the end of
Figure 10. Distributions of outer and inner shaft resist-
pile driving.
ances at the end of 3rd blow.

Inner shear stress (kPa)


Excess pore pressure (kPa)
-10 -8 -6 -4 -2 0 2 4 6 8
0 100 200 300 400 0
0
t=0
t =0
t=1
t = 1 day 2
2 t=5
t = 5 day
Depth from G.L. (m)

t = 30
Depth from G.L. (m)

t = 30 day
4 t =100
4 t = 100 day day

6 6

8 8

10 10

Figure 11. Change in distribution of pore pressures in the Figure 13. Change in distribution of inner shaft resistance
soil plug with elapsed time after the end of pile driving. with elapsed time after the end of pile driving.

The distributions of the outer and the inner shaft It is seen from Figure 11 that the pore pressures in the
resistances at the end of the 3rd blow are shown in soil plug dissipate rapidly than expected. Note that it
Figure 10. At this time, the inner shaft resistance acts takes more than 2000 days to reach 90% consolida-
downward against the whole length of the soil plug. tion if the inner shaft resistance were negligible dur-
This causes the pore pressures in the soil plug shown ing consolidation.
in Figure 9. The increase in the effective vertical stresses in the
The changes in the pore pressures, the effective soil plug is significantly lower than the dissipation of
vertical stresses and the inner shaft resistance down the excess pore pressures (Figure 12). This is caused by
soil plug with elapsed time from the end of pile driv- the development of the negative inner shaft resistance
ing are shown in Figures 11, 12 and 13, respectively. that acts upward against the soil plug (Figure 13).

834

Copyright 2005 Taylor & Francis Group plc, London, UK


Pile head load (kN) REFERENCES
0 200 400 600 800
0 Deeks, A.J. & Randolph, M.F. 1995. A simple model for
inelastic footing response to transient loading. Int. Jour.
for Num. and Anal. Methods in Geomech., 19: 307329.
5 Gibson, G. & Coyle, H.M. 1968. Soil damping constant
Pile head settlement (mm)

related to common soil properties in sands and clays.


10 Report No. 1251, Texas Transport Inst., Texas A & M
University.
Heerema, E.P. 1979. Relationships between wall friction,
15 displacement, velocity and horizontal stress in clay and
in sand for pile driveability analysis. Ground Engrg.,
12(1).
20 Heerema, E.P. & de Jong, A. 1980. An advanced wave equa-
tion computer program which simulates dynamic pile
25 plugging through a coupled mass-spring system. Proc.
Int. Conf. on Num. Methods in Offshore Piling, ICE:
3742.
30 Kitiyodom, P., Matsumoto, T., Hayashi, M., Kawabata, N.,
Hashimoto, O., Ohtsuki, M. & Noji, M. 2004. Experiment
Figure 14. Load-displacement curves at just the end of pile on soil plugging of driven open-ended steel pipe piles in
driving and after 100 days from the pile driving. sand and its analysis. Proc. 7th Int. Conf. on the Appl.
of Stress-Wave Theory to Piles, Selangor, Malaysia:
447458.
The load-displacement curves at just the end of Leong, E.C. & Randolph, M.F. 1991. Finite element analy-
pile driving and after 100 days from the pile driving ses of soil plug response. Int. Jour. for Num. and
were calculated, assuming consolidation of the soil Analytical Methods in Geomech. 15: 121141.
Litkouhi, S. & Poskitt, T.J. 1980. Damping constant for pile
plug yet not radial consolidation of the soil around the driveability calculations. Geotechnique 30(1): 7786.
pile. The maximum inner shaft resistance of the latter Liyanapathirana, D.S., Deeks, A.J. & Randolph M.F. 2001.
case was estimated as 0.3 times v after the rest period Numerical modelling of the driving response of thin-
of 100 days shown in Figure 12. walled open-ended piles. Int. Jour. for Num. and
The negative the inner shaft resistance exists after Analytical Methods in Geomech., 25: 933953.
consolidation as shown in Figure 13. This negative Matsumoto, T. & Takei, M. 1991. Effects of soil plug on
inner shaft resistance drags the pile downward, result- behaviour of driven pipe piles. Soils and Foundations,
ing in the increase in the outer shaft resistance during 31(2): 1434.
consolidation. However, this phenomenon was not Matsumoto, T., Michi, Y. & Hirano, T. 1995: Performance of
axially loaded steel pipe piles driven in soft rock. Jour. of
taken into account in the consolidation analysis, i.e., Geotech. Eng., ASCE, 121(4): 305315.
the pile was assumed as rigid and was assumed to be Matsumoto, T. 1998: A FEM analysis of s STATNAMIC test
stationary during consolidation. Hence, the outer and on open-ended steel pipe piles. Proc. 2nd Int. Statnamic
inner shaft resistance at the beginning of static load- Seminar, Tokyo, Japan: 287294.
ing was assumed to be 0 in the analysis of the static Matsumoto, T., Kitiyodom, P., Wakisaka, T. & Nishimura, S.
loading. 2004. Research on plugging of open-ended steel pipe
There is no distinct between the calculated load- piles and practice in Japan. Proc. 7th Int. Conf. on the
displacement curves as shown in Figure 14, clearly Appl. of Stress-Wave Theory to Piles, Selangor, Malaysia:
suggesting that the set-up phenomena due to the con- 133152.
Paikowsky, S.G., Whitman, V.R. & Baligh, M.M. 1989.
solidation of the soil plug is negligible in this particu- A new look at the phenomenon of offshore pile plugging.
lar example case. Marine Technology 8: 213230.
Paikowsky, S.G. & Whitman, V.R. 1990. The effects of plug-
ging on pile performance and design. Canadian
4 CONCLUDING REMARKS Geotechnical Journal, 27: 429440.
Randolph, M.F. 1987. Modelling of the soil plug response
The influence of the generation of pore pressures in during pile driving. Proc. 20th South East Asian Conf. on
the soil plug during pile driving and of the consolida- Soil Mech., Bangkok : 117.
Randolph, M.F., Leong, E.C. & Houlsby, G.T. 1991. One-
tion of the soil plug after pile driving on the behaviour
dimensional analysis of soil plugs in pipe piles.
of an example open-ended steel pipe pile has been dis- Gotechnique 41(4): 587598.
cussed, using the proposed analytical methods. The Randolph, M.F., May, M., Leong, E.C., Hyden, A.M. &
calculation results of the example case suggested that Murff, J.D. 1992. Soil plug response in open-ended pipe
the set-up phenomena due to the consolidation of the piles. Journal of Geotechnical Engineering, ASCE,
soil plug after pile driving could not be expected. 118(5): 743759.

835

Copyright 2005 Taylor & Francis Group plc, London, UK


Randolph, M.F. & Simons, H.A. 1986. An improved soil dynamic and static pile load tests. Proc. 7th Int. Conf. on
model for one-dimensional pile driving analysis. Proc.of the Appl. of Stress-Wave Theory to Piles, Selangor,
3rd Int. Conf. on Num. Methods in Offshore Piling: 117. Malaysia: 341350.
Sekiguchi, H., Nishida, Y. & Kanai, F. 1981. Analysis of par- Yamahara, H. 1964a, b. Plugging effects and bearing mech-
tially-drained triaxial testing of clay. Soils and anism of steel pipe piles. Trans. of the Architectural Inst.
Foundations, 21(3): 5366. of Japan, 96: 2835.
Taylor, D.W. 1948. Fundamentals of Soil Mechanics, Yamahara, H. 1964b. Plugging effects and bearing mech-
Modern Asian Edition, John Wiley and Sons. anism of steel pipe piles (Part 2). ditto, 97: 3441.
Wakisaka, T., Matsumoto, T., Kojima, E. & Kuwayama, S.
2004. Development of a new computer program for

836

Copyright 2005 Taylor & Francis Group plc, London, UK


Simplified analysis of single pile subjected to dynamic active and
passive loadings

P. Kitiyodom, R. Sonoda & T. Matsumoto


Kanazawa University, Kanazawa, Japan

ABSTRACT: In this paper, a three-dimensional simplified analysis of pile foundations subjected to dynamic
active and passive loadings using a hybrid model is presented. In the method, the pile is modelled as elastic beam
elements, the pile cap is modelled as thin plate elements, and the soil is treated as springs and dashpots. Reasonable
agreements are found between the solutions calculated using the proposed method and those calculated using more
rigorous analyses.

1 INTRODUCTION external forces at each pile node in the dynamic


analysis of the pile foundation.
Piles are widely used for foundations in both on land In order to examine the validity of the extended
and offshore in order to support mainly the static verti- PRAB, called D-PRAB hereafter, the solutions of
cal load from the superstructure. However in some D-PRAB are compared with the solutions of more
cases, a pile foundation may be also subjected to rigorous finite difference program FLAC3D (Itasca
dynamic loads due to ocean waves or winds (active 2002) and finite element program ADINA (ADINA
load) as well as earthquakes (passive load). Considering 2004). Reasonable agreements are shown between
current trends toward the limit state design or perform- these solutions.
ance based design in the area of foundation engineer-
ing, precise estimation of deformation of a whole
foundation and of stresses of its structural members is 2 METHOD OF ANALYSIS
an important aspect in the framework of these new
design criteria. As a simple design tool for piled raft Figure 1 illustrates the analytical model for piled raft
foundation subjected to active static combination load, foundation subjected to static active and passive loads
a computer program PRAB (Piled Raft Analysis with used in the previous works (Kitiyodom & Matsumoto
Batter piles) has been developed (Kitiyodom & 2002, 2003a, 2003b, Kitiyodom et al. 2004). The flex-
Matsumoto 2002, 2003a). Later, the program was ible raft is modelled as thin plate elements, the pile is
extended to accommodate three-dimensional simpli-
fied analysis of piled raft foundation subjected to static
ground movements (Kitiyodom & Matsumoto 2003b,
Kitiyodom et al. 2004).
In this work, PRAB was extended so that it can be
used to analyse the behaviour of pile foundation sub-
jected to dynamic loadings. In this method, the pile is
modelled as elastic beam elements, the pile cap is
modelled as thin plate elements and the soil is treated
as springs and dashpots. Dynamic load due to ocean
wave or wind is treated directly as an external force by
using the site measurement values as well as empirical
design values. In the case of pile foundation subjected
to an earthquake, first a dynamic analysis of the
response of the ground alone is conducted. The calcu-
lated free-field ground movement profiles at each
position of pile nodes are then used to obtain the Figure 1. Plate-beam-spring modelling of a piled raft.

837

Copyright 2005 Taylor & Francis Group plc, London, UK


modelled as beam elements and the soil is treated as The soil resistance can be written in matrix form as
interactive springs. The interaction between structural follows:
members is calculated based on Mindlins solutions
(Mindlin 1936) for both vertical and lateral forces. (8)
In this work, the hybrid model was extended for
dynamic analysis of pile foundations. The load- where [Ks], [Cs], and [Ms] are the stiffness matrix, the
displacement relationship of the piles and the pile damping matrix and the mass matrix of the soil.
cap can be written as Finally, from Equations (1), (2) and (8), we get

(1) (9)

where [K]  [Kp]  [Kc]  [Ks], [C]  [Cs], [M] 


(2) [Mp]  [Mc]  [Ms], and {F}  {Fp}  {Fc}. Note
that in the case of the pile foundation subjected to
where [Kp] is the pile stiffness matrix, [Kc] the pile cap dynamic passive loading the external force vector {F}
stiffness matrix, [Mp] the pile mass matrix, [Mc] the can be calculated by Equation (10)
pile cap mass matrix, {w} the displacement vector,
(10)
{Fp} the external force vector acting on the pile, {Fc}
the external force vector acting on the pile cap, and
{P} is the internal force vector. where {w0} is the free field ground movement vector.
The hybrid modelling of the pile and the soil as It is convenient to express Equation (9) in finite
shown in Figure 2 is adopted. The values of the vertical difference form with respect to time t.
spring, k, the horizontal spring, kh, the vertical radi-
ation damping, c, and the horizontal radiation damp-
ing, ch, per unit shaft area are approximated by means
of Equations (3) and (4), based on the work of Novak (11)
et al. (1978).

(3)

(4)

where Gs and Vs are the shear modulus and the shear


wave velocity of the surrounding soil respectively, and
d is the outer diameter of the pile.
The values of the soil spring at the pile base, kb, the
damping, cb, and the lumped soil mass, mb, per unit
base area can be estimated as follows (Randolph &
Deeks 1992):

(5)

(6)

(7)

in which s and s are the Poissons ratio and the dens-


ity of the soil respectively. Figure 2. Hybrid modelling of the pile and the soil.

838

Copyright 2005 Taylor & Francis Group plc, London, UK


Equation (11) is solved for pile deformations using 3.2 Impacts on pile without soil resistances
the Newmarks method.
Figure 3 shows the impact force applied to the pile head
In this paper, the dynamic interaction between
in vertical or lateral direction. Figures 4 and 5 show the
structural nodes is not taken into account in the analy-
pile responses calculated using D-PRAB compared
sis. This dynamic interaction may be incorporated into
with those of FLAC3D in vertical and lateral direction,
the method by using dynamic form of Mindlins solu-
respectively. Theoretically, the front of the force wave
tions suggested by Matsuoka & Yahata (1980a, b).
reaches the pile base at t  5.25 ms in the case of the
Note also that D-PRAB can also be used for the estima-
vertical impact and at t  8 ms in the case of the lateral
tion of non-linear deformation of the foundations sub-
impact. It can be seen from the figures that, the solu-
jected to dynamic loading, by employing the bi-linear
tions of both programs match very well with these the-
response of the soil springs associated with the viscous
oretical values and there are good agreements between
damping. These extensions are left for future work.
the two solutions.

3 COMPARISON ANALYSIS 3.3 Capped pile subjected to active loading


In this section, the analyses of a capped single pile sub-
In order to examine the validity of the proposed
jected to active dynamic loadings (as shown in Figure 6)
method, the results calculated using D-PRAB are com-
pared with the corresponding results calculated using
ADINA and FLAC3D.

3.1 Analytical condition


In this particular paper, only elastic responses of soil
and pile foundations are concerned. The soil, the pile
and the pile cap were treated as elastic homogeneous
materials in the analysis. The soil, the pile and the pile
cap properties are summarised in Table 1. All analy-
ses were carried out using these properties. In the
analyses using FLAC3D, the piles are modelled as
Pile Structural Elements and the pile cap as Shell Figure 3. Pile head force.
Structural Elements. For details of these structural
elements, refer to Itasca (2002).

Table 1. Material properties used in analyses.

Property Value

Soil:
Youngs modulus 5.96  104 kPa
Poissons ratio 0.49
Density 2 ton/m3
Longitudinal wave velocity 173 m/s
Shear wave velocity 100 m/s
Pile:
Youngs modulus 3.84  107 kPa
Poissons ratio 0.16
Density 2.4 ton/m3
Longitudinal wave velocity 4000 m/s
Shear wave velocity 2626 m/s
Length 21 m
Diameter 0.5 m
Square Pile Cap:
Youngs modulus 3.84  107 kPa
Poissons ratio 0.16
Density 2.4 ton/m3
Width 2m
Thickness 1m
Figure 4. Calculated vertical pile responses.

839

Copyright 2005 Taylor & Francis Group plc, London, UK


Figure 7. Input dynamic active load.

estimated taking into account the influence of the pile


size. Two types of calculations were performed using
ADINA. In the first analysis, the pile and the pile cap
were modelled as beam and shell elements, respect-
ively, while the pile and the pile cap were modelled as
solid elements in the second analysis. The ground was
modelled so that there was no reflection of outward
propagating waves back into the model pile founda-
tion during the time concerned. The top of soil model
(at z  0 m) is a free surface. The base of the model
Figure 5. Calculated lateral pile responses.
(at z  500 m) and the sides of the model (at
|x|  300 m and | y|  300 m) are fixed in all direc-
tions. The model pile foundation was located at the
centre of the model ground.
Figure 7 shows the input sinusoidal dynamic load
acting at the pile head in vertical or lateral direction.
Figures 8 and 9 show the vertical movement at the pile
head and the pile responses at time t  1.25 s, calcu-
lated using the proposed method, compared with the
responses calculated using ADINA. Both the calculated
results of ADINA where the foundation was modelled
as solid elements and those where the foundation was
modelled as beam and shell elements are shown in the
figures. There are good agreements among the solu-
tions. The results calculated using the proposed method
match well with the results calculated by ADINA where
the foundation was modelled as solid elements.
Figures 10 and 11 show the lateral movement at the
pile head and the pile responses at time t  1.25 s,
calculated using the proposed method, compared with
the responses calculated using ADINA. The results
calculated by ADINA where the foundation was mod-
elled as beam and shell elements are much higher
than the results calculated by ADINA where the foun-
dation was modelled as solid elements. The results
calculated using the proposed method lies between
these two solutions.
Figure 6. Ground and capped pile analysed (active load).
3.4 Capped pile subjected to passive loading
were carried out using the proposed method and the In this section, comparison analyses of a capped single
finite element program ADINA. Although in the pro- pile subjected to dynamic passive loading (seismic
posed method, the pile is modelled using beam elem- wave) were carried out using the proposed method
ents, the soil spring values and damping values are and FLAC3D. Analytical conditions in these analyses

840

Copyright 2005 Taylor & Francis Group plc, London, UK


Figure 8. Calculated vertical movement at pile head.

Figure 9. Calculated vertical pile responses (t  1.25 s).

Figure 12. Ground and capped pile analysed (passive load).

In the analysis using FLAC3D, free-field bound-


aries were applied at the sides of the model ground, in
order to make sure that there is no reflection of out-
ward propagating waves. The input acceleration was
applied at the base of the model (at z  50 m).
Figure 10. Calculated lateral movement at pile head. In the analysis using D-PRAB, the analysis is
divided into two stages. The response free field ground
movement which would occur if the pile was not pre-
sent is calculated first. The calculated free field ground
movement profiles at each position of pile nodes are
then used to obtain the external forces in the dynamic
analysis of the pile foundation as described in Equation
(10). The response of the pile to these computed exter-
nal forces induced by ground movements is computed.
Figure 13 shows the input accelerations that are
calculated using the following equation:
(12)
Figure 11. Calculated lateral pile responses (t  1.25 s).
where   2.2;  0.375;   8.0; f  0.5 and
were the same as those in Section 3.3, except for load- 1.0 Hz.
ing condition in which two kinds of time history of Figures 14 and 15 show the free field ground
horizontal acceleration were applied at the depth of movements at the depth z  0 m and z  20 m for the
50 m as shown in Figure 12, instead of dynamic load case of input frequency f  0.5 Hz and 1.0 Hz,
at the pile head. respectively. Very large lateral ground movement

841

Copyright 2005 Taylor & Francis Group plc, London, UK


Figure 13. Input acceleration. Figure 16. Calculated response at pile head ( f  0.5 Hz).

Figure 14. Free field ground responses ( f  0.5 Hz).


Figure 17. Calculated pile responses ( f  0.5 Hz).

Figure 15. Free field ground responses ( f  1.0 Hz).

responses were obtained in the case of f  0.5 Hz.


Because this frequency is equal to the natural fre- Figure 18. Calculated response at pile head ( f  1.0 Hz).
quency, fn, of the ground, which is calculated based
on Equation (13).

(13)

where H is the depth of the ground.


Figures 16 and 17 show the lateral displacements at
the pile head and the pile responses at a time when the
lateral movement at the pile head shows the maximum
absolute value, calculated using the proposed method
compared with those calculated using FLAC3D for
the case of input frequency f  0.5 Hz. It can be seen
from the figures that the results calculated using the
proposed method match very well with those calcu-
lated using FLAC3D. Figure 19. Calculated pile responses ( f  1.0 Hz).

842

Copyright 2005 Taylor & Francis Group plc, London, UK


The lateral displacements at the pile head and the Kitiyodom, P. & Matsumoto, T. 2002. A simplified analysis
pile responses calculated using the proposed method method for piled raft and pile group foundations with
and FLAC3D for the case of input frequency f  1.0 Hz batter piles. International Journal for Numerical and
are shown in Figures 18 and 19. Good agreements Analytical Methods in Geomechanics 26: 13491369.
Kitiyodom, P. & Matsumoto, T. 2003a. A simplified analysis
between the two solutions can be seen again. method for piled raft foundations in non-homogeneous
soils. International Journal for Numerical and Analytical
4 CONCLUDING REMARKS Methods in Geomechanics 27: 85109.
Kitiyodom, P. & Matsumoto, T. 2003b. Extension of a com-
A three-dimensional simplified analytical method has puter program PRAB for deformation analysis of piled
been proposed for the analysis of the deformation and rafts subjected to ground movements. Proc. of the Sino-
the load distribution of pile foundations subjected to Japanese Symposium on Geotechnical Engineering,
Beijing: 7479.
dynamic active and passive loadings. The proposed Kitiyodom, P., Matsumoto, T. & Kawaguchi, K. 2004.
method was examined through comparisons with the Analysis of piled raft foundation subjected to ground
results calculated from FLAC3D and ADINA. movement induced by tunnelling. Proc. of 15th Southeast
The advantage of the use of D-PRAB in dynamic Asia Geotechnical Conference, Bangkok: 183188.
analysis of pile responses is that the time needed to Matsuoka, O. & Yahata, K. 1981a. Basic analyses on prob-
analyse the problem is much less than the time needed lems of a three-dimensional homogeneous, isotropic,
to analyse the problem by using finite difference or elastic medium, and the applications (Part II).
finite element method. The proposed type of analysis Transaction of the Architectural Institute of Japan 293:
method may be very useful in routine design of pile 3544 (in Japanese).
Matsuoka, O. & Yahata, K. 1981b. Basic analyses on prob-
foundations. lems of a three-dimensional homogeneous, isotropic,
More comparative and parametric analyses of pile elastic medium, and the applications (Part III).
group subjected to dynamic loads can be found in Transaction of the Architectural Institute of Japan 298:
Sonoda et al. (2005). 4353 (in Japanese).
Mindlin, R.D. 1936. Force at a point in the interior of a semi-
ACKNOWLEDGEMENT infinite solid. Physics 7: 195202.
Novak, M., Nogami, T. & Aboul-Ella, F. 1978. Dynamic soil
reactions for plane strain case. Journal of Mechanical
The authors would like to express their special appre- Engineering, ASCE 104(EM4): 953959.
ciate to Professor H. Masuya and Miss S. Tachibana, Randolph, M.F. & Deeks, A.J. 1992. Dynamic and static soil
Kanazawa University, for their support in carrying models for axial pile response. Proc. 4th Int. Conf. on the
out finite element analyses. Application of Stress-Wave Theory to Piles, The Hague:
314.
Sonoda, R., Kitiyodom, P. & Matsumoto, M. 2005.
REFERENCES Simplified dynamic analysis of pile group subjected
to horizontal loading. Proc. International Symposium
ADINA. 2004. ADINA User Interface Primer, ADINA on Frontiers in Offshore Geotechnics, Perth (to be
R&D, Inc., USA. presented).
Itasca. 2002. FLAC3D Users Guide, Itasca Consulting
Group, USA.

843

Copyright 2005 Taylor & Francis Group plc, London, UK


Simplified dynamic analysis of pile group subjected to horizontal loading

R. Sonoda, P. Kitiyodom & T. Matsumoto


Kanazawa University, Kanazawa, Japan

ABSTRACT: In the present paper, the behaviour of a pile group foundation subjected to horizontal dynamic
loads is analysed by using D-PRAB. The solutions obtained using D-PRAB are compared with the solutions
obtained using ADINA and FLAC3D in order to examine the validity of D-PRAB for pile groups. Two kinds of
dynamic analysis are conducted for a 4-pile pile group with a rigid cap. In the first type of analysis, external
dynamic horizontal load supposing ocean wave in an approximate way is directly applied to the pile cap. In the other
type of analysis, earthquakes are applied at seismic bedrock to calculate the responses of the foundation structure.

1 INTRODUCTION Several three-dimensional FEM codes are available


for dynamic analysis of foundation structures. How-
Pile foundations supporting offshore structures may ever, such analysis takes relatively much calculation
be subjected to dynamic horizontal loads due to ocean work and time, hence it could not be used properly as
wave or wind as well as earthquake. Hence, dynamic a routine design tool.
design of the pile-supported offshore structures consider- In the companion paper (Kitiyodom et al. 2005), a
ing kinetic interaction between the structure and the simplified dynamic three-dimensional deformation
ground will be required. analysis method, D-PRAB, of pile foundations is pro-
In seismic areas such as Japan, earthquake is the posed. In Kitiyodom et al. (2005), the applicability of
most critical design load in the design of foundation of D-PRAB to single piles is discussed.
a high-rise building. However, an approximate design In the present paper, the applicability of D-PRAB
method has been adopted in usual design of a high-rise to pile group foundations is discussed. Two kinds of
building in Japan. In the approximate design method, dynamic analysis were conducted for a 4-pile pile
an input design acceleration wave is applied to the base group with a rigid cap using D-PRAB. In the first type
of the building directly without consideration of the of analysis, external dynamic horizontal load supposing
existence of the substructure and the ground, or the ocean wave in an approximate way is directly applied
foundation structure including piles and the ground are to the pile cap. In another analysis, earthquakes are
modelled by a combination of a horizontal spring and applied at seismic bedrock to calculate the responses
a rotational spring to estimate horizontal displacement of the foundation structure. The results calculated using
and rocking motion. D-PRAB are compared with those computed using
Currently, the foundation design in Japan is chan finite element program ADINA (ADINA 2004) and
ging from the conventional allowable stress design to finite difference program FLAC3D (Itasca 2002).
the limit state design or the performance-based design. A good performance of D-PRAB will be demon-
In these new design methods, dynamic analysis of the strated especially for calculation of the pile founda-
whole structure including the super structure, the sub- tion subjected to earthquake loading.
structure and the ground is required to correctly esti-
mate the deformation of the whole structure, and the
deformation and the stresses generated in each struc- 2 ANALYSIS METHOD
tural member.
Behaviours of pile foundation models subjected to Figure 1 illustrates the analytical model for pile foun-
static horizontal loads or dynamic loads were investi- dations subjected to dynamic loads used in D-PRAB
gated in centrifuge modelling (Horikoshi et al. 2003a, b) (Kitiyodom et al. 2005). The flexible pile cap is
and in 1-g field modelling (Matsumoto et al. 2004a, b). modelled as thin plates, the pile is modelled as beam
The results of these modellings suggest that the response elements and the soil is treated as springs and dashpots
of the whole structure is significantly controlled by as shown in Figure 2.
kinetic interaction between the foundation structure In the dynamic analysis, the values of the soil spring
and the ground. and the radiation damping at the pile shaft nodes in

845

Copyright 2005 Taylor & Francis Group plc, London, UK


the vertical and horizontal directions are estimated In the dynamic analysis, the dynamic interaction
according to Novak et al. (1978) while the soil resist- between structural nodes is not taken into account at the
ance parameters at the pile base are estimated according present stage. This dynamic interaction may be incorp-
to Randolph & Deeks (1992). orated into the method by using dynamic form of
Mindlins solutions proposed by Matsuoka & Yahata
(1981a, b). Note that D-PRAB can also be used for the
y estimation of non-linear deformation of the founda-
tions subjected to dynamic loading by employing the
bi-linear response of soil springs and viscous damping.
These extensions are left for further work.
Note that although the dynamic interaction between
the piles is not taken into account at the present stage,
x reasonable agreement between the results calculated
using D-PRAB and those calculated using more rigor-
ous methods are shown especially in the case of pile
group subjected to passive loading such as earthquakes.
soil spring
3 ANALYSIS OF PILE GROUP SUBJECTED
TO WAVE LOAD
Radiation
damping
3.1 Analytical conditions
Figure 1. Plate-beam-mass-spring modelling used in Figure 3 shows the problem analysed in this section.
D-PRAB. A 4-pile freestanding pile group with a rigid cap is sub-
jected to horizontal sinusoidal wave load as shown in
Figure 4. Comparison analysis of the problem using
kh0 D-PRAB and ADINA was carried out. The soil, the pile
and the pile cap were treated as elastic homogeneous
m0
c0
ch0 4m
k0
(Kp)1
kh1 4m 2m
m1
c1 Dynamic
ch1 load
k1
(Kp)2 1m
kh2
1m pile head
m2
c2 (K )
p 3
k2 ch2 Concrete pile
rp = 2.4 t/m3
khn np = 0.16
(Kp)n mn
Ep = 38.4 GPa
cn L = 20 m
kn d = 0.5 m
chn
pile 1 pile 2
Soil
cb cb
rs = 2.0 t/m3
kb ns = 0.49
2m
mb pile base Gs = 20 MPa

Figure 2. Hybrid modelling of the pile and the soil. Figure 3. Ground and pile group models analysed.

846

Copyright 2005 Taylor & Francis Group plc, London, UK


materials in the analysis. The properties of the soil, calculated using ADINA. As expected, the lateral
the pile and the pile cap are summarised in Table 1. movement calculated using D-PRAB shows the small-
Two types of calculations were performed using est value due to the neglecting of the dynamic inter-
ADINA. In the first analysis, the pile and the pile cap action between piles.
were modelled as beam and shell elements, respect- Figure 6 shows the lateral deflections, bending
ively, while both the pile and the pile cap were mod- moments, vertical movements and axial forces along
elled as solid elements in the second analysis. The the piles at a time t  1.25 s, calculated using D-PRAB
ground was modelled so that there is no reflection of compared with those calculated using ADINA. It can
outward travelling waves back into the model founda- be seen from the figures that except for the lateral
tion during the calculation period concerned. Consider- deflection, the results calculated using D-PRAB match
ing the condition of plane symmetry, one-half of the very well with the results calculated using ADINA.
entire system was modelled. The top of the model (at For comparison, the responses of a capped single
z  0 m) is a free surface. The base of the model and pile presented in Kitiyodom et al. (2005) are plotted
three sides of the model (at |x|  300 m and y  300 m) again in Figures 7 and 8. The amount of dynamic load
are fixed in all directions. The side of the model at applied on a pile is the same. It can be seen from the
y  0 m is fixed in y-direction. figure that the lateral displacement at the pile head is
higher in the case of single pile than that in the case of
3.2 Analytical results pile group. This is thought to be due to the difference in
Figure 5 shows the lateral movement at the pile head
calculated using D-PRAB compared with the solutions 20
Lateral Disp. (mm)

10
2000

1000 0
Force (kN)

D-PRAB
0 -10 ADINA(beam)
ADINA(solid)
-1000
-20
0 1 2 3 4 5
-2000
Time (s)
0 1 2 3 4 5
Time (s) Figure 5. Calculated lateral movement at pile head.
Figure 4. Input force at pile head.

Table 1. Material properties used in analyses. Lateral Deflection of Pile (mm) Bending Moment (kNm)
-3 0 3 6 9 12 -600 -400 -200 0 200
0 0
Property Value
5 5
Soil:
Depth (m)

Depth (m)

Youngs modulus 5.96  104 kPa 10


10
Poissons ratio 0.49
Density 2 ton/m3 15 D-PRAB 15
D-PRAB
Longitudinal wave velocity 173 m/s ADINA(beam)
ADINA(solid) ADINA(beam)
Shear wave velocity 100 m/s 20 20
Pile: (a) Lateral Deflection (b) Bending Moment
Youngs modulus 3.84  107 kPa
Vertical Movement of Pile (mm) Axial Force (kN)
Poissons ratio 0.16 -2 -1 0 1 2 -600 -400 -200 0 200 400 600
Density 2.4 ton/m3 0 0
Longitudinal wave velocity 4000 m/s
Shear wave velocity 2626 m/s 5 5
Depth (m)

Depth (m)

Length 21 m
pile 2
Diameter 0.5 m 10 pile 1 10
Square Pile Cap: D-PRAB
ADINA
pile 1 pile 2
Youngs modulus 3.84  107 kPa 15 (beam)
15 D-PRAB
ADINA
ADINA
Poissons ratio 0.16 (solid) (beam)
20 20
Density 2.4 ton/m3
(c) Vertical Movement (d) Axial Force
Width 4m
Thickness 1m
Figure 6. Calculated pile group responses (t  1.25 s).

847

Copyright 2005 Taylor & Francis Group plc, London, UK


20 1m
Lateral Disp. (mm)

1m
10 rs1, ns1 Gs1
1m rs2, ns2 Gs2 1m
0
rs3, ns3 Gs3
D-PRAB pile head
-10 ADINA(beam) pile node
Soil
ADINA(solid) rsi = 2.0 t/m3 L = 20 m d = 0.5 m
-20 50 m nsi = 0.49
0 1 2 3 4 5 Gsi = 20MPa
pile 1 pile 2
Time (s) free field ground

Figure 7. Calculated lateral movement at pile head (single rs48, ns48, Gs48
pile). 1 m rs49, ns49, Gs49 pile base
rs50, ns50, Gs50 Input dynamic external forces
Lateral Deflection of Pile (mm) Bending Moment (kNm)
Input acceleration
-5 0 5 10 15 20 -100 0 100 200 300 400 500 600
0 0
Seismic bedrock

5 5 Figure 9. Analysis procedure in D-PRAB for earthquake.


Depth (m)

Depth (m)

10 10 2
Acceleration (m/s2)
15 D-PRAB
ADINA(beam)
15
D-PRAB
1
ADINA(solid) ADINA(beam)
20 20
0
(a) Lateral Deflection (b) Bending Moment

-1
Figure 8. Calculated single pile responses (t  1.25 s). Input Acceleration (f = 0.5 Hz)
-2
the pile head conditions, free in single pile and rigid in 0 2 4 6 8 10 12 14
pile group. Time (s)

4 ANALYSES OF PILE GROUP SUBJECTED Figure 10. Input acceleration with a frequency of 0.5 Hz.
TO EARTHQUAKES
2
Acceleration (m/s2)

4.1 Analytical conditions


1
Analytical conditions in these analyses were the same as
Figure 3, except for loading condition. In these analyses, 0
two kinds of time history of horizontal acceleration were
applied at the depth of 50 m, instead of dynamic load at -1
the pile cap. Comparison analyses of the problem using Input Acceleration (f = 1.0 Hz)
D-PRAB and FLAC3D were carried out. -2
In the analysis using FLAC3D, the pile and the pile 0 2 4 6 8 10 12 14
cap were modelled as beam and shell elements. Free- Time (s)
field boundaries are applied at the side of the model
ground, in order to make sure that there is no reflec-
Figure 11. Input acceleration with a frequency of 1 Hz.
tion of outward travelling wave.
In the analysis using D-PRAB, the response free
field ground movements are calculated first. The cal- where   2.2;  0.375;   8.0; f  0.5 and 1 Hz.
culated free field ground movement profiles at each Figures 10 and 11 show the input acceleration at
position of pile nodes are used to estimate the forces bedrock at the depth z  50 m for the case of input
at each pile node. These forces are then used as the frequency f  0.5 Hz and 1 Hz, respectively.
external forces in the dynamic analysis of the pile Figures 12 and 13 show the free field ground
foundation (see Figure 9). movements at the depth z  0 m and z  20 m for the
Input acceleration was calculated using the follow- case of input frequency f  0.5 Hz and 1 Hz, respect-
ing equation: ively. Very large lateral soil movement responses were
obtained in the case of f  0.5 Hz. This frequency is
equal to the natural frequency, fn, of the ground, which
(1) is calculated based on Equation (2).

848

Copyright 2005 Taylor & Francis Group plc, London, UK


3000 3000
pile head
Lateral Disp. (mm)

z=0m D-PRAB

Lateral Disp. (mm)


2000 2000
z = 20 m FLAC3D
1000 1000
0 0
-1000 -1000
-2000 -2000
Max. lat. disp.
-3000 -3000
0 2 4 6 8 10 12 14 0 2 4 6 8 10 12 14
Time (s) Time (s)
(a) Pile Head (GL+1.0m)
Figure 12. Free field ground responses (f  0.5 Hz).
3000

Lateral Disp. (mm)


D-PRAB pile base
2000
75 FLAC3D
Lateral Disp. (mm)

1000
50
0
25
-1000
0
-2000
-25
z=0m -3000
-50 z = 20 m 0 2 4 6 8 10 12 14
-75 Time (s)
0 2 4 6 8 10 12 14 (b) Pile Base (GL-20m)
Time (s) 3000
Lateral Disp. (mm)

Figure 13. Free field ground responses ( f  1 Hz). 2000


1000
0
(2)
-1000

where H is the depth of the ground. -2000


displacement at depth = 50 m
-3000
0 2 4 6 8 10 12 14
4.2 Analytical results Time (s)
Figures 14 and 15 show the lateral displacements at the (c) Input motion at bedrock (GL-50m)
pile head and the pile base calculated using D-PRAB
compared with those calculated using FLAC3D for Figure 14. Calculated lateral displacement ( f  0.5 Hz).
the case of input frequency f  0.5 Hz and 1 Hz, respect-
ively. It can be seen from the figures that both solutions
match very well in both input acceleration frequencies.
The lateral deflections, bending moments, vertical
movements and axial forces along the piles at a time 5 CONCLUDING REMARKS
when the lateral displacement at the pile head shows the
maximum absolute value, calculated using D-PRAB The applicability of D-PRAB proposed by Kitiyodom
compared with those calculated using FLAC3D, are et al. (2005) to analyse pile group foundations was
plotted in Figures 16 and 17. Although D-PRAB tends examined. Two kinds of dynamic analysis were con-
to underestimate the vertical movements and axial ducted for a 4-pile pile group with a rigid cap using
forces along the pile, and the bending moment at the D-PRAB. In the first type of analysis, external
pile head compared with the results of FLAC3D, there dynamic horizontal load was applied. In the other
are good agreements between the two solutions. case the responses of the pile foundation subjected to
In Kitiyodom et al. (2005), the single pile responses earthquakes were analysed.
to the same input accelerations are presented. For The results calculated using D-PRAB were com-
comparison, these single pile responses at the max- pared with those computed using ADINA and FLAC3D.
imum lateral pile head displacement are plotted in A good performance of D-PRAB was demonstrated
Figure 18 and 19. It can be seen from the figures that especially for the calculation of the pile foundation
the lateral deflections along the pile are almost the same subjected to earthquake loading.
in the cases of the single pile and pile group. However, It should be noticed that the dynamic interaction
the induced bending moments along the pile are totally between the piles should be incorporated in dynamic
different between the single pile and the pile group. analysis of pile foundation subjected to external load.

849

Copyright 2005 Taylor & Francis Group plc, London, UK


Lateral Disp. (mm) 75 Lateral Deflection of Pile (mm) Bending Moment (kNm)
Max. lat. disp. 0 25 50 75 -50 -40 -30 -20 -10 0
50
0 0
25 D-PRAB
FLAC3D
0

Depth (m)

Depth (m)
5 5
-25
D-PRAB 10 10
-50 pile head FLAC3D
-75 15 15
0 2 4 6 8 10 12 14 D-PRAB
FLAC3D
Time (s) 20 20
(a) Pile Head (GL+1.0m) (a) Lateral Deflection (b) Bending Moment

75
Vertical Movement of Pile (mm) Axial Force (kN)
Lateral Disp. (mm)

50 pile base
-0.150 -0.075 0.000 0.075 0.150 -50 -25 0 25 50
25 0 0
0
5 5
-25

Depth (m)

Depth (m)
D-PRAB
-50 FLAC3D 10 10
-75 pile 1 pile 2 pile 1 pile 2
0 2 4 6 8 10 12 14 15 15
Time (s) D-PRAB D-PRAB
FLAC3D FLAC3D
(b) Pile Base (GL-20m) 20 20
75
Calculated pile group responses ( f  1 Hz).
Lateral Disp. (mm)

50 Figure 17.
25
0
-25
-50
displacement at depth = 50 m
-75 Lateral Deflection of Pile (mm) Bending Moment (kNm)
0 2 4 6 8 10 12 14 -2000 -1500 -1000 -500 0 -160 -80 0 80 160 240
Time (s) 0 0
D-PRAB
(c) Input motion at bedrock (GL-50m) FLAC3D
Depth (m)

Depth (m)

5 5
D-PRAB
Figure 15. Calculated response displacement ( f  1 Hz). 10 10
FLAC3D

Lateral Deflection of Pile (mm) Bending Moment (kNm) 15 15


-2000 -1500 -1000 -500 0 0 100 200 300 400 500
0 0 20 20
D-PRAB
FLAC3D (a) Lateral Deflection (b) Bending Moment
Depth (m)

5 5
Depth (m)

10 10 Figure 18. Calculated single pile responses ( f  0.5 Hz).

15 15 D-PRAB
FLAC3D
20 20
(a) Lateral Deflection (b) Bending Moment

Vertical Movement of Pile (mm) Axial Force (kN) Lateral Deflection of Pile (mm) Bending Moment (kNm)
-1.50 -0.75 0.00 0.75 1.50 -500 -250 0 250 500 0 25 50 75 -30 -20 -10 0 10 20 30
0 0 0 0
D-PRAB
FLAC3D
Depth (m)

5 5 5 5
Depth (m)

Depth (m)
Depth (m)

10 10 10 10
pile 2
pile 1 pile 2 pile 1 15 15
15 15 D-PRAB
D-PRAB D-PRAB FLAC3D
FLAC3D FLAC3D
20 20 20 20
(c) Vertical Movement (d) Axial Force (a) Lateral Deflection (b) Bending Moment

Figure 16. Calculated pile group responses ( f  0.5 Hz). Figure 19. Calculated single pile responses ( f  1 Hz).

850

Copyright 2005 Taylor & Francis Group plc, London, UK


ACKNOWLEDGEMENT Symposium on Frontiers in Offshore Geotechnics. Perth
(to be presented).
The authors would like to express their special appre- Matsuoka, O. & Yahata, K. 1981a. Basic analyses on prob-
ciate to Professor H. Masuya and Miss S. Tachibana, lems of a three dimensional homogeneous, isotropic,
elastic medium, and the applications (Part II).
Kanazawa University, for their support in carrying Transaction of the Architectural Institute of Japan 293:
out finite element analyses. 3544 (in Japanese).
Matsuoka, O. & Yahata K. 1981b. Basic analyses on prob-
lems of a three dimensional homogeneous, isotropic,
elastic medium, and the applications (Part III).
REFERENCES Transaction of the Architectural Institute of Japan 298:
4353 (in Japanese).
ADINA. 2004. ADINA User Interface Primer, ADINA Matsumoto, T., Fukumura, K., Kitiyodom, P., Oki, A. &
R&D, Inc., USA. Horikoshi, K. 2004a. Experimental and analytical study
Horikoshi, K., Matsumoto, T., Hashizume, Y., Watanabe, T. on behaviour of model piled rafts in sand subjected to
& Fukuyama, H. 2003a. Performance of piled raft foun- horizontal and moment loading. Int. Journal of Physical
dations subjected to static vertical loading and horizontal Modelling in Geotechnics 3: 119.
loading. Int. Journal of Physical Modelling in Matsumoto, T., Fukumura, K., Oki, A. & Horikoshi, K.
Geotechnics 2: 3750. 2004b. Shaking table tests on model piled rafts in sand
Horikoshi, K., Matsumoto, T., Hashizume, Y. & Watanabe, T. considering influence of superstructures. Int. Journal of
2003b. Performance of piled raft foundations subjected Physical Modelling in Geotechnics 3: 2037.
to dynamic loading. Int. Journal of Physical Modelling in Novak, M., Nogami, T. & Aboul-Ella, F. 1978. Dynamic soil
Geotechnics 2: 5162. reactions for plane strain case. Journal of Mechanical
Itasca. 2002. FLAC3D Users Guide, Itasca Consulting Engineering, ASCE 104(EM4): 953959.
Group, USA. Randolph, M.F. & Deeks, A.J. 1992. Dynamic and static soil
Kitiyodom, P., Sonoda, R. & Matsumoto, T. 2005. models for axial pile response. Proc. 4th Int. Conf. on the
Simplified analysis of single pile subjected to dynamic Application of Stress-Wave Theory to Piles, The Hague:
active and passive loadings. Proc. International 314.

851

Copyright 2005 Taylor & Francis Group plc, London, UK


Reliability analysis on axially loaded pile foundation of offshore platforms

Yan Shuwang, Zhou Hongjie & Liu Run


School of Civil Engineering, Tianjin University, Tianjin, China

ABSTRACT: An approach based on Advanced First-Order Second-Moment (AFOSM) is proposed for the
reliability analysis on the pile foundation subject to axial loads. Analysis results show that the dominant factors
that contribute to the reliability index are the uncertainties in the extreme environmental load and the soil
strength. As the mean value and/or the coefficient of variation of the axial load increase, both the reliability
index and safety factor decrease and more resistance of the subsoil surrounding the piles will be mobilized in
order to meet the requirements of the limit state equation. And when the variability of soil resistance rises, all
the partial factors will fall simultaneously. In this analysis, the limit state equation can only be satisfied by
reducing the designed load, which leads to a decrease in the global reliability of the pile foundation.

1 INTRODUCTION different from those in the Mexico Bay and California


coastal area, the reliability analysis should be per-
Pile foundation is the most commonly used founda- formed before using the API Code. In this paper, a
tion form for offshore platforms. It has been found procedure is presented for the reliability analysis on
that the Work Stress Design (WSD) method, which the pile foundations subject to axial loads in Bohai
has being used for a long time in the conventional pile Bay region. In particular, the objectives of the study
foundations design, has some disadvantages. For exam- include the following aspects:
ple, the WSD method merely depends on the safety fac-
tor, which is determined by comparing the resistance To assess and quantify uncertainties of input
with the predicted load, without taking into account parameters (soil data mainly).
of the uncertainties in both loads and resistances. To estimate the sensitivity of input parameters
Consequently, the pile foundation with greater uncer- (mainly soil strength) and obtain quantitative infor-
tainties in resistance and/or applied load may be judged mation about the effects of various uncertain fac-
as safe as that with smaller uncertainties in resistance tors on the reliability of pile foundation.
and load, as long as the means match each other. To analyze and evaluate the effects of soils param-
In order to avoid the disadvantages of WSD, the eters variations (soil strength mainly) on reliability
Load and Resistance Factor Design (LRFD) Method and partial factors.
has been introduced into pile foundation design, Several reliability analysis methods have been
which is firstly used in the reliability analysis of developed in recent years. Ruiz (1984) presented a
structure design. With the LRFD method, the uncer- First-Order Second-Moment approach for estimating
tainties in resistance and applied load can be con- the lateral deflections and moments in piles driven in
sidered in assessment of the safety of pile foundation soft clays. This paper uses a similar approach to esti-
in accordance with that of superstructure. The LRFD mate the reliability index and partial factors of safety
design method has been recommended in the API for designing pile foundation in Bohai Bay region.
Code of RP 2A-LRFD. Although the principles of the
LRFD method have been made clear in the API Code,
the design parameters, such as the partial safety fac- 2 PROCEDURE FOR RELIABILITY ANALYSIS
tors of resistance and the applied load, should be
checked with the reliability theory when the environ- 2.1 Partial factors
mental conditions and structural frame are totally dif-
ferent from the regions covered by the API Code. In the reliability analysis, soil strength and applied
Bohai Bay region is a highly developed area in load are taken as random variables, which are denoted
China for offshore oil industry. Since the geological as xi (i  1, 2, , n). The objective of the reliability
and environmental conditions in this area are quite analysis is to estimate the designed values of material

853

Copyright 2005 Taylor & Francis Group plc, London, UK


strength and applied load when the pile foundation is where  Xi is the standard deviation of variable Xi. The
at the ultimate state. With the mean and designed magnitude of i indicates the weight of the random
data, the partial factor can be expressed as: variable Xi, i.e., the greater the absolute value of the
sensitivity coefficient is, the more important the ran-
dom variable will be. If a random variable has a
(1) relatively big sensitivity coefficient, it can induce sig-
nificant change in the reliability index with a small
where variation. According to sensitivity analysis results,
i  partial factor of material i or load i; designers can find out the important factors that have
xi*  designed value of material i or load i; significant effects on the reliability index . Thus the
mi  mean value for variable xi. pile foundation can be designed to have sufficient
reliability with less cost.
From Equation (1) it can be seen that the partial
factor of a material is the degree of mobilization of
the strength of the material; and the partial factor of a 2.3 Ultimate state equation
load is the ratio of the load, which satisfies the ulti-
The calculation formula for the axial capacity of a
mate state equation, to the mean of the load. With the
axial loaded pile can be found in API RP 2A-LRFD
partial factors, designers can comprehensively assess
(1993) and can be expressed below:
the safety degree of the pile foundation.

2.2 Reliability index


(6)
The reliability index, , can be obtained with
where
(2) Qc  bearing capacity of the pile;
P  axial load applied on pile head under the
extreme environmental condition;
where z and Z are the mean and the standard devi- QS1  the outer friction of the pile;
ation of the target function Z, respectively. Equation QS2  the inner friction of the pile;
Z  0 is called the ultimate state equation, which can QP1  the soil resistance to the pile tip when the plug
be expressed as is formed;
QP2  the soil resistance to the steel ring at the pile
(3) pit;
W  the submerged weight of the pile including
where Xi (i  1, 2, , n) is the basic variables includ- the pile plug.
ing applied load and soil strength. The failure prob-
ability Pf can be expressed as The ultimate state equation (6) can be solved with the
first-order-second-moment method to find the design
(4) values of the variables, which is an iterative proce-
dure to calculate all the possible values for the vari-
ables until Z  0 is reached.
where )() is the standard normal distribution function.
Since the Pf is derived from , the value could
be adopted to evaluate the safety degree of the pile 2.4 Random field model
foundation.
In reliability analysis, it is necessary to estimate In order to calculate the reliability index, a database
the weight of each random variable. Therefore the involving plenty of soil profiles has been established
sensitivity coefficient i is introduced to quantify the by collecting abundant borehole logs in the Bohai
effect of each random variable on the reliability index. Bay region. Then based on the spatial random field
The sensitivity coefficient i, is defined as follows: theory (Yan 1995), the stationarity and ergodicity of a
soil profile in the Bohai Bay have been examined and
a random field modeling is established.
In the random field theory, the correlation function
is defined as
(5)

(7)

854

Copyright 2005 Taylor & Francis Group plc, London, UK


where 3 CASE STUDY
u  the correlation function;
z  the considered depth; 3.1 Statistical analysis
z  the interval of the depth;
For a typical platform pile foundation in the Bohai Bay
u(z)  the soil strength at depth z.
region, the adopted piles are 1.3716 m in diameter and
It can be approximately calculated with the the penetration length is 75 m in 11 soil layers. Based on
observed data using the following equation: the environmental data, the mean estimated vertical
load on pile foundation is 19000 kN under the extreme
condition with a coefficient of variation of 0.25. With
the established soil database and the spatial random
(8) field model, the means and the standard variations of
the random variables can be calculated and listed in
Table 1. All parameters including the friction angle
The variation function is related to the correlation for sand layer and undrained shear strength for clay
function in the following form: layer are assumed to follow the normal distribution.

3.2 Results and sensitivity


(9) The reliability index, partial factors and sensitivity
factors are calculated using the data presented in
Section 3.1 with the AFOSM method. The results are
Functions pu and *u are the most important param- listed in Table 24.
eters to characterize a random process and can be In Table 2 and Table 3, it can be seen that for the
converted to each other with Equation (9). Vanmarcke upper soil layers, the resistance partial factors are
(1977) presented several ways to obtain the correl- approximate to 1. This means that the skin friction
ation function. In this analysis, the correlation function resistance is fully mobilized in upper soil layers. As
is derived from an analytical expression, which can the penetration increases, the degree of mobilized pit
simulate the soil profile very well. resistance reduces, implying the skin friction is
Then the correlation distance and variation reduc- always fully mobilized before the pit resistance
tion function are obtained. With the database and the reaches its peak value.
established random model, the standard deviation of For the sensitivity efficient i of each random vari-
the soil strength is corrected with the variation reduc- able in Table 2 and Table 3, combining the correspond-
tion function. Subsequently the reliability analysis ing soil layer in Table 1, it can be seen that the thicker
can be carried out. and/or deeper the soil layer is, the bigger the sensitivity

Table 1. Statistical results of soil profile.

From To  * V* Cu** VC**

Soil layer (m) (m) (kN/m3) (deg.) (kPa)

1 0 1.7 8.0 0 0 26.3 0.29


2 1.7 3.8 8.5 17.5 0.14 0 0
3 3.8 5.7 8.7 27.5 0.09 0 0
4 5.7 13.3 8.9 0 0 53.8 0.28
5 13.3 16.0 8.9 26.7 0.09 0 0
6 16.0 20.5 9.0 0 0 64.4 0.28
7 20.5 40.7 9.3 29.7 0.16 0 0
8 40.7 57.2 9.5 0 0 104.1 0.26
9 57.2 61.7 10 25.0 0.15 0 0
10 61.7 69.1 9.8 0 0 151.6 0.28
11 69.1 75.0 9.8 28 0.13 0 0

*  and V are the mean value and coefficient of variation of friction angle between the soil and the
pile respectively.
** Cu and VC are the mean value and coefficient of variation of undrained shear strength
respectively.

855

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 2. Results of .

Variable *1 *2 *3 *4 *5 *6 *7 *8 *9 *10 *11 **
b

Design value 0 17.5 27.5 0 26.7 0 29.7 0 25.0 0 28.3 28.3


Partial factor 0 0.999 0.999 0 0.997 0 0.841 0 0.949 0 0.927 0.725
i 0 0.001 0.002 0 0.010 0 0.307 0 0.106 0 0.173 0.652

* i denotes the friction angle between soil and pile wall in i-th soil layer.
** b denotes the external friction angle between soil and pile wall located at the pile point.

Table 3. Results of CU.

Variable C*1 C*2 C*3 C*4 C*5 C*6 C*7 C*8 C*9 C*10 C*11 C*b

Design value 26.3 0 0 53.3 0 64.4 0 104.1 0 151.6 0 0


Partial factor 0.998 0 0 0.965 0 0.968 0 0.798 0 0.847 0 0
i 0.003 0 0 0.039 0 0.035 0 0.240 0 0.169 0 0

* Ci denotes undrained shear strength in i-th soil layer.


** Cb denotes undrained shear strength of soil layer located at the pile pit.

Table 4. Other results. 10


reliability index/safety

Reliability index 3.24 8


Failure probability Pf 7.8  10
4
6
factor

Design load P 28120 kN


Load partial factor p 1.48 4
reliability
Sensitivity factor of the load P
0.59 index
Safety factor* 2.2 2 safety
factor
* The safety factor is estimated by the WSD method. 0
0 10 20 30 40 50
mean value of load P(MN)
coefficient will be. The cause can be found in the cal-
culation formula of the bearing capacity of the pile Figure 1. Reliability index, safety factor vs. mean value of
foundation (API 1993). Since the thick and/or deep load P.
soil layers and the soil layer in the pile pit contribute
more resistance than others, their sensitivity coeffi-
cients are bigger than others. Therefore, the strength 6
of these soil layers has significant effect on the relia- 5
reliability index

bility of the pile foundation. 4


3
3.3 Sensitivity analysis
2
It is very important for the designers to know the 1
sensitivity of reliability to the random variables.
Therefore in this subsection, the effects of each ran- 0
0 0.15 0.3 0.45 0.6 0.75
dom variable on reliability are examined.
Firstly the mean of the load P is changed to inves- coefficient of variation of load P
tigate its influence on reliability index. The variation
Figure 2. Reliability index vs. coefficient of variation of
coefficient of load P remains a typical value of 0.25.
load P.
Others are same with those given in section 3.1. The
relationships between load P and resulting reliability
index, safety factor are shown in Figure 1. Secondly the effects of variability of applied load
According to Equation (6), the pile foundation has on reliability index and partial factors are investigated
an ultimate capacity of 41920 kN. As shown in Figure by changing the coefficient of variation of load P.
1, as load P increases to the ultimate capacity, reli- (See Figure 2 and Figure 3)
ability index and safety factor reduce to 0 and 1.0, In Figure 2, it can be seen that as variability of
respectively. the load P increases, the global reliability of pile

856

Copyright 2005 Taylor & Francis Group plc, London, UK


2 Comparing Figure 4 with Figure 2, it can be seen
that the effect of variability of soil strength on the
partial factors

1.5 reliability index is similar to that of the applied load.


load However, by comparing Figure 5 with Figure 3, it can
1 be seen that the partial factors go in different ways.
resistance That implies when the variability of bearing capacity
0.5
of subsoil rises, the global reliability drops.
0
0 0.15 0.3 0.45 0.6 0.75
coefficient of variation of load P 4 CONCLUSIONS

Figure 3. Partial factors vs. coefficient of variation of load P. In this paper, a procedure is proposed for the reliabil-
ity analysis on the axial loaded pile foundation. A
random field model of soil profile in the Bohai Bay
4 region of China is built and the AFOSM method is
3.5 used to calculate the reliability index , partial factors
reliability index

3 and sensitivity coefficient i. The analysis results can


2.5 be summarized as follows:
2
1.5 From the sensitivity analysis, it may be concluded
1 that the factors that most severely influence the
0.5 performance of bearing capacity of the pile foun-
0
dation are the applied axial load and strength of
0 0.15 0.3 0.45 0.6 0.75 thick and deep soil layers along the pile.
When the mean value of the axial load increases,
coefficient of variation of soil strength
the safety degree of the pile foundation will
Figure 4. Reliability index vs. coefficient of variability of decrease and more resistance of the subsoil will be
soil strength. mobilized. When the variability of the load
increases, the ultimate state equation can only be
satisfied by mobilizing more resistance.
2 When the variability of soil strength increases, the
partial factor of resistance will decrease. In the
partial factors

1.5 analysis, the ultimate state equation can only be


load satisfied by reducing the design load, which causes
1 the partial factor of the axial load and reliability to
decrease.
resistance
0.5

0 REFERENCES
0 0.15 0.3 0.45 0.6 0.75
coefficient of variation of soil strength
American Petroleum Institute. 1993. Recommended
Practice for Planning, Designing and Constructing Fixed
Figure 5. Partial factors vs. coefficient of variability of Offshore Platforms-Load and Resistance Factor Design,
soil strength. Washington DC: American Petroleum Institute.
Gao, D. 1989. Reliability Theory of Soil Mechanics.
Beijing: China Architecture & Building Press.
foundation decreases and hence the corresponding Ronald, K.O. 1990. Random field modeling of foundation
failure probability increases. failure modes. Journal of Geotechnical Engineering,
Figure 3 shows that if the variability of the load P 116(4): 554570.
rises, the partial factor of resistance will increase, too. Ruiz, S.E. 1984. Reliability index for offshore piles sub-
This indicates that more strength of soil has to be jected to bending. Structural Safety, 2(2): 8390.
mobilized in order to meet the ultimate state, which Yan, S., Jia, X. & Guo, H. 1995. Examination of stationarity
leads to the rise of resistance partial factor. and ergodicity on soil profile. Chinese Journal of
Geotechnical Engineering, 17(3): 19.
Finally, the effects of the variability of soil strength Vanmarcke, E.H. 1977. Probabilistic modeling of soil pro-
are studied by changing the coefficient of variation of files. Journal of the Geotechnical Engineering Division,
the soil strength from 0.1 to 0.7 while keep other vari- 103(11): 12271246.
ables unchanged. The results are illustrated in Figures
4 and Figure 5.

857

Copyright 2005 Taylor & Francis Group plc, London, UK


Numerical analysis of pile axial loading test on Ryukyu calcareous
sediments

M. Ohuchi, M. Kiyosumi & N. Umeda


Shiraishi Corporation, Tokyo, Japan

F.L. Peng
Tongji University, Shanghai, China

O. Kusakabe
Tokyo Institute of Technology, Tokyo, Japan

ABSTRACT: In order to develop a rational numerical method for predicting the bearing capacity and deform-
ation of the pile foundation on Ryukyu calcareous sediments, a series of nonlinear elasto-plastic FEM analyses
were conducted to simulate pile vertical loading tests. In the FEM analyses, the calcareous soil was modeled as an
elastic perfectly plastic material. The load and displacement relation of pile foundation was analyzed with the
different strength reduction factors for the interface between the pile and calcareous soil. The analytical results
indicated that the strength reduction factor significantly affects the load and displacement of pile foundation on
Ryukyu calcareous sediments, and that the proposed FEM could reasonably simulate the experimental results
when an appropriate strength reduction factor was used. In addition, the soil plugging effect for open-ended piles
was also examined from the FEM results.

1 INTRODUCTION A number of studies have been conducted on the pile


capacity in calcareous sands, but it is still difficult to
Calcareous materials are widely spread throughout design pile foundation properly in calcareous sediments
the worlds oceans, and Ryukyu calcareous sediments due to its complexity of geotechnical properties (Murff,
are widely distributed in the Okinawa region of Japan 1987; Poulos et al., 1988; Alba & Audibert, 1999). The
(Murff, 1987; Poulos et al., 1988; Jewell, 1993). Ryukyu authors have used elasto-plasticity FEM analysis for a
calcareous sediments are 20 to 50-m thick and overlie numerical simulation of a compressive load test for
a base layer of Shimajiri mudstones. The main charac- steel-pipe piles driven into Ryukyu calcareous sedi-
teristics of Ryukyu calcareous sediments are: 1) highly ments. The required soil constants for Ryukyu calcar-
irregular N-value distribution with depth and in hori- eous sediments were determined using SPT N-values
zontal directions, 2) multi-layered ground consisting of according to the Japanese Specifications for Highway
alternating layers of cemented and un-cemented layers, Bridges, and axial symmetry was assumed in elasto-
and 3) abundant pores and cavities. These factors have plasticity FEM analysis using a Mohr-Coulomb fail-
led to the conclusion that Ryukyu calcareous sediments ure criterion.
are unsuitable for the bearing layers for structural foun- The load-displacement curves for the pile head
dations. However, the deep base layer of Shimajiri mud- obtained based on the results of the simulations were
stones is also unsuitable for a bearing layer for structural consistent with field loading test data. Discussion will
foundations in terms of cost. Thus the establishment focus on the results of the analysis, with particular
of a practical and economical foundation design that attention to the vertical bearing capacity of the pile.
makes use of Ryukyu calcareous sediments is required Specifically, the effect of differences in modeling of
for the Okinawa region. The purpose of this study is the pile shape on pile-skin friction and pile-tip bear-
to propose such a foundation design as well as a ing capacity will be discussed in detail. The mechanism
method of numerical analysis of ground deformation, of the soil plugging effect with open-ended piles will
and also to understand ground failure mechanisms. also be examined.

859

Copyright 2005 Taylor & Francis Group plc, London, UK


N-value Pile Loading
Steel-pipe pile
0 25 50 (1,000 t19 L40,000) EL-2.6 m
-2.6 As : Gravel-sand mixtures -4.6
-5.0 Ag

29 m
-16.1
La -22.2
Ag : Silty sand and gravel
Lb
-15.0 -34.3
29 m
-16.1
La : Silty sand and gravel
EL (m)

-22.2
Sb : Sandy Silt
-25.0 -23.5 Lc
Lb : Silty sand and gravel

-35.0 -34.3 -82.6


Lc : Base
Gravel-sandy silt mixtures rock
-102.6
-45.0
100 m
Figure 1. Specifications of test pile and soil profile.
Figure 2. Finite element mesh for the present simulation.

2 PILE LOAD TEST AND FEM ANALYSIS

2.1 Pile compressive load test


Compressive load tests for a steel-pipe pile for simu-
lation were performed at the Kouri Ohashi Bridge in
Okinawa prefecture (Kamimura et al., 2001). Figure 1
shows the specifications of the test pile and the soil
profile of the test site. Layers La, Lb, and Lc in the soil
profile correspond to Ryukyu calcareous sediments.
Note that the Shimajiri mudstones are distributed below
a depth of approximately 80 m, and are not appeared
in the section.
The test pile was driven in by the impact method
using a hydraulic hammer, and terminated in a layer
mainly composed of un-cemented Ryukyu calcareous Figure 3. Modeling of the steel-pipe pile.
sediments. In practice, the pile tips were driven into
the Lb layer, where the N-value is approximately 20; the known, two models were assumed for the state of the
amount of penetration at the end of driving was approxi- steel pipe: a hollow-pile model and a solid-pile model.
mately 15 mm. To evaluate the effects of long-term In the simulations, the plate thickness of the steel-pipe
behavior, compressive load tests were performed on a piles was set at 80 mm in the case of the hollow-pile
single test pile one month after installation (Month-1 model despite an actual thickness of 17 mm, due to
test) and again after five months (Month-5 test). restrictions of the program used (PLAXIS). In the
hollow-pile model, interface elements were placed
both on the interior and exterior surfaces of the pile
2.2 FEM Model (Matsumoto & Nishijima, 1996). Figure 3 shows the
Figure 2 shows the axisymmetric finite element mesh configuration of the steel-pipe pile model. The strength
for the present simulation. The layer boundary in the characteristics of the interface elements that follow
vertical direction was set based on the soil elevation the Coulombs failure criterion shown in Eq. (1) were
of Fig. 1. Although Ryukyu calcareous sediments also reduced by the product of soil shear stress multiplied
feature highly irregular distribution in the lateral direc- by a strength reduction factor R. The symbols N,
tion, a uniform horizontal boundary was assumed soil, and csoil in Eq. (1) represent normal stress acting
for purposes of simplicity. Each element consists of upon the interface elements, the soils internal friction
15-node triangles; there are 812 elements, 6,689 nodes, angle, and the soils adhesive strength, respectively.
and 9,744 Gauss points. As boundary conditions for
the domain, the vertical and horizontal displacements at (1)
the bottom were fixed, as were the horizontal displace-
ments at the sides. Since the condition of the soil inside The soil constants were calculated assuming a Mohr-
the steel pipe at the moment the pile was driven in is not Coulomb model. Except for the base rock layer, the

860

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 1. Physical properties used in FEM analysis.

Ag La Lb Lc Base rock Hollow pile Solid pile

Average SPT N-values 6 22 26 37


Submerged unit weight s(kN/m3) 7 9 10 10 9 17.8 5.3
Poissonratio v 0.3 0.3 0.3 0.3 0.3 0.1 0.1
Youngs Modulus E (kN/m2) 1.7  104 5.9  104 7.1  104 1.0  105 4.9  106 5.5  107 1.5  107
Internal friction angle  (deg) 24.6 25.9 25.5 24.8 60
Adhesive strength c (kN/m2) 0 0 0 0 3.2  103
Dilatancy angle  (deg) 0 0 0 0 30

physical properties of the soil layers were selected Table 2. FEM analysis case.
according to the recommended methods in the Japanese
Specifications for Highway Bridges (2002). The phys- Case R Pile shape
ical properties of the base rock were evaluated to be a
P-1 0.8
layer belonging to class A in the rock classification P-2 0.6 Hollow pile
scheme presented in the Japanese Handbook of Civil P-3 0.4
Engineering (1989). Note that the dilatancy angle 
P-4 0.8
was calculated from Eq. (2). The physical properties of P-5 0.6 Solid pile
the steel-pipe pile were selected, assuming that the pile P-6 0.4
was a linear elastic body. The values of the physical
properties used in the simulation are listed in Table 1. R: Strength Reduction Factor.
Here, the weight of the steel-pipe pile per unit volume
in water and its Youngs modulus were calculated from two models for the pile shape hollow and solid and
Eq. (3), since the thickness of the steel pile in the simu- by varying the strength reduction factor R from 0.4 to
lation differs from its actual thickness in the tests. 0.6 to 0.8.
However, the contact pressure will differ in practice
even when this converted value is used, due to this
difference in thickness. In Eq. (3), A is the cross-sec- (4)
tional area of the actual steel pile, while As is that of
the model. E is the actual Youngs modulus and  is the
(5)
weight of the pile per unit volume in water (Michi &
Matsumoto, 1994).
3 RESULTS AND DISCUSSIONS
(2)
3.1 Pile-load and displacement
(3) Figure 4 shows load-displacement relations for the pile
head obtained from the tests and FEM analysis. It is
The failure criterion of the soil will follow the shown that the load at a displacement level correspond-
Mohr-Coulomb model, and the plastic strain will be ing to that of the yield load is larger in the Month-5 test
calculated according to the non-associated flow rule than in the Month-1 test. This provides confirmation
represented by Eq. (4). In the equation, symbols g and of the soil set-up phenomenon, in which an increase in
 represent the plastic potential and non-negative scalar load-bearing capacity over time is observed. In the
coefficient, respectively. Eq. (5) is the formula for plas- present simulation, the set-up phenomenon was repro-
tic potential and the symbol  stands for the dilatancy duced by increasing the strength reduction factor R.
angle. In the simulation, an initial stress analysis was The yield load increases along with an increase in
performed, after which a uniformly distributed load was strength reduction factor R. Furthermore, since the load
placed on the entire surface of the pile head, as shown at a displacement level corresponding to that of yield
in Fig. 2. The end of the simulation was set to corres- load also increases, it can be seen that initial rigidity
pond to the point at which load reached 6,000 kN. The also increases. In other words, with the hollow pile
effects of pile-shape modeling and the strength reduc- (Cases P-1, P-2 and P-3), the yield load is approxi-
tion factor R were also evaluated in the simulation. mately 1,800 kN for R  0.4 (P-3), while it increases
Table 2 shows the combination of analysis conditions to approximately 3,000 kN for R  0.8 (P-1) with
used in the simulation. An attempt was made to repro- the same pile shape. Furthermore, the initial rigidity
duce the test results in the simulations by adopting until yielding is also larger for P-1 than for P-3. This

861

Copyright 2005 Taylor & Francis Group plc, London, UK


5000 P-4 (Solid, R = 0.8) 5000
P-1 (pile-head)
P-1 (Hollow, 0.8) P-1 (pile-skin)
Pile head load P(kN)

Pile head load P(kN)


4000 P-5 (Solid, 0.6) 4000 P-1 (pile-tip)

3000 3000 Month-5 test (pile-head)

P-3 (Hollow, R = 0.4)


2000 P-6 (Solid, 0.4)
2000 Month-5 test (pile-skin)
P-2 (Hollow, 0.6)
1000 Month-1 test (pile head) 1000
Month-5 test (pile head) Month-5 test (pile-tip)
0 0
0 20 40 60 80 100 0 20 40 60 80 100
Pile head displacement S(mm) Pile head displacement S(mm)

Figure 4. Load-pile head displacement relations obtained Figure 5. Load-pile head displacement relations (Month-5
from the tests and FEM analyses. test vs. FEM for Case P-1).

5000
relationship also holds true in the cases of solid pile. P-5 (pile-head)
Month-5 test (pile-head)
From the above results, it can be seen that the strength P-5 (pile-skin)

Pile head load P(kN)


4000
reduction factor R affects both the yield load and initial P-5 (pile-tip)
rigidity before yielding. Simulation results that mostly 3000 Month-5 test (pile-skin)
consist with the Month-1 test results correspond to
Case P-2 for the hollow pile and P-6 for the solid pile. 2000
Those that best reproduced the Month-5 test results
were P-1 and P-5. In cases in which the test results were 1000
successfully reproduced, the simulated yield load and Month-5 test (pile-tip)
0
the initial rigidity before yielding showed excellent 0 20 40 60 80 100
matches with the test results. However, resistance at the Pile head displacement S(mm)
pile head can be viewed as the sum of tip resistance
and pile-skin resistance, phenomena that result from Figure 6. Load-pile head displacement relations (Month-5
two entirely different mechanisms. Therefore, the simu- test vs. FEM for Case P-5).
lation may not necessarily have reproduced the behav-
ior of the entire pile, even if the behavior of the pile
head seems consistent when comparing the simula- fairly consistent with test results. Based on the above,
tion and the test results. Thus, the simulation results it can be concluded that highly precise reproduction
were reviewed to isolate pile-skin resistance and tip of pile-head behavior, which is represented as the
resistance for comparison to the test results. Cases P-1 sum of pile-skin and pile-tip behaviors, results from
and P-5 were analyzed, as these were found to be most the precise match between the simulation results for
consistent with the Month-5 test results for the pile Case P-1 and the test results for pile-skin and pile-tip
head (Fig. 4). behaviors.
Figure 5 shows the load-pile head displacement Figure 6 shows the load-pile head displacement rela-
curves for the loads at the pile head, pile skin, and pile tions on the pile head, pile skin, and pile tip for Case
tip for Case P-1 (hollow pile). The pile-tip load was P-5, in which the pile shape is defined as the solid pile.
calculated by subtracting the pile-skin load from the The behavior of the pile in the simulation of Case P-5
pile-head load. The results of simulation for Case P-1 and the test performed 5 months after installation are
and the Month-5 test coincide precisely in terms of pile- consistent only for the pile head. The pile skin and
skin and pile-tip behaviors. In the test, the pile-head pile tip behaviors show inconsistency between the P-5
displacement upon yielding of skin and tip resistance simulation and the test results. The yield load for the
was approximately 10 mm (1% of the pile diameter) skin resistance was approximately 2,000 kN in the test
for skin resistance and 15 mm (1.5% of pile diameter) and 1,500 kN in the P-5 simulation. The yield load of
for tip resistance. This difference in mobilization was the tip resistance was approximately 1,000 kN in the
also observed in the simulation (P-1), in which tip test, while there seems to be no obvious yield point
resistance yielded after skin resistance. The behavior within a pile head displacement of 100 mm for P-5.
of the pile skin in P-1 coincides extremely well with the From the above, it is believed that the consistency
test results, and the yield load is nearly identical. In con- between the pile-head behavior in the P-5 simulation
trast, the yield load and yield displacement for the pile and in the test results was due to the canceling out of
tip are slightly larger for P-1, although these values are the differences in behavior of the pile skin and tip.

862

Copyright 2005 Taylor & Francis Group plc, London, UK


Axial force P(kN) Axial force P(kN)
0 1000 2000 3000 4000 EL=-2.6m 0 1000 2000 3000 4000 EL=-2.6 m
0 0
Ag Ag
Depth y(m)

10

Depth y(m)
10
La La
20 20
Lb
Lb
30
30
FEM (P-1) Month-5 test
FEM (P-5) Month-5 test
Figure 7. Axial load distribution of pile (Month-5 test vs.
FEM for Case P-1). Figure 8. Axial load distribution of pile (Month-5 test vs.
FEM for Case P-5).

The tip resistance estimated from two kinds of end


bearing capacity formulas for driving in the calcareous case of P-1, the differences between the axial load
sediments. One formula is in proportion to the overbur- distributions determined from the P-5 simulation and
den pressure (Poulos, 1989); the end bearing capacity from the test become larger as the load increases. The
was calculated at 1,900 kN/m2. The other formula is the difference between the two becomes significant at
exponential function of overburden pressure (Houlsby the pile tip, reaching approximately 500 kN at load of
et al., 1988); the end bearing capacity was calculated at 3,000 kN and approximately 1,000 kN at load of
5,200 kN/m2. In contrast, the tip resistance in the test at 4,000 kN. This difference also increases as the load
the pile head displacement of 100 mm was 1,400 kN/m2. increases. This is also reflected in the results of load-
The test value was smaller than two kinds of formulas displacement curves shown in Figure 6.
value. The axial load distributions of test results above
loads of 3,000 kN show that the axial load decreases
drastically from the top of the Lb layer to the pile tip.
3.2 Effect of pile shape modeling
This trend is reproduced fairly well in the P-1 simula-
We will draw a comparison between the axial load dis- tions, while the actual axial load at the pile tip increases
tributions seen in the results of the Month-5 test and more than the test results with increasing load in the
those seen in the simulations. The axial load distribu- P-5 simulation. The pile-skin resistance for Case P-1
tions have been determined for loads of 500 kN, with a hollow-pile model is the sum of the force act-
1,000 kN, 2,000 kN, 3,000 kN, and 4,000 kN. We will ing on the interior and exterior surfaces of the pile. In
compare the simulation results for Cases P-1 and P-5 contrast, only the force exerted on the exterior surface
with test results, as these cases showed successful accounts for the skin resistance of a solid pile (Case
reproduction of load-displacement behavior of the pile P-5). Thus, it is believed that the large difference in
head seen in the Month-5 test results (Fig. 4). Figure 7 the axial load at the pile tip for Cases P-1 and P-5 is
shows the axial load distribution results for the due to a difference in the mechanisms that generate
Month-5 test and the P-1 simulation. The differences skin resistance.
between these results increase as load increases. Below
2,000 kN, the axial load distribution for the P-1 simu-
3.3 Soil plugging effect for the open-ended pile
lation and the test nearly coincide in the range from
below a depth of 20 m down to the pile tip. However, Figure 9 shows the changes in load and effective nor-
when the load exceeds 3,000 kN, the test and simula- mal stress inside the pile. The indicated nodes on the
tion axial loads coincide only at the pile tip. For loads interior of the pile are situated 5.2 m above the pile tip
exceeding 4,000 kN, the axial load of the entire pile at EL 
26.4 m (point A) and 0.2 m above the pile
calculated in the P-1 simulation exceeds that of the tip at EL 
31.4 m (point B). The dashed line in the
test results. All of the above results can also be seen in figure represents the first critical resistance value of
Fig. 5. The axial load at the pile tip is equivalent to the the pile skin and pile tip, determined by plotting load
tip resistance, and the simulated behavior of the pile and displacement in a double log-scale. The effective
tip up to a pile-head load of 3,000 kN coincides pre- normal stress at point A does not deviate from the ini-
cisely with test behavior. However, at greater pile-head tial stress value, even with an increase in load. In con-
loads, the results of P-1 simulation and actual testing trast, the effective normal stress at point B does not
diverge slightly. deviate from the initial stress value until the load
The axial load distributions for the Month-5 test reaches the first critical skin resistance, but, once this
and P-5 simulations are shown in Figure 8. As in the value is exceeded, effective normal stress begins to

863

Copyright 2005 Taylor & Francis Group plc, London, UK


6000 steel-pipe piles driven into a bearing layer of Ryukyu
calcareous sediments. The following conclusions were
Pile head load P(kN)

obtained based on the results.


4000
Pile-tip: First critical resistance (1) The set-up phenomenon, in which bearing capa-
city recovers over time, was confirmed in the
2000 Pile-skin: First critical resistance
load test. In the simulation, the set-up phenome-
A point (5.2 m from pile-tip) non was reproduced by increasing the strength
B point (0.2 m from pile-tip)
0
reduction factor R.
0 300 600 900 1200 1500 (2) The results of simulation using a steel hollow-
Normalstress (kN/m2) pile model with strength reduction factor R set to
0.8 produced skin resistance and tip resistance
Figure 9. Variations in load and the effective normal stress values that were consistent with load test results
on the interior of the pile (P-1). on the load-displacement curves. Thus, pile-head
behavior (as the sum of skin and tip resistances)
4000 was also consistent.
P-1 (inside pile)
P-5 (pile-tip)
(3) The soil plugging effect was confirmed in the
3000 simulation using a hollow-pile model. The soil
Force P(kN)

plugging effect in the present simulation is


2000 exerted when the soil yields along the entire pile
shaft. Thus, incomplete soil plugging was con-
1000 firmed here.

0
0 100 200 300 400
Pile head displacement S(mm) REFERENCES

Figure 10. Load-pile head displacement curve (FEM). Alba, J. L. and Audibert, J. M. E. 1999. Pile design in
calcareous and carbonaceous granular materials: An histor-
ical overview. Engineering for Calcareous Sediments,
increase slightly. Then, after the load exceeds the first Al-Shafei(ed.), 2943. Rotterdam: Balkema.
critical tip resistance, a drastic increase is seen in Houlsby, G. T., Evans, K. and Sweeney, M. A. 1988. End bear-
effective normal stress. From the above, it can be con- ing capacity of model piles in layered carbonate soils.
cluded that the soil plugging effect appears after the Proc. Int. Cong. On Calcareous Sediments, Perth, 209214.
load exceeds the first critical resistances for both the Jewell, R. J. 1993. An introduction to calcareous sediments.
pile skin and pile tip. In other words, the skin resist- Proceeding of International Seminar in Kagoshima93,
ance on the interior of the pile due to the soil plugging 145. Kogoshima University.
Japan Society of Civil Engineers, 1989. Handbook of Civil
effect is not exerted until all of the soil surrounding Engineering I, 4th Edition, Gihodo Shuppan, 395423
the pile skin yields. (in Japanese).
The resistance produced by the soil plugging effect Kamimura,Y., Oyakawa, K. and Matayoshi, Y. 2001. Load test
is the frictional force acting on the interior surface of on steel-pipe piles with a bearing layer of Ryukyu calcare-
the pile. When this frictional force is equal to or greater ous sediments. Abstracts of the 14th Okinawa Geotechnical
than the tip-bearing capacity of a close-ended pile with Engineering Seminar, 5257 (in Japanese).
the same dimensions, the soil plugging effect is exerted Matsumoto, T. and Nishijima , Y. 1996. FEM analysis of ver-
at 100%. In contrast, when the frictional force is lower tical load test of steel-pipe pile driven into diatomaceous
than the tip-bearing capacity soil plugging is incom- mudstone. Proceedings of the 31st Geotechnical
Engineering Conf, 2:16231624 (in Japanese).
plete. Figure 10 shows the load-displacement curves Michi, Y. and Matsumoto, T. 1994. FEM analysis on static load
for the skin resistance inside the pile for Case P-1 and test of steel-pipe pile driven into diatomaceous mudstone.
the tip resistance for Case P-5. The skin resistance for The 29th Soil Engineering Conf, 2:14411442 (in
P-1 always features a smaller load relative to the tip Japanese).
resistance for P-5 at the same level of displacement. Murff, J. D. 1987. Pile capacity in calcareous sands: State of
Thus, it can be concluded that the soil blocking effect the art. Journal of Geotechnical Engineering 113(5):
is incomplete for P-1. 490507.
Poulos, H. G. 1989. Pile behaviour-theory and application.
Geotechnique 39, No.3, 365415.
4 CONCLUSION Poulos, H. G., Randolph, M. E. and Semple, R. M. 1988.
Evaluation of pile friction from conductor tests. Proc. Int.
Simulations were performed by axisymmetric elasto- Cong. On Calcareous Sediments, Perth, 599605.
plastic FEM analysis based on the axial load test of

864

Copyright 2005 Taylor & Francis Group plc, London, UK


A preliminary investigation into the effect of axial load on piles subjected to
lateral soil movement

W.D. Guo & E.H. Ghee


Griffith University, Gold Coast, Australia

ABSTRACT: A new experimental apparatus was developed, which allows lateral soil movements and vertical
load to be applied simultaneously to a pile. With this apparatus, a large number of tests were undertaken on instru-
mented piles embedded in sand. In this paper, based on four test results, preliminary analysis is presented to
demonstrate the increase in bending moment, shear force, soil reaction, and the change of pile deflection mode
due to axial load on the pile.

1 INTRODUCTION strain gauges, and LVDTs, which are then recorded


via a data acquisition system.
Pile foundations designed to support offshore plat- In this paper, typical test results are presented for
forms, structures and services are often subjected to instrumented piles that were embedded in sand and
lateral soil movements and axial loads simultan- subjected to uniform lateral soil movement and axial
eously. There have been many studies on piles sub- load. The influence of the thickness of the movable
jected to vertical loads, and some research on piles soil layer, and axial load on the pile response was
subjected to lateral soil movements. However, little investigated.
information is available for evaluating the response of
vertically loaded piles due to soil movement. This
response is important. In particular, for offshore foun- 2 TEST DESCRIPTION
dations, API (2000) specifies that the possibility of soil
movement against foundations should be investigated Information regarding the tests has been reported pre-
and the forces caused by such movements, if antici- viously by Guo & Ghee (2004). Thus only relevant
pated, should be considered in the design. Studies on parts are briefly introduced herein.
piles due to liquefaction induced lateral soil move-
ment have found that axial load on the piles can cause
(1) additional bending moment; (2) additional com-
2.1 Shear box
pression stress, and (3) additional lateral displace-
ment (e.g. Bhattacharya, 2003). On the other hand, The shear box has internal dimensions of 1 m by 1 m,
lateral soil movement may increase the axial capacity and 0.8 m in height as shown in Figure 1. The upper
of piles (Chen et al, 2002). The increase is similar to movable part of the box consists of the desired num-
that noted for a laterally loaded pile subjected to axial ber of 25 mm thick square laminar aluminium frames
load in experiments by Anagnostopoulos & to achieve a thickness of Lm ( 400 mm). They are
Georgiadis (1993) and numerical simulations by moved together by a rectangular loading block (for
Trochanis et al (1991). this research) in order to generate a uniform lateral
With the support of the Australian Research Council, soil movement. The lower fixed section of the box is a
a new apparatus was developed to investigate the timber box 400 mm in height, and a desired number
response of a pile subjected to axial load and lateral soil of laminar aluminium frames to achieve a stable sand
movement. The apparatus mainly consists of a shear layer of thickness Ls ( 400 mm). The rate of the
box, and a loading system that allow different soil movement of the upper shear box (thus the soil) is
movement profiles and vertical loading to be applied controlled by a hydraulic pump, and a flow control
simultaneously. The responses are measured using valve.

865

Copyright 2005 Taylor & Francis Group plc, London, UK


Vertical
column Height adjustable
bridge
Roller bearing
Vertical jack Figure 2. An instrumented model pile.

Loading block
Weight Axial Load
LVDT Laminar aluminium Test Lm Ls Axial
Model frames ws
pile
(mm) (mm) load (N)

Lm
1 200 600 0 Lm
Lateral 2 200 600 294 L Soil

Ls
jack 3 400 400 0 Ls Movement
400

100
4 400 400 294 Pile
Model pile Fixed timber box
(a) Sectional view Figure 3. Tests on a pile subjected to a uniform lateral soil
movement together with different axial load.

fabricated from a fixed timber piece of 18 mm in


thickness underlaid by a moveable 6 mm thick plastic
Height adjustable piece. Both pieces were perforated with 5 mm diameter
bridge for vertical jack holes in a 35 mm  50 mm grid pattern. Sand density
Vertical column was controlled by selecting different falling heights.
For the current test, a dry density of 16.27 kN/m3 (a
Data acquisition falling height of 600 mm) was used, which corre-
system
Secondary frame sponds to a relative density index of 0.89, and an inter-
for LVDTs nal frictional angle of 38 (Guo and Ghee, 2004). The
Laminar aluminium shear modulus at the middle depth of the shear box
frames (shear was about 220 kPa as determined from oedometer test.
box upper section) The model pile used in the tests was made of alu-
minium tube, 1200 mm in length, 32 mm in outer
Fixed timber box diameter, and 1.5 mm in wall thickness. The calculated
bending stiffness of the pile was 1.28  106 kNmm2.
(b) Photograph of sectional view The pile was instrumented with ten pairs of strain
gauges mounted along the shaft as shown in Figure 2.
Figure 1. Testing apparatus for the model pile tests.
2.3 Loading conditions in the tests
As shown in Figure 1a, a vertical jack was used to
drive the pile into the shear box. In order to provide a A number of tests have been conducted on the model
free-headed pile condition, axial load was applied pile subjected to a uniform lateral soil movement and
using three weights of 10 kg each. The load selected axial load. Presented here are four tests on the pile
was less than 60% the ultimate capacity of the pile with an embedment length, L of 700 mm. Figure 3
determined using the recorded jack-in pressure of provides a summary of the tests.
3.45 MPa. These weights were secured on the pile The pile was tested without any constraint but soil
head by using a connector, and they were about resistance. The uniform soil movement was applied
500 mm above the sand surface. using the rectangular loading block at an increment of
10 mm until a total of 60 mm was reached. Strain
gauge readings have been used to obtain a bending
2.2 Sand, model pile & experimental procedure
moment profile for each increment of soil movement.
The sand used in this study was an oven dried
medium grained quartz, Queensland sand. It was well
graded with little to no fines. In order to maintain a 3 TEST RESULTS
reasonably uniform density of the sand within the
shear box and from one test to another, a sand rainer A spreadsheet program was written to analyse and
was used. The rainer was fabricated from plywood, process the data obtained from strain gauges. The
having internal dimensions of 1 m by 1 m, and a inclination and deflection profiles along the pile were
height of 150 mm. The base of the sand rainer was derived, respectively from 1st and 2nd order numerical

866

Copyright 2005 Taylor & Francis Group plc, London, UK


Bending moment (Nmm) Shear force (N)
0 10000 20000 30000 -150 -100 -50 0 50 100 150
0 0

4(a) Bending moment


200
200 4(b) Shear force

Depth (mm)
Depth (mm)

400
400

10
20 600
600 10 30
20
30 60
60 800
800
Soil reaction (N/mm) Rotation (dy/dx)
-1 -0.5 0 0.5 0 0.005 0.01 0.015 0.02
0 0

4(c) Soil reaction


200 200
4(d) Rotation
Depth (mm)

Depth, x (mm)

400 400

10
600 10
20 600 20
30
30
60
800 60
800

Deflection, y (mm)
-10 -5 0 5 10
0

200
Depth, x (mm)

ws = 10

400
4(e) Deflection

10
ws = 20 600 20
30
60
800

Figure 4. Response of the pile during Test 1 (No axial load).

integration (trapezoidal rule) of the bending moment reaction, rotation and deflection were presented sub-
profiles. The profiles of shear force, and the soil reac- sequently for each test.
tion were derived by single and double numerical
differentiation (finite difference method) of the bend-
3.1 Test 1 (Lm  200 mm, no axial load)
ing moment profiles, respectively. The numerical
integration and differentiation methods were found to For Test 1, Figure 4 presented the fiveprofiles for a soil
offer consistent results presented in this paper. The movement, ws of 10, 20, 30 and 60 mm. The maximum
five profiles of bending moment, shear force, soil bending moment, Mmax occurs at a depth of 460 mm

867

Copyright 2005 Taylor & Francis Group plc, London, UK


Bending moment (Nmm) Shear force (N)
0 10000 20000 30000 40000 50000 -200 -100 0 100 200
0 0

5(a) Bending moment


5(b) Shear force
200
200

Depth (mm)
Depth (mm)

400
400

10
10 20 600
600
20 30
30 60
60 800
800

Soil reaction (N/mm) Rotation (dy/dx)


-1.5 -1 -0.5 0 0.5 1 0 0.008 0.016 0.024
0 0

5(c) Soil reaction


200 200
Depth, x (mm)

5(d) Rotation
Depth (mm)

400 400

10 10
600 600
20 20
30 30
60 60
800 800

Deflection, y (mm)
-2 0 2 4 6 8 10
0

5(e) Deflection
200
Depth, x (mm)

400 ws = 10, 20, 30, 60

10
600 20
30
60
800

Figure 5. Response of the pile during Test 2 (Axial load  294 N).

(Fig. 4a) [At this depth, the shear force is indeed zero
3.2 Test 2 (Lm  200 mm, axial load  294 N)
(Fig. 4b)]. The bending moment, shear force and soil
reaction reach maxima and remain constant beyond ws Test 2 was performed under identical conditions to Test
of 30 mm (ws/d  0.94, where d  pile diameter) (Fig. 1, but with an axial load of 294 N applied at the head.
4a, b, c). Figure 4e indicates that the pile rotates about a Figure 5 provides the same five profiles for the four
depth of 190 mm at ws  010 mm, then translates at different values of soil movement. Figure 5a indicates
ws  1030 mm, and finally remains stationary for that the depth of Mmax is 350 mm, against 460 mm
ws  3060 mm. The pile head deflection of 6.6 mm is observed in Test 1 (Table 1). The bending moment,
well below the corresponding ws of 60 mm. Thus the soil shear force and soil reaction reach maximum values at
in the sliding layer (Lm) flowed around the pile. ws of 20 mm (ws/d  0.625) compared with 30 mm

868

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 1. Summary of test results (ws  60 mm).

Max. soil reaction (N/mm)


Test Depth of Max. shear Max. pile
no. Mmax. (mm) Mmax (Nm) force (N) def. (mm) In layer Lm In layer Ls

1 460 22.53 92.0 6.6 0.34


0.70
2 350 35.86 145.0 9.6 0.63
1.10
3 520 70.00 399.0 67.6 2.25
3.15
4 475 99.48 421.0 72.2 2.25
3.80

Bending moment (Nmm) Shear Force (N)


-40000 0 40000 80000 -500 -300 -100 100 300 500
0 0
10 10
20 20
30 30
40 40
200 50 200 50
60 60

Depth (mm)
Depth (mm)

400 400 6(b) Shear force

600 600

6(a) Bending moment


800 800

Soil Reaction (N/mm)


Rotation (dy/dx)
-4 -3 -2 -1 0 1 2 3
0 0 0.05 0.1 0.15 0.2 0.25
10 0
20
30
40
50 200
60 200
6(d) Rotation
Depth, x (mm)
Depth (mm)

400 400

6(c) Soil reaction


10
600 600 20
30
40
50
60
800 800

Deflection,y (mm)
-120 -80 -40 0 40 80
0
6(e) Deflection

200
Depth, x (mm)

400
ws = 10, 20, 30,
40,50, 60
10
600 20
30
40
50
60
800

Figure 6. Response of the pile during Test 3 (No axial load).

869

Copyright 2005 Taylor & Francis Group plc, London, UK


noted in Test 1. The pile rotates about a depth of The bending moment, shear force and soil reaction
580640 mm. At ground level, the deflection is increase with the soil movement, even at the maximum
89.5 mm, which is higher than for Test 1. soil displacement of 60 mm. Similar to the failure
mode proposed by Viggiani (1981), the rotation and
deflection profiles indicate that the pile deflects in a
3.3 Test 3 (Lm  400 mm, axial load  0 N)
rigid mode. It rotates (ws  020 mm), translates
The effect of varying the sliding depth, Lm was exam- (2030 mm), and then translates and rotates
ined in Test 3 by increasing Lm to 400 mm from (3060 mm). The pile deflects more at the surface
200 mm in Tests 1 & 2. The same five profiles are pre- than the soil movement, thus shear resistance in the
sented in Figure 6. Mmax occurs at a depth of 520 mm. opposite direction is noted.

Bending moment (Nmm) Shear Force (N)


0 40000 80000 120000 -500 -300 -100 100 300 500
0 0
10 10
20 20
30 30
40 200 40
200 50
50
60 60

Depth (mm)
Depth (mm)

400
400

600
600
7(b) Shear force
7(a) Bending moment
800
800

Soil Reaction (N/mm) Rotation (dy/dx)


-4 -3 -2 -1 0 1 2 3 0 0.04 0.08 0.12 0.16
0
10 0
20
30
40 200
50 200
Depth (mm)

60
Depth, x (mm)

400
400 7(d) Rotation

10
600
20
600 30
7(c) Soil reaction
40
800 50
60
800
Deflection, y (mm)
-20 0 20 40 60 80
0

200
Depth, x (mm)

ws = 10, 20,
400 30, 40, 50, 60

10
20
600 30
40
7(e) Deflection 50
60
800

Figure 7. Response of the pile during Test 4 (Axial load  294 N).

870

Copyright 2005 Taylor & Francis Group plc, London, UK


3.4 Test 4 (Lm  400 mm, axial load  294 N) 4 DISCUSSION
Test 4 was also performed with an Lm value of 400 mm, Table 1 compares the critical responses of maximum
but with an axial load of 294 N. The responses are bending moment, shear force, and maximum pile
shown in Figure 7. The depth of maximum bending deflection. Figure 8 shows the development of the
moment reduces to 475 mm against 520 mm for Test 3. responses with soil movement. An increase in the
The bending moment, shear force and soil reaction axial load from 0 to 294 N on the pile-head leads to:
increase with the soil movement even at ws of 60 mm. (1) a 50% increase in Mmax; (2) a 25% decrease in the
The pile rotates rigidly above a depth of 590 mm (Fig. depth of Mmax; (3) a 30% lower soil movement at
7e). The pile displacement at the surface is greater than which the soil reactions reach maximum. When an
the soil movement ws once ws  40 mm. axial load is applied, (i) at Lm/Ls  0.5, the pile
behaves as if rigid; (ii) the pile head may deflect more
100 than the soil movement, and (iii) the pile deflects
Test 1
Test 2 mainly by rotation, rather than by translation as was
Test 3
80 Test 4
the case with no axial load.
Stable soil layer, Ls
Mmax (Nm)

60
5 CONCLUSIONS
40
Tests on model piles subjected to simultaneous uniform
lateral soil movement, and axial load were undertaken.
20
Due to an axial load on the pile-head, preliminary analy-
sis shows an increase in bending moment, shear force,
0 soil reaction, and the change of pile deflection mode.
0 10 20 30 40 50 60 More research is currently under way.
Soil movement, ws (mm)
450
ACKNOWLEDGEMENTS
Stable soil layer, Ls
300
Max. shear force (N/mm)

The work reported herein is currently supported by


150 Australian Research Council (DP0209027). The finan-
cial assistance is gratefully acknowledged. The authors
0 would like to thank Maria Jansson and Hongyu Qin for
0 10 20 30 40 50 60 their contribution to the experiment.
-150
Moving soil layer, Lm
-300 Test 1
Test 2
REFERENCES
Test 3
-450
Test 4 American Petroleum Institute, 2000. Recommended prac-
Soil movement, ws (mm) tice for planning, designing and constructing fixed off-
shore platforms-working stress design, API RP 2A-WSD.
2.4 Anagnostopoulos, C. & Georgiadis, M., 1993, Interaction of
1.6 Moving soil layer, Lm axial and lateral pile responses. J. of Geotechnical
Engineering Vol. 119, No.4: 793798.
Max. soil reaction (N/mm)

0.8 Bhattacharya, S., 2003. Pile instability during earthquake


liquefaction. PhD Thesis, University of Cambridge.
0 Chen C.C., Takahashi, A. & Kusakabe O., 2002 Change in
0 10 20 30 40 50 60 vertical bearing capacity of pile due to horizontal ground
-0.8
movement. Physical Modelling in Geotechnics: 459464.
Stable soil layer, Ls
-1.6 Guo, W. D. & Ghee, E. H., 2004. Model tests on single piles
in sand subjected to lateral soil movement. Proc. of the
-2.4 Test 1 18th Australasian Conf. on the Mechanics of Structures
Test 2
Test 3 and Materials, Vol. 2: 9971003.
-3.2 Test 4 Trochanis, A. M., Bielak, J. & Christiano, P., 1991 Three
-4 dimensional nonlinear study of piles, J. of Geotechnical
Soil movement, ws (mm) Engineering, Vol. 117: No. 3, 448466.

Figure 8. Comparison of responses in each test.

871

Copyright 2005 Taylor & Francis Group plc, London, UK


Modelling combined loading of piles with local interacting
yield surfaces

N.H. Levy, I. Einav & M.F. Randolph


Centre for Offshore Foundation Systems, University of Western Australia, Nedlands, Australia

ABSTRACT: The current design practice for piles subjected to inclined loads is to consider the axial and lateral
pile behaviour separately. Each of these directions is represented by a separate differential equation that is solved
independently. However, this procedure does not consider the combined effects of the different load directions.
In this paper these differential equations are coupled using a concept of local interacting plasticity models and
are thus solved simultaneously to determine the behaviour of a pile subjected to combined loading. The coupling
is introduced using the concept of local elasto-plasticity yield surfaces, simulating the interaction between the soil
responses in the two perpendicular directions. At this stage of the work elliptical yield surfaces have been incorp-
orated. Although an ellipse may not be the ideal shape, it still enables several well-known physical observations
to be captured. This model can be refined in the future to represent the local response of soil more accurately.

1 INTRODUCTION 2 BACKGROUND

Large diameter steel pipes are used offshore for pile The basic governing equation for laterally loaded
foundations and also for conductors. The latter are piles was defined by Hetenyi (1946) for a free-ended
installed vertically into the seabed to guide the drilling beam of infinite or finite length. Adopting the
and contain drilling and hydrocarbon fluids. The loads assumption that small displacements are imposed for
applied to these quasi-vertical pipes differ greatly each increment, the influence of a vertical load can be
from typical onshore situations, with more substantial neglected and the equation expressed as follows:
lateral and torsional loads. Current design practice
involves separate analysis of the axial and lateral
responses of piles and conductors, and thus does not (1)
consider the effects of interaction between the different
load directions. The overall objective of the work
where Ep is the Youngs modulus of the pile; Ip repre-
described here is to develop a new design approach
sents the second moment of area of the pile; kr(z) is
that considers the interaction between each mode of
the horizontal Winkler modulus of subgrade reaction;
loading in terms of the local soil response.
ur(z) denotes the horizontal displacement of the pile;
The behaviour of a pile subject to axial and lateral
and z is the depth below pile head. The modulus of
loading has been investigated experimentally (Sastry &
subgrade reaction is a function of both the pile and
Meyerhof 1990, Anagnostopoulos & Georgiadis 1993)
soil characteristics.
and numerically (Shahrour & Meimon 1991). These
The basic governing equation for axially loaded
results suggest that the axial response is strongly influ-
piles, assuming that radial pile displacements are
enced by lateral loading but the lateral response is
minor, is given by Scott (1981):
much less affected by axial loading.
This paper presents a model that predicts the
behaviour of a single vertical pile under combined (2)
axial and lateral loading. The model considers the
interaction between the two directions using local
elasto-plasticity yield surfaces that allows coupling of where Ap is the cross sectional area of the pile; kz(z)
the differential equations for axial and lateral loading; is the vertical Winkler modulus of subgrade reaction;
thus, the solution of these two equations is obtained and uz(z) denotes the vertical displacement of the
simultaneously. soil.

873

Copyright 2005 Taylor & Francis Group plc, London, UK


Equations (1) and (2) can be solved independently directions are expressed as functions of the change in
subject to a set of boundary conditions that are the elastic displacements from Equations (5), as:
defined for the system. The example considered for
the current model is a free headed pile with an end (6a)
bearing capacity and subject to displacement control
at the pile head. Therefore, this system must satisfy
the conditions in the lateral direction ur(0)  ur0; ur (6b)
(0)  0; ur (L)  0; u
r (L)  0 and in the axial direc-
tion uz(0)  uz0; EpApuz(L) 
fzbase(uz(L)) where
It should be remembered that all displacements,
L is the pile length and ur0 and uz0 designate the
moduli of subgrade reaction and ultimate soil reaction
applied lateral and axial displacements at the pile
forces are a function of depth below the pile head. The
head. The base force, fzbase(uz(L)), is given by
soil reaction forces along the pile shaft are functions
of depth and the displacements defined in Equations
(3) (4a), (4b), (6a) and (6b).
The local resistance forces in the soil in the two
with a Winkler modulus of kzbase and an ultimate bear- loading directions are coupled by an elliptical yield
ing force value of fuzbase. function given by the following:
Note that in the above models the soil resistance in
the lateral and axial directions respectively is a func-
tion of the total displacement at a point z: (7)

(4a)
It should be noted that at each depth the value of y
is a function of z, ure, and uze. Under purely horizon-
(4b) tal or vertical displacement the force required for
yield will correspond to the ultimate lateral or axial
soil resistance respectively. The model of the soil-pile
3 THE MODEL interaction system can be represented schematically
as in Figure 1.
In the current model the soil behaviour is assumed The bold lines represent the typical local reaction
linearly elastic perfectly plastic, incorporating elliptical force (per unit length) paths along the two perpendicu-
local yield surfaces. This behaviour is defined by two lar directions; where yielding occurs, local plastic
parameters (one for the elastic part and one for the flow alters the trend of these lines. The plastic flow is
plastic part) in each direction: the modulus of sub-
grade reaction (kr(z) or kz(z)) and the ultimate soil
resistance force (fur(z) or fuz(z)). In this section the
model and the concept of local interaction yield sur- z
faces will be described. fuz(z)
The elasto-plasticity treatment of the local inter-
fur(z)
action models requires an incremental solution of the z
two differential equations. For that purpose, Equations
(1) and (2) may be expressed in incremental form as
follows:
L

(5a)

(5b)

where the superscript p denotes the plastic displace-


ment and  represents the change in the variable over
a time step. Comparing with Equations (4), the
change in the soil resistances in the lateral and axial Figure 1. Schematic representation of the interaction model.

874

Copyright 2005 Taylor & Francis Group plc, London, UK


the product of the local consistency condition and numerical scheme was incorporated in Mathematica
flow rule. 4.2, making use of its built-in numerical differential
For the yield surface defined by Equation (7) the solver. Due to computer memory limitations the
consistency condition is given by: continuous displacement functions calculated by
Mathematica at each time step could not be retained.
Therefore, points at a spacing of L/40 were used to
(8)
interpolate the plastic strains in each direction and from
these the elastic strains required for the next time step.
and the associated flow rule is defined as:
4 PARAMETRIC STUDY
(9a)
A study was carried out to investigate the influence of
the load inclination on the pile behaviour using the pile
properties defined as Case 1 in Table 1. Displacement
(9b) control was used and displacements were applied at 0,
30, 45, 60 and 90 degrees to the horizontal.
A further analysis was then completed on the effect
for the two components, where  is the non-negative of varying the pile diameter, flexibility and length for
plasticity multiplier. Combining Equations (7), (8) displacements at 0, 45 and 90 degrees to the horizon-
and (9) and solving for  gives the expression: tal. The pile properties considered for the analysis are
summarised in Table 1.
The soil stiffness and ultimate capacity values esti-
mated for the springs in the lateral and axial direc-
(10) tions were calculated according to the equations
given in Table 2.
The pile and soil properties referred to in Table 2
are the pile diameter D (see Table 1); the undrained
shear strength su, soil shear modulus Gs and soil
When Equations (6)(10) are incremented over
Youngs modulus Es, each being a function of depth;
finite time steps, some drifting from the true yield sur-
the soil shear modulus Gb at base of pile; the shaft
faces may occur, resulting in non-zero values of y. To
interface roughness factor  (assumed as 0.5); the
minimise this numerical effect it is possible to expand
the Taylors series for yield surfaces around zero,
neglecting second order terms and above, to give: Table 1. Pile properties.

Diameter Wall Thickness Length


(11) Case m m m

1 2 solid 20
in which case: 2 1 solid 20
3 2 0.02 20
4 2 solid 10

(12)
Table 2. Formulae for determining soil properties.

Property Formula Reference


where y* is the value calculated from Equation (7).
kr ** (Ashford & Juirnarongrit 2003)
The system of differential equations can then be fur 10.8 Dsu (Randolph & Houlsby 1984)
solved using the elastic displacement values from the kz DGs (Fleming et al. 1992)
previous time step to calculate fr and fz. These values fuz Dsu (Fleming et al. 1992)
can then be substituted into Equation (7) to determine kzbase 2DGb/(1-v) (Fleming et al. 1992)
where the soil is yielding along the length of the pile. If fuzbase D2Ncsu (Fleming et al. 1992)
the soil is at yield then  is expressed as a function of
the changes in displacement for the current time step
and consequently both differential equations contain
the change in displacements in both directions. This

875

Copyright 2005 Taylor & Francis Group plc, London, UK


Poissons ratio of the soil  (assumed as 0.4); and the The elliptical model shows a minor difference (less
end bearing capacity factor Nc (assumed as 9). The than 1% at the pile base) in the displacement profile
reference diameter Dref in the expression for kr was along the pile for different loading inclinations. The
taken as 1 m. An undrained shear strength of 20 kPa axial force in the pile (V(z)) as a ratio of the axial
was assumed at the ground surface increasing at a rate force applied at the pile head (V(0)) does vary with
of 4 kPa/m with depth and the soil shear modulus was the load inclination as illustrated in Figure 4. The
taken as 10700 kPa (Youngs modulus of 30000 kPa) (normalised) axial force in the upper part of the pile
at the soil surface increasing by 357 kPa/m (Youngs increases as more horizontal load is applied, exceeding
modulus gradient of 1000 kPa/m). the pile head force for a  value of 30. This implies
that the soil is providing reduced resistance due to the
disturbance caused by lateral loading. It has been sug-
5 RESULTS gested that the increase in the axial force is due to the
friction forces caused by the upward movement of the
A displacement of 30 mm was applied at the top of soil in front of the pile (Shahrour & Meimon 1991).
the pile in the direction of interest, e.g. for a resultant This is consistent with the wedge failure mechanism
displacement at 30 degrees to the horizontal a combin- proposed by Murff & Hamilton (1993).
ation of 17 mm horizontally and 30 mm vertically was In the present analysis, although there is no concept
considered to explore the influence of lateral loading of a continuum wedge of soil moving upwards as the
on the axial pile behaviour. pile fails under lateral loading, upward motion of each
A set of five yield surfaces along the pile (at inter- soil element occurs because of the elliptical yield sur-
vals of L/4 along the pile) are superimposed together face in lateral:axial load space. Thus as an element
with the corresponding relative horizontal and verti-
cal forces as they change with time in Figure 2.
Figure 3 shows the yield surface and force path at -100
the pile head. Note that at this point only the dis-
placement rate is controlled such that during purely
axial force (kN)

elastic displacement the angle shown in Figure 3 -600 600


corresponds to the angle of the applied load. The

equilibrium point on the yield surface, where the pile
head behaviour becomes fully plastic, is reached
when the following condition is satisfied:

100
(13)
lateral force (kN)

where  is the angle of the specified displacement at Figure 3. Local yield surface and force progression at the
the pile head and  is the angle of the soil resistance pile cap.
force at the final equilibrium point.
normalised axial force, V(z)/V(0)
-400 0.4 0.6 0.8 1
0

 = 90
axial force (kN)

5
-2500 2500
 = 60
depth, z (m)

0 L/4 10

L/2
3L/4 15
L  = 45
400
 = 30
lateral force (kN) 20

Figure 2. Typical yield surfaces at intervals along the pile. Figure 4. Normalised axial force versus depth.

876

Copyright 2005 Taylor & Francis Group plc, London, UK


reaches the yield surface, the displacement vector will axial load behaviour for all cases is plotted against the
initially contain a significant component of axial depth normalised by pile length in Figure 7 for verti-
motion, gradually reducing as the load path travels cal loading and a displacement inclination of 45. The
around the ellipse (Fig. 2). In addition, the interaction vertical loads imposed to achieve yielding along the
approach leads to reduced axial resistance for elements entire pile shaft are given in Table 3 for the two load-
in the upper part of the pile that are loaded primarily ing angles, along with the percentage reduction for
in the lateral direction. inclined loading.
The pile head displacement in the vertical direction The pile diameter was reduced to 1 m for Case 2
versus the applied axial force is illustrated in Figure 5. and the results of this analysis showed a reduction in
This plot illustrates that for a lower inclination angle the axial loads and deflections along the pile. As for
the vertical force required to displace the pile is
reduced. This is due to the reduced axial load carrying 3500
capacity of the soil in the top part of the pile, as shown  = 0
= 30
in Figure 4. 3000
The point at which the curves in Figure 5 rapidly
2500

lateral force (kN)


change slope corresponds to the time step where the
incremental shaft capacity of the piles becomes zero.  = 45
2000
This means that the remaining stiffness is due to the  = 60
end bearing resistance of the pile (which is purely 1500
elastic until the end bearing capacity is reached).
When an inclined load is applied the axial force at 1000
which this transition occurs is similar for all inclin-
ations (in the current study) and this force is a reduc- 500
tion of 11.5% from the value for purely axial loading. 0
From this point the curves for inclined loading appear 0 0.01 0.02 0.03
to diverge, with a reduction of almost 1000 kN (13%) lateral deflection (m)
between 90 and 30 at 30 mm displacement.
The results of the analysis showed no significant Figure 6. Pile head lateral force versus deflection.
change in the deflection and bending moment pro-
files with depth when the inclination of the load was
normalised axial load, V(z)/V(0)
varied. The lateral force versus displacement profile
at the pile head is shown by Figure 6. As for axial 0 0.2 0.4 0.6 0.8 1
loading the curves diverge with different inclinations 0
with a reduction of 336 kN (10%) from 0 to 60 at  = 45
30 mm displacement. 0.2
normalised depth, z/L

A parametric study was carried out on the pile prop-  = 90 Case 1


erties for the cases listed in Table 1. The (normalised) 0.4
Case 2
0.6
8000
 = 60 Case 3
0.8
6000 Case 4
 = 90  = 30 1
axial force (kN)

Figure 7. Normalised axial load versus normalised depth


4000 for Case 1 to 4.

 = 45
Table 3. Axial forces at pile head for shaft yielding.
2000
Force (  90) Force (  45)
Case kN kN % reduction
0
0 0.01 0.02 0.03 1 4121 4656 11.5
2 2089 2272 8.0
axial deflection (m)
3 4017 4278 6.1
4 1507 1771 14.9
Figure 5. Pile head axial force versus deflection.

877

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 4. Lateral forces required to move pile head 30 mm. the pile head for inclined loading. Previous research
suggests that this influence should be minimal and as
Force (  0) Force (  45) discussed below this model can be modified to reflect
Case kN KN % reduction this condition better.
A parametric study completed, varying the pile
1 3019 3222 6.3
2 841 884 4.9 dimensions, suggests that the stiffer the pile the
3 1508 1595 5.4 greater the influence of lateral load on the axial pile
4 2115 2303 8.2 behaviour and vice versa. For a stiffer pile a longer
zone of influence was observed due to lateral loading
and also a greater reduction in the vertical pile head
Case 1 the deflection profile was similar for the verti- force required to displace the pile vertically.
cal and 45 inclined loading. The axial loads were Previous research suggests that a greater influence
similar for purely vertical loading in the two cases but is expected from the lateral loading on the axial pile
for the inclined load smaller loads were predicted for behaviour and less influence on the lateral behaviour
the smaller diameter pile, as illustrated in Figure 7. from axial loading (Shahrour & Meimon 1991,
This suggests reduced disturbance of the soil due to Anagnostopoulos & Georgiadis 1993). The model
lateral loading. A reduction in the pile head vertical described in this paper is presented as a starting point
force required for shaft yielding of 8.0% was observed from which a more rigorous model may be developed.
for Case 2. This model has the following limiting assumtions:
In Case 3 the pile stiffness is decreased while the second order geometric effects (the P- effect)
same outer diameter is maintained by considering a have not been included in the differential equation
thin-walled pile with wall thickness of 20 mm. In this for lateral loading;
case more pile compression is experienced and there- for a short pile, shearing at the pile tip has not been
fore smaller displacements and internal forces occur considered;
in the vertical direction. A 5% decrease was observed strength degradation in the axial direction due to
in the pile base deflection in Case 3 between the ver- softening of the soil after yield has not been taken
tical and inclined loading. The applied vertical load into consideration;
required to yield along the shaft was reduced by 6.1% the yield surface has been considered as an ellipse
for the thin-walled pile. where this shape would be flatter near the vertical
The pile considered in Case 4 had a shorter length axis and more convex approaching the lateral
of 10 m, which makes the pile stiffer. A higher load direction (i.e. a change in the lateral force would
was calculated for both the vertical and inclined loads have a greater influence on the axial force but not
and the zone of influence of the lateral loading extended vice versa);
deeper than for the other three cases, as illustrated in a linearly elastic perfectly plastic spring has been
Figure 5. For the stiffer pile the reduction in the verti- considered in both directions, which considers
cal force at shaft yield was reduced by 14.9%. only one yield surface. The inclusion of multiple
The lateral deflections and bending moment pro- yield surfaces would model gradual yielding of the
files with depth were similar for purely horizontal and soil more accurately;
inclined loads for all four cases. The pile head forces coupling between the two directions of loading has
required to displace the pile 30 mm horizontally and been modelled only based on plasticity using the
the percentage reduction due to inclined loading are concept of local yield surfaces. Further coupling
given in Table 4. These results suggest that for greater may be attributed to the interaction between the
pile stiffness the reduction in force due to an inclined soil springs due to the local moments in the soil
load is increased. resulting from inclined distortion of the pile;
torsional loading has not been considered; and
the model does not allow for cyclic loading.
6 CONCLUSIONS
The above limitations could all potentially be
The results discussed above show that the procedure addressed by modifying the differential equations or the
has the potential to yield useful data. A definite effect local yield surface criteria. With the aid of the results
on the axial behaviour of the pile is observed when of previous investigations and physical test results the
the load inclination tends towards the horizontal. A real pile behaviour could be modelled more accur-
reduction in the axial force required to displace the ately. These additional effects could conveniently be
pile head vertically was observed for inclined loads. described using the energy based variational methods
The influence of axial loading on the lateral pile of (Einav 2005a, 2005b), in which the dissipative
behaviour was not as prominent. A reduction was nature of soil is addressed. The differential equations
observed in the horizontal load required to displace given as Equations (5) can be recovered directly from

878

Copyright 2005 Taylor & Francis Group plc, London, UK


these equations and this model can be more easily Einav, I. 2005a. Energy and Variational Principles for Piles
solved while also providing an excellent approach for in Disspative Soil. Geotechnique (accepted).
cyclic loading (Einav 2005). Einav, I. 2005b. A second look at strain space plasticity and
latest applications. 18th Australiasian Conference on the
Mechanics of Structures and Materials (ACMSM), Perth,
(1), 225231.
ACKNOWLEDGEMENTS Fleming, W. G., Weltman, A. J., Randolph, M. F., &
Elson, W. K. 1992. Piling Engineering, E & FN Spon,
The work described here forms part of the activities of New York & London.
the Special Research Centre for Offshore Foundation Hetenyi, M. 1946. Beams on Elastic Foundations, The
Systems, established and supported under the Australian University of Michigan Press.
Research Councils Research Centres Program. Murff, J. D., & Hamilton, J. M. 1993. P-ultimate for
undrained analysis of laterally loaded piles. Journal of
Geotechnical Engineering, 119(1), 91107.
Randolph, M. F., & Houlsby, G. T. 1984. Limiting Pressure
REFERENCES on a Circular Pile Loaded Laterally in Cohesive Soil.
Geotechnique, 34(4), 613623.
Anagnostopoulos, C., & Georgiadis, M. 1993. Interaction of Sastry, V. V. R. N., & Meyerhof, G. G. 1990. Behaviour of
axial and lateral pile responses. Journal of Geotechnical flexible piles under inclined loads. Canadian
Engineering, 119(4), 793798. Geotechnical Journal, 27(1), 1928.
Ashford, S. A., & Juirnarongrit, T. 2003. Evaluation of pile Scott, R. F. 1981. Foundation Analysis, Prentice-Hall.
diameter effect on initial modulus of subgrade reaction. Shahrour, I., & Meimon, Y. 1991. Analysis of the behaviour
Journal of Geotechnical and Geoenvironmental Engine- of offshore piles under inclined loads. International
ering, 129(3), 234242. Conference on Deep Foundations, 277284.

879

Copyright 2005 Taylor & Francis Group plc, London, UK


Assessing geohazards

Copyright 2005 Taylor & Francis Group plc, London, UK


Assessment of the hydrate geohazard

A.J. Digby
Fugro Ltd, Wallingford, UK

ABSTRACT: Recent hydrate assessments from the Ocean Drilling Programme (ODP) and the Mallik Test site
have advanced the techniques of hydrate detection and evaluation. Transfer, and some modification, of these
techniques to the routine evaluation of hydrates as a deepwater geohazard is now needed. Early detection of a
hydrate geohazard can be achieved by the integration of seismic, wireline and sampling data before full evalu-
ation proceeds. Proxy identifiers of hydrates are established and seismic and wireline data, particularly where
used in tandem, are proving effective at locating hydrate risk. Recent infra-red spectral core logging has shown
the potential for identifying a hydrate signature in non-pressurized cores. No single proxy method of measure-
ment is totally effective at identification and none, as yet, appears reliable at quantifying the absolute values of
hydrate concentration and critical porosity. At our present depth of knowledge, a pressurized core recovery pro-
gramme is required to obtain this objective data.

1 INTRODUCTION
2 LOCATING THE HYDRATES
Indications of hydrates have now been found worldwide
in practically all deepwater sites with suitable tempera- 2.1 Bottom simulating reflectors
ture and pressure regimes. Production from hydrates The primary proxy indicator of hydrates is the identi-
has been achieved below permafrost caps in Siberia and fication and recognition of a Bottom Simulating
at the test site at Mallik in the McKenzie Delta. Reflector (BSR). A BSR, potentially, indicates the
Dealing with a potential geohazard starts with the interface of the gas hydrate stability zone (GHSZ) with
decision of whether to mitigate or avoid the hazard. the free gas zone beneath.
With the growing evidence for the widespread occur- For assessment of the hydrate hazard the BSR has
rence of hydrates, avoidance may not always be an four drawbacks. Firstly, the hydrate BSR must be sepa-
option. rated from other potential BSRs. Secondly, a number of
Locating and evaluating a hydrate hazard will con- observations have confirmed that hydrates can be pre-
tinue to depend initially on proxy methods of meas- sent without a BSR. Thirdly, the BSR is just the bottom
urement and identification. The challenge now is to limit of the GHSZ so the vertical extent of the hazard
calibrate and integrate these proxy measurements with can only be inferred by estimating temperature/pressure
absolute values. conditions and from theoretical hydrate stability
The shallow gas risk of hydrates is present at all envelopes. Lastly, a travel time seismic survey provides
concentrations but the other major risks caused by the no absolute values that can be used to approximate the
development of overpressures occurs when hydrate hydrate content of the sediment.
dissociation increases the effective porosity of the for- It has been observed that BSRs appear to be sharper
mation above the critical porosity. The critical porosity in some areas than others but this need not reflect
is the level above which the pore fluid becomes load varying concentrations of hydrate. The sharpness of
bearing. the BSR may be more indicative of gas concentrations
Porosity, !, and the fraction of the pore space occu- below the BSR which are controlled by lithology and
pied by hydrate (the hydrate concentration +) are thus structure and potentially irrelevant to the conditions in
the key measurements for assessing the hydrate hazard. the GHSZ.

883

Copyright 2005 Taylor & Francis Group plc, London, UK


2.2 Discriminating hydrate BSRs 3 RECOGNITION AND ASSESMENT USING
WIRELINE AND LOG WHILE DRILLING
Other temperature/pressure delimited phenomena also
(LWD) DATA
capable of generating BSRs include opal A/Ct transi-
tion and smectite/illite changes. A BSR with a nega-
3.1 Wireline properties of a massive hydrate
tive polarity and occurring at or close to the predicted
base of a hydrate stability zone can now, with experi- The recovery of over a meter of intact hydrate repre-
ence, be identified with some confidence. senting potentially a 34 m thick massive hydrate
off Guatemala in 1984 Collett & Ladd (2000) was
hugely significant in attempts to understand and iden-
2.3 Hydrates without BSRs tify true in situ responses of wireline logs to hydrate.
The first, and as yet thickest, recovery of marine
hydrate was 1.05 m from a 34 m massive layer came 3.1.1 Resistivity
from a boring on ODP leg 84 offshore Guatemala. The resistivity value of a massive hydrate is very
This location lacked a BSR and one of the three Blake high, (155185 Ohm/m) and, given that the resistivity
Ridge occurrences of hydrate had no associated BSR, values in a shallow marine clay may be expected to be
Collett (1995). Simply expressed the BSR may require around 1 Ohm/m, any significant hydrate saturation
a free gas zone below the hydrate, but the existence of should be identified as a positive resistivity anomaly.
hydrates does not require an existing free gas zone to Correlations between resistivity anomalies and BSRs
be present beneath. are convincing enough to suggest that this is the pri-
mary wireline and LWD function for identifying
hydrate presence. Figure 1 illustrates a clear anomaly in
2.4 Gas hydrate stability zone (GHSZ) 100 m section with a baseline shift in resistance of only
1.15 Ohm/m. The Resistivity-at-Bit LWD log is likely
Theoretical extents of the gas stability zone based
to be the first indication of hydrates in the borehole.
on temperature and pressure assumptions may be
misleading. Structural and stratigraphic controls may 3.1.2 Natural gamma
restrict hydrate occurrence. The acoustic impedance Hydrates should not affect natural gamma logs: it is
of hydrated sediments is the most likely property that this total lack of sensitivity that makes the measure-
could be exploited to provide a full thickness of the ment essential. Any resistivity anomaly associated
GHSZ. Seismic Inversion modelling appears to have with unchanged gamma can be viewed with suspicion
real potential for assessing the vertical and lateral as a potential hydrate indication. An anomaly associ-
extent of a hydrate risk and enabling a risk gradient to ated with a gamma change is likely to be interpreted as
be applied to a GHSZ. lithologically induced. The ragged hole potential of
hydrated sediments has least effect on the natural
2.5 Hydrate saturation gamma log of all the wireline measurements.
Recent research, Reister (2003), has cast some doubt 3.1.3 P Wave velocity
on the ability of acoustic impedance alone to go Laboratory and some field data (Reister 2003) sug-
beyond providing a gradient of concentrations and gest an appropriate value of around 3,600 m/sec for
achieving absolute saturation values. With low hydrate massive hydrate. In deepwater, the marine shallow
saturation, a lower limit of 20% has been suggested section without hydrate has p wave velocity values of
for accurate assessment of gross saturation (by % vol- 15001800 m/sec. The contrast is still considerable but
ume). The hydrate hazard, rather than a hydrate the ratio of around 2 times base line is not as clear as the
resource, is often likely to have saturation signifi- likely resistivity contrast, especially when dealing with
cantly below this figure where dissociation would low hydrate saturation. If gas is actively dissociating
still result in porosity values being increased above during logging, gas bubbles can reduce the apparent
the critical level. P-wave velocity, providing a typical gas signature.
Whether this pessimism is correct will only be
confirmed when sufficient absolute values are avail- Table 1. Wireline properties of massive hydrate (after
able to compare with proxy estimates based on Collett [USGS] & Ladd [Lamont Doherty] 2000).
acoustic impedance. The vastly different laboratory
properties for dissociation rates (Kirby et al. 2004) Density Slightly low 0.9881.055 gm/cm3
which depend upon the presence of disseminated or SP Less ve
massive hydrate and the nature of the host material Calliper Usually ragged and oversize
Resistivity High 155185 /m
i.e. the soil geology, makes it likely that a single %
P wave velocity High 3,690 m/sec
volume and in situ porosity will not be sufficient to Neutron porosity Slightly high 67%
model all likely behavior.

884

Copyright 2005 Taylor & Francis Group plc, London, UK


3.1.4 Density anomaly showed a 12% increase in apparent neutron
Theoretical density contrasts are negligible. However porosity. The difficulty in assuming hydrate concentra-
a pattern of very variable density log curves has been tions from the baseline shift in the neutron log involves
observed with base lines as low as 1.80 g/cm3 against an assumption that, with or without hydrate, the true
background of 2.05 g/cm3. The explanation can be water filled porosity is the same above and below the
found in the calliper characteristics which indicate hydrate zone. This is a chicken and egg argument as
ragged or extremely ragged holes. This raggedness, increased existing porosity may have allowed hydrate
possibly the result of localized loss of soil strength to be deposited and the apparent hydrate zone may
caused by hydrate dissociation results in the variable not be controlled simply by the temperature pressure
and low density log. regime.

3.1.5 Calliper
The raggedness of a borehole through a GHSZ has 3.1.7 Nuclear magnetic resonance
been noted several times. Although providing another Nuclear Magnetic Resonance has been used for recent
qualitative hydrate indicator, the ragged borehole ODP legs but these have been on accretionary com-
has significant knock on effects on the accuracy and plexes and may be atypical of hydrate occurrence as a
quantitative estimate of hydrate concentration from geohazard in production areas. No non-propriety results
resistivity, neutron or velocity measurements. have been available to the author for this assessment.

3.1.6 Neutron 3.1.8 Fluid temperature


Neutron porosity increases with hydrate formation, Temperature is an obvious measurement to make,
despite the slightly reduced density of hydrate com- given the temperature/pressure control on the exist-
pared to water. The example in Figure 1 of a resistivity ence of hydrates. Formation temperature measure-
ments have been performed on research programs to
establish in situ methane solubility, but the potential of
fluid temperature logs has yet to be fully tested.
The dissociation of hydrates is endothermic, therefore a
reduced, zero or negative gradient would be anticipated.

3.2 Assessing hydrate concentration


As with the identification of hydrate BSRs, there is a
sufficient number of independent indicators within the

Figure 1. Resistivity anomaly 11751275 mbsl; typical of


low hydrate concentrations anomaly peaking at 1 /m above Figure 2. Density, neutron and calliper curves for the same
baseline resistivity. resistivity anomaly as Figure 1. 11751275 mbsl.

885

Copyright 2005 Taylor & Francis Group plc, London, UK


wireline and LWD arsenal of measurements to be rea- derived from 100% quartz with 100% hydrate and
sonably confident of identifying hydrates from these 100% quartz with 100% water respectively. The lower
measurements. Gamma and resistivity may be the sim- moduli for clay compared to quartz means the plots,
plest employable combination for identification but to when being used to assess hydrate concentrations, are
be certain of the interpretation, especially with poten- still sensitive to soil mineralogy. This can be seen in
tially low concentrations, a full suite of measurements the population distribution of the derived values.
is desirable. Though not stated in the paper, the lack of sensitivity
The assessment of the quantity of hydrate and the of the determination for concentrations below 20%
geotechnical impact of hydrate dissociation is far more may result directly from a lack of differentiation of
challenging. soil type. With additional geological information,
From Table 1 of wireline properties, sonic velocities separate sand and clay plots may improve the accur-
and electrical resistivity of soils are seen to be signifi- acy of the velocity determined concentrations.
cantly affected by hydrate formation. These properties
are the basis of two mathematical assessments avail- 3.2.2 Hydrate concentration using resistivity
able from wireline data for hydrate concentration Work by Collett & Ladd (2000) shows the effective-
determination. ness of resistivity measurements for determining
relatively low concentrations at Blake Ridge. The cal-
3.2.1 Hydrate concentration using measured culations compared favorably with chlorite anomaly
velocities assessments of hydrate concentration. The assessments
Measured velocities are determined by the volume are all based on the standard Archie equation
fractions of the three phases soil matrix , hydrate and
water. Bulk density for a three phases soil, water, and (2)
hydrate mixture is a linear function of the volume frac-
tions which presented in terms of porosity and hydrate where R is measured resistivity, Rw is the resistivity of
concentration becomes: the pore water, and ,  and are parameters. Collett
et al. (1999) used values of   0.62,   2.15, and
(1)   1.9386 ( and  are called the humble values and
 is from Pearson et al. (1983)).
The resistivity values are very sensitive to hydrate
where   Bulk density; w  water density; h  concentrations but less so to porosity, thus justifying the
hydrate density; m  grain (matrix) density; use of both velocity and resistivity in assessing hydrate
!  porosity; +  hydrate concentration. concentration and in situ porosity.
Density is very sensitive to porosity changes due to The resistivity methods can use the standard
the contrast between hydrate/water and grain dens- Archie equation or a number of refinements and short
ities, but insensitive to changes in concentration of cuts. Three methods have been found to provide simi-
hydrate due to closeness of the two densities of lar results and their relative merits are discussed in
hydrate and water. The velocities are very different Collett & Ladd (2000).
between hydrate and water and Reister (2003) uses The measured resistivity shows great sensitivity to
the density, velocity and moduli values in Table 2 for hydrate presence and the main control on the accuracy
gas hydrate concentration determinations at Mallik. of this method is the correct choice of the Rw resistiv-
Reister (2003) uses plots of the three elastic mod- ity base line. With the Blake Ridge work similar soils
uli (Bulk K, Shear G and P-Wave M) against porosity, occur above and below the hydrate so that a reasonably
and has described a method of inverse modelling of accurate base line can be drawn using resistivity val-
the hydrate concentration from these cross plots. The ues from these sections. The method is thus vulnerable
upper and lower bounds of these plots join the moduli to any formation controlled hydrate occurrence where
the strata above and below may have significantly dif-
ferent electrical properties to the hydrated sediment.
Table 2. Velocities, density and elastic moduli Reister (2003).

Quartz Water Hydrate Clay

Density (gm/cm3) 2.65 1.03 0.90 2.58 4 DIRECT MEASUREMENTS


P wave velocity (m/sec) 6,038 1,558 3,690 3,400
S wave velocity (m/sec) 4,120 0 1,915 1,623 4.1 Pressurized core recovery
K bulk moduli (GPa) 36.6 2.5 7.9 20.9
G shear moduli (GPa) 45.0 0 3.3 6.8 Three coring systems were used on recent ODP legs
M P-wave moduli (GPa) 96.6 2.5 12.3 30.0 the DP Pressure Core Sampler (PCS), the Hyacinth
Fugro Pressure Corer (FPC) and the Hyacinth (Hyace)

886

Copyright 2005 Taylor & Francis Group plc, London, UK


Rotary Corer to obtain pressurized samples. The thought this could reflect gas coming out of solution as
two Hyacinth corers can be used directly with the the pressure is released. The slow dissociation of finely
GEOTEK V multi-sensor core logger (MSCL). disseminated hydrate in a clay matrix modeled in the
laboratory could mean this overall lower density may
4.1.1 Difficulties of using recovered volumes still be hydrate related and two small additional low
Absolute volumes of gas from a recovered core in density potential hydrate beds are tentatively identi-
hydrate still need consideration before an accurate fied in Figure 3. The total gas extracted from this
figure for hydrate concentration can be reached. The particular core was unfortunately not accurately
most likely inaccuracy in quantifying concentrations measured but with future runs it should be possible to
of hydrate from recovered cores is accounting for core calibrate the density log against actual production.
loss. Logging using a MSCL before depressurizing the The endothermic nature of the methane dissoci-
core is required. Repeated passes of the MSCL during ation allows infra-red scans to be employed to indi-
depressurization ensures post retrieval changes can be cate temperature lows related to discrete bands of
monitored. The MSCL log also provides an opportun- hydrate. Trehu et al. (2004) have shown how these
ity of comparing the core directly with in situ wireline low temperature spikes can be used to identify the
measurements of the cored section. The suite of high extent of the gas hydrate occurrence zone and have
vertical resolution logs with a very small vertical sam- compared these spikes to chlorine anomalies to esti-
ple interval can be directly compared to the MSCL mate hydrate concentration.
log. Natural gamma correlation is the simplest log to
use. Another potential error in calculating concentra-
tion from the total of recovered gas is accounting for 5.2 Chlorine anomalies in core pore water
the solubility of the recovered gas at the in situ tem- The pressurized cores have also been used to obtain
peratures and pressures. Formation pressures and tem- salinity measurements of the pore water. The forma-
peratures must be assessed accurately and even then tion of hydrates excludes salt from the hydrated water.
there has been some debate over the accuracy of exist- This effect has been known about for many years but
ing solubility charts, Laherrere (2000). has mainly attracted the attention of those studying its
There are inevitably difficulties in coring intervals potential for desalination projects. The low salinity of
with widely variable elastic properties, an unavoidable water from dissociating hydrates can be collected from
result of uneven distribution of hydrate. This has core samples and, using a baseline salinity from sedi-
prompted the development of both piston and rotary ments above the hydrate zone or using an empirical
coring techniques.
The recent research on rates of dissociation, Kirby
et al. (2004), illustrates how difficult it is to predeter-
mine the rate of methane dissociation. The MSCL dens-
ity function is used for repeated passes to ensure gas
extraction continues until a steady state density pro-
file is obtained and it is safe to assume full extraction
has been achieved.

5 PROXY ESTIMATES FROM CORE SAMPLES

5.1 The multi sensor core logger (MSCL)


In addition to the essential quality control function pro-
vided by the MSCL, the MSCL density curve and the
infra-red temperature log can both be used for semi-
quantitative assessments. They are useful cross checks
against measured volumes. The first use of the MSCL
with a pressurized core is shown as Figure 3. The log
shows a density low near the top, which was equated
with a solid hydrate layer of 5 mm thickness. Log 2
shows very low density at this part of the core, pre-
sumed to indicate the hydrate dissociating and a second
gas layer near the base that is presumed to be the prod-
uct of a thinner initially un-resolvable hydrate layer. Figure 3. Density and infra-red images of recovered pres-
The overall core density is less on run 2 and it is surized core (courtesy GEOTEK /Spectra Map).

887

Copyright 2005 Taylor & Francis Group plc, London, UK


approach, a hydrate estimate can be made. The assess- high quality wireline logs of the shallow section and
ment of dilution from the Blake Ridge cores showed hence quality well control for high resolution seismic
good correlation with the resistivity assessments of inversion modelling are available but relatively
hydrate content from the wireline logs, but both meth- untried.
ods do rely on an accurate estimate of the non-hydrate Absolute values for low but still potentially critical
baseline. hydrate concentrations will need sampling and methane
extraction.
5.3 Infra-red thermal scans and hydrate Correct core handling procedures and interpreta-
indicators in non pressurized cores tion of core loss, will be critical if any accurate assess-
ment is to be made. Downhole conditions also need to
The use of infra-red scanning has lead to the discovery be accurately modeled to account for the recovery of
of the potential of using infra-red spectral mapping of gas in solution against gas from hydrate.
recovered cores irrespective of the method of recov- The MSCL logs of natural gamma, density and
ery. The two core images at the left of Figure 3 show a infra-red spectroscopy offer the facility for quality
spectral map with the spectra set to identify salt control that will be needed for confidence in any
deposits on the core. The image, originally in colour, assessment.
shows that the two tentative bands of hydrate corres-
pond to a clear positive anomaly. The large anomalies
at the top and bottom were also identified by a salt
REFERENCES
signature but this is not so apparent on the image.
These spectral maps were confirmed on the Archie, G.E. 1942. The electrical resistivity log as an aid
depressurized cores after storage and this suggests in determining some reservoir characteristics: J. Pet.
that similar signatures could remain on convention- Technol., 5:18.
ally recovered cores. If stored, cores from any known Collett, T.S. 1995. Gas hydrate resources in the USA. 1995
hydrate locations could be made available this National assessment of United States oil and gas
hypothesis could be tested. resources: U. S. Geol Survey Digital Data Ser., 30.
Collett, T.S. & Ladd, J. 2000. Detection of gas hydrate with
downhole logs and assessment of gas hydrate concentra-
6 CONCLUSIONS tions (saturations) and gas volumes on the Blake Ridge
with electrical resistivity log data: In 2000 Proc. ODP,
Sci Results, 164: College Station, TX (Ocean Drilling
The research into the resource potential of hydrates Program).
has resulted in the development of a number of proxy Kirby, S.K. Stern, L.A., Circone, S., Pinkston, J. 2004.
identifiers and semi-quantitative analysis methods for Hydrate decomposition under scrutiny, The National
detecting and assessing the hydrate geohazard. energy Technology Laboratory Methane Hydrate
The hydrate geohazard assessment must identify Newsletter, Winter 2004.
the critical porosity of a formation and the concentra- Laherrere, J. 2000, Oceanic hydrates: more questions than
tion of hydrate existing, and whether dissociation will answers: Energy Exploration and Exploitation. 2000.
increase porosity to above the critical level. Pearson, C.F. Halleck, P.M. McGuire, P.L. Heermes, R. &
The proxy methods of assessment all rely to a sig- Matthews, M. 1983. Natural gas hydrate deposits: A
review of in situ properties. J Physical Chemistry, 87,
nificant extent on establishing a baseline property of 41804185.
the formation without hydrate and comparing this to Reister, D.B. 2003, Using measured velocity to estimate gas
the measured in situ properties. In accretionary zones hydrates concentration. Geophysics 68(3): 884891.
such as the Cascadia and Blake Ridge these may Schlumberger 1989, log interpretation principles/applica-
be reliable estimates. It is unproven, but likely, that tions. Schlumberger Educational Services.
hydrates occur with significant lithological controls Trehu, A.M. Long, P.E. Torres, M.E. Bohrmann, G. Rack,
where baselines beyond the hydrate occurrence may F.R. Collett, T.S. Goldberg, D.S. Milkov, A.V. Riedel, M.
represent formations not suitable for hydrate forma- Schultheiss, P. Bangs, N.L. Barr, S.R. Botrowski, W.S.
tion and preservation, hence making these baseline Claypool, G.E. Delwiche, M.E. Dickens, G.R. Gracia, E.
Guerin, G. Holland, M. Johnson, J.E. Lee, Y.-J. Liu, C.-S.
assessments numerically invalid. Su, X. Teichert, B. Tomaru, H. Vanneste, M. Watanabe,
Notwithstanding these difficulties, the early detec- M. Weinberger, J.L. 2004, Three-dimensional distribu-
tion of hydrate presence and several semi-quantitative tion of gas hydrate beneath southern hydrate ridge: con-
methods of assessment can be used in parallel to create straints from ODP Leg 204, Earth and Planetary Science
confidence in the model. These methods including Letters 222 (2004) 845862 Elsevier.

888

Copyright 2005 Taylor & Francis Group plc, London, UK


Gas hydrates and their potential effects on deep water
exploration activities

J.A. Priest & C.R.I. Clayton


School of Civil Engineering and the Environment, University of Southampton, UK

A.I. Best
Challenger Division for Seafloor Processes, Southampton Oceanography Centre, UK

ABSTRACT: Methane gas hydrates are solid, ice-like, compounds that form from methane gas and water in
deep water sediments. These compounds are metastable. Changes in temperature and pressure can lead to dis-
sociation of hydrates back to gas and water, which will alter the geotechnical properties of the sediment, per-
haps leading to slope instability and oil exploration infrastructure losses. Laboratory experiments have been
conducted on synthetic methane gas hydrate-bearing sand specimens in a specially constructed resonant col-
umn apparatus, in order to investigate the effects of bonding and dissociation, under a range of isotropic effect-
ive stress. The results show that the stiffness moduli are highly sensitive to hydrate inclusion, with large
increases in moduli resulting from low hydrate contents. Damping also showed considerable sensitivity to
hydrate content, which was lost upon dissociation.

1 INTRODUCTION may have triggered large seafloor failures that have


occurred in our geological past (Ashi 1999, Meinert
Oil and gas exploration activities have begun to extend et al. 1998) causing devastating tsunamis and wide-
into significant water depths (greater than 1000 m) on spread flooding of coastal regions.
our continental margins and seas where ice-like com- To avoid the potential risks associated with drilling
pounds, termed gas hydrates, are known to exist at through hydrates, a detailed knowledge of the occur-
shallow depths below the seafloor. These crystalline rence and concentration of hydrate is required along
compounds are formed from hydrocarbon gases (pre- with its potential effect on sediment properties. Marine
dominantly methane) that become encaged by hydro- seismic geophysical testing offers the potential to
gen bonded water molecules within the sediment pore detect hydrates, however the design and optimization
space. of seismic techniques also requires an understanding
Gas hydrate is a metastable compound that exists of the relationship between hydrate concentration and
under restricted temperature and pressure conditions. the physical properties of the sediment.
Dissociation can be induced by oil drilling activities. This paper therefore reports the results of labora-
This can change the ice-like structure back to its con- tory experiments that have been undertaken to inves-
stituent parts of gas and water possibly leading to dra- tigate the effects of disseminated synthetic methane gas
matic changes in the physical properties of the hydrate on the physical properties of sand specimens
sediment. Also the gas hydrate structure enables gas in terms of dynamic stiffness and damping, under the
molecules to be held in a more dense configuration than low frequency conditions relevant to seismic geo-
that which it would exist as a gas (1 m3 of gas hydrate physical testing.
contains 164 m3 of gas at Standard Temperature and
Pressure (Sloan 1998)), so an increase in pore pres-
sure could result from hydrate dissociation.
2 LABORATORY APPARATUS AND
It has been suggested that this phenomena has
HYDRATE FORMATION PROCEDURE
resulted in oil platform blowouts, casing failures and
at the extreme, platform subsidence (Bily & Dick 1974,
2.1 Gas hydrate resonant column
Franklin 1980). It has also been suggested that wide
scale hydrate dissociation, brought about by changes The need to correlate the physical properties of labora-
in sea level, or rises in ocean bottom temperatures, tory specimens with those determined from in-situ

889

Copyright 2005 Taylor & Francis Group plc, London, UK


torsional 0.14
excitation
magnet accelerometer 0.12
resonant frequency

Accelerometer output (V)


0.10
drive
coils
top cap 0.08

0.06

buty1 0.04
rubber specimen
membrane 0.02

0.00
134 135 136 137 138 139 140
base
pedestal Excitation frequency (Hz)

Figure 2. A dynamic response curve for a Leighton


Buzzard sand obtained during a frequency sweep for tor-
Figure 1. General layout of the resonant column apparatus. sional vibration.
The pressure cell, which surrounds the apparatus is omitted
for clarity along with the various electrical and pneumatic
connections. Drawing modified from Stokoe et al. (1999).
0.4
0.3
Accelerometer output, v

measurements, using remote marine seismic tech- 0.2


niques, required the laboratory testing to be undertaken
both within the seismic frequency range 50500 Hz 0.1
and at low shear strains ( 10
6), since they affect the 0.0
dynamic properties of a soil. The restricted conditions -0.1
under which methane gas hydrates exist required high
back pressures (up to 20 MPa) and low temperatures -0.2
(down to
20C) to be applied to the specimen to -0.3
enable the effective formation of hydrate (Stern et al.
-0.4
1996). Therefore a resonant column apparatus (RC), 0 200 400 600 800 1000 1200 1400 1600
which would as standard operate under the frequency Number of samples, n
and strain conditions required, was constructed in an
environmental chamber. Figure 1 shows the general Figure 3. Typical free vibration decay response curve
layout of the apparatus, for a more detailed descrip- obtained for a Leighton Buzzard sand.
tion of the apparatus the reader is referred to Clayton
et al. (in press).
In a standard Stokoe RC, torsional vibrations are were made to the RC so that flexural excitation could
induced in the specimen and, from the resonant fre- be induced in the specimen (by switching the voltage
quency of the system (specimen and drive mechanism), through the coils). This would allow Youngs modu-
the shear modulus of the soil, G, is calculated. Figure 2 lus, Eflex to be calculated (Cascante et al. 1998).
shows the response of a dry Leighton Buzzard sand Material damping in the resonant column test is eval-
during a frequency sweep under torsional excitation uated from the free vibration decay curve recorded
(achieved by applying a sinusoidal voltage through when the power to the coils is shut off when the speci-
the coils which induce motion in the magnets (Fig. 1) men is vibrating at its resonant frequency (Fig. 3).
and hence the specimen). The resonance frequency is Using this standard method, the calculated damping is
clearly identified by the peak in output from the influenced by apparatus damping caused by a back
accelerometer. e.m.f effect (Wang et al. 2003). Therefore an electrical
In the laboratory, differing excitation modes need open circuit arrangement was implemented to remove
to be used in order to derive the range of parameters this aspect of the apparatus damping. For a detailed dis-
applicable to shear and compressional wave propaga- cussion on the data reduction and various calibration
tion in geophysical surveys. To this end modifications issues the reader is referred to Priest et al. (in press).

890

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 1. Hydrate saturation, densities and moisture con- 18
tent of sand specimens. Rise in pore pressure due to temperature
increase after inlet valve is closed
Hydrate Moisture 16

Methane pore pressure (MPa)


saturation* Dry density content
Sample (% of pore space) (kg/m3) (%)
14
H0L 0 1417.2 0 Reduction in methane gas pore
H0D 0 1548.0 0 12 pressure during hydrate formation
H1 1.1 1589.4 0.25 as bulk melting of ice occurs
H2 2.2 1538.8 0.54 Time
H3-1 3.0 1538.9 0.75 10
Application of
H3-2 2.7 1481.7 0.73
methane gas
H4-1 3.8 1529.0 0.94
H4-2 3.9 1511.0 0.99 8
-20 -15 -10 -5 0 5 10
H5-1 4.9 1538.2 1.20
H5-2 4.9 1528.7 1.22 Temperature (C)
H10 9.7 1505.5 2.50
H20 18.1 1520.7 4.58 Figure 4. Methane gas pore pressure rise with temperature
H40 35.3 1516.4 8.08 for specimen H40 (35.3% of pore space occupied by hydrates).

* calculated from added water.


of 250 kPa before the interstitial water in the speci-
men was refrozen, to help prevent water migration.
2.2 Specimen preparation and hydrate formation
Once frozen, methane gas was injected into the pore
Due to the difficulty in forming hydrates in fine space of the specimen and slowly raised to 15 MPa
grained cohesive soils (Clennell et al. 1999) all tests whilst simultaneously maintaining an effective stress
were conducted using a fine Leighton Buzzard sand of 250 kPa on the specimen. Hydrate formation was
(Grade E). This is a uniform fine sand with a grain then encouraged by slowly raising the specimen tem-
size between 90150 m (85% by weight). Its specific perature to 10C, which was then maintained for a
gravity was found to be 2.65. Minimum and max- minimum of 15 hrs. Figure 4 shows the change in
imum dry densities were 1331 kg/m3 and 1624 kg/m3 methane back pressure with temperature for a speci-
respectively (Rad and Tumay 1987, Cresswell et al. men initially containing 35% ice in the pore space. It
1999). Minimum and maximum voids ratios were cal- can be seen that as the temperature approached 0C
culated as 0.633 and 0.993. there was a dramatic drop in the methane gas pres-
Table 1 shows the initial properties of hydrate- sure. This was due to the rapid formation of gas
bearing specimens. Each specimen was formed by hydrates as the ice in the pore space started to melt.
compacting moist sand in 8 equal layers to form a Once full hydrate conversion had occurred (inferred
dense specimen (specimen length L  140 mm, diam- from constant pore pressure, since continuing hydrate
eter d  70 mm). The mass of water in each specimen formation would cause a reduction in the methane pore
was accurately controlled by riffling a known mass of pressure), the specimen temperature was lowered to
air-dried sand and sieved, ground ice (125300 m), 3C and the back pressure reduced to 5 MPa (whilst
which was subsequently allowed to melt in air-tight maintaining an effective stress of 250 kPa). The reson-
bags. The riffling of the sand and ice was shown to ant column measurements were then undertaken.
give a reasonably homogenous distribution of water
throughout the specimen. This was investigated using
2.3 Testing procedure
an analysis on thin sections obtained from sand speci-
mens containing plastic beads, which were used as a Resonant column testing, using both torsional and
substitute for the ice. The hydrate saturation is directly flexural excitation, was undertaken to determine the
related to the volume of added water (methane gas effect of differing volumes of disseminated methane
encaged by water molecules leading to a lower dens- gas hydrate on the dynamic properties of a sand. After
ity 900 kg/m3 compared to 1000 kg/m3 for water) formation of the hydrate, isotropic effective stresses
and is therefore calculated from the initial mass of were increased from 250 kPa to 2000 kPa, in 250 kPa
sand, added water and specimen dimensions. In add- steps. Unloading followed with the same increments.
ition to the hydrate-bearing specimens, loose and Each load step was maintained for thirty minutes to
dense specimens of dry Leighton Buzzard sand were allow for any initial consolidation of the specimen to
also produced to provide comparative data. occur before resonant testing was undertaken. This
Once the specimen was formed and set up in the was observed by monitoring the output of an axial
apparatus, it was subjected to an initial effective stress LVDT, which measured the displacement of the drive

891

Copyright 2005 Taylor & Francis Group plc, London, UK


mechanism. The measured changes in height were function of isotropic confining stress,  for a selec-
subsequently used to update specimen geometry used tion of the specimens listed in Table 1. The hydrated
in the reduction of data. specimens are identified by their initial hydrate satur-
All tests were conducted under constant tempera- ation. As a reference the curves for the loose and dense
ture and back pressure and at strain levels below the unbonded specimens are also shown. Specimens H3-1,
elastic threshold,  et, which is defined as the point H4-1 and H5-1 are omitted from Figure 5 for clarity.
below which the modulus value is independent of From graphs of Gmax and Eflex vs.  several dis-
applied strain. Tests on the loose and dense sands had tinct behaviours can be observed. For specimens with
shown that  et occurred at strains of 10
5 at a confin- low hydrate saturation ( 3%) of the pore space, both
ing pressure of 250 kPa (Saxena et al. 1988, suggest Gmax and Eflex increased with effective confining
that  et is largely unaffected by cement bonding). Free stress. As small amounts of hydrate were added into
vibration decay curves were acquired for each load the pore space Gmax and Eflex increased rapidly, with
step and excitation mode, from which material damp- stiffness increasing by nearly 5 fold from 0 to 3%
ing was determined. hydrate in the pore space. A tenfold increase in sat-
uration from 3% to 35%, only increased stiffness by
about 3.5. For hydrate saturation of 3% and above,
3 RESULTS AND DISCUSSION both moduli exhibited little stress dependency. Figure 6,
which compares Gmax and Eflex as a function of hydrate
3.1 Small strain stiffness saturation for an effective confining pressure of
500 kPa, clearly shows the differing rates of increase
Figure 5 shows the relationship between the small strain of Gmax and Eflex for hydrate saturations up to 3% and
shear modulus, Gmax and small strain Youngs modu- the lower rate of increase at higher saturations.
lus from flexure, Eflex (plotted on a log scale) as a In unconsolidated soils subjected to isotropic load-
ing a simple exponential relationship has been shown
10000 to exist between stiffness and mean effective stress of
the form
Small strain shear modulus, Gmax (MPa)

(1)

where A and b are constants (Hardin & Black 1968).


1000 The power exponent b represents both the nature of
the contact stiffness and fabric change as a function
of isotropic stress (Cascante et al. 1998). In clean plu-
viated laboratory specimens the value of b is typically
of the order of 0.5. To investigate the effect of hydrate
on the stiffness of sands, best fit curves of the form
100 given by Equation 1 were applied to the data in Figure
5. Figure 7 presents the results of this analysis for
105
Small strain Young's modulus, Eflex (MPa)

35% 12000
18%
104 10% 10000 Gmax
5% Eflex
4%
Moduli (MPa)

3% 8000
2%
1% 6000
103 Dense
Loose
4000

2000
102
0 500 1000 1500 2000 2500 0
Isotropic effective confining pressure, ' (kPa) 0 10 20 30 40
Hydrate saturation (%)
Figure 5. Relationship between (a) Gmax and (b) Eflex as a
function of isotropic confining pressure, . The percent- Figure 6. Relationship between Gmax, Eflex and hydrate
ages given in (b) relate to both graphs. saturation.

892

Copyright 2005 Taylor & Francis Group plc, London, UK


both Gmax and Eflex. The power exponent for the loose damping decreased. Upon dissociation the damping
and dense specimens is 0.46 and 0.54 respectively. for all previously hydrated specimens fell within a
These values are consistent with published data on narrow band, although still higher than for the loose
remoulded clean sands (Cascante et al. 1998). The and dense samples which were air dry.
reduction in stress dependency from the loose to It may be noted that for both DS and DE the max-
dense specimens can be attributed to more stable con- imum damping level corresponds to the region in
tacts, and smaller changes in specimen fabric during Figure 6 where the rate of increase in Gmax and Eflex
isotropic consolidation. abruptly changes. It is also evident that at this hydrate
As hydrate saturation was increased a sharp reduc- saturation, variations occurred (Fig. 6 & Fig. 8) in the
tion in the b value occurred, with its value falling to computed values for both stiffness and damping of
0.05 at 3% hydrate saturation. Thereafter the b value specimens with approximately the same hydrate sat-
remained very low. The reduction in b value and the uration. It is suggested that at this critical hydrate sat-
associated increase in stiffness occurred as a result of uration, when the sand becomes fully bonded, stiffness
cementing the sand grain contacts. Similar results have and damping are sensitive to minor non-homogeneities
been observed for small amounts of differing forms in the water distribution at particle contacts.
of cement, such as frozen capillary water or epoxy Damping in particulate materials at small strains is
resin (Dvorkin & Nur 1993), Portland cement (Saxena independent of frictional losses (Winkler et al. 1979)
et al. 1988) and calcium carbonate in chalk (Clayton and may result from viscous fluid flow in cracks, or at
et al. 2002). For higher hydrate saturations the increase grain contacts as they are deformed during stress
in specimen stiffness is probably attributable to stiff- changes (Mavko & Nur 1979). As hydrate has similar
ening of existing cement bridges. characteristics to ice, and soil particles remain water-
Resonant column tests were also carried out on ini- wet in the presence of water ice (Valiullin & Furo
tially hydrated specimens after the hydrate had been 2002), it is suggested that small volumes of water
allowed to dissociate. The results showed a remark- remain after hydrate formation. Figure 8 show that the
able correspondence with those of the unbonded air-dry dense sand specimen had the lowest degree of
dense sand, suggesting that for this sand the effects of
hydrate bonding were reversible.
1.2
(a)
3.2 Material damping 1.0
Figure 8 shows damping obtained from torsional
Damping DS (%)

excitation, DS, and flexural excitation, DE, as a func- 0.8


tion of hydrate saturation obtained during free vibra-
tion decay. It can be seen that the introduction of 0.6
hydrate into the pore space initially caused a rise in
material damping, reaching a maximum between 0.4
35% hydrate saturation. For higher saturation,
0.2
0.6
0.0
0.5 1.6
(b)
0.4 1.4 Hydrate
Gmax
Power exponent, b

Eflex dissociated
1.2
Damping DE (%)

0.3 dry
1.0
0.2
0.8
0.1
0.6
0.0
0.4

0 10 20 30 40 0.2
Hydrate saturation (%)
0 10 20 30 40
Hydrate saturation (%)
Figure 7. Effect of hydrate cementation on the stress
dependent power exponent for shear modulus and Youngs Figure 8. Free vibration decay damping of dry, hydrated and
modulus. dissociated specimens as a function of hydrate saturation.

893

Copyright 2005 Taylor & Francis Group plc, London, UK


damping, with the loose sand being slightly higher. under the EU HYDRATECH project. The authors
Although air dried, both sands had minor volumes of gratefully acknowledge the contributions of col-
absorbed water which may have gathered at grain leagues at Southampton Oceanography Centre,
contacts. As hydrate saturation was initially increased Professor Clive McCann (Reading University, UK)
the effective grain contact area will also have increased, and Professor Kenneth Stokoe II (University of Texas
leading to greater volumes of viscous fluid flow and at Austin, USA).
damping. With additional cementation grain contact
deformation will have reduced, and the pathway of
adsorbed water to areas of low stress may have become REFERENCES
blocked, leading to a reduction in damping.
After dissociation all specimens exhibited a slight Ashi, J. 1999. Large submarine landslides associated with
increase in damping compared to the dense sand, decomposition of gas hydrates. Landslide News, 12:1719.
which may have been caused by saturation of the Bily, C. and Dick, J.W.L. 1974. Natural occurring gas
grain contacts. hydrates in the Mackenzie Delta, Northwest Territories.
Bulletin of Canadian Petroleum Geology, 22(3):340352.
Cascante, G., Santamarina, C., and Yassir, N. 1998. Flexural
excitation in a standard torsional-resonant column. Can.
4 CONCLUSIONS Geotech. J., 35:47849.
Clayton, C.R.I., Matthews, M.C. and Heymann, G. 2002.
Results of tests into the affect of disseminated gas The Chalk Invited Paper. International Workshop on
hydrates on the dynamic properties of a Leighton Characterisation and Engineering Properties of Natural
Buzzard sand show that their introduction has a pro- Soils, ISSMGE Committee TC24, Singapore, November
found effect on both the shear modulus, Gmax and 2002.
Youngs modulus, Eflex of the sand. At around 35% Clayton, C.R.I., Priest, J.A., and Best, A.I. (in press). The
hydrate saturation the sand became fully bonded by effects of disseminated methane gas hydrate on the
the hydrate, and stiffness showed no dependency on dynamic stiffness and damping of a sand. Geotechnique.
Clennell, B.M., Hovland, M., Booth, J., Henry, P., and
effective stress. Upon dissociation, previously hydrated Winters, W.J. 1999. Formation of natural hydrates in
specimens exhibited the same behaviour as the non- marine sediments: 1. Conceptual model of gas hydrate
hydrated dense specimen. growth conditioned by host sediment properties. Journal
Damping was found initially to increase with of Geophysical Research, 104:22,98523,003.
hydrate content. The maximum value occurred at Cresswell, A., Barton, M.E., and Brown, M.R. 1999.
hydrate saturation similar to that at which the sand Determining the maximum dry density of sands by plu-
became fully bonded. Damping upon dissociation was viation. Geotechnical Testing Journal, 22(4):324328.
found to fall in a narrow band, which was slightly Dvorkin, J. and Nur, A. 1993. Rock physics for the character-
higher than that for dry specimens but much less than isation of gas hydrates. In The Future of Energy Gases, vol-
ume 84, pages 699703. Geol. Surv. Prof. Pap., US., 1570.
the observed maximum. Franklin, L.J. 1980. In-situ gas hydrates a potential energy
The results show that gas hydrates can have a major source. Petroleum Engineer International, Nov:112122.
influence on soil behaviour. Hydrate dissociation, due Hardin, B.O. and Black, W.L. 1968. Vibration modulus of
to oil drilling activities, may bring about a dramatic normally consolidated clays. J. of Soil Mech. and Found.
reduction in soil strength of sand enhancing drilling Div., ASCE, Vol 94, No. SM2, Proc. Paper 5833, pages
risks and possible slope instabilities. However, the envi- 353369.
ronments in which this type of behaviour is relevant Helgerud, M.B., Dvorkin, J., Nur, A., Sakai, A., and Collett, T.
may be limited to gas rich environments such as those 1999. Elastic-wave velocity in marine sediments with gas
at active margins. This is because the formation proce- hydrates: Effective medium modeling. Geophys. Res.
Lett., 26:20212024.
dure used in this work forces gas hydrate to form at Mavko, G. and Nur, A. 1979. Wave attenuation in partially
grain contacts. It has been suggested that this growth saturated rocks. Geophysics, 44:161178.
morphology may not be representative in regions where Meinert, J., Posewang, J., and Baumann, M. 1998. Gas
gas hydrate grows from solution, or in water saturated hydrates along the northeastern Atlantic margin: possible
environments, since in these areas hydrates may not hydrate-bound margin instabilities and possible release
be cementing (Helgerud et al. 1999). Further work is of methane. In Henriet, J.P. and Meinert, J., editors, Gas
currently being undertaken into the effect of different hydrates: Relevance to world margin stability and cli-
hydrate growth morphologies on sediment properties. mate change. Geological Society, London, Special
Publications, 137, 275291.
Rad, N.S. and Tumay, M.T. 1987. Factors affecting sand
specimen preparation by raining. Geotechnical Testing
ACKNOWLEDGEMENTS Journal, 10(1):3137.
Priest, J.A., Clayton, C.R.I., and Best, A.I. (in press). A labora-
The work herein described was partly funded by an tory investigation into the seismic velocities of methane
EPSRC research studentship and was carried out gas hydrate-bearing sand. J. Geophys. Research.

894

Copyright 2005 Taylor & Francis Group plc, London, UK


Saxena, S.K., Avramidis, A.S., and Reddy, K.R. 1988. Valiullin, R. and Furo, I. 2002. The morphology of coexist-
Dynamic moduli and damping ratios for cemented sands ing liquid and frozen phases in porous material as
at low strains. Can. Geotech. J., 25:353368. revealed by exchange of nuclear spin magnetization fol-
Stern, L.A., Kirby, S.H., and Durham, W.B. 1996. lowed by 1H nuclear magnetic resonance. J. Chem. Phys.,
Peculiarities of methane clathrate hydrate formation and 117:23072316.
solid-state deformation, including possible superheating Wang, Y.H., Cascante, G., and Santamarina, J.C. 2003.
of water ice. Science, 273:18431848. Resonant column testing: the inherent counter EMF
Stokoe, K.H., Darendeli, M.B., Andrus, R.D., and Brown, effect. Geotech. Testing. J., 26(03):342352.
L.T. 1999. Dynamic soil properties: laboratory, field and Winkler, K. and Nur, A. 1979. Pore fluids and seismic attenu-
correlation studies. In 2nd International conference on ation in rocks. Geophys. Res. Lett., 6:14.
Earthquake Geotech. Eng., volume 3, Lisbon, Portugal.

895

Copyright 2005 Taylor & Francis Group plc, London, UK


OpenSees modeling of the 3D plastic behavior of underwater
slopes: achievements and limitations

H.G. Brandes & S. Wang


Department of Civil & Environmental Engineering, University of Hawaii, USA

ABSTRACT: The plastic response of two hypothetical slopes in response to surface loading from sediment
deposition is examined with reference to two constitutive models. Predictions using a cap model are judged to
be more reasonable than those from a non-cap model and significant differences are noted between 2D and 3D
geometries. Limitations are also discussed as far as realistic modeling of large offshore slopes is concerned.

1 INTRODUCTION Of course, techniques other than the finite element


method have been used extensively to analyze slopes,
Modeling and simulation of complex nonlinear geo- including limit analysis (bearing capacity, limit equi-
technical problems has usually meant the use of large librium) and various probabilistic approaches
commercial finite element programs. Although some (Abramson 2002, TRB 1996). However, they do not
of these are quite powerful, the underlying codes are provide the level of stress and strain detail that a
generally not accessible to benefit from community- numerical solution can provide and therefore gener-
wide development. Realizing that major advances in ally provide less insight into the mechanics of defor-
modeling and simulation were necessary in specific mation and failure of slopes.
fields such as earthquake engineering, the National
Science Foundation of the United States has supported
the development of OpenSees (Open System for 2 CONSTITUTIVE MODELS
Earthquake Engineering Simulation; Pacific Earth-
quake Engineering Research Center 2004). OpenSees A hierarchy of constitutive models has been proposed
is an open-source software framework consisting of a for frictional soil materials, ranging from the elemen-
set of interrelated classes, designed as independent as tary Mohr-Coulomb model to sophisticated hardening
possible for maximum flexibility, which control data plasticity models that are able to account for the most
structure, material behavior models, hierarchical elem- important aspects of soil behavior observed through
ent types, solution algorithms, integrators, equation experimentation. The perfectly-plastic Mohr-Coulomb
solvers, databases, visualization and communication model is adequate for simple force limit equilibrium
procedures. OpenSees has seen enormous growth in calculations regarding slope stability, bearing capacity
the last few years as developers and researchers have of footings, and active/passive pressures on retaining
made use of the softwares inherent flexibility to add walls. However, generalization of the Mohr-Coulomb
additional capability (Elgamal et al. 2003). OpenSees criteria to include the effect of all principal stresses and
has been adapted to solve a wide range of problems plastic yield deformations leads to unrealistic volume
involving soil, soil-structure interaction, earthquake change predictions (Drucker et al. 1957) and numerical
shaking, and reliability assessment (Jeremic 2003). It difficulties. The latter ones arise from the discontinuity
has also been adopted by the Network for Earthquake of the hexagonal yield surface in stress space along
Engineering Simulation as its platform for simulation principal stress planes. Drucker & Prager (1952) pro-
and visualization. posed a generalization of the Mohr-Coulomb model
In this article we report on two of a series of specific that takes into account the effects of all the principal
soil plasticity models that we are in the process of stresses and that is more amenable to numerical compu-
adding to OpenSees and examine their prediction with tation. Nonetheless, the Drucker-Prager model in its
regard to simplified two- and three-dimensional hypo- original form assumes similar behavior in compression
thetical slopes, which may be useful for understanding and tension and does not correctly predict volume com-
the behavior of slopes in the offshore environment. pression that accompanies increases in mean pressure.

897

Copyright 2005 Taylor & Francis Group plc, London, UK


'1
'3 t
ft
an
'1='2='3 (d+pa tan )
di

fs d+pa tan
eri
eM

n
siv

dia
eri
res

eM '2 
mp

iv
Co

ns
Te fc
'1 '3 '2

d
'2 2c cos
pa R(d+pa tan )
1-sin Compression line pb p
q

Tension line
Figure 2. Modified Drucker Prager model (MDP-C).
'2 k=qt/qc
qt qc
p in the EDP-I model, hence equation 1 applies. The
2c cos
1+sin
elliptical cap is given by the equation:

Figure 1. Extended Drucker Prager model (EDP-I).

An improvement upon the original model pro-


posed by Drucker and Prager, referred to herein as the
Extended Drucker Prager Model I (EDP-I), has been
coded into OpenSees as part of this study. This model (2)
is conceptually similar to the one that is available in
the code ABAQUS (HBS 2002). Its chief improve- where R  a cap shape constant;   a small number
ment upon the original Drucker-Prager model is that (typically 0.01 to 0.05); pb  the hydrostatic com-
the resulting yield surface is non-circular, hence pression yield stress; and pa  an evolution param-
allowing for a compressive yield strength that is eter that is a function of plastic volumetric strain
greater than the tensile yield strength. Soil is allowed controlling the amount of hardening or softening
to harden and/or soften isotropically. The flow rule (HBS 2002). In this equation d is the t-intercept at
allows simultaneous inelastic dilation and inelastic p  0 (Fig. 2). The parameter  is used to define a
shearing (Fig. 1). A linear yield criterion is assumed: transition yield surface between the one given by
equations 1 and 2. This transition yield surface has
(1) the following functional form:

where p  volumetric stress; is the slope of the lin-


ear yield surface in the p
t stress plane (i.e. friction
angle); d represents the materials cohesion; and t is
the deviatoric stress (HBS 2002).
(3)
Frictional soil models with a single surface, such
as the one just described, are unable to correctly
account for volume compression upon yielding. It is The material parameters assumed for the numerical
well known that normally consolidated clays and analysis are listed in Table 1. In addition to the EDP-I
loose sands exhibit a decrease in volume during and MDP-C models, we are also in the process of
shear, often almost from the beginning of loading. To adding additional constitutive models into OpenSees,
correctly account for this type of behavior, a range of such as a new and more versatile form of the modi-
more versatile models have been proposed that incorp- fied Cam clay model.
orate a second yield surface, or cap, which grows or
shrinks depending upon stress history. One such 3 SLOPE GEOMETRY AND MODELING
model, referred to herein as the Modified Drucker- APPROACH
Prager cap model (MDP-C), has also been coded into
OpenSees (Fig. 2). Again, this model is similar to one Computations were carried out using 20-node brick
available in ABAQUS. Frictional failure is defined as elements arranged to simulate 2D and 3D semi-infinite

898

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 1. Soil properties.

Slope Surcharge

c (kPa) 5 100
 (Deg.) 28 37
pd (kPa) 150 250
E (kPa) 17,400 17,400
E0 (kPa) 17,400 17,400
 0.35 0.35
(kg/m3) 1,478 3,000
k 1.0 1.0
 0.1 0.1
R 0.6 0.6
d 10.5 199.7
tan 1.11 1.51

Figure 4. 3D slope geometry before and after self-weight


consolidation.

Figure 3. 2D slope geometry before and after self-weight


consolidation.

slopes. By 2D we mean hypothetical slopes of uni- Figure 5. 2D slope with surface surcharge.
form depth and width (Fig. 3), whereas by 3D we
mean equivalent slopes that also vary in depth along as shown in Fig. 4. The slope itself was discretized
the cross-slope direction (Fig. 4). The intention in using 12  6  15 elements, whereas the surcharge
selecting these two geometries was to explore differ- thickness consisted of 6  6  3 elements.
ences between conventional 2D modeling and poten- The parameters listed in Table 1 are typical for fine-
tially more realistic 3D modeling. Another aspect that grained marine sediments, with the exception of the
is of interest is the loading of the sloping seafloor by density for the surcharge load. This density was chosen
sediment deposits (surcharge loading) that may have to be much higher than that of the slope sediment to
been transported by mass wasting or other means. focus attention on the behavior of the slope sediment
This was modeled in a very simple fashion by adding and not on the surcharge material. In fact, the surcharge
sediment thickness over a portion of the 2D and 3D was assumed to have purely elastic properties.
slopes and examining the resulting behavior (Fig. 5). In all cases, time-dependency and associated pri-
The intention was to illustrate, in a very simple- mary consolidation were ignored. In other words,
minded way, the effects of placing a sediment sur- drained and/or long-term conditions were assumed to
charge on the seabed that is of limited extent and prevail. This may be adequate for medium and deep
varies in load intensity in the down-slope direction. water slopes where sedimentation processes often
The slopes considered were 5 m wide, 10 m high and occur over very long periods of time. In each case, the
20 m long. The down-slope angle was 25. The 3D semi-infinite slope geometry was allowed to undergo
geometry varies from the 2D one only in the sense that self-weight compression first, followed by excess
the bottom is slanted 40 in the cross-slope direction, loading induced by embankment-type sediment heaps

899

Copyright 2005 Taylor & Francis Group plc, London, UK


as shown in Figure 5. Self-weight compression was
forced to occur in the vertical direction only by
imposing roller boundary conditions on all the verti-
cal faces and rigid conditions along the bottom. The
seabed surface was allowed to deform freely. During
excess loading the vertical faces on the upper and
lower ends of the boxes were changed from rollers
(which allowed only vertical deformations) to stress
boundaries with initial lateral pressures equal to those
at the end of the prior self-weight stage. Stress-strain
behavior was modeled using the two constitutive
models described previously. These models were
implemented using a forward Euler stress incremen-
tation technique for optimum stability.

4 RESULTS

The first stage of the analysis involves modeling of the


self-weight consolidation process. Roller boundary
conditions on the vertical faces lead to vertical dis- Figure 6. Yield points for EDP-I model.
placements (Figs. 3 and 4) and vertical stress gradients
proportional to the sediment weight. Although the dis-
tribution of resulting displacements and stresses is
reasonable, it can hardly be said that conventional
numerical modeling of the process of self-weight con-
solidation is an accurate representation of what occurs
in the field. After all, it is well known that complex
sedimentary and post-depositional processes, which
are difficult to understand and model, play an import-
ant role in subsequent slope behavior. Nonetheless, the
conventional numerical approach allows the setting of
initial stresses that satisfy equilibrium and boundary
conditions.
Following self-weight consolidation, the surcharge
load shown in Figure 5 was added to the surface of the
2D and 3D slope geometries, and the boundary condi-
tions along the upper and lower vertical faces were
relaxed to allow for global shearing. In addition to gen-
eral down-slope deformation, localized distortion in
the slope region beneath the surcharge load occurs. The
overall response though is very much dependent on the
plastic constitutive model used. In terms of stresses,
yielding is limited to the region immediately beneath Figure 7. Yield points for MDP-C model.
the surcharge load for the non-cap EDP-I model,
whereas yielding is much more widespread (but does
not occur everywhere) when the MDP-C cap model is This difference in behavior can be seen by follow-
used (Figs. 6 and 7). The reason for this of course is the ing the stress path of points A and B, identified in
presence or absence of the non-frictional yield cap. Figure 3. Point A, close to the surface of the slope and
Stress points in the slope near the surcharge yield immediately beneath the surcharge, is seen to yield as
because they reach the frictional surface, i.e. equation it reaches the frictional surface, regardless of which
(1), regardless of which of the two models are used. On model is used, whereas the deeper point B remains in
the other hand, stress points further away from the sur- an elastic state in the EDP-I model but yields by
charge remain in an elastic state in the EDP-I model, reaching the cap in the MDP-C model (Figs. 8 and 9).
but in many instances are seen to yield with the MDP-C It should be noted that the surcharge itself was given
model as they reach either the general cap, (Eqn. 3) or material parameters such that it would not yield in
the transition yield surface (Eqn. 5). order to focus on the response of the slope itself.

900

Copyright 2005 Taylor & Francis Group plc, London, UK


120

Compressive strength (MPa)


16
Stress Path at Point A 14 MH
100 Stress Path at Point B
12 Ice
Deviatoric Stress J2 (kPa)

10
80
8
6
60
4
2
40
0
0.01 0.1 1 10
20
Strain rate (%/min)

0 Figure 10. Volumetric strains for model MDP-C (2D).


-100 0 100 200 300 400 500
Stress Invariant I1 (kPa)
10
3D MDP-C Model
Figure 8. Selected stress paths for model EDP-I (2D). 3D EDP-I Model
2D MDP-C Model
8 2D EDP-I Model
120
Stress Path at Point A Depth (m)
Stress Path at Point B 6
100
Deviatoric Stress J2 (kPa)

80 4

60
2
40
0
20 0.16 0.14 0.12 0.1 0.08 0.06 0.04 0.02 0
Displacement Along Slope (m)
0
0 100 200 300 400 500 Figure 11. Downslope displacement profiles.
Stress Invariant I1 (kPa)
slopes should always employ a cap model. Naturally,
Figure 9. Selected stress paths for model MDP-C (2D). the MDP-C model chosen for this study is only one
such model and many others can be considered.
Among these are the widely-used modified Cam Clay
As already mentioned, the cap model has a distinct model (Roscoe & Burland 1968) and many others.
advantage over the non-cap model in that it is able to One important difference between the MDP-C and
predict either compressive or expansive plastic volu- the modified Cam Clay model is that the former one
metric strains, whereas the non-cap model is limited provides a seamless transition from the cap to the fric-
to the prediction of expansive inelastic strains only. tional shear surface by means of the transition surface
For example, Figure 10 shows plastic and total volu- (Fig. 2, Eqn. 5). While this may seem desirable since
metric strains at points A and B for the MDP-C it allows for a smooth transition in terms of the direc-
model. Point A reaches the frictional surface, such tion of the incremental plastic strain at the intersec-
that the normal incremental plastic strain points tion of the frictional and cap surfaces, it can lead to
upward and to the left, thus indicating plastic dilation. difficulties with certain stress integration algorithms
On the other hand, point B first reaches the cap, (such as the backward Euler method) in the region of
where the incremental plastic strain point upward and the transition yield surface when spanning the elastic-
to the right. As a result, plastic volumetric strain is inelastic divide.
compressive. In other words, shallow slope sediment The contrast in predictions of global shear deform-
subjected to loading by the surcharge will tend to ation from the two models can be considered with ref-
expand and shear under relatively low confining pres- erence to Figure 11. This figure shows downslope
sure, whereas deeper sediment will undergo less deformations along profile C, located at the lower end
shearing and some volume compression at higher of the slope (Figs. 3 and 4). Interestingly, the choice
confining pressure. This makes intuitive sense and of constitutive model makes little difference for the
suggests that modeling of elasto-plastic behavior of 3D slope geometry, but is very important for the 2D

901

Copyright 2005 Taylor & Francis Group plc, London, UK


slope processes involve sediment sequences that are
tens or hundreds of meters deep and hundreds of
meters to many kilometers in horizontal extent. This
would require massive meshes and therefore enor-
mous computational resources. For reference, the
nonlinear analysis conducted for this study involved
on the order of about 1,100 elements with 20 nodes
each and some 40 or so load steps. Each run took
about 8 hours when using the EDP-I model and about
28 hours when using the MDP-C model, all on a top
of the line PC platform. Even if more optimum mod-
eling setups can be conceived of, current computa-
tional capabilities are still unable to cope with
the very large slope geometries that need to be con-
sidered to investigate the type of large scale mass
wasting and sediment transport processes that are
found on the continental slope and rise.
Figure 12. Cross section of deformed 2D mesh. It is therefore important that quantum advances be
made in computational resources and code optimiza-
tion. A simple switch to supercomputer platforms is
geometry. In the latter case, more widespread yield- unlikely to be sufficient without a parallelization of
ing leads to more significant shearing when using the codes. Promising in this regard is that the OpenSees
MDP-C model. It is likely that if a deeper slope had program used herein is now in the process of being
been selected, the difference in predictions between parallelized by its original developers. The intention
the two constitutive models may have been more dra- is to continue our own modeling work by developing
matic in the 3D case as well. The 3D configuration the requisite parallel constitutive models to comple-
leads to smaller overall displacements due to a ment these other efforts.
smaller volume of material that is subject to deform-
ation. It is obvious that 2D simplification of real
slopes may be inadequate to capture the true behavior REFERENCES
of submarine slopes, particularly in cases where there
are significant geometric and material variations in Abramson, L.W., Lee, T.S., Sharma, S. and Boyce, G.M.
the cross-slope direction. 2002. Soil stability and stabilization methods. New York:
Another view of overall deformation is captured in John Wiley & Sons, Inc.
Figure 12, which shows displacements on a cross sec- Drucker, D.C. and Prager, W. 1952. Soil mechanics and plastic
tion of the 2D slope. As expected, the effect of the sur- analysis or limit design. Quart. Appl. Math. 10(2):157165.
charge is to produce a curved indentation in the Drucker, D.C., Gibson, R.E. and Henkel, D.J. 1957. Soil
underlying slope sediment, along with a passive bulge mechanics and work-hardening theories of plasticity.
Transactions Amer. Soc. Civ. Engs. Paper No.
of material in the front of the surcharge. As already 2864:338346.
noted, this occurs as stresses along the interface reach Elgamal, A., Yan, L. and Yang, Z. 2003. OpenSees modeling
frictional yielding and significant plastic shearing. of bridge-foundation-foil interaction. 2003 OpenSees
User Workshop, University of California at Davis,
August 2122.
5 FINAL COMMENTS HBS 2002. ABAQUS theory manual, version 6.3. Hibbitt,
Karlsson & Sorensen, Inc. Main St., Pawtucket, RI.
The boundary value problem chosen for this study is Jeremic, B. 2003. Computational geomechanics: simula-
not meant to be an accurate representation of any par- tions of static and dynamic behavior of foundation sys-
ticular field slope. Instead, it was selected to investi- tem. 2003 OpenSees User Workshop, University of
gate generic modeling issues related to constitutive California at Davis, August 2122.
soil behavior, slope geometry and surcharge loading. Muir, D.M. 1990. Soil behavior and critical state soil
Monumental obstacles remain in achieving satisfac- mechanics. Cambridge: University Press.
tory models of offshore slopes such as may be found Pacific Earthquake Engineering Research Center 2004.
OpenSees web site, http://opensees.berkeley.edu/.
along the continental slope and rise. Obstacles Roscoe, K.H. and Burland, J.B. 1968. On the generalized
include accurate time-history simulation of sediment stress-strain behavior of wet clay. Engineering Plasticity,
transport and deposition processes, generating rea- pp. 535609.
sonable geometries that incorporate typical field TRB 1996. Landslides, investigation and mitigation, Special
dimensions, as well as reliable and sufficient sediment Report 247. Washington, DC: Transportation Research
characterization. It is not uncommon that gravity-driven Board, National Research Council.

902

Copyright 2005 Taylor & Francis Group plc, London, UK


3DSTAB: a history of 3D stability analysis applied in offshore
geotechnics

H.J. Luger, J.L. Bijnagte & J.A.M. Teunissen


GeoDelft, The Netherlands

ABSTRACT: About 20 years ago GeoDelft made the first 3 dimensional stability analyses regarding sub sea
slope stability. Later the the initial computer code was reused and if necessary adapted whenever relevant new
cases arose. The increased availability of detailed digital terrain- and sub sea models as well as the increased
computing power of personal computers made this kind of analyses both desirable as well as possible in a cost
effective manner. The paper explains the basis and implementation of the calculation method and shows results
of several practical cases.

1 INTRODUCTION

In 1985 a question about sub sea slope stability led to


the creation of 3DSTAB: a program for 3 dimensional
stability analyses. Given the available computer
power at that time a somewhat different approach was
needed then the usually adopted strategy for 2 dimen-
sional analyses. The analyses therefore focused ini-
tially on a simple soil profile consisting of normally
consolidated clays with a linear strength increase with
depth. For such cases a semi analytical solution of the
stability factor of spherical and ellipsoidal surfaces
against rotational failure was derived. In this way a
detailed integration of driving and resisting forces
over the volume and the slip surface was avoided.
Instead a rapid assessment of the safety factor of the
= 1 3 4
whole body was possible. This opened the opportun-
ity to include uncertainties with regard to the spatial Figure 1. Map with stability regions of sea bottom
distribution of soil strength into the analyses based on canyons.
basic geostatistical methods such as Kriging interpol-
ation. As a result risk maps could be made of the
investigated area with lines of equal probability of 2 FIRST ADAPTATION OF THE PROGRAM
failure. One of the maps is presented in Figure 1.
Hard to imagine as it may be now, at the time colour Although the first project was a success the limita-
plots were almost impossible to produce so the tions of the soil profile that could be analysed pre-
colours of the presented map had to be added by mak- vented the program from being used intensively.
ing overlays by cutting coloured plastic sheets into the About ten years later the situation was completely
right shape! As a special feature this version included changed by the availability of sufficient computer
so-called retrogressive slope failure: after failure of power. Since this was matched by the increased detail
a slip surface the material of this body was assumed in digital terrain modelling a more generally applic-
to be completely removed thus creating a new soil able 3D stability program became feasible. At that time
surface and possibly new failure. In this way a failure the analytical solution was replaced by the more gen-
could grow to extend backwards into the slope. erally applicable column method, a 3D extension of

903

Copyright 2005 Taylor & Francis Group plc, London, UK


= 1 3 4

Figure 3. Stability of pipeline landfall.

14000

12000

10000

8000
[m]

6000

4000

Figure 2. Calculated safety levels Hycosi project. 2000

0
0 2000 4000 6000 8000 10000 12000 14000
the 2D slice method. Columns were used to model [m]
seabed and subsoil geometry. The slip surface of this
version was limited to a spherical surface only, but a -5 3 10 20 30 40 50 60
layered soil profile was possible. In order to avoid
development of tailor made pre- and post processing Figure 4. 1st sediment layer thickness (in m).
parts of the program the in- and output formats were
made to comply with the requirements of the com-
mercial contouring program Surfer. The input of the positioned between areas that may have some superfi-
terrain consisted of a Surfer grid file while output was cial instability ( 1).
generated to allow processing with Surfer as well.
This greatly increased the manageability of the pro-
gram and thereby the field of application, which ranged 3 EARTHQUAKE ANALYSES
from offshore, e.g. rock dump stability, to onshore
landslide analysis as performed together with BRGM Shortly after this period a dynamic module with
(Bureau de Recherches Geologiques et Minieres) in which earthquake analyses were possible, based on a
the HYCOSI project. The latter was a European pro- Newmark method, was included as well. The method
ject on the Impact of HYdrometereologic Changes On allowed for the incorporation of the effect of remould-
Slope Instability. ing: during the analysis the shear strength could be
Figure 2 shows calculated safety levels for a region reduced based on the displacements caused by the
in the French Alps in both in a 3D view and a top earthquake.
view. Figure 3 also dates from this period and shows the This was used intensively for the analysis of sea
calculated stability factors for a rock dump ramp con- bottom stability and prediction of seabed displace-
structed for a pipeline landfall. The support shown in ments in the event of an earthquake. Some results of
figure 3 has a sufficient safety factor (  3) and is these analyses are presented in Figures 4 to 7. The

904

Copyright 2005 Taylor & Francis Group plc, London, UK


14000 14000

12000
12000

10000
10000

8000
[m]

8000

[m]
6000
6000
4000

4000
2000

2000
0
0 2000 4000 6000 8000 10000 12000 14000
[m]
0
0 2000 4000 6000 8000 10000 12000 14000
1.0 1.3 1.5 1.8 2.0 2.5 4.0 [m]

Figure 5. Calculated safety factors. 0.0 0.0 0.7 1.5 1.7 1.9 2.3 2.7 3.0

Figure 7. Yield acceleration (in m/s2).

14000 In Figure 6 the thickness of the soil layer that is


possibly unstable for a 1000 year earthquake is pre-
12000
sented based on slope angle, soil layer thickness and
soil strength. Figure 7 shows the calculated yield accel-
10000
erations necessary for the generation of instability.
8000
[m]

4 CRITICAL FAILURE DIRECTION


6000

4000
A special point of interest is the fact that it is possible
to investigate the critical failure direction. In contrast to
2000 2D stability analyses in a 3D analysis it is not suffi-
cient to simply specify the slip surfaces and rotation
0 centres. The actual failure direction also depends on the
0 2000 4000 6000 8000 10000 12000 14000 16000 horizontal angle of the rotation axis through the slip
[m]
centre. This angle may be influenced by local loads or
0.0 1.0 1.5 2.0 2.5 3.0 3.5 4.0 5.0
local variation in soil strength. This is illustrated in
Figures 8 and 9 where the axis of rotation is given for
Figure 6. Thickness unstable layer for earthquake (in m). the case of homogeneous soil and loading conditions
and in case a local vertical load is present. In these fig-
ures the projected centre of gravity (c.o.g.) is shown
as well as the projected centre of the sphere (c.o.s.).
scale of these figures is 15  15 km. In Figure 4 the Since a point load generates an additional moment,
thickness of the first, soft, top sediment layer is pre- around a different rotation axis, the final result is a
sented. The white areas represent the areas where no contribution to the driving moment and a rotation of
soft soil is present. This is at the top of the elevated the rotation axis. Clearly a non uniform distribution
seabed as well as in the bottom of some gullies. of weights will have similar effects. In these cases the
In Figure 5 the calculated safety factors are pre- rotation axis remains in a horizontal plane, for a hori-
sented. As can be expected only the combination of a zontal load however the result will generally be that
steep slope and a soft sediment layer results in low the rotation axis will tilt out of a horizontal plane.
safety factors. This is typically the case for the slopes This may be caused by external horizontal loads
of the gullies, whereas the gully bottoms itself are as well as local differences in shear resistance, e.g.
mostly stable due to lack of the soft soil. caused by rock outcrops.

905

Copyright 2005 Taylor & Francis Group plc, London, UK


120

100

c.o.g. 80
Rotation axis
c.o.s. 60

40

Figure 8. Rotation axis (homogeneous conditions). 20

-20

-40
FF n axis
Ro tatio -20 0 20 40 60 80 100 120 140
c.o.g.
Figure 10. Critical slip surfaces (length scale in m).
c.o.s.

120
Figure 9. Rotated axis of rotation due to vertical load.
100

5 PIPELINE LOADS ON SUB-SEA ROCK 80


BERMS
60
An interesting recent application is the analyses of
the effect of multiple (pipeline) loads acting on a rock 40
berm. The critical failure surface for a 2D analysis
would usually be taken as the cross section with the 20
highest load, or is determined based on an assumed
load distribution over the height and width of the rock 0
berm. In the case considered here 3 pipelines were
present and the single pipeline with the highest load -20
was considered to be critical. 3D analyses showed
that a slip surface which includes the 2 lower pipeline -40
loads is more critical. The results of the 3D analyses -20 0 20 40 60 80 100 120 140
are presented in Figure 10 and 11.
For this application a special adaptation was made Figure 11. Critical failure directions (length scale m).
to be able to generate vertical pipeline loads. Each
pipeline load is modelled as a small wedge shaped soil
layer with a unit weight that matches the prescribed
pipeline load. In Figure 9 these loads are visible. clay and sand were compared with semi analytical
solutions and found to be satisfying.
The second step was to compare the program with
6 VALIDATION AND COMPARISON WITH other available 3D stability programmes. Comparison
OTHER METHODS has been done with Clara-W, (Hungr 2001). Because
of the limitations of Clara-W regarding size of the
For 3D stability calculations no generally accepted surface input and pre-defined failure direction, the
method is present and validation of the program was bench marks were limited to some rather simple
therefore limited. As a start the results of slopes in cases. The results of those cases proved to be similar.

906

Copyright 2005 Taylor & Francis Group plc, London, UK


the interior of a slip surface is shown. If, for example,
a rock outcrop is present in area A this means that the
resistance of that part of the slip surface is very high
which increasing the overall safety factor to an unreal-
istically high level.
By checking the internal failure along the surface
A X-X (see Figure 13) the part of the original failure body
on the left of X-X can be replaced by the maximum
B horizontal and vertical force that can be developed in
the surface X-X.
It is still believed that by evaluating a large amount
of slip surfaces the error in the end result will be rela-
Figure 12. Multi area slip surface. tively small. Another slip surface, that does not include
the rock outcrop will then be found to be critical.
Another limitation is that no check is made for the
Rotation axis size of the stresses. There is no limitation to passive
stresses in the passive part of the slip surfaces and the
X possibility of crack formation due to exceeding the
tensile stress in the active part of the slip surface is
also not checked.
This means that even with advanced analyses
X methods as 3DSTAB an experienced engineer is still
needed.

8 SAFETY FACTORS IN 2D AND 3D

A point of interest in the use of 2D and 3D stability


programmes is the value of the safety factors. Since
the methods are generally not completely comparable
Figure 13. Internal body failure due to rock outcrop. the calculated values of the safety factors may differ
even for those cases that are/should be quite similar.
Still 3D analyses may be very valuable since they may
7 LIMITATIONS OF THE PRESENT reveal unexpected failure mechanisms and directions.
PROGRAM In Figure 11 for example the critical direction for the
left side of the central rock berm is to the right!
It should be noted that the program aims at providing
a tool for determination of critical areas of a larger
region. In some cases the area that had to be analysed 9 FUTURE DEVELOPMENTS
was 75  75 km! As a result the choice was made to
take no inter column forces into account since it was A possible near future development includes a graph-
preferred to evaluate of a larger number of slip sur- ical user interface. Further additions that are being
faces rather than a small number of surfaces wth a considered are the introduction of:
check of the internal stability. For the effect of surface
loads this will generally not be a problem since for a horizontal loads
slip surface through a frictional material all of the load load distribution over the columns
will eventually act on the slip surface and thus generate inter column forces
the same amount of resistance while in a non frictional
material the resistance is independent of the load.
However some special situations can occur where REFERENCES
the program may not provide a correct result. For
Hungr, O. 2001. Users Manual CLARA-W, Slope stability
example in case of a multi area slip surface, as pre-
analysis in two or three dimensions for microcomputers.
sented in Figure 12. Clearly the safety factor should O.Hungr Geotechnical research Inc. West Vancouver
be based on the soil body in area A while area B Canada.
should not be taken into account. HYCOSI Impact of Hydrometereologic Changes On
More generally it should be checked if the internal Slope Instability 1997. Final report: BGRM Report R
stability of the slip surface is ensured. In Figure 13 39946.

907

Copyright 2005 Taylor & Francis Group plc, London, UK


A study of ice as an analog of methane hydrate on the basis
of static shear strength

Y.Nabeshima & Y.Takai


Osaka University, Suita, Osaka, Japan

ABSTRACT: A series of the triaxial compression tests on specimens of artificial methane hydrate and ice
were carried out, and the effect of various test conditions such as the temperature, confining pressure, strain rate
and density on the compressive strength of the specimens was investigated. The results suggest that the com-
pressive strength of both methane hydrate and the ice at low density was dependent on all test conditions inves-
tigated, although that of ice at high density was independent of confining pressure. The applicability of ice as an
analog for methane hydrate was discussed through the comparison of both test results. In relation to compres-
sive strength, ice at low density can be substituted for methane hydrate.

1 INTRODUCTION clarified, and the applicability of ice as an analog for


methane hydrate is discussed through the comparison
Methane hydrate is a crystalline structure of water and of both test results.
methane molecules, where the methane molecules are
encaged by water molecules under low temperature and
high pressure. In recent years, hydrate has been con- 2 TEST PROCEDURE
sidered as an encouraging new energy resource possibly
replacing oil and natural gas. Estimates suggest that the 2.1 Sample preparation
volume of methane gas stored in hydrates are sufficient
to last over 100 years based on present levels of natural Methane hydrate samples used in this study are artifi-
gas consumption. Gas hydrates can be found in ocean cially synthesized by mixing methane gas and pure ice
sediments at the edge of the tectonic plates around powder at
2 degrees Celsius (C) and high pressure
Japanese islands under the ocean floor in the Pacific from 8.5 to 10 MPa. In this study, this method of mix-
Ocean (Matsumoto et al. 1993). Methane hydrate is ing methane gas and ice powder was adopted because
stable under low temperature and high pressure. How- it could be synthesized faster than another method
ever, extraction of methane gas from hydrate might which involves mixing methane gas and water under
cause serious disasters such as sliding or mass move- low temperature and high pressure. Powder ice passed
ment of unstable seabed sediments. To prevent such through a 0.425 mm sieve was used in order to corres-
disasters occurring, it is necessary to understand the pond to the powder size of methane hydrate. Both sam-
mechanical properties of gas hydrate from a geotech- ples were compacted statically under high pressure
nical viewpoint. Furthermore gas hydrates are (100 MPa) and formed into columnar specimen. The
metastable and therefore recovery of samples of size of columnar specimen is 15 mm in diameter and
methane hydrate are precious and difficult to obtain 30 mm in height. Figure 1 shows a typical methane
under in-situ conditions. It may be economical to sub- hydrate specimen.
stitute ice for methane hydrate to investigate its
mechanical properties, but to achieve this it is also
2.2 Triaxial compression test apparatus
necessary to clarify the mechanical properties of ice
and compare these with those of methane hydrate. To examine static shear behavior of different speci-
In this study, a series of the triaxial compression mens of methane hydrate and ice, a triaxial test appar-
tests on specimens of artificial methane hydrate and atus which can perform under low temperature and
ice were carried out, and the effect of various test con- high confining pressure is used in this study. Figure 2
ditions such as the temperature, confining pressure, shows external and internal views of this apparatus. A
strain rate and density on each compressive strength is triaxial cell and a reservoir cell water tank are installed

909

Copyright 2005 Taylor & Francis Group plc, London, UK


100

test condition

Pressure (MPa)
10

Approximaton line of
stability boundary
Figure 1. Methane hydrate specimen.
0.1
-60 -30 0 30
Temperature (degree C.)

Figure 3. Test conditions and stability boundary of methane


hydrate.

Standard strain rate is 0.5%/min, and additional com-


pression tests are carried out at 0.1, 1, 2%/min to exam-
ine the strain rate dependency of the compressive
strength.

3 RESULTS AND DISCUSION

3.1 Compressive strength of methane hydrate


Figure 2. Low temperature and high confining pressure Figure 4 shows the relationship between deviator stress
triaxial test apparatus.
and axial strain. It can be seen that the differences in
deviator stress are due to the differences among the
in a freeze chamber. Anti-freeze liquid mixture is used conditions of temperature and confining pressure.
as the cell water. Confining pressure can be applied Here, the maximum deviator stress up to 15% of axial
by high pressure nitrogen gas up to 10 MPa and tempe- strain is defined as the compressive strength.
rature can be reduced to
34C. Axial load is applied Figure 5 shows the relationship between com-
from outside the freeze chamber by a screw jack with pressive strength and density. It can be seen that com-
constant rate of displacement. pressive strength increases with increase in density
Temperature in the triaxial cell is measured by three regardless of temperature and confining pressure. The
thermostats which are installed at the top, center and increase in compressive strength with increase in
bottom positions of the specimen. confining pressure at 0.9 g/cm3 is bigger than that at
0.7 g/cm3 regardless of temperature. Therefore, the
dependency of compressive strength on the density is
2.3 Test conditions recognized. In this study, for convenience, 0.9 g/cm3
A series of triaxial compression tests for methane is defined as high density and 0.7 g/cm3 as low density.
hydrate and ice is carried out to investigate the effect of In the following sentence, we discuss the compressive
various test conditions on their compressive strengths. strength at high density.
Figure 3 shows the positions of all test conditions and Figure 6 shows the relationship between the com-
approximation line of the stability boundary of methane pressive strength and confining pressure at different
hydrate which is summarized from previous chemical temperatures for methane hydrate. It can be seen that
experimental results by Hyodo et al. (2002). Based on compressive strength increases with increase in con-
the stability conditions of methane hydrate, we decided fining pressure regardless of temperature, and also
to use temperatures of
10,
20,
30C and confin- compressive strength increases with decrease in tem-
ing pressures of 2, 3, 6 MPa. Standard density is about perature under the same confining pressure.
0.90 g/cm3, and additional compression tests are car- Figure 7 shows the relationship between the com-
ried out at around 0.80, 0.70 g/cm3 to examine the pressive strength and strain rate at
30C and 6 MPa
influence on the compressive strength at low density. conditions. It can be seen that compressive strength

910

Copyright 2005 Taylor & Francis Group plc, London, UK


12

Compressive strength (MPa)


15
Deviator stress (MPa)

10 12
8
9
6
6
4
2 -30C.,6MPa -30C.,3MPa 3
-10C.,6MPa -10C.,3MPa
0 0
0 5 10 15 20 0.01 0.1 1 10
Axial strain (%) Strain rate (%/min)

Figure 4. Relationship between deviator stress and axial Figure 7. Relationship between compressive strength and
strain for methane hydrate. strain rate at
30C and 6 MPa for methane hydrate.

Compressive strength (MPa)


12 8
Compressive strength (MPa)

-30C.,6MPa -30C.,3MPa 7
10 -10C.,6MPa -10C.,3MPa 6
8 5
4
6
3
4 2 -30C.,6MPa -30C.,3MPa
1 -10C.,6MPa -10C.,3MPa
2
0
0 0.65 0.7 0.75 0.8 0.85 0.9 0.95
0.65 0.7 0.75 0.8 0.85 0.9 0.95 Density (g/cm3)
Density (g/cm3)
Figure 8. Relationship between compressive strength and
Figure 5. Relationship between compressive strength and density for ice.
density for methane hydrate.

3.2 Compressive strength of ice


Figure 8 shows the relationship between compressive
Compressive strength (MPa)

12
strength and density for ice. It can be seen that com-
-30C.
10 -20C. pressive strength increases with increase in density
8 -10C. regardless of temperature and confining pressure,
similar to that of methane hydrate. As the density of
6 ice increases, the difference of compressive strength
4 between 3 and 6 MPa at any temperature decreases. In
contrast, the difference between
30 and
10C at the
2 same confining pressure increases. Therefore, a sep-
0 arate discussion with each density is necessary to dis-
0 1 2 3 4 5 6 7 cuss the compressive strength of ice.
Confining pressure (MPa) Figure 9 shows the relationship between compres-
sive strength and confining pressure. Solid lines show
Figure 6. Relationship between compressive strength and the regression lines of experiment results at high dens-
confining pressure for methane hydrate. ity while the dotted lines show the regression lines of
experiment results at low density. From this figure, it
can be seen that compressive strength at high density
linearly increases with increase in strain rate when is independent of confining pressure but compressive
strain rate is plotted on a semilogarithmic scale. The strength at low density is dependent on confining pres-
increase in compressive strength for methane hydrate sure. Since ice specimens at high density were com-
is 80% (per log cycle of strain rate) and it is much pacted to almost maximum density with high pressure,
larger than that of clay, which has an increase in com- their compressive strength is insensitive to confining
pressive strength around 5 to 10% for the same given pressure. On the other hand, ice specimens at low dens-
strain range. ity may be porous and have some voids, and therefore,

911

Copyright 2005 Taylor & Francis Group plc, London, UK


7

Compressive strength (MPa)


-30C. High
14 -20C. 6
density -30C.
Compressive strength (MPa)

-10C.
12 -30C. Low 5 -20C. MH
-20C. -10C.
10 -10C. density 4
-30C.
8 3 -20C. Ice
6 2 -10C.
4 1

2 0
0 1 2 3 4 5 6 7
0 Confining stress (MPa)
0 1 2 3 4 5 6 7
Confining pressure (MPa) Figure 11. Relationship between compressive strength and
confining pressure for methane hydrate and ice.
Figure 9. Relationship between compressive strength and
confining pressure for ice.
confining pressure, therefore the applicability of ice
at low density as an analog for methane hydrate is
examined. Figure 11 shows the relationships between
Compressive strength (MPa)

16
14 MH compressive strength of methane hydrate and ice with
12 Ice confining pressure at low density. It can be seen from
10 this figure that compressive strength of ice at low dens-
8 ity of 0.7 g/cm3 is close to that of methane hydrate
regardless of confining pressure, therefore, the appli-
6
cability of ice as an analog for methane hydrate is
4
confirmed.
2
0
0.01 0.1 1 10
4 CONCLUSIONS
Strain rate (%/min)
A series of the triaxial compression tests on specimens
Figure 10. Relationship between compressive strength and
strain rate at
30C and 6 MPa for methane hydrate and ice. of artificial methane hydrate and ice were carried out to
investigate the compressive strength of methane hydrate
and ice, and the following results were obtained.
their compressive strength is dependent on confining
1 The compressive strength of methane hydrate is
pressure. The dependency of temperature on compres-
dependent on temperature, confining pressure, strain
sive strength on temperature is due to the increase in
rate and density.
strength of the ice crystal.
Figure 10 shows the relationship between the com- 2 The compressive strength of ice is dependent on
pressive strength and strain rate at
30C and 6 MPa temperature, strain rate and density. Although com-
conditions. It can be seen that compressive strength of pressive strength of ice at low density is dependent
ice linearly increases with increase in strain rate when on confining pressure, that of ice at high density is
plotted on a semilogarithmic scale. This is similar to independent of confining pressure.
methane hydrate, but increase in compressive strength
3 Ice at low density (0.70 g/cm3) is applicable as a
for ice is smaller than that of methane hydrate.
methane hydrate analog because their compressive
strength are similar regardless of confining pressure.
3.3 Applicability of ice as analog methane hydrate
From discussion in previous sections, although com-
pressive strength of methane hydrate and ice is REFERENCES
dependent on temperature, strain rate and density, the
Hyodo, M et al. 2002. Triaxial compressive strength of
compressive strength of ice at high density is inde-
methane hydrate. Proceedings of the Twelfth International
pendent of confining pressure. This is different from Offshore and Polar Engineering Conference: 422428.
the case of methane hydrate. In other words, the com- Matsumoto, R et al. 1993. Methane hydrate - Huge 21st
pressive strength behavior of methane hydrate can not century natural gas resources, Nikkei Science: 39144.
be represented by ice at high density. However, the com-
pressive strength of ice at low density is dependent on

912

Copyright 2005 Taylor & Francis Group plc, London, UK


Tackling geohazards a case study from the Turkmenistan shelf,
Caspian Sea

J. Wegerif & M. Galavazi


Fugro Engineers BV, Leidschendam, The Netherlands

I. Hamilton
Fugro Survey Limited, Aberdeen, U.K.

Z.B.A. Razak
Petronas Carigali SDN. BHD., Kuala Lumpur, Malaysia

ABSTRACT: This paper presents a case study on geohazards affecting hydrocarbon field development in
Block1, offshore Turkmenistan and the approach taken to minimize the financial and environmental risks asso-
ciated with these hazards. The field development area covers approximately 2000 square kilometers and is
located in water depths ranging from 40 m to 100 m. The geohazard assessment in Block1 has identified mul-
tiple geohazards based on exploration seismic data, geotechnical data and several rig site surveys. The most sig-
nificant hazards include active faulting, mud volcanoes, shallow gas and seismic activity. Less dramatic but
equally important features are seabed ridges and cemented authigenic carbonates. The identification of these
hazards in the conceptual phase of development allowed their incorporation in the field design without major
cost implications, and led to a change in development concept from multiple directionally drilled wells from a
single platform to mostly vertical wells from different platforms. The resulting field layout allows mud volca-
noes, fault scarps and shallow gas occurrences to be avoided, and active faults were considered for well plan-
ning, thus minimizing the financial and environmental risks.

1 INTRODUCTION there was no detailed geotechnical or high-resolution


survey data available and exact locations of facilities
Petronas Carigali intends to develop the Livanov, were not yet decided upon. Therefore, a solution had
Barinov and Gubkin fields on the Turkmenistan shelf to be found that would fit the tight timeframe as well
of the Caspian Sea. The fields are located in the as allow for a realistic conceptual engineering study.
Block1 area, approximately 65 km to 95 km offshore As no further data acquisition could be performed
Turkmenistan (Fig. 1). prior to the CES, the available data for the study was
Block1 covers an area of 2000 km2 and the water restricted to the existing exploration 2D and 3D seis-
depth ranges from 40 m in the east to 100 m in the mic data, three small-scale 2DHR surveys and some
west of the area. Little is known about the detailed shallow soil borings. Based on this limited data, a
geology of the Turkmenistan shelf and the Turkmenistan geohazard map was produced of the Block1 area,
part of the Apsheron Ridge. However, during previous which allowed for siting of surface facilities in low
developments in Soviet times, many wells, platforms risk areas.
and pipelines have collapsed as a result of geohazards,
indicating that the area is prone to risks associated
with geohazards. 2 GEOLOGICAL SETTING
Performing a Conceptual Engineering Study (CES)
in this type of environment is challenging as risk and The Caspian Sea lies near the boundary between the
uncertainties are ample. The main challenge was that Arabian and Eurasian continental plates. The relative
an in-depth analysis was required of a 2000 km2 area motion of these plates results in crustal compression
in one of the most geohazard-prone areas in the world of this region. The mountain ranges surrounding the
within a time frame of three months. At this stage, southern and central Caspian Sea are fold and thrust

913

Copyright 2005 Taylor & Francis Group plc, London, UK


sedimentation rate of 2 km/My (Brunet et al. 2003).
This rapid, clay dominated sedimentation combined
with hydrocarbon generation and tectonic forces has
led to the formation of highly overpressured, under-
consolidated clays at Maikopian (Oligocene to Early
Miocene) level (Brunet et al. 2003). Mud volcanism
occurs where these overpressured mud and fluids
escape to the seabed.

3 METHODOLOGY

Interpretation of the available data for the Block 1


geohazard study included the following steps:

1 regional geological modelling and probabilistic


seismic hazard analysis
2 seabed reflection interpretation on 3D seismic data
to produce pseudo-bathymetry map
3 geohazards interpretation on seismic data sets and
screening of the 3D seismic data for drilling hazards
4 correlation of geotechnical data and seismic data
to define geotechnical soil zones

The first step in the geohazard interpretation for


Block 1 was to produce a geological model of the
Figure 1. Location of Block1 on the Turkmenistan shelf, Turkmenistan shelf and eastern Apsheron Ridge area
Caspian Sea. and to collect information on historical seismicity in a
radius of approximately 500 km around the Block1 site.
Jackson et al (2002) showed that the Apsheron Ridge
belts formed as a result of the crustal shortening from historical seismicity is dominated by deep (focal depth,
plate collision (Yilmaz 1997, Jackson et al. 2002). d  30 km) seismic events, whereas in the surrounding
The South Caspian Basin is thought to be a remnant fault block mountains, seismicity is dominated by rela-
of oceanic crust from the ancient Tethys Sea from the tively shallow (d 30 km) seismic events. For the
late Mesozoic or early Tertiary. This basin is being analysis, the historical seismicity was grouped into
subducted beneath both the Eurasian and Arabian deep events of the Apsheron Ridge region and shallow
plates at the margins of the basin. Of importance to the events of the surrounding regions. The Gutenberg-
Block1 area is the subduction zone at the northern Richter formulation (Gutenberg & Richter 1956) was
edge of the South Caspian Basin, in the centre of the adopted to define earthquake recurrence within each
Caspian Sea. This subduction zone, manifested as the of the seismic sources, which served as input for a
Apsheron Ridge (Jackson et al. 2002, Brunet et al. seismogenic model. Based on this model, a probabilis-
2003), is in the direct vicinity of Block1 and is one of tic seismic hazard analysis (PSHA) was carried out to
the main controls on the seismic hazards at the site. determine the expected peak ground accelerations
The Apsheron Ridge stretches across the Caspian required for preliminary design.
Sea, from the Caucasus Mountains in the west to the Next, existing data was interpreted for shallow
Kopet Dag Mountains in the east. Thrust faulting in geohazards and drilling hazards. This interpretation
the deeper Mesozoic strata caused by subduction of was based on the following available data sets in the
the South Caspian Basin, resulted in the formation of Block1 area:
an anticline in the overlying Tertiary and Quaternary
strata. Extensive, deep and shallow seated faulting total coverage of Block1 with 3D exploration seis-
occurs along the crest of the ridge and many of these mic data to 1.5 second two-way travel time (TWT)
faults are active and exposed at seabed. a regional 2D exploration seismic data set covering
Although the South Caspian Basin is thought to the western half of Block1
have originated in Mesozoic times, the bulk of the high resolution 2D site surveys covering three
sedimentary infill consists of Oligocene and younger. 2 km  2 km areas
Approximately 10 km of Plio-Pleistocene sediments interpretive maps of regional analogue survey side
have accumulated in the basin, representing an average scan sonar and shallow seismic profiler data

914

Copyright 2005 Taylor & Francis Group plc, London, UK


Figure 2. Regional shallow geohazard map (to 200 m sub-seabed).

four detailed geotechnical site investigations vary- of the seismic data, correlating actual soil units across
ing from 40 m to 80 m depth the site proved difficult. However, changes in layer
thickness and changes in character could be determined
The seabed reflection interpretation from 3D seismic from the seismic data, which were used together with
data was used to produce a bathymetry map. The seis- the available geotechnical data to provide a basic
mic TWT in milliseconds to the seabed was converted geotechnical stratigraphy.
to water depth in metres using a calibration based on
comparison with echo sounder data acquired as part
of the site surveys. 4 RESULTS
The bathymetric map produced from the 3D seabed
interpretation provides a useful overview of water Several seabed and sub-seabed features that may
depth variations, details of seabed features and relief. impact development were identified in Block1. These
Seabed features such as fault scarps, mud volcanoes, features are presented on the geohazard map shown in
pockmarks and carbonate mounds are clearly visible Figure 2. A detailed description of the geohazards is
on the bathymetry. A 3D seabed interpretation bathym- given below.
etry map, is therefore the key to a first understanding
of seabed features and geohazards. It is noted, how- 4.1 Seismicity
ever, that such a bathymetry map has some limitations The Caspian Sea, and particularly the Apsheron Ridge,
related to the limited vertical resolution of the 3D is considered a highly seismic area. The probabilistic
data at seabed level, especially in water depths shal- seismic hazard analysis indicated that peak ground
lower than 50 m. Seeing through soft seabed sedi- accelerations of 0.20 g should be taken into consider-
ments and noisy data in areas of irregular seabed ation for preliminary design. This preliminary value
reduce the accuracy of the seabed pick. represents a conservative estimate for the complete
Utilising an appropriate combination of vertical soil column, as at the time of study, no detailed fault
sections, time-slices, horizon slices, amplitude maps analysis was as yet available. Further study may result
and horizon-based attribute maps, the foundation in a more precise estimate of peak ground acceleration.
zone to approximately 200 m sub-seabed was inter-
preted for geohazards. Additionally, the 3D seismic
4.2 Faults
volume was screened for evidence of drilling hazards
to approximately 1000 m sub-seabed. An extensive transtensional fault system was identi-
Detailed geotechnical site investigations varying fied in the Block1 area. The faults form a series of lin-
from 40 m to 80 m in depth were available for four ear features with curved ends, forming bends in the
platform sites in the Block1 area. Using geophysical otherwise linear fault traces (Fig. 2). Strike-slip dis-
data and borehole data, generalised geotechnical soil placement along such curvi-linear faults produces a
zones were determined for the Block1 area. Due to complex zone of deformation, resulting in a strike-
the complex geology and the limited vertical resolution slip duplex (Twiss & Moore 1999). On vertical sections

915

Copyright 2005 Taylor & Francis Group plc, London, UK


Figure 3. Vertical section showing normal displacement Figure 4. Mud mound type mud volcano.
along faults and ridges at seabed.

(Fig. 3), the faults display normal displacement, indi-


cating an extensional dextral strike-slip system. This
transtensional system is associated with thin skin tec-
tonics at the crest of the Apsheron Ridge.
The shaded relief image of the bathymetry shows
that many fault traces have significant offsets at seabed.
Sea level curves for the Caspian Sea (Kosarev &
Yablonskaya. 1994) indicate that the Block1 area has
been exposed above sea level during the Mangyshlak
regression approximately 10,000 years ago. Sub-aerial
erosion would probably have removed any previously
formed fault scarps. It is therefore probable that seabed
fault scarps observed at presents represent fault move-
ment over a maximum period of about 10,000 years.
Fault scarps at seabed in excess of 5 m, and locally up Figure 5. Seabed depression type mud volcano.
to 12 m, suggest a high level of fault activity over the last
10,000 years, and fault movement may be considered
a serious hazard to structures near a fault. depression type mud volcanoes, also known as salsas
(Jakubov et al. 1971). The mud mound type mud
volcanoes (Fig. 4) are associated with the release of
4.3 Shallow gas hydrocarbon fluids and mud at seabed, producing a
A broad zone of disseminated shallow gas is observed significant mound around the venting area. Seabed
along the central axis of the Block1 area. The gas is depression type mud volcanoes (Fig. 5) probably emit
probably not pressurized and is unlikely to be a water and hydrocarbon fluids, but no mud. Except for
drilling hazard, as no problems were encountered in a slight bulge due to fluid escape, no mound is
the area when drilling the geotechnical boreholes. formed at seabed (Jakubov et al. 1971, Hovland et al.
However, within the interpreted 1000 m sub-seabed, 1997, Planke et al. 2003).
localised potentially pressurized pockets of shallow gas The mud volcanoes in the area are interpreted to
were observed on 3D seismic data. The gas pockets may actively seep gas, water and locally mud. A mud out-
be a hazard to drilling. A number of wrecked platforms, flow feature approximately 3 km long and 500 m to
observed on side scan sonar data from the regional ana- 900 m wide at seabed, suggesting a recent eruption
logue survey, may be attributed to gas blow-outs due to was observed at one of the features. Coarse fragments
drilling such pressurized pockets of shallow gas. of claystone, characteristic of mud volcano deposits,
were observed in samples from the flanks of a num-
ber of mud volcanoes in the Block1 area. In addition,
4.4 Expulsion features
evidence of slumping on the flanks of the mud mound
In Block1, two types of mud volcanoes were identi- type mud volcanoes is clearly visible on side scan
fied: mud mound type mud volcanoes and seabed sonar records, suggesting unstable slopes.

916

Copyright 2005 Taylor & Francis Group plc, London, UK


Further hazards associated with mud volcanoes of influence of the identified hazards, was a useful tool
may be over-pressured mud and fluids and disturbed in assessing the impact on field development, and in
soil conditions. A wrecked platform has been observed finding the solution to manage the associated risks.
on the flank of one of the mud volcanoes, which por- As the geohazard assessment was performed in the
trays the unstable nature of these features. conceptual design phase of the project, full integra-
Buried mud volcanoes are also observed in the tion of the hazards into the conceptual field design
Block1 area. The buried mud volcanoes appear to be was possible. Some examples of geohazard-driven
cut-off from the source of fluids and mud. However, changes to the preliminary design are presented in
disturbed soil conditions may result in hole instability more detail below.
and form a drilling hazard.
5.1 Impacts of faulting
4.5 Gas seepage and authigenic carbonates
A zone of very closely spaced faults was identified in
Gas seepage along faults and from isolated vents is the center of Block1. These faults are active and spac-
extensive in the Block1 area. Many faults appear to ing is typically less than 1 to 2 km. The presence of
act as migration routes for gas. The seepage is mostly these faults inhibited the extensive use of directional
concentrated in the areas of extensive faulting along drilling techniques, as the wells would have crossed
the center of the Apsheron Ridge. On 2D seismic data active faults thereby risking the drilling operations as
from the regional survey plumes of gas and suspended well as the integrity of the wells. Therefore the decision
material are often visible in the water column around was taken to change the development concept from
areas of faulting, indicating active seepage at present. multiple directionally drilled wells from a single plat-
Associated with gas seepage are carbonate concre- form to mostly vertical wells from different platforms.
tions, interpreted as methane derived authigenic car- This decision significantly increased the required
bonates. These carbonates form due to bacterial activity number of platform sites, but significantly reduced
at the gas and seawater interface. The authigenic car- risks for drilling. Based on the hazard chart, drilling
bonates have a rock like character and form rough sites were selected that allowed unfaulted well paths.
patches on the seabed. Extensive carbonate concretions The fault offsets at seabed in Block1 can be as
are present along faults outcropping at seabed and in much as 5 m to locally 12 m. Besides the fact that
vents and mud volcano craters (Hovland 1990, Johnson platforms should not be sited over these active faults,
et al. 2003). there is also a strong impact on pipeline routing.
Areas of gas seepage and carbonate concretions Steep ridges are difficult, if not impossible to cross
are considered a risk for siting of a platform. The auth- due to the limitations for pipeline curvature, and fault
igenic carbonates in the Block1 area form isolated pat- movement during the life of the field may potentially
ches and ridges, which may be underlain by soils with cause rupture of the pipes. Some platform sites had to
potentially reduced shear strengths due to their high be repositioned to minimise the crossing of fault
gas content. This may result in a punch-through risk. ridges by pipelines.

4.6 Mass transport features 5.2 Impacts of shallow gas


Mass transport complexes as a result of slope failures Several areas were identified which showed a high
are common throughout the South Caspian Basin. In risk of over-pressured shallow gas occurrences in the
the Block1 area, they occur at various depths and are upper 1000 m sub-seabed. These suspected gas occur-
generally very thick. The mass transport complexes rences form a serious drilling hazard as they could
are mostly buried features which have no seabed potentially cause uncontrolled gas blow-outs and
expression. There is no direct evidence of drilling unstable hole conditions leading to the loss of a hole.
hazards associated with these features; however, these Mapping these suspect zones and avoiding them in
disturbed sediments may be more prone to unstable the field layout prevented unexpected and costly
hole conditions and loss of circulation than undis- changes to the layout at a later stage in development.
turbed sediments.
5.3 Impacts of seismicity
5 IMPACTS ON FIELD DEVELOPMENT The preliminary conservative value of expected max-
imum ground acceleration of 0.20 g is a challenge for
Several of the above mentioned geohazards proved engineering design of the planned platforms and
to have significant impact on the preliminary field facilities. At a later stage of development, more detailed
design. The regional hazard chart (Fig. 2), which pro- modelling of ground acceleration should provide a
vides an overview of the spatial distribution and zone more accurate value.

917

Copyright 2005 Taylor & Francis Group plc, London, UK


Unfortunately, unlike most of the other hazards in drilling and platform sites and relocation of wells and
Block1, seismicity cannot be avoided. The only way platform sites to low-risk areas before spending large
to deal with seismicity is to incorporate the maximum sums on detailed surveys in high-risk areas.
expected ground acceleration into the design of the pro-
posed facilities and thereby minimizing the cost effects.
REFERENCES
5.4 Impacts of expulsion features Brunet, M-F., Korotaev, M.V., Ershov, A.V. and Nikishin,
The expulsion features in Block1 comprise mud vol- A.M. 2003. The South Caspian Basin: a review of its evo-
lution from subsidence modeling. Sedimentary Geology
canoes and pockmarks, which are associated with the 156: 119148.
expulsion of mud, water and gasses or a combination Gutenberg, B. and Richter, C.F. 1956. Earthquake Magnitude,
of these. These processes are active and adversely Intensity, Energy and Acceleration. B. Seism. Soc. Am.
affect foundation stability, drilling operations and 46 (2): 143145.
platform and pipeline siting. By mapping the max- Hovland, M. 1990. Do carbonate reefs form due to fluid
imum extend of mud flows and slope failures, it was seepage? Terra Nova 2: 818.
possible to avoid these features in the field design and Hovland, M., Hill, A. and Stokes, D. 1997. The structure and
to determine safe stand-off distances. Also, by map- geomorphology of the Dashgil mud volcano, Azerbaijan.
ping mud volcano and pockmark vents, drilling prob- Geomorphology 21: 115.
Jackson, J., Priestley, K., Allen, M. and Berberian, M. 2002.
lems could be minimised. Expulsion features are
Active Tectonics of the South Caspian Basin. Geophys.
generally isolated and the impact on the total field lay- J. Int. 148: 214245.
out was minimal. Only a few platform sites required Jakubov, A.A., Ali-Zade, A.A. and Zeinalov, M.M. 1971.
relocation due to close proximity of expulsion fea- Mud volcanoes of the Azerbaijan SSR: atlas (in Russian).
tures, and relocation distances never exceeded 1 km. Azerbaijan Academy of Sciences, Baku.
Johnson, J.E., Goldfinger, C. and Suess, E. 2003. Geophysical
constraints on the surface distribution of authigenic car-
6 CONCLUSIONS bonates across the Hydrate Ridge region, Cascadia margin.
Marine Geology 202: 79120.
This study shows that with the right approach and Kosarev, A.N. and Yablonskaya, E.A. 1994. The Caspian Sea.
timely recognition, geohazards need not have a major Planke, S., Svensen, H., Hovland, M., Banks, D.A. and
Jamtveit, B. 2003. Mud and fluid migration in active mud
impact on offshore hydrocarbon field development, volcanoes in Azerbaijan. Geomarine Letters 23: 258268.
not even in one of the worlds most geohazard prone Twiss, R.J. and Moore, E.M. 1992. Structural Geology.
and geologically complex areas. Freeman, W.H. and Company, 1992.
Assessing geohazards as early as the conceptual Yilmaz, Y. 1997. Techtonics of the East Anatolian-Caspian
phase allows upfront risk assessments for proposed Regions. The Leading Edge: 889891.

918

Copyright 2005 Taylor & Francis Group plc, London, UK


Site investigation techniques

Copyright 2005 Taylor & Francis Group plc, London, UK


Deepwater geotechnical site investigation practice in the Gulf of Guinea

D. Borel & A. Puech


Fugro France, Nanterre, France

H. Dendani & J.L. Colliat


Total, La Dfense, Paris, France

ABSTRACT: The development of deepwater fields offshore West Africa has been accompanied by a major
evolution of the techniques used for geotechnical site investigations. The foundation engineering of the deep
water structures installed in this part of the world generally requires soil parameters only over the first 20 to 40
meters below seabed. The traditional down-hole sampling and in situ testing techniques have been replaced by
a combination of in-situ testing from seabed modules and high quality piston coring. These geotechnical tech-
niques are described and their performance is highlighted. The need for new tools dedicated to specific require-
ments is emphasized.

1 INTRODUCTION heave compensated and stabilized by means of a heavy


seabed frame. All sampling and in-situ testing oper-
A major trend in the industry is that most of the struc- ations are carried out by the wireline technique with
tures considered and built in deep waters only require specialized tools operated through the drillstring.
geotechnical design parameters to a depth of a few The most efficient geotechnical vessels active in
meters to a few tens of meters. This is the case for lines West Europe and West Africa (M/V Bavenit and M/V
in contact with seabed (pipelines, flowlines, umbilicals, Bucentaur) have the capability to drill in water depths
Steel Catenary Risers) and for suction caissons used of up to 1100 m when using standard steel drillstrings.
for anchoring or for supporting subsidiary structures. This capability has been extended to about 2 000 m by
The first deep water geotechnical site investigation introducing aluminium drillstrings. This technique
in the Gulf of Guinea (GoG) was conducted in 1998 was used in 1998 on the Girassol field in 1300 m water
on the Girassol field. Since that date geotechnical depth for the first deepwater site investigation in the
data have been gathered in a number of sites located Gulf of Guinea.
offshore Nigeria, Equatorial Guinea, Congo or Angola Deep penetration borings may be required for
in water depth ranging from 450 to 1 500 m. geohazard assessment (potential slumping, faulting
The main objective of this paper is to describe the features, etc.) and for conductor pipes or casing
geotechnical site investigation procedures which are installation studies. Certain types of structures con-
currently applied in this part of the world. They are sidered for deepwater like Tension Leg Platforms also
based on a combination of soil sampling by giant piston need deep investigations. However most of the struc-
coring and in-situ testing from seabed modules. This tures designed and built in deep waters for the oil and
is a major change when compared to traditional site gas industry only require geotechnical parameters to
investigations based on wireline sampling and testing a depth of a few meters to a few tens of meters. This
procedures through 5( API drillstrings. Operation is the case for lines in contact with the seabed (e.g.
times and costs are drastically reduced. pipelines, flowlines, umbilicals, Steel Catenary Risers)
or for suction caissons used for anchoring floating
structures (e.g. FPSO) or as foundations for support-
2 GEOTECHNICAL INVESTIGATION NEEDS ing seabed subsidiary structures (e.g. templates,
IN DEEP WATERS PLEMs). Once the regional geology is known through
geophysical surveys and, where applicable, deep
In shallow water depths ( 450 m) geotechnical boreholes the actual collection of sediment character-
investigations are traditionally performed by conven- istics required for establishing design parameters can
tional drilling using 5( API drillstrings which are often be limited to the top 20 to 40 m only.

921

Copyright 2005 Taylor & Francis Group plc, London, UK


Figure 1. Deep water seacalf penetrometer.

Characterizing the top 20 to 40 m of deep water


sediments for geotechnical engineering purposes can
Figure 2. Deep water seascout penetrometer.
be achieved via two complementary methods:
performing in-situ testing using seabed modules;
collecting soil samples using giant piston corers.
These techniques are more attractive and econom-
ical in deep waters than deploying a drillstring. This
explains their large success over the last five years.
Most of the deepwater site investigations in the Gulf
of Guinea are performed by combining in situ testing
with seabed modules and giant piston coring. This
should apply also for future ultra-deep water projects
as prevailing technical and economical aspects remain
the same.

3 IN- SITU TESTING IN DEEP WATERS

In-situ tests in deep waters can be performed from Figure 3. T-bar 250-40 penetrometer.
heavy or light frames lowered to the sea bottom.
The piezocone test (CPTU) is a standardized test Shallow penetration accurate testing can be add-
consisting of pushing at constant speed into the soil a itionally performed by using light weight seabed equip-
rod equipped with sensors allowing a measurement of ment e.g. the Fugro DeepWaterSeascout (Fig. 2).
three parameters called cone resistance (qc), local These seabed modules allow the implementation
friction (fs) and excess pore water pressure (u). of other in situ testing techniques like vane shear test-
These parameters alone or in combination provide an ing (VST) or T-bar testing.
identification of the soil types. They also give access The vane shear test provides a direct measurement
to the undrained shear strength of the clays Su via an of the peak undrained shear strength of the soil at pre-
empirical correlation factor Nk expressed as: qn  Nk determined intervals (typically 1 m) by rotating blades
Su, qn being the net cone resistance (measured cone into the soil at a controlled speed. A measurement of
resistance corrected for hydrostatic and transient pore the remolded shear strength Sur is possible by repeat-
pressures, in-situ stress and cone construction). ing the test after several fast revolutions of the blades.
Deep penetration tests in the Gulf of Guinea are per- The T-bar test was recently introduced (Randolph
formed with heavy seabed equipment, as the Fugro et al. 1998). The test is quite similar in principle to the
Deepwater Seacalf (Fig. 1) or the Ifremer Penfeld, CPTU test except that the cone is replaced by a short
capable to push the standard 10 cm2 cone to respect- cylindrical bar attached perpendicularly to the pen-
ively 40 and 30 m penetration. etrometer rods (Fig. 3). Compared to the CPTU test,

922

Copyright 2005 Taylor & Francis Group plc, London, UK


Undrained shear strength Su (kPa)
0 5 10 15 20 25 30 weight-stand
0

upper pulleys
2

piston cables
4

corer tubes
Depth [m]

piston
8
base plate
ML

10
T-bar
VST
lower pulleys
12

Figure 4. T-bar results with Nt  11.5.

this equipment presents the major advantage that the Figure 5. Stacor principle.
ambient stress level is self-equilibrated across the bar
at top and bottom surfaces. The bearing resistance qT
is then obtained directly by the load measurement
divided by the bar area with no corrections for hydro-
static and transient pore pressure and for overburden known Kullenberg corer widely used for oceanologi-
stress. The flow of soil around the bar can be modeled cal purposes in theoretically unlimited water depth.
with an exact plasticity solution and closed form The basic principles and the main drawbacks of these
solutions to give a Nt factor (Nt is equivalent to the systems are described in Borel et al. 2002.
cone factor Nk in CPT test) fin the range 9.4 to 11.9 The STACOR system is a satisfactory comprom-
depending on bar roughness. The T-bar currently used ise between the operational flexibility of gravity
in the Gulf of Guinea is a 250  40 mm cylinder corers and the absolute requirement of obtaining undis-
which can be pushed down to 20 m of penetration and turbed samples. A detailed description can be found in
at a controlled rate of 20 mm/s. Figure 4 discloses an Borel et al. 2002. The STACOR is a seabed piston
example of undrained shear strength profile obtained corer that differentiates itself from others in the sense
from T-bar measurements using a Nt factor of 11.5. that it has a truly stationary piston. During the freefall,
The match with VST data is remarkable. the piston is maintained stationary by a system of
cables and pulleys connected to a base plate standing
The T-bar test is however more subject to lateral
drift and potential damage than a cone penetrometer. on the seabed. The principle is shown in Figure 5.
It is therefore preferable to use this system as a com- The STACOR is permanently installed on the M/V
Bavenit (Fig. 6) and a total of more than 100 samples
plementary test to investigate accurately the very soft
have been obtained in the Gulf of Guinea over the last
clays within the first 5 m of penetration.
5 years in water depths ranging from 300 to 1500 m.
The average working time to take a core in 1000 to
4 CORING FOR OIL AND GAS 1500 m water depth is 8 to 10 hours.
APPLICATIONS A detailed analysis of the data obtained to date
shows that:
4.1 The STACOR piston corer the corer as installed on the M/V Bavenit can rou-
Giant piston corers capable of sampling up to 50 m of tinely reach penetrations of about 20 m in deep sea
soft sediments have been derived from the well soft materials (Fig. 7);

923

Copyright 2005 Taylor & Francis Group plc, London, UK


25

20

15

Number
10

70%

75%

80%

85%

90%

95%

100%

105%
Recovery (%)

Figure 8. Stacor recovery offshore Africa.

the material should be as undisturbed as possible:


what is the degree of remoulding acceptable to
derive representative mechanical properties?
Whereas the second point is a permanent concern for
geotechnical engineers, the first point is not as well
perceived as stratigraphical continuity is often a priori
assumed.
Figure 6. Giant Stacor stationary piston corer.
The assessment of the core integrity should be based
on a combination of a detailed geological description
of the core, X-ray photographs and implementation of
Water depth (m) logging techniques such as the multi-sensor core log-
0 200 400 600 800 1000 1200 1400 1600 ger (MSCL) which provides a continuous profile of
10
soil unit weight (gamma-densimetry), compressive
wave velocity, and magnetic susceptibility. This
allows the detection of potential anomalies (cracks,
Recovery (m)

15
remolded zones) before opening of the liner and
aids in selecting representative sections for mechan-
ical laboratory testing. It is now established that
20
STACOR samples provide an accurate picture of
the stratigraphy over the recovery height.
Remoulding of samples taken in deep waters is due
25
Ivory Coast Nigeria Eq. Guinea to several reasons, the most commonly cited being:
Congo Angola Egypt
the mode of sampling: push sampling is generally
considered as one of the least disturbing modes in
Figure 7. Stacor performance offshore Africa. offshore geotechnics; at the opposite end, gravity
coring in Kullenberg mode is known to cause
severe remoulding.
recovery rates are very high, typically in excess of variations in confining pressure: in 1500 m water
90% (Fig. 8); depth, the natural volumetric expansion of a sam-
the quality of the samples is acceptable for engineer- ple (in the absence of gas) is 0.7%;
ing purposes. This point is developed hereafter. possible presence of dissolved gas in pore water,
including sublimation of hydrates.
4.2 Quality of STACOR samples The degree of remoulding can be conveniently
expressed by a disturbance index Id:
Two aspects are of importance in assessing the qual-
ity of a piston corer: (1)
the litho-stratigraphic integrity must be preserved: where: e  variation of void ratio when a sample is
is the sampled material an exact picture of the sedi- reconsolidated to its in situ confining pressure,
ment column in place? eo  initial void ratio

924

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 1. Assessment of sample quality from disturbance Undrained shear strength (kPa)
index (after Lunne et al. 1998). 0 5 10 15 20
0
Sample quality Disturbance index Id (%)
0.2
Very good to excellent 4
Fair to good 47

Depth (m)
Poor 714 0.4
Very poor 14
Box core - Fallcone
0.6
test
Box Core Lab. Vane
Undrained shear strength (kPa)
0.8 test
0 10 20 30 40 50
Deep water
0
Seascout CPTU
CPT A with Nk = 14 1
2 CPT B with Nk = 14
DSS tests Figure 10. Laboratory tests on box core samples.
4

6 5 HIGH QUALITY SUPERFICIAL SAMPLING

A detailed characterization of the geotechnical prop-


8
erties of the soils immediately below the seabed, is
Depth (m)

essential for a proper assessment of soil-pipe inter-


10 actions. This is particularly critical in the deepwater
sediments of the Gulf of Guinea which present some
12 unusual properties in terms of water contents, shear
strength profiles and yield strength ratios (Puech et al.
14 2005, De Gennaro et al. 2005).
The STACOR tool is certainly well adapted for
16 sampling down to about 20 m of penetration but some
concerns have arisen on its capacity to capture the
18 subtle changes in the geotechnical parameters in the
very tens of centimeters below the seabed due to pos-
sible washing of the extremely soft materials at the
20
mudline in the initial phase of the free fall penetration.
Figure 9. CPTU and DSS results. Box corer sampling has been recently introduced.
The box corer is a very simple devise that envelops
an area of seabed then seals the base of its box to retain
Id can be obtained through oedometer tests, advanced the sample from further disturbance during recovery.
triaxial tests or direct simple shear tests. According to The standard box corer consists of a steel frame incorp-
Lunne et al. (1998), the degree of disturbance of nor- orating the sample box surmounted by a 200300 kg
mally consolidated samples (OCR 2) can be rated as weight. The sample box is first pushed 40 to 50 cm
indicated in Table 1. These criteria applied to more than into the seabed by the action of the weight. The trig-
200 samples recovered from Gulf of Guinea deep ger mechanism releases a latch that allows the swivel
waters indicate that samples taken by the STACOR base to close off the captured sample, before the whole
rate typically between 2 and 7% (fair to excellent). unit is recovered to the surface where the sample is
Laboratory testing procedures have been adapted removed for examination. The sample size is gener-
to accommodate the preparation of very soft mater- ally 0.45  0.3  0.2 m.
ials and the monitoring of tests under very low con- The recovered sample is of extremely high quality
fining pressures reproducing the in situ stress state. It and allows the performance of any type of laboratory
is now considered that advanced laboratory tests per- testing. Figure 10 presents fallcone and laboratory vane
formed on high quality STACOR samples provide shear strength measurements performed at very close
representative shear strength data which can be used intervals (of the order of 5 to 10 cm) on a box core.
with confidence in the design of suction anchors. They clearly confirm the presence of a peak in
Figure 9 presents a comparison of shear strength undrained shear strength encountered on some sites
profiles obtained from DSS tests performed on a within the first meter of penetration and revealed by
STACOR corer and from a neighboring CPTU test. CPTU profiles (Puech et al. 2005)

925

Copyright 2005 Taylor & Francis Group plc, London, UK


6 OPTIMIZING DEEPWATER penetration below the seabed. It is recognized that for
GEOTECHNICAL SITE INVESTIGATIONS such applications there is a need for a tool capable to
IN THE GOG take 3 to 5 m undisturbed samples and easier to oper-
ate than the giant STACOR.
CPTU tests provide a continuous measurement of the
soil resistance and their synthesis allows the appraisal
of the soil homogeneity over the whole site area. 7 CONCLUSIONS
CPTUs penetrating between 20 and 40 m will remain
the basic in-situ test in the future. The introduction of the giant STACOR piston corer
CPTU testing should be performed systematically has modified the strategy of geotechnical investiga-
at the following locations: tions in the deepwater sites of the Gulf of Guinea.
Combining STACOR sampling and CPTU testing
FPSO and buoy mooring anchors;
from seabed units has proved to be a cost effective
riser towers;
and technically satisfactory solution to obtain the
heavy manifolds, PLEM structures.
parameters required for the geotechnical engineering
It is recommended to couple some deep CPTU of most types of deepwater structures. This basic
tests by in-situ vane tests to at least 20 m penetration solution is extendable to deeper waters, although deep
in order to acquire direct in situ measurements of the boreholes may remain indispensable for some spe-
undrained shear strength and calibrate the Nk cone cific applications.
factor. VST also provides valuable information on the In order to optimize the geotechnical site investi-
soil sensitivity. It should be recognized that VST test- gations, it is strongly recommended to:
ing takes significant time but the additional informa-
combine different in-situ measurement techniques
tion is highly valuable.
for a better appraisal of complex soil conditions;
High quality piston coring should be performed at
adapt the tools most appropriate for each type of
a selected number of locations where advanced labora-
engineering problem;
tory tests are required. Once again it is recommended
diversify the sampling tools according to the differ-
to couple piston cores and CPTUs in order to provide
ent objectives.
a sound basis for extrapolating design parameters to
the whole site area. As a minimum giant piston cores
should be taken at the following locations:
REFERENCES
one per anchor cluster;
one at the riser tower location; Borel, D., Puech, A., Dendani, H. and de Ruijter, M. 2002.
at a selected number of Manifold or PLEM loca- High quality sampling for deepwater engineering: the
tions, covering the whole site area. Stacor experience. Ultra Deep Engineering and
Technology (UDET), Brest, France.
The pipeline and flowline routes as well as the De Gennaro, V., Puech, A. and Delage, P. 2005. On the com-
SCR touch down points (where applicable) should be pressibility of deepwater sediments of the Gulf of
investigated with specialized tools adapted to the very Guinea. Proc. Intern. Symp. on Frontiers in Offshore
soft soils encountered at very shallow penetration. Geotechnics, Perth, Australia.
Light CPTU modules, preferably equipped with large Lunne, T., Berre, T. and Strandvik, S. 1998. Sample disturb-
cones, can provide valuable data down to 5 m below ance effect in deep water soil investigations. Proc.
mudline. They should be performed at regular spa- Offshore site investigation and foundation behaviour,
SUT, London.
cing along the routes, in order to assess the spatial vari- Puech, A., Colliat, J.L, Nauroy, J.F. and Meunier, J. 2005.
ability of the superficial soil conditions. T-bar testing Some geotechnical specificities of Gulf of Guinea deep-
should also be considered because the tool has a great water sediments. Proc. Intern. Symp. on Frontiers in
potential to capture the very low undrained shear Offshore Geotechnics, Perth, Australia.
strength values which characterize the very top sedi- Quiros, G.W. and Little, R.L. 2003. Deepwater soil proper-
ments and govern soil-pipe interactions. ties and their impact on the geotechnical program. Proc.
In addition to shallow CPTU and T-bar tests, intact Offshore Technology Conference, Houston, Tx, OTC
samples should be recovered along the flowline paper 15262.
routes. This can be done using a box corer or any Randolph, M.F., Hefer, P.A., Geise, J.M. and Watson, P.G.,
1998. Improved seabed strength profiling using T-bar
other system capable of sampling a certain volume of penetrometer. Proc. Offshore site investigation and
undisturbed sediment near the surface. foundation behaviour, SUT, London.
A certain number of engineering problems (e.g.
soil-riser interactions at touch down point (TDP),
bearing capacity of small seabed structures) require
high quality sampling over the first few meters of

926

Copyright 2005 Taylor & Francis Group plc, London, UK


The origin of near-seafloor crust zones in deepwater

C.J. Ehlers & J. Chen


Chevron Texaco Energy Technology Company, Houston, Texas, U.S.A.

H.H. Roberts
Coastal Studies Institute, Baton Rouge, Louisiana, U.S.A

Y.C. Lee
Fugro-McClelland Marine Geosciences, Inc., Houston, Texas, U.S.A.

ABSTRACT: In recent years, geotechnical site investigations for deepwater oil and gas developments have
disclosed the presence of a thin crust zone of anomalous high strength immediately below the seabed in sev-
eral deepwater regions around the world. The origin of the crust zone, however, has not been satisfactorily
explained with conventional soil mechanics. The sediments within the crust do not have physical property val-
ues or definable trends in water content, density, plasticity and grain size distribution that explain why the crust
exists. Previous marine biological studies have shown that biological activity or bioturbation affects geotechni-
cal properties including soil shear strength. A variety of sedimentological and geochemical tests together with
geotechnical tests were performed on cores recovered at two deepwater sites offshore Nigeria where a crust is
present. This paper presents the test results that suggest that the crust zone is both a product of intense bio-
turbation and early geochemical alteration of burrow walls and pelletized burrow fills. The paper begins with a
literature review on the effects of biological activity on sediment physical properties.

1 INTRODUCTION physical explanation that takes into account the effects


of ionic concentration changes.
With the increase of subsea oil and gas development This paper examines bioturbation and early geo-
in deepwater, more attention has been given to the geo- chemical changes in the sediment as possible causes
technical properties in the first few meters below the for the crust zone. Bioturbation has been shown in pre-
seabed, because the soil properties have a very large vious studies to affect geotechnical engineering prop-
impact on the design and installation of shallows foun- erties. In this paper, geotechnical properties and data
dation for subsea equipment, flowlines, and risers. In on bioturbation are presented for two deepwater loca-
recent site investigations, a crust zone of anomalous tions offshore West Africa in an attempt to establish a
high strength immediately below the seabed has been connection between bioturbation and the crust zone.
reported in several deepwater regions around the world.
Accurate definition of the thickness and strength of
such a crust is possible today with the use of conven-
2 EFFECTS OF BIOLOGICAL ACTIVITY ON
tional piezocone penetrometers and through strength
SEDIMENT PHYSICAL PROPERTIES
testing on high quality samples recovered with a box
core sampler. The origin of the crust zone, however, has
2.1 Bioturbation
not been satisfactorily explained with conventional soil
mechanics. The sediments within the crust do not have Most near seafloor deepwater sedimentary environ-
definable trends in water content, density, plasticity ments are subject to a variety of post-depositional
and grain size distribution that explain why the crust processes that serve to remobilize and alter the sedi-
exists. Others (Sultan et al. 2001) have suggested that mentary deposit before permanent burial. Biological
the over consolidation observed in a crust zone at a processes can have important effects on the physical
deepwater site in the Gulf of Guinea offshore West properties and behavior of sediments. Physical and
Africa has no mechanical origin and have presented a chemical alteration of sediments occurs when benthic

927

Copyright 2005 Taylor & Francis Group plc, London, UK


animals or fauna move about and feed. Bacterial deg- (3) biological interactions such as competition for food
radation of organic matter and chemical reactions and predator-prey interactions, and (4) pressure. In
change the nature of the sediment and create chemical addition, low oxygen has profound effects on the bot-
gradients that induce solute transport through the pore tom fauna, causing the zonation of fauna to be highly
fluids. Invertebrate macrofauna and smaller fauna modified (Rowe & Haedrich 1979).
called meiofauna (less than 1.0 mm in size) are abun-
dant and alter the sediment fabric in many different
2.3 Depths and zones of sediments mixing by
ways. Some organisms construct rigid tubes within the
macrofauna
sediment to allow free exchange of water and solutes
between the sediments and the overlying water (Aller Deepwater (3500 m) hemipelagic samples recovered
1980) or inject water directly into the sediment habi- in the Venezuela Basin (Richardson et al. 1985) dis-
tats during feeding, burrowing, and locomotory life closed that the upper 58 cm of cores was characterized
processes (McCall et al. 1979). Altogether, these by a uniform dark yellowish-brown, with numerous
processes are termed biological stirring or mixing, or very fine burrows less than 1 mm in diameter. This
simply bioturbation (Matisoff & Robbins 1987). layer was less opaque to X-rays and corresponds to
The rate of bioturbation varies with temperature, the Berger et al. (1979) definition of a well-mixed zone.
amount and input rate of organic matter in the seabed, Below this zone to depths of 1519 cm, the sediment
type of benthic fauna, and abundance of organisms. was characterized by a yellowish-brown color with
Bioturbation rates decrease with sediment depth, due increasing amounts of dark olive gay sediments
to decreased abundance of organisms. The effects of occurring mostly as mottles or streaks. As indicated by
biological activity are generally important only within X-radiographs, this layer had the greatest amount of
the upper 1 m of sediment. biologically derived structure with numerous horizontal
In addition to the physical and chemical alteration and vertical burrows and feeding voids up to 0.5 cm
of sediments, the redistribution of sediment particles by in diameter. This layer corresponds to the Berger et al.
organisms can obscure primary stratigraphic features (1979) transition zone. The transition zone is the limit
and create other secondary structures. Bioturbation of advective transport of oxygen to sediments and
invariably leads to poorer resolution of microfossil dat- removal of metabolic wastes (Richardson et al. 1985).
ing of deep sea sediments by lengthening the sediment Below the transition zone, the sediments were a
section in which index fossils are found and by over- uniform olive gray with little evidence of horizontal
lapping the horizons of fossils indicative of distinct mixing. Sediment structure in this layer was primarily
time intervals (Berger & Heath 1968). either sedimentological in origin or larger structures
of biological origin (1.0 cm in size) including infilled
tubes and burrows. This layer corresponds to the Berger
2.2 Zonation of benthic fauna with water depth
et al. (1979) definition of a historical zone.
A universal feature in the distribution patterns of ben-
thic fauna on the continental slope is the tight zonation
2.4 Effects of bioturbation and organic carbon on
of the fauna. Slight changes in water depth across con-
sediment geotechnical properties
tinental slopes result in radical changes in what lives
there, but over extensive distances along a depth con- Bioturbation has been shown to alter geotechnical prop-
tour the composition of a community is altered very erties including: (a) porosity/void ratio/density, (b)
little. More recent work has shown that some animal grain size distribution, (c) shear strength, and (d) con-
groups are congregated in such a way that the slope solidation characteristics. The study of Reimers (1982)
fauna can be subdivided by statistical criteria into stresses that diagenetic processes, interdependent with
groups of species arranged in vertical depth zones porosity, significantly affect geotechnical properties.
(Rowe & Haedrich 1979). The studies of Busch & Keller (1981, 1982, 1983)
Why animals are narrowly zoned on the continen- show, as a first approximation, good agreement between
tal slope is not known. As temperature variations and organic carbon content, sediment age, and geotechnical
gradients are slight below the permanent thermocline, properties, particularly porosity. What remains is to
a gradient of temperature seems unlikely to affect the elucidate the integrated relationship between physical
distribution of fauna. Other parameters, with the excep- properties and organic phases as they change with
tion of pressure, do not appear to be arranged in a gra- space, time, and burial (Reimers 1982).
dient along which the organisms range themselves.
Consistent trends over the depth gradient can seldom 2.4.1 Forms of organic matter
be seen in sediment particle size, organic matter or sedi- Organic matter occurs in two forms: organic sheaths
ment physical properties (Rowe & Haedrich 1979). and organo-mineral aggregates. Sheaths have been
Suggestions advanced to explain zonation include identified as slimy substances formed in situ by fila-
(1) bottom currents, (2) variations in food sources, mentous bacteria by Gallardo (1977) and as membranes

928

Copyright 2005 Taylor & Francis Group plc, London, UK


enclosing zooplankton fecal pellets in the Baltic Sea the consolidation characteristics are different. The
and at several deep ocean stations by Schrader (1971) extent of these differences depends upon the source,
and Honjo (1980), respectively. decomposition, and depositional history of the organic
Organo-mineral aggregates appear to be the biode- matter.
graded remains of the interior of fecal pellets (Honjo & Many benthic animals produce mucous secretions
Roman 1978). Invertebrate fecal material usually in as an aid in locomotion and feeding, whereas micro-
the form of pellets is a biogenous feature of deepwater algae and bacteria produce secretions as an anchorage
sediments (Kuenen 1961 & Young 1971). Reimers or protective mechanism. The resulting effect is
(1982) found that aggregates combine a mixture of increased soil shear strength that is strongest near the
mineral and amorphous phases, including fecal matter, seabed where most biological activity is concentrated
opal, and clay minerals and that everything in the (National Research Council 1989).
2 m size fraction appears packaged in aggregates, Richardson et al. (1985) suggest that the shear
which is further evidence that humic substances are ini- strength of the surface sediments is largely the result of
tially synthesized while clay and organic materials are biological activities coupled with biologically mediated
being thoroughly mixed in the guts of benthic organ- chemical changes. The average oxidation state is
isms such as zooplankton (Greenland 1965). Deposit perhaps the most critical parameter in predicting the
feeders, such as the various kinds of worms that live chemical reactions in sediment. In general, deepwater
in the surface sediments, annually pass large quantities sediment on the continental slopes is oxidizing,
of sediment through their digestive tracts to remove although there are exceptions. The amount of available
organic nutrient, whether detritus or the microbiota oxygen affects the abundance and species composition
living on the detritus and in the process making bur- of benthic organisms, which can strongly modify the
rows and dumps of fecal pellets (Emery 1963). biogeochemical processes that occur (Aller 1982).
Rhoads (1974) stated that high shear strengths were
2.4.2 Effect on soil porosity the result of extensive burrowing and tube-building.
The amount of reduction in porosity with increasing Burrow abundance and geometry can considerably alter
depth of burial varies considerably among sediment distribution of the principal oxidants within sediments,
types and appears to be predominantly controlled by and the average size and distribution of burrows can
physico-chemical conditions and the depositional envir- strongly influence fluxes of solutes across the sediment-
onment rather than by overburdern stress (Bennett water interface (National Research Council 1989).
1976). Organo-mineral aggregate has the properties of Richardson et al. (1985) states that enough organic
a cementing material, creating a network of local bod- carbon was present to create anaerobic conditions for
ies separated by irregular large voids/pores. With nitrate, manganese, iron and sulfate reduction to occur
increasing sediment burial and biochemical stripping at hemipelagic deepwater sites. Mobilization and
of organic sheaths, the volume of these irregular large reprecipitation of these and other oxidants may result
voids and the porosity is decreased, but micropore in chemical bonding that would lead to higher shear
space ( 2 m) within aggregates persists. There is a strengths in the transition zone. Distinctive changes
general tendency for the microstructure to change in pore size and water content may be correlated with
towards an increasingly uniform system of organo- rapid remineralization of sheath-like organic matter
mineral aggregates. This effect can be attributed to the by sulfate reduction (Reimers 1982).
decomposition and increasing humification of organic
compounds and consolidation (Reimers 1982).
3 TEST RESULTS ON CORES FROM
2.4.3 Effect on soil strength and compressibility DEEPWATER OFFSHORE NIGERIA
Busch & Keller (1981, 1982, 1983) describe sediments
typical of the upper slope of the Peru-Chile continental At two deepwater sites designated A and B (1390
margin that have extremely high water contents and and 1493 m water depth) offshore Nigeria, piezocone
plasticity and higher than expected undrained shear penetration tests and strength tests performed on
strengths and sensitivities. Consolidation tests show a recovered 100-mm-diameter jumbo piston drop cores
combination of apparent over consolidation with high showed the presence of a crust zone approximately
compressibility and pronounced secondary compres- 1.21.5 m thick at the seafloor. Figure 1 shows the
sion. Busch and Keller concluded that interparticle anomalous high strength profile of the crust at the
bonding due to organic matter and aggregate formation two locations based on the piezocone penetration test
was probably responsible for the initial strength and the (PCPT) data.
magnitude of the apparent over consolidation. Although The upper about 1.5 m of the two cores that included
the mineralogy and organic carbon content may be the crust zone were analyzed using X-ray radi-
fairly equal in surface sediments and in sediments ography (to determine sediment structure and inclu-
buried to several meters, the pore size distribution and sions), X-ray diffractometry (to determine types and

929

Copyright 2005 Taylor & Francis Group plc, London, UK


relative abundances of minerals), inductively coupled 3.2 X-ray radiography
plasma emission spectrometer (to determine elemen-
X-ray radiographs (Figure 2) show that both A &
tal concentrations), total organic carbon analysis, and
B location cores are fine grained, thoroughly bur-
index property tests (water content, unit weight,
rowed at a variety of scales, and contain evidence of
Atterberg limits, and grain size determinations). The
early diagenetic alterations, primarily in association
test results are discussed below.
with burrow structures. The top portions of both cores
display the effects of large burrows, probably poly-
3.1 Physical/index properties chaete worms that are oriented roughly parallel to the
seabed and generally filled with less dense sediment
Physical/index property testing performed on the two
as compared to the surrounding sediment matrix.
cores from Nigeria included water content, unit wet
Both cores displayed a contact defining two slightly
weight, Atterberg limits, and grain size determination.
different sediment types based on structure and inclu-
The results of these tests (Table 1) indicate high-plas-
sions observed in the X-ray radiographs. The contact
ticity clay, with decreasing water content and corr-
occurs at about 2225 cm and 20 cm from the top of
esponding increasing unit weight down the crust zone
the two cores and is interpreted to be the transition-
which is similar to normally consolidated hemipelagic
historical zone interface. Above the contact, burrows
sediments without a crust. There was no observed peak
are up to 1.5 cm diameter with very little evidence of
or trough of the water content and unit weight at the
small-scale burrowing. Below the contact, burrows of
middle of the crust zone, where the shear strength is the
0.2 cm diameter and smaller are common to the bot-
largest. The sediments within the crust do not have
tom of the core sections at a depth of 1.5 m below the
physical property values or definable trends that
seabed. In addition, high density (to X-radiation) fil-
explain why the crust exists. All these data are similar
aments, small nodular masses, and possible small
to those obtained using identical PCPT and soil sam-
shell fragments are present.
pling methods from a West African site designated
A scanning electron photomicrograph of one of the
Site X in a water depth of about 900 m without a
high density filaments show that the high density
crust (Table 1).
material filling these filaments is pyrite in spherical
framboidal morphologies that are thought to represent
the pyritized remains of Beggiatoa bacterial strands,

Figure 1. Evidence of crust zone from piezocone data. Figure 2: X-Ray radiographs for topmost seabed soil.

Table 1. Summary of physical/index property data.

Water Submerged unit Plastic Liquid limit Clay content


Location content [%] weight [kN/m3] index [%] [%] [%]

Nigeria Sites A & B 132211 2.43.1 97109 142147 9991


Site X 160240 2.13.0 110130 150160 Not Available

930

Copyright 2005 Taylor & Francis Group plc, London, UK


which is the largest bacterium currently known to All other quantitative data sets (components of
exist and can be seen with the unaided eye (to 200 mineralogy and elemental concentrations) are inter-
microns diameter). Beggiatoa sequesters sulfur gran- preted as showing no definable trends.
ules in its tube-like body. When buried in marine sedi-
ments, the transition from sulfur to iron sulfide (pyrite)
is a common reaction. Abundant iron is available from 4 SUMMARY OF TEST RESULTS
clay minerals. Iron combines with the sulfur to produce
iron monosulfied. In the presence of hydrogen sulfide, A broad spectrum of sedimentological and geochemi-
pyrite and hydrogen are produced. Also, iron mono- cal testing methods were focused on the problem of
sulfide with microbial mediation can combine with sul- determining the origin of an anomalous zone of high
fur to form pyrite. The abundance of these pyritized sediment strength commonly referred to as the crust
filaments in the cores analyzed is not unusual as com- zone. Opposing trends in mineralogical constituents,
pared to sediments of the Gulf of Mexico slope. specifically quartz increasing down-core and calcite
increasing up-core at both coring sites, are to be
expected in deepwater settings where the source of flu-
3.3 X-ray diffraction vially-derived sediment is modulated by sea level. The
interpretation of these results is that at lowered sea
Table 2 summarizes the X-ray diffraction data for min-
level when a source of fluvial sediment (Niger River)
eralogy for the two cores. The data show quartz increas-
was closer to the shelf edge, more quartz-rich fluvial
ing down-core and calcite increasing up-core in both
sediment was deposited on the slope, causing the high-
cores and is interpreted as a sea-level dependent
est quartz content to be deepest in the core sections
trend, representing the last sea level rise and a reduc-
analyzed. As sea level rose to present, the source of flu-
tion of fluvial sediment input to the continental slope.
vial sediment retreated to its current position on the
These trends extend through the crust zone and are
inner shelf and left the slope to be impacted by micro-
not obscured by the crust zone.
fossil-rich hemipelagic sediments as reflected in the
Another trend in the data is that pyrite generally
high calcite content at the tops of each core. No dating
increases down-core at both sites with exception of a
was performed on these cores, but the fine-grained
slight decrease in pyrite (0.58% and 1.7%) within the
nature of the sediment, abundant microfossil tests, and
central part of the crust zone. Iron and sulfur in the
lack of turbid flow units in the stratigraphy suggest that
elemental data follow the same trend.
the crust zone is no older than about 1012 kyr BP.
The X-ray radiographs provide perhaps the most
diagnostic data. These images of sediment structure and
3.4 Total organic carbon
inclusions indicate a correlation between the most
Total organic carbon (% by weight) was greatest in intensely burrowed sections and the highest soil shear
the upper 20 cm of the two cores, ranging from 3.07 strength. At a depth of 20 to 25 cm below the top of
to 3.70 in one core and from 3.67 to 3.75 in the other the two cores, burrows tend to be smaller ( 0.2 cm
core. Below about 20 cm, the organic carbon content diameter) in the crust than at the top of the cores at
decreased and was relatively constant with depth, ran- each site. The larger burrows in the top of the cores are
ging from about 2.5 to 2.8 in one core and from about filled with sediment that is low density to X-radiation
2.6 to 2.9 in the other core. and possibly low density surficial sediment introduced

Table 2. Summary of X-ray diffraction data; Cores A & B.

Depth BML Quartz Calcite K-Feldspar Plagio-clase Pyrite Clay


[cm] [%] [%] [%] [%] [%] [%]

Core A
15 8.0 9.2 1.6 1.6 0.7 78.9
65 8.3 2.0 1.5 1.6 2.1 84.5
100 9.0 0.6 1.8 1.7 0.6 86.2
130 10.1 0.6 1.5 1.7 3.4 82.5
Core B
15 7.7 9.4 1.3 1.8 1.1 78.7
65 8.1 2.6 2.2 1.7 1.9 83.5
100 8.1 2.3 2.0 1.7 1.7 84.1
130 10.2 0.5 2.0 1.4 2.9 83.0

931

Copyright 2005 Taylor & Francis Group plc, London, UK


into the burrows. The underlying zones of smaller bur- the effects of biological activity on deepwater sedi-
rows that correlate with the crust zone are filled with ments and providing references.
higher density sediment than the surrounding matrix.
This increase in density to X-radiation (lighter tone)
is subtle, but present throughout most of the crust. REFERENCES
These intensely burrowed parts of the cores from both Aller, R. C. 1980. Relationships of tube-dwelling benthos with
sites are overprinted with filaments/burrows filled or sediment and overlying water chemistry. In Tenore, K. R.
lined with very high density material, as interpreted and Coull (eds), Marine benthic dynamics: 285308.
from the X-ray radiographs. Scanning electron photo- Columbia: University of South Carolina Press.
micrographs identify this high density material as Aller, R. C. 1982. The effects of macrobenthos on chemical
pyrite. No definitive diagenetic minerals could be dis- properties of marine sediment and overlying water. In:
tinguished in the burrow fills that had a subtle increase McCall, P. L. and Tevesz, M. J. S. (eds), Topics in geobi-
in density above matrix values on X-ray radiographs. ology, Vol. 2, Animal-sediment relations: 53102. New
York: Plenum.
Bennett, R. H. 1976. Clay fabric and geotechnical properties
5 CONCLUSIONS of selected submarine sediment cores from the Mississippi
delta. Ph.D. Thesis, Texas A&M University: 269 pp.
The results of this study on two cores from deepwater Berger, W. H., Ekdale, A. A. and Bryant, P. F. 1979. Selective
offshore Nigeria suggest that the crust zone is both a preservation of burrows in deep-sea carbonates. Marine
geology, 32: 205230.
product of intense bioturbation and early geochemical
Berger, W. H. and Heath, G. R. 1968. Vertical mixing in pelagic
alteration of burrow walls and pelletized burrow fills. sediments. Journal of marine research, 26: 134143.
Both the burrow walls as well as sediment within the Busch, W. H. and Keller, G. H. 1981. Physical properties of
burrows appear to be undergoing alteration, primarily Peru-Chile continental margin sediments the influence
to pyrite. The intricate architecture of the burrows, the of coastal upwelling on sediment properties. Journal of
frequency of their occurrence, and early diagenetic Sedimentary Petrology, 51, Issue 3: 705719.
products contained within the burrows all appear to Busch, W. H. and Keller, G. H. 1982. Consolidation charac-
contribute to the development/origin of the crust teristics of sediments from the Peru-Chile continental
zone with the anomalous high strength profile. margin and implications for past sediment instability.
Marine geology, 45:1739.
The above findings and conclusions are in agree-
Busch, W. H. and Keller, G. H. 1983. Analysis of sediment
ment with the those of previous studies reported in the stability on the Peru-Chile continental slope. Marine
literature and discussed herein, namely, (1) Richardson Geotechology, 5, Issue 2: 181211.
et al. (1985) shear strength of the surface sediments Emery, K. O. 1963. Organic transportation of marine sedi-
is largely the result of biological activities (bioturba- ments. M. N. Hill (ed.), New York: Interscience
tion) coupled with biologically mediated chemical Publishers. 3, 776793.
changes, (2) Rhoads (1974) high shear strengths are Gallardo, V. A. 1977. Large benthic microbial communities
the result of extensive burrowing and tube-building, in sulphide biota under Peru-Chile subsurface counter-
and (3) Richardson et al. (1985) the presence of current. Nature, 268: 331332.
Greenland, D. J. 1965. Interaction between clays and organic
organic carbon (in seabed sediments) creates anaer-
compounds in soils. Part II, Adsorption of soil organic
obic conditions for nitrate, manganese, iron and sulfate compounds and its effects on soil properties. Soils and
reduction to occur, and the mobilization and reprecipi- fertilizers, 28: 521532.
tation of these and other oxidants may result in chemi- Honjo, S. 1980. Material fluxes and modes of sedimentation
cal bonding that would lead to higher shear strengths. in the mesopelagic and bathypelagic zone. Journal of
This research is considered a pilot study that took marine research, 38: 5397.
a megascopic view of the crust zone. Future study of Honjo, S. and Roman, M. R. 1978. Marine copepod fecal
the crust zone should take a microscopic view and pellets: production, preservation and sedimentation.
focus on the burrows and their filled interiors as well Journal of marine research, 36: 4557.
Kuenen, H. 1961. Some arched and spiral structures in sedi-
as the matrix material in which the burrows occur. In
ments. Geology en mignbouw, 40: 7174.
addition, future study is needed to develop an under- Matisoff, Gerald and Robbins, John A. 1987. A model for
standing of why (1) the shear strength increases and biological mixing of sediments. Journal of geological
then decreases in a more or less linear manner within education: 35, 144149.
the crust zone, and (2) the crust disappears and is not McCall, P. L., Tevesz, M. J. S., and Schweigien, S. F. 1979.
buried with sediment deposition. Sediment mixing by lampsilis radiata siliquoidea (mol-
lusca) from western lake erie. Journal of great lakes
research, 5: 105111.
ACKNOWLEDGEMENTS National Research Council 1989. Our seabed frontier: chal-
lenges and choices. Report of the committee on seabed
The authors wish to thank James Hooper of Fugro- utilization in the exclusive economic zone, 1415.
McClelland Marine Geosciences for awakening us to Washington, D.C.: National Academy Press.

932

Copyright 2005 Taylor & Francis Group plc, London, UK


Reimers, C. E. 1982. Organic matter in anoxic sediments off Rowe, G. T. 1974. The effects of the benthic fauna on the
central peru: Relations of porosity, microbial decom- physical properties of deep-sea sediments. In
position and deformation properties. Marine geology, A. L. Inderbitzen (ed.), Deep-sea sediments: physical and
46: 175197. mechanical properties: 381400. New York: Plenum.
Richardson, M. D., Briggs, K. B. and Young, D. K. 1985. Schrader, H. J. 1971. Fecal pellets: role in sedimentation of
Effects of biological activity by abyssal benthic macroin- pelagic diatoms. Science, 174: 5557.
vertebrates on a sedimentary structure in the venezuela Sultan, N., Cochonat, P., Cauquil, E. and Colliat, J. L. 2001.
basin. Marine geology, 68: 243267. Apparent over-consolidation and failure mechanisms in
Rhoads, D. C. 1974. Organism-sediment relations on the marine sediment. In Aubeny, C. and Briaud, J-L (eds),
muddy sea floor. Oceanography and marine biology: an Proceedings of OTRC 2001, geophysical conference hon-
annual review, 12: 263300. oring prof. Wayne A. Dunlap: 86102.
Rowe, G. T. and Haedrich, R. L. 1979. The biota and biolog- Young, D. K. 1971. Effects of infauna on the sediment and
ical processes of the continental slope. The society of seston of a subtidal environment. Vie et milieu, 22:
economic paleontologists and mineralogists special pub- 557572.
lication no. 27, 4959.

933

Copyright 2005 Taylor & Francis Group plc, London, UK


Casing mounted method used for geotechnical drilling in Lavan

A. Fakher
Associate Professor, Geotechnical Group, Civil Engineering Department, Tehran University &
SAHEL Consultant Engineers, Tehran, Iran

A. Cheshomi
Head of Geotechnical Investigation Department, SAHEL Consulting Engineers, Tehran, Iran &
PhD Candidate, Tarbiyat-Moddares Univ.

ABSTRACT: The paper aims to present a modified casing mounted method for geotechnical drilling used in
Persian Gulf. The present geotechnical investigation was undertaken for a SBM project which is located at
4.5 km South-West of Lavan Island in a water depth of 55 m. The geotechnical investigation of the project
includes the drilling of 5 boreholes, undisturbed sampling and also CPT probing adjacent to boreholes. The
depth of boreholes into the seabed was 60 m.
A floating barge of 12000 kN was utilized for geotechnical investigation. The horizontal movements of the
barge were minimized by means of mooring anchors. Tension forces of mooring cables were measured during
the drilling operation to predict any horizontal movement of the barge. The vertical movements were compen-
sated by installing the drilling machine on a modified vertical casing pipe.
The following results were achieved out of successful implementation of the project:
Accurate active control on horizontal movements of the barge by four anchors and force measurements were
successfully experienced.
A light casing mounted method for installation of drilling system was designed and experienced.

1 INTRODUCTION
due to fluctuation of water. When a floating vessel is
used, drilling is usually undertaken through a moon
1.1 Marine geotechnical investigation in Iran
pool at the centre of vessels. A seabed frame is also
Considering the existence of offshore resources of oil & used to perform CPT. For the presented geotechnical
gas in Iran, a large number of geotechnical investi- investigation; it was not possible to adopt common
gations have been implemented in the Persian Gulf. methods of practice to isolate drilling bits and sam-
Borehole drilling and undisturbed sampling in water pling devices from the movement of the vessel. The
depth of less than 20 m is generally implemented using use of a conventional seabed frame was not also prac-
jack-up barges (Fakher & Pahlavan 2000) in Persian tical due to a very soft surface layer encountered.
Gulf and Caspian Sea for near shore projects. Lessons Geotechnical investigations at the site of the SBM
learned from using jack-up barges for geotechnical project, 4.5 km on South-West of Lavan Island, a
drilling in Iranian waters have been described by famous island for its oil and gas related projects in
Fakher and Pahlavan (2000). Qualitative investigations the area, were due to be performed by Falat-e-Ghareh
and disturbed sampling into very deep water have been Oil Company, a subsidiary of National Iranian Oil
also performed in Caspian Sea in a water depth of Company. These investigations consist of reconnais-
about 700 m using shallow samplers. sance survey of pipeline route (4.5 km long from
For high quality geotechnical drilling and sampling Lavan to the site), and geotechnical investigation of
in water depth of more than 20 m, it is common prac- SBM installation site (1.5 by 1.5 square km). Water
tice to install the drilling machine on a floating vessel depth was 55 m at SBM installation site and the depth
and to use an appropriate heave-compensator. Heave of boreholes into the seabed was 60 m. The use of
compensators can isolate drilling bits and sampling underwater piles was considered as an alternative to
devices from the vertical movement of the floating vessel anchor SBM.

935

Copyright 2005 Taylor & Francis Group plc, London, UK


1.2 Installation methods for drilling
equipments in sea
The installation of geotechnical drilling equipment in
sea must be normally carried out over a kind of fixed
or floating platform. The applied platforms are of
three types (Mori 1981):
Platforms supported by ground which are jack-up
barges, fixed platform or temporary scaffoldings.
Immersed platforms which are placed under water
and drilling are controlled by divers or carried out
by means of a remote control system.
Figure 1. The barge used in Lavan SBM Project.
Floating vessels which are used with a heave com-
pensator and also an appropriate mooring system
to control vertical and horizontal movements. In LAVAN Project, 4 anchors of 700 ton were
The use of floating vessels is the most common applied to control the horizontal movements of the
method of installation for drilling equipments in deep barge. Horizontal motion were measured with resol-
waters. Heave compensators are used to isolate drilling ution of 10 mm using differential positioning system
bits and sampling devices for the vessel. A heave com- (DGPS). Total horizontal limits of the platform
pensator also minimizes the pressure variations on the were 0.5 m during the operation.The drilling had to
drilling bit and the disturbance of soil under the bit. A stop when the horizontal limits were reached. Consider-
heave compensator is, in fact, a combination of some ing very soft seabed encountered, the shape and weight
springs and dashpots (Chaney 1991), which are con- of the anchors were designed in a way that horizontal
nected to drilling rod to exert a certain tension force. component of the mooring cables forces could be sup-
This causes the transfer of a part of drilling rod weight ported. The design of the anchor shape was out of the
to the deck of the vessel through an A-frame. scope of presented paper (SAHEL Consultant Engineers
1998). To install anchors, a tug boat was utilized.
However, the installed anchors and cables were in a
2 THE APPLIED SYSTEM IN LAVAN loose condition before wind, wave or current forces to
SBM PROJECT be applied on the vessel. A fairly large horizontal dis-
placement of the vessel would be required to convert
2.1 Used barge mooring cables from loose to tight condition. Such a
The local availability of equipments and environmen- horizontal movement should be minimized to prevent
tal conditions, including wave climate and seabed any damage to drilling system. Therefore the pre-
condition could substantially affect applied system in stressing of mooring cables was considered after the
geotechnical investigations. The applied method in the anchors installation. To create pre-stressing force
geotechnical investigation of Lavan SBM project was through mooring cables, an electric winch was used.
based upon a general feasibility study conducted one However the length of the cables should be continu-
year before the commencement of Lavan project for ously adjusted within tide periods. Therefore, meas-
similar conditions in the region (SAHEL Consultant urements of the cable forces were undertaken to
Engineers 1997). perform appropriate actions for the length adjustments
To implement the geotechnical investigation of the of the cables. Considering the measured forces, any
project, a barge of 12000 kN, as shown in Figure 1, was considerable barge movements were predicted and
utilized. The barge is 42.2 m long, 12.1 m wide, 3 m loosening or tightening of the cable were implemented
high; it has a draft of 0.5 m. The required geotechnical before any considerable horizontal movement. In order
equipments, power supply and personnels recre- to monitor displacement of the barge, differential
ational facilities were provided on the barge. A crane Position System (DGPS) was also used on the barge.
was also placed on the barge to perform required lift- Using the above mentioned method, it became pos-
ing for the drilling process and supporting activities. sible to control the horizontal movements of the barge;
in a way that drilling from seabed surface down to the
depth of 60 m, into the ground, came to practicability.
2.2 The control of horizontal movements
Using 4 or 6 anchors and mooring cables, with suffi-
2.3 Drilling system and control of vertical
cient length, it is possible to fix drilling floating vessels
movement
on sea in water depths of maximum 300 m (DeFruiter &
Richard 1983). For very deep waters, it is more com- Wire line boring rods were utilized for drilling. The
mon to take the advantage of dynamic positioning. external diameter of the rod is 88.9 mm and the internal

936

Copyright 2005 Taylor & Francis Group plc, London, UK


Figure 2. The side drilling platform on the top of the verti- Figure 3. Cylindrical steel tank fixed to the casing pipe to
cal casing pipe. prevent buckling.

diameter, 77.8 mm. Wire-line system proved to be Platform


successful to facilitate the drilling process. A borehole (weight=2500 kgf)
Sea Level
of 60 m deep was drilled within 48 hours.
The hired barge did not any moon pool so the drilling
machine was installed on the side of the barge, Figure 2. Steel Tank
(up-ward vertical force=3500 kgf)
When a heave compensator is used, the drilling
should be stopped even in a moderately rough sea
conditions to prevent any harm to drilling operations.
Available heave compensators could not isolate drilling Casing
rod from floating barges vertical fluctuations in a mod- (Total Weight=2500 kgf)

erately rough sea conditions. To isolate fully the drilling


machine from the vertical movement of the barge, the
drilling machine was installed on a modified vertical Seabed frame (weight=3500 kgf)
casing pipe (Diameter  300 mm). The casing was
6.0m

sited on the seabed using an special footing, designed


for very soft seabed. It would be very difficult to han-
dle large casings in offshore conditions so any diameter
larger than 300 mm was not accepted for the casing Figure 4. The schematic of steel tank and seabed frame.
because of practical handling difficulties involved.
However, such a small cross section of casing is not
adequate to support the weight of drilling equipment down through the 12 inches casing. A crane was used
and also the self-weight of the casing in a water depth to help the erection and drilling processes.
of 55 m. The casing is too long. To increase buckling
capacity of the casing, a cylindrical steel tank, shown
2.4 Sampling and in-situ tests
in Figure 3, was attached to the casing to prevent
buckling. The submerged steel tank was fixed at the Borehole drilling was carried out by continuous coring.
level of 10 m below water surface and creates an The ground materials were mostly silt and clay, with
upward vertical force on the casing. The system is high percentage of carbonate, increasing in stiffness
schematically shown in Figure 4. The buckling of cas- by depth. Thin-walled sampler went down through HQ
ing was prevented because the main length of the casing tubes and took samples, using the exerted pressure.
was in tension when the tank was fixed. This caused Seabed frame was difficult to be used for CPT prob-
that the pipes of 30 cm/300 mm with a thickness of ing because of the very soft seabed condition. Therefore,
6 mm could support vertical loads. CPTs were undertaken from the platform fixed on the
The drilling machine was installed on a platform top of the casing. To provide required reaction force
on the top of the casing but the power unit was placed for drilling and CPT, a seabed reaction footing with
on the barge. The power was transferred via the linking the area of about 4 m2 and weight of 3.5 ton, shown in
cables/hoses to the drilling motor. The drilling per- Figure 5, was built. Due to the soft condition of the
sonnel worked on the platform, as shown in Figure 2, seabed, the (300 mm) casing passed through the foot-
on the top of the casing. Drilling rods were pushed ing to a depth of 6 m below seabed.

937

Copyright 2005 Taylor & Francis Group plc, London, UK


vessels. It is also essential to continuously measure
and adjust the length of the cables by means of
electric winches within tide periods.
Using the above mentioned casing mounted method,
it became possible to control the horizontal and verti-
cal movements of the barge; in a way that drilling
from seabed surface down to the depth of 60 m came
to practicability. It is a cost effective method and very
easy to be mobilized.

ACKNOWLEDGEMENT

Figure 5. The seabed frame. The authors would like to express their appreciation
to all colleagues in SAHEL Consulting Engineers who
CPT probing was carried out adjacent to the bore- brought the described experience to reality.
holes. To avoid buckling the rod of the mechanical
CPT equipment, the rod was passed through BQ rods.
Shear-Vane and Pocket Penetrometer tests were REFERENCES
also done on the samples at site.
Chaney, R.C. 1991. Sampling and preparation of marine sedi-
ments, Foundation Engineering Handbook, Fang (ed),
3 CONCLUSION Chapman and Hall, pp. 7287.
De Ruiter, J. and Richards, A.F. 1983. Marine geotechnical
investigation, A mature technology, Proceeding of the
Implementation of the mentioned operations resulted
Conference on Geotechnical Practice In Offshore
the following practical achievements: Engineering, Wright (ed). Austin, Texas, April 2729,
Drilling bits and sampling devices can be practically ASCE, pp. 124.
isolated from the vertical fluctuation of drilling float- Fakher, A. and Pahlavan, B. 2000. Lessons learned from using
jack-up barges for geotechnical drilling, Geo Engineering
ing vessels by means of casing mounted method,
in Arid Lands, Mohammed & Hosane (eds) Al-Ain,
described in the paper. UAE, Nov. 47, A. A. Balkema, pp. 8995.
Buckling is the major criterion of structural design Mori, H. 1981. Soil sampling in site investigation for coastal
of casing when casing mounted method is used. To structures, Proceedings of the Symposium on Geotechnical
reduce the size of casing, it is practical to exert a ten- Aspects of Coastal and Offshore Structures, Bangkok,
sion forced by connecting an air cylindrical steel December 1418, 1981, Yudhbir and Balasubramaniam
tank to the casing pipe. (eds.), pp. 110.
Wire-Line System proved to be successful to facili- SAHEL Consultant Engineers 1997. Technical-Executive
tate the drilling process when the mentioned cas- requirements and budgetary for offshore geotechnical
drilling in Iranian waters in Persian Gulf, Autumn 1997,
ing mounted method is used.
Managerial report prepared by Geotechnical Investigation
The described 4-anchor system is practically suc- Department, 52 pages (In Farsi Language).
cessful to control horizontal movements of drilling SAHEL Consultant Engineers 1998. Mooring geotechnical
vessels for geotechnical borehole drilling. drilling floats, Spring 1998, Joint Report of Marine
Pre-stressing of mooring cables is essential for suc- Structures Department and Geotechnical Investigation,
cessful control of horizontal movements of floating 32 pages (In Farsi Language).

938

Copyright 2005 Taylor & Francis Group plc, London, UK


Self/barge Installing Platform, SIP II: the Calder experience

R.J. van den Heuvel & M.E. Riemers


SPT Offshore, Woerden, The Netherlands

ABSTRACT: This paper describes in detail SPT Offshores suction pile based SIP II platform concept that
was successfully used and installed for Burlington Resources (Irish Sea) Ltd in their Calder field located in the
East Irish Sea in block 110/7 of the UKCS. Calder suction foundations consist of four 9.25 m diameter cans,
penetrating up to 5.5 m into the seabed. Penetrated soils were an approximately 1.5 m thick loose silty sand layer
underlain by a very dense fine to medium sand layer. Penetration of the gravelly layer starting at 5.7 m below
seabed was to be avoided in order to prevent expected suction installation refusal. Local water depth was 29 m
plus 8 m tide.

1 INTRODUCTION

1.1 SPT Offshore, company introduction


Self and barge Installed Platforms are giving new
considerations for the development of marginal fields.
SPT Offshore from The Netherlands have developed
four (4) Self (and barge) Installing Platform concepts
named SIP I, II, III and IV based on suction piles. All Figure 1. SIP II installation: (a) transportation, (b) legs
these concepts make marginal to medium sized fields lowered, buckets installed & deck lifted, (c) as-installed.
economically viable, as the installation costs are gen-
erally less compared with using traditional heavy lift Following arrival in the field and mooring the barge
crane barges. For the installation of these self and barge to the pre-installed anchors, the seafastening is cut to
installed platforms, local equipment is used, i.e. tugs, lower the legs using pre-tensioned strand jacks. Extra
barges and other rented equipment, such as strand strand jacks are available in order to partially lift the
and/or chain jacks and suction pumps. The schedule barge to obtain more pre-load on the suction pile
for the development of the offshore field is therefore foundation during penetration. After the suction piles
flexible. SPT experienced on two projects in the North are installed the deck is jacked up using another set of
Sea that re-use of these suction pile based platforms is strand jacks to the final elevation. The legs are then
easy and cheap as the suction pile foundation can sim- clamped and welded to the deck and thereafter the
ply be pressed out of the soil without damaging or barge will be demobilised.
leaving behind the suction piles. Suction piles can be Installation of a SIP II platform is illustrated in the
used in a range of soils from soft to stiff clays, in loose Figures 1a through 1c above.
and dense sands and in layered soils.
This paper describes SPT Offshores successful
installation of the SIP II platform concept for Burlington 2 SUCTION FOUNDATION DESIGN
Resources (Irish Sea) Ltd in their Calder field located
in the East Irish Sea in block 110/7 of the UKCS. 2.1 Design criteria
Detailed geotechnical design of the suction pile foun-
dation in subject was performed by Fugro Engineers
1.2 SIP II platform, basic principle
BV, Leidschendam, The Netherlands [4]. A summary of
The SIP II concept is a barge installed platform, which the Calder soil design profile is given in Table 1 below.
is supported by 4 legs and 4 suction piles as the founda- The challenge for design of the suction pile founda-
tion. During the tow to the field, the barge supports the tion in above-presented sandy soil was soil unit weight
platform deck and the legs are in its highest position. as well as soil permeability increasing with depth ending

939

Copyright 2005 Taylor & Francis Group plc, London, UK


in a gravel layer at approx. 6 m depth. The following were essentially flush inside, i.e. neither stiffeners nor
problems were assessed during the detailed design: cross plates were required.
the loose/lightweight and low permeable character The connection of the suction pile to the leg is
of the top soil resulting in relative low critical suc- external, which is needed to avoid a clash with the
tion pressures with regard to soil plug heave and/or installation barge, but also proves to be advantageous
liquefaction as the platform overturning moment is spread out over
gravel layer that needed to be avoided, as pile a larger distance.
refusal was expected
Though this paper is mainly focused on suction 3 SUCTION INSTALLATION: ANALYSES
installation of the Calder platform foundation, for the PERFORMED
sake of completeness governing suction pile design
load cases are presented in Table 2 above. Calder suction foundations consist of four 9.25 m
From above-presented load cases it may promptly diameter cans, penetrating up to 5.5 m into the seabed.
be concluded that moment loading was mainly gov- Penetrated soils were an approximately 1.5 m thick
erning suction pile design regarding pile capacity. loose silty sand layer underlain by a very dense fine to
Detailed design resulted in suction piles with an medium sand layer. Penetration of the gravelly layer
OD of 9.25 m and 5.5 m penetration. The suction piles starting at 5.7 m below seabed was to be avoided in
order to prevent expected suction installation refusal.
Laboratory testing performed by Fugro Ltd. (1999)
Table 1. Summary soil design profile. [5] was primarily focused on determination of particle
sizing distribution (PSD-)curves, maximum/minimum
Depth density and specific gravity. Subsequent labtesting by
range
Fugro Engineers bv (2001) [4] included permeability
[m below Thickness CPT qc
Unit Description mudline] [m] [MPa] (parameter and triaxial constant head) tests performed
on the soils ranging from mudline to 5.7 m below
1 very loose 01.5 1.5 0.5 mudline. No permeability test data were available on
SAND (01 m depth) the underlying gravelly material; based on indirect
0.522 methods, i.e. Hazen [6], Kozeny/Carman [2,7] and
(11.5 m depth) Aubertin et al. [1], a permeability value k  1*10
2 m/s
3 very dense 1.55.7 4.2 2238 was selected. Design soil parameter values for suction
SAND installation are presented in Table 3 above.
3A GRAVELLY 5.78.5 2.8 45
One of the main parameters in Table 3 above is the
layer
critical hydraulic gradient ic. Assuming pure vertical
flow and zero cohesion, sands will liquefy when the
Table 2. LRFD factored design loads (50 yr. storm). effective stress becomes zero, leading to the follow-
ing expression for the critical hydraulic gradient:
Load case V [MN] H [MN] M [MNm]

LC3A 1.18 1.44 35.06


(maximum M) where   soil submerged unit weight [kN/m3]
LC2A
0.58 1.48 30.03 w  water unit weight [kN/m3]
(maximum V, tension)
LC1A 2.24 1.74 30.57
(maximum H) Installation of suction foundations commences with
self-weight penetration, i.e. penetration into the seabed

Table 3. Design soil parameters (suction installation).

Interpr. Critical
Min. Max. Specific relative Subm. hydr.
Soil density density gravity density unit Wt. gradient Permeability
Unit type [kN/m3] [kN/m3] [kN/m3] Dr []  [kN/m3] ic [] k [m/s]

1 silty 12.45 20.01 26.43 0.30 9.0 0.9 1.10


53.10
5
SAND
3 fine-med. 14.12 18.83 26.38 1.00 12.0 1.2 2.10
56.10
5
SAND
3A GRAVEL 1.10
2

940

Copyright 2005 Taylor & Francis Group plc, London, UK


with only suction can submerged weight and any
added platform weight as the driving force. Obviously,
during self-weight penetration the suction can water
escape valves need to be opened in order for the water
entrapped inside the can to flow out with increasing
penetration depth. Self-weight penetration terminates
when the driving force is in equilibrium with soil pene- Figure 2. Scenario I: (a) self-weight penetration, (b) lique-
tration resistance. At this time the water escape valves faction & boiling of upper sand, (c) liquefaction (& boiling)
are closed and suction-assisted penetration commences. of upper & underlaying sand.
In clean sand profiles the suction (under)pressure inside
the can assists in further suction can penetration, as:
1. the driving force for penetration is increased by the
differential pressure across the can topplate
2. the suction-induced waterflow into the can decreases
the soil penetration resistance. This consists of a
slight increase of the outer skirt-soil friction and
significant reduction of both inner skirt-soil fric-
Figure 3. Scenario II: (a) self-weight penetration, (b) lique-
tion and skirt tip penetration resistance.
faction of upper sand and water bell formation, (c) upper
As mentioned before, one of the main challenges in soil pumped out through pump, liquefaction (& boiling) of
suction can design for the Calder Field was soil per- underlaying sand.
meability as well as soil unit weight increasing with
depth. If during suction-assisted penetration upward Scenario II
water flow gradients are sufficiently high in the internal
soil plug, the soil will liquefy. The top 1.5 m of soil, (1) and (2) as Scenario I above
both looser/lighter and less permeable than the under- (3) formation of a water bell at the interface between
lying soil, will liquefy first. the upper 1.5 m of loose sand and the underlying
During detailed geotechnical design and based on 4 m of dense sand. The water bell formation rate
soil parameters listed in Tables 1 and 3, it was envis- is equal to the difference in unit flow rates (q2
q1)
aged that suction-assisted penetration would only be per layer:
possible with pressures and associated flow rates caus- q2
q1  k2 i2
k1 ic,1
ing liquefaction of the upper 1.5 m of the internal soil
where ic,1  (critical) hydraulic gradient in upper
plug. The following scenarios for internal soil plug
layer
behaviour were assumed:
i2  hydraulic gradient in layer 2
Scenario I (4) the upper soil layer is pumped out through the
suction pump
(1) self weight penetration. (5) liquefaction (eventually followed by boiling) of
(2) suction-assisted penetration until liquefaction occurs the underlying 4 m of dense sand will occur. This
within the upper 1.5 m of soil, leading to the follow- is dependent on e.g. pumping rate and it occurs
ing critical hydraulic gradient for the upper layer: when the critical hydraulic gradient for this dense
ic,1  1/w  0.9 (see Table 3). sand layer (ic,2) is first reached.
(3) boiling of the upper soil occurs via near vertical Above-mentioned stages in Scenario II are illustrated
erosion channels from the base to the top surface. in Figure 3 above.
Water flow will be either turbulent or laminar, Scenario III
dependent on the channel flow velocity. However
the flow rate will essentially be equal to that in (1), (2) and (3) as Scenario II above
the underlying 4 m of dense sand, where laminar (4) the upper soil layer meets the can topplate.
flow is still occurring. Surcharge increases the effective stresses and
(4) liquefaction (eventually followed by boiling) of increases water flow through the upper soil layer.
the underlying 4 m of dense sand will occur. This At relatively low pump flow rates the upper layer
is dependent on e.g. pumping rate and it occurs will remain stationary (relative to mudline); alter-
when the critical hydraulic gradient for this soil natively, at higher pump flow rates, the water
layer (ic,2) is first reached. bell closes again
(5) liquefaction (eventually followed by boiling) of
Above-mentioned stages in Scenario I are illustrated the underlying 4 m of dense sand will occur. This
in Figure 2 above. is dependent on e.g. pumping rate and it occurs

941

Copyright 2005 Taylor & Francis Group plc, London, UK


p [kPa] (3.6MN preload) p [kPa] (3.6MN preload)
0 50 100 150 200 250 0 50 100 150 200 250
0 0
1 1

Depth [m]

Depth [m]
2 2
3 3
Figure 4. Scenario III: (a) self-weight penetration, (b) lique- 4 4
faction of upper sand and water bell formation, (c) upper
soil meets topplate, liquefaction (& boiling) of underlaying 5 5
sand. 6 6

p [kPa] (8.3MN preload)


when the critical hydraulic gradient for this soil 0 50 100 150 200 250
layer (ic,2) is first reached. 0
1
Above-mentioned stages in Scenario III are illus- qc - method (most probable)

Depth [m]
trated in Figure 4 above. 2 qc - method (highest expected)
Allowable (design)
Constant head upwards flow laboratory permeame- 3
Allowable (maximum)
ter tests [4] suggested that Scenario I was more likely 4
to occur than the two remaining scenarios. The final 5
phase of Scenario I implies liquefaction of the whole 6
soil plug, which was not recommended. Though lique- REQUIRED SUCTION VERSUS PENETRATION DEPTH
faction is a temporary situation, which would stop
when hydraulic gradients became less than critical, Figure 5. Allowable versus SPT expected suction installa-
soil will only partially recompress and thus regain tion pressures.
strength once liquefaction is halted. This was accept-
able for the upper 1.5 m of loose sand, however a
strength reduction of the underlying dense sand would friction breakers, when applicable, etc. Expected suc-
decrease in-place suction can capacity under com- tion installation pressures can then be found by dividing
bined (VHM) storm loading. the pile penetration resistance minus (submerged) suc-
During detailed geotechnical design installation tion pile weight and any added weight by the gross
analyses were made of Scenario I phases (1) through suction pile area.
(3) where only the upper 1.5 m of soil would liquefy, Figure 5 shows SPT expected suction installation
the underlying 4 m of dense sand would experience pressures versus allowable installation pressures, of
hydraulic gradients less than critical and effective which the maximum value was considered a true
stresses within the soil plug were reduced sufficiently upper bound; above this boundary the internal soil plug
to enable suction installation. After platform installa- was supposed to liquefy completely. Three preload
tion the upper 1.5 m of soil will return to almost the values were considered, i.e. 3.6 MN/can, 6.6 MN/can
original (loose) relative density condition; any (small) and 8.3 MN/can.
reduction in density will hardly impair suction can in- Following design it was concluded that in general,
place capacity as most of the capacity originates from installation seemed marginally feasible; if econom-
the underlying dense sand (and gravel). ically possible, preload should be maximised during
Predictions for expected suction installation pres- suction installation.
sures are preferably based on CPT-data following the
DNV-method [3]: 4 ACTUAL SUCTION PILE INSTALLATION
Suction pile penetration resistance is a summation
Suction pile installation was performed using the flow-
of internal and external pile-soil friction and, for
rate controlled SPT suction pumps (instead of pressure
installation pile tip end bearing. Pile-soil friction
controlled), which eliminated the danger of uncon-
tip end bearing are calculated using CPT-data (see
trolled soil liquefaction, should this occur. In addit-
Table 1) in conjunction with the DNV skirt pene-
ion plug heave was constantly being monitored using
tration factors [3]. Two sets of skirt penetration fac-
echosounders in the suction cans.
tors in sand are provided by DNV, giving most
Figure 6 shows actual installation pressures for all
probable and highest expected values for pene-
4 cans versus SPT expected suction installation pres-
tration resistance.
sures. It should be noted that maximum preload during
Based on SPT installation experience above analyses installation was 6.6 MN/can. Final penetration depths
are tuned accounting for e.g. soil sensitivity (St) in for all 4 cans ranged from 5.55 m to 5.75 m, all exceed-
clay, suction-induced waterflow in sand, presence of ing minimum required penetration depth of 5.5 m.

942

Copyright 2005 Taylor & Francis Group plc, London, UK


p [kPa] (CAN A1) p [kPa] (CAN A3)
0 50 100 150 200 250 300 350 0 50 100 150 200 250 300 350
0 0

1 1

2 2

Depth [m]
Depth [m]

3 3

4 4

0
5 5

6 6

p [kPa] (CAN B1) p [kPa] (CAN B3)


0 50 100 150 200 250 300 350 0 50 100 150 200 250 300 350
0 0

1 1

2 2
Depth [m]

Depth [m]

3 3

4 4

5 5

6 6

Actual Installation Data


qc - method (best estimate)
qc - method (highest expected)

ACTUAL SUCTION VERSUS PENETRATION DEPTH

Figure 6. Actual installation data.

Based on carefully monitoring the suction installa- 2. Carman, P.C. 1939. Permeability of saturated sands,
tion process, e.g. pump flow rates & pump electrical soils and clays. Journal of Agricultural Science, 29:
current gauges, it is concluded that no soil plug lique- 263273.
faction and/or excessive soil heave occurred during 3. Det Norske Veritas. Foundations, Classification Notes
No. 30.4, February 1992.
installation. Also the suction pump flow rates, 4. Fugro Engineers BV. 2002. Geotechnical Engineering
observed during installation, were approximately 1/3 Suction Can Foundations Calder Platform, Irish Sea. SPT
of anticipated in design. One might conclude that the Doc. No. 93020/CALC/301 Rev.C1, dated 24-Jun-02.
soil plug liquefaction scenarios investigated were too 5. Fugro Ltd. 1999. Laboratory and In-situ Testing Report
conservative for this project. Geotechnical Site Investigation, Calder Field, Block
110/7, UK Sector, Irish Sea, Report No. 93309-3 to
Burlington Resources (Irish Sea) Limited, Jul-1999.
REFERENCES 6. Hazen, A. 1892. Some physical properties of sand and
gravel, with special reference to their use in filtration, in
1. Aubertin, M., Busiere B. and Chapuis, R.P. 1996. 24th Annual Report to the State Board of Health,
Hydraulic conductivity of homogenized tailings from Commonwealth of Massachusetts, Boston, pp. 539556.
hard rock mines. Canadian Geotechnical Journal, 33: 7. Kozeny, J. 1927. Uber Kapillare Leitung des Wassers in
470482. Boden, Ber. Wien Akademie, 136a271.

943

Copyright 2005 Taylor & Francis Group plc, London, UK


Assessment of sand quality using seismic techniques at
Fisherman Islands, Brisbane

R.J. Whiteley
Coffey Geosciences Pty Ltd, Sydney, NSW, Australia

J. Ameratunga
Coffey Geosciences Pty Ltd, Brisbane, Queensland, Australia

P.J. Boyle
Port of Brisbane Corporation, Brisbane, Queensland, Australia

ABSTRACT: The Future Port Expansion (FPE) Project at Fisherman Islands involves the reclamation of 230 ha
of submerged land. Stage 1 of the project required the design and construction of the perimeter FPE Seawall of
4.5 km in length. The most economical option for construction of this seawall was to use sand in its construc-
tion from within the reclamation if sufficient was available. As marine sediments in the near-shore environment
can be highly variable, Underwater Seismic Refraction (USR) was used to measure the seismic velocity in the
shallow marine sediments. The results were used to assist in locating boreholes and vibracores for direct testing
and to allow correlation for determination of sand quality. The study enabled the assessment of the quality and
distribution of sand over the FPE site to be made. This provided the necessary information for informed com-
mercial decisions to be made on the FPE seawall project.

1 INTRODUCTION During the Target Cost Estimate phase of this Alli-


ance project, several design options for the seawall and
The Port of Brisbane is Queenslands main port and one reclamation were considered while benefits and risks
of the fastest growing ports in Australia. It is located at to design and construction were assessed. Design
the mouth of the Brisbane River at Fisherman Islands options ranged from sand, rock, geosynthetics or a com-
(Fig. 1). bination of these materials. As the soil profile at the
The Port and surrounding areas have been progres- site consists of very soft to soft clay to depths as great
sively developed during the past 25 years to cater for the as 30 m, the most economical option was considered
rapidly increasing trade activities in Brisbane and South to be construction of the seawall bund using sand.
East Queensland. It is anticipated that this trend will However, this option was economically feasible
continue for the next 25 years and beyond. In order to only if sufficient clean sand was locally available. It
ensure that sufficient Port capacity is available to cater was therefore decided to investigate available sand
for this increasing demand, the Future Port Expansion resources from within the FPE area to be reclaimed,
(FPE) Project has been developed. bounded by the perimeter seawall footprint. Drilling
The first stage consists of the FPE Seawall Project, information indicated the presence and thickness of
which involves the design and construction of a sea- sand, but its cleanliness (i.e. the sand quality) and extent
wall to the east of the existing reclaimed areas at the was unknown.
Port. Construction of the perimeter seawall is the first As this area is relatively large and time constraints
step in allowing Port of Brisbane Corporation (PBC) existed, a marine geophysical method called con-
to reclaim an additional 230 ha of submerged land for tinuous underwater seismic refraction (USR), was
future expansion. The rock and sand seawall extends employed to supplement conventional direct testing
from the existing reclamation for 1.8 km into Moreton methods at a limited number of locations. The object-
Bay and forms a horseshoe shape, reconnecting with ive of this exercise was to guide rational assessment
the mainland. The project commenced in late 2002 and of sand quality, based on the thickness information
practical completion was achieved in December 2004. obtained from the drilling investigation, and extend

945

Copyright 2005 Taylor & Francis Group plc, London, UK


Figure 1. Fisherman Islands locality plan showing proposed seawall expansion area.

to assist commercial decisions on the FPE Seawall range of sea floor materials (from Hamilton 1980)
Project. plotted against porosity. The correlation coefficient
(R2) of 0.87 is sufficiently high to demonstrate that in
situ seismic velocity measurements can assist in the
2 USR STUDY classification of the marine materials and in the
assessment of sand quality for the purposes of foun-
Marine sediments in the near-shore environment such dation or embankment design.
as Fisherman Islands can be highly variable, ranging From Figure 2,
from very soft mud to coarse sands with shells. These
can be difficult to classify and quantify for engineer- (2)
ing design purposes with a limited number of widely
spaced boreholes. (3)
The seismic (P-wave or sound) velocity (V) in
these sediments can range from the water velocity USR methods can be used to measure the seismic
(Vw)  1500 m/s for the soft mud to 1900 m/s in the velocity in shallow marine sediments. The continuous
very coarse sands (Whiteley 2002). USR method was applied to determine seismic veloci-
Extensive measurements of seismic velocity in ties from the upper 5 m of sediments. These results
marine sediments at many locations in the world have were used to assist in location of further boreholes
demonstrated a good correlation between seismic and vibracores to allow correlation and direct deter-
velocity (V) and porosity (n) or void ratio (e): mination of sand quality.
(1)
3 EQUIPMENT AND FIELD PROCEDURES
where V/Vw is the seismic velocity ratio.
Figure 2 shows a second degree polynomial fit to The continuous USR method is equivalent to the
measured values of velocity ratio, obtained in a wide single-ended land seismic refraction method (Whiteley

946

Copyright 2005 Taylor & Francis Group plc, London, UK


Figure 2. Measured values of seismic velocity/water velocity in marine sediments plotted against porosity.

C
MARINE REFRACTION USR

WATER
X X X X X X

X SEISMIC SOURCE LOCATIONS

Figure 3. Sketch illustrating USR field procedures.

1994). It uses a bottom towed, non-explosive seismic vessels position was recorded using a differential GPS
source and hydrophone detector array. This involves unit, which was used to locate the mid point of the
lowering a purpose-built sled housing, an air gun source hydrophone array.
and the hydrophone array to the sea floor, in this case, The seismic data was recorded digitally on an
in water depths from about 1 to 5 m. The airgun is engineering seismograph. In general, the seismic data
powered by gas bottles housed on the vessel. The was of good quality and adequate for analysis and
hydrophone array contains 12 seismic detectors at 2 m interpretation.
intervals. This was towed at a fixed distance of 2.6 m
from the source so that the sub-bottom depth of inves-
tigation is at least 5 m (see Figure 3). 4 RESULTS AND INTERPRETATION
This system was deployed from the PBCs vessel,
the Barku, operating at approximately 2 knots. Once the The USR data was analysed using procedures that are in
overboard equipment was deployed and towing straight accordance with accepted practice. Initially, the digital
along the sea floor on the selected profiles, the air gun refraction data is examined and first arrivals are
was triggered and seismic data was recorded. This pro- picked interactively and timed using REFRACT soft-
cedure was repeated at regular intervals along prede- ware (Leung et al 1997). Only clearly observed first
termined survey lines (typically at 20 seconds or arrivals are timed. The strong arriving water wave is
20 to 30 m intervals depending on vessel speed). The identified on selected records and a water velocity is

947

Copyright 2005 Taylor & Francis Group plc, London, UK


determined by plotting this arrival against distance. sub-bottom refractors i.e. velocity/density increases
First-arrival travel-times are then plotted against dis- with depth were observed. Initial drilling within the area
tance along the detector array. Straight-line segments indicated that the general geology within the Fisher-
are fitted to the travel-time data to identify sub-bottom man Islands survey area is sands over clays and muds.
refractors. Apparent seismic velocities are computed This could also create a situation where the velocity
for these line segments and the location of any break- decreases rather than increases with depth (i.e. veloc-
in slope-point is identified. This represents the critical ity reversal occurs). In such a situation sand thick-
distance point. Assuming a horizontally layered model nesses cannot be determined from the refraction data.
over the length of the array, the apparent velocity seg- At the Seawall Project site no velocity inversion (e.g.
ments are taken as true velocities and the depth to any a stepping in the first arrival travel-time curve) was
interface is computed from the standard critical dis- observed in the seismic refraction data. As only a sin-
tance formula. A horizontally layered model is consid- gle refractor was observed, the refraction data were
ered appropriate for the Moreton Bay because sea taken to represent a composite layer whose velocity
transgressions and regressions over the past 120,000 varied laterally according to the amount of sand pre-
years have created a shallow sediment section (mainly sent. Sand thicknesses were obtained directly by con-
muds, sands with more restricted gravels and shells) touring the information obtained from boreholes and
that is grossly horizontally layered with laterally grad- vibracores.
ational boundaries. This is certainly the case in respect The seismic velocities determined from the seismic
of the seismic velocity distribution. data were taken as an average seismic velocity over
Normally only one straight-line segment was fitted the upper 5 m of sediment. These seismic velocities
to the seismic first-arrival data from this site. This ranged from about 1480 to 1640 m/s, i.e. from soft muds
represents the direct wave through the sub-bottom to shelly sands. Repeat measurements at some locat-
sediments and the slope of this line represents the ions demonstrated that these velocities were accurate,
average seismic velocity to the depth of investigation on average, to within a few percent.
on the seismic system (approx. 5 m). Depth determi- Figure 4 shows locations of the USR profiles and a
nations from the seismic data were not possible as no contour map of these average seismic velocities on an

Figure 4. Average seismic velocity contour map.

948

Copyright 2005 Taylor & Francis Group plc, London, UK


AMG survey grid. Using these velocities and the water where T is the sand thickness derived from contours of
velocity in equations 2 and 3 allowed void ratios of the borehole data, V is the measured seismic velocity
the shallow sub-bottom sediments including suspended and Vw is the water velocity (1480 m/s).
material at the seafloor to be computed. These ranged This index incorporates all available data and allows
of about 1.0 to 2.6 and averaged about 1.8. the materials to be classified as poor, intermediate and
Recognising that average velocity can be expected good according to Table 1.
to increase with both increasing sand content and sand Figure 5 shows a contour plot of SQI on an AMG
layer thicknesses, allowed the seismic and borehole data survey grid. The thickest and best quality sand is
to be directly combined to produce a Sand Quality located in the centre and western sides of the FPE area.
Index (SQI) that represents the distribution of thin/ The sand significantly decreases in quality to the north
poor quality and thick/good quality sand. along the proposed seawall alignment and to the south
The SQI is computed according to: near the northeast and southeast corners of Fisherman
Islands.

(4)
5 CONCLUSIONS

Underwater seismic refraction method was used to


Table 1. Sand Quality Index and Classification. determine seismic velocities in the shallow marine
sediments in the upper 5 m below the seabed within
Range of the FPE Project site. The results were used to assist in
SQI values Classification Interpretation locating boreholes and vibracores for direct testing
and to allow for correlation and determination of sand
00.4 Poor Limited sand present quality. Relationships between seismic velocity, poros-
0.40.8 Intermediate Some sand present, average ity and void ratio enabled the assessment of the qual-
condition at site
ity and distribution of sand over the FPE area. This
0.81.3 Good Thick/good quality sand
present information allowed for commercial decisions to be
made on the FPE Seawall project.

Figure 5. Sand Quality Index (SQI) contour map.

949

Copyright 2005 Taylor & Francis Group plc, London, UK


ACKNOWLEDGEMENTS Leung, T.M., Win M.A., Walker, C.S. and Whiteley, R.J.
1997. A flexible algorithm for seismic refraction inter-
The design and construction for the FPE Seawall pretation using program REFRACT, In Eddleston et al.
project was carried out as an Alliance between the (eds), Modern Geophysics in Engineering Geology,
Geological Society Engineering Geology Special Publi-
client (PBC) and a team consisting of Coffey Geo- cation. No. 12: 399407.
sciences (Geotechnical Consultant), Leighton Con- Whiteley, R.J. 1994. Seismic refraction testing A tutorial.
tractors (Contractor), Parsons Brinckerhoff Australia In R.C. Woods (ed.), Geophysical characterization of
(Civil Consultant) and WBM Oceanics (Hydraulic sites, ISSMFE, Tech. Comm. No. 10. New Delhi, India.
Consultant). Whiteley, R.J. 2002. Integrating geophysical and geotech-
nical technologies for improved site assessment of ports
and harbours. Proc. PIANC 2002, 30th International
Navigation Congress, Sydney, Sept. 2002
REFERENCES

Hamilton, E.L. 1980. Geoacoustic modelling of the sea floor.


Jnl. Acoust. Soc. Am. 68: 13131340.

950

Copyright 2005 Taylor & Francis Group plc, London, UK


Site characterization of Bootlegger Cove clay for Port of Anchorage

Paul W. Mayne
Georgia Institute of Technology, Atlanta, GA USA

Richard A. Pearce
Terracon Incorporated, Charlotte, NC USA

ABSTRACT: A geotechnical site investigation for the Port of Anchorage Expansion involved use of the
SeaCore jack-up platform for conducting soil test borings, cone soundings, vane shear, and downhole shear
wave velocities to characterize the subsurface profiles. As the daily tides vary some 10 m in elevation, the CPTu
soundings proved to better suit the field testing timetable in the expedient collection of continuous data. With
the mudline located about 14 m deep, the primary soil formation consists of two lower facies of Bootlegger
Cove Formation clay in the upper 50 m, with occasional shallower layers of dense sands. Series of consolidation
tests showed that the BCF is overconsolidated, with an approximate prestress value of 400 kPa. Similar profiles
of OCR are derived from supporting piezocone and vane data. Lab strength tests included series of triaxial com-
pression and simple shear tests that agree well within a critical-state soil mechanics framework. Interpretations
of the field and laboratory data are shown to be internally consistent and compatible.

1 INTRODUCTION is rather short (May to October) because of sea ice


and freezing temperatures.
The Port of Anchorage (POA) is expanding its exist-
ing dock facilities in order to accommodate a greater
number and size of ships for commerce and military 2 IN-SITU TESTING
purposes. The new wharf will extend out 80 to 120 m
seaward with a total length of 2500 m along Knik A jack up platform situated on four legs with spuds
Arm that is located at the northeast edge of Cook (SeaCore) was used for drilling and in-situ testing
Inlet. The area is in a complex geologic setting with operations. A total of 20 soil test borings, 38 cone
sand and clay sediments of glacial, fluvial, lacustrine, penetration tests (CPT), three seismic velocity profiles,
and colluvial origin overlying dense gravelly sands and one series of vane shear tests (VST) were per-
and bedrock. formed for the initial phase of site characterization.
Of particular interest here at POA is the presence The borings were advanced to depths of 31 to 65 m
of the Bootlegger Cove Formation (BCF) that relates to deep and utilized a 134 mm diameter Geobarrel system
the seismic-induced landsliding at nearby Government with a modified Pitcher coring attachment to obtain
Hill and Turnagain Heights during the 1964 Alaskan undisturbed samples for the laboratory program.
earthquake. The BCF is a complex glacio-estuarine Generally, clay of the BCF was encountered at the
silty clay formed about 13000 years ago and com- mudline depth of 14 meters, however at certain loca-
posed of eight separate facies. The 1964 problems are tions, a 5- to 15-m thick layer of sand capped above
related to facies 3 that are at higher elevations approx- the clay. At select locations, standard penetration test-
imately 5 to 15 m above sealevel. For the new POA ing (SPT) with energy measurements were taken. The
expansion, the lowest two units of the BCF are predominance of in-situ data were collected by cone
encountered, generally beneath elevation
14 meters penetration tests (CPT) using a shoulder filter ele-
MSL. As the daily tidal variation here is rather extreme, ment (piezocone, CPTu) to obtain continuous profiles
the change is about 10 m between low and high tide, of tip stress (qt), sleeve friction (fs), and porewater
thus complicating the geotechnical site investigation pressures (u2) with depth. At 11 soundings, dissipation
program, as well as routine construction activities in tests were performed at select depths to record the
near shore environment. Moreover, due to the harsh time rate of decay of the induced porewater pressures
northern climate of Anchorage, the construction season with time. The time for 50% dissipation (t50) is a

951

Copyright 2005 Taylor & Francis Group plc, London, UK


Tip Stress, qt (MPa) Friction, fs (kPa) Porewater, ub (MPa) Dissipation,t50 (min) Shear Wave, Vs (m/s)
0 1 2 3 4 0 50 100 150 0 1 2 3 1 10 100 0 100 200 300 400 500
10
Mudline
15

20
Depth (meters)

25

30

35

40

45

Figure 1. Composite seismic piezocone test with dissipations (SCPT) in Bootlegger Cove clay at Port of Anchorage.

characteristic measure of the coefficient of consolida- 500


Shear Stress, q = 0.5(1 - 3) kPa

tion (cvh) and hydraulic conductivity (k) of the geo-


material. Downhole geophysical surveys were made Bootlegger Cove Clay (Lower Facies):
400 Effective Strength Envelope
during three CPTs to obtain the shear wave velocity (Overconsolidated Range)
(Vs) profiles.
Results from a representative piezocone sounding 300 ' = 27
(Number C28) taken to 42 m depth are presented in c' = 20 kPa
Figure 1. A 3-m capping of sand overlies the BCF 200
clays with sandy layers between 35 and 38 m. Also
shown on the profile are the superimposed results of 100
dissipation tests from nearby soundings (C16 and
C22) and shear wave velocity profile (C21). These
0
data form a composite seismic piezocone with dissi- 0 100 200 300 400 500 600
pation phases (designated SCPT) that provides up to Effective Stress, p' = 0.5(1'+ 3') kPa
five independent readings on soil behavior: qt, fs, ub,
t50, and Vs. Figure 2. Summary of CIUC triaxials on BCF clay.

3 LABORATORY TESTING
rate-of-strain devices. Results from 7 tests showed that
Laboratory index testing was conducted onboard on the effective preconsolidation stress (p) increased
the recovered tube samples that were later subjected approximately linearly from around 450 kPa at 10 m
to one-dimensional consolidation, triaxial, and sim- below mudline to about 700 kPa at a depth of 32 meters.
ple shear testing (Pearce & Hale 2004). As this line is parallel with the current effective over-
burden stress (vo ), it corresponds to an equivalent
3.1 Index parameters prestress (p  pvo   400 kPa), assuming simple
mechanical erosion causing the overconsolidation.
Basic mean index parameters on the BCF clay
include: natural water content wn  20 to 31%, liq-
uid limit LL  45%, and plasticity index PI  24%.
3.3 Triaxial testing
Nineteen triaxial specimens were subjected to
3.2 Consolidation testing
isotropically-consolidated type triaxial compression
One-dimensional consolidation testing was performed with porewater pressure measurements (CIUC). Stress
using both incremental load oedometer and constant paths are summarized in Figure 2.

952

Copyright 2005 Taylor & Francis Group plc, London, UK


Overconsolidation Ratio, OCR 10
0 1 2 3 4 5 6 7 8 9 10
0 Critical State Soil Mechanics
(Modified Cam Clay)
Depth Below Mudline (m)

' = 27

Strength Ratio, su/vo'


10 = 0.75

20
1

30 CIUC Data
Prestress = 400 kPa
DSS Data
Triaxial
40 MCC Pred CIUC
Oedometer
MCC Pred DSS

50 0.1
1 10 100
Figure 3. OCR profiles from triaxial, oedometer, and Overconsolidation Ratio, OCR
prestress.
Figure 4. Summary of undrained strength data for BCF clay.

The CIUC triaxials were valuable as they help


define both the effective stress strength envelope (i.e. employed to represent these relationships (Kulhawy &
c and ), and total stress strength (i.e. undrained Mayne 1990):
shear strength suCIUC), as well as be utilized to assess
the in-situ overconsolidation ratio (OCR). The proce- (2)
dure is based on an inverted Cam-clay approach
(Mayne 1988) whereby:

(3)
(1)

As evidenced by Figure 3, the Modified Cam-clay


where Mc  6 sin (3-sin ),   effective stress (MCC) predictions are quite good for   27 and
friction angle, su  measured undrained shear strength, ,  0.75 in comparison with the measured results
and ,  1
Cs/Cc. From the consolidation tests, the over the testing range: 1 OCRs 20.
mean value of virgin compression index Cc  0.242,
swelling index Cs  0.060, giving ,  0.75, which
is reasonable for low to medium sensitive clays 4 FIELD DATA INTERPRETATION
(Kulhawy & Mayne 1990). An alternate procedure is
to plot log of the normalized undrained strength vs. log 4.1 Preconsolidation stress profiles
of reciprocal of the effective confining stress and the
The results from the in-situ field data are useful for
slope equals , (Mayne 1980), here giving ,  0.71.
the interpretation of soil engineering parameters as
The CIUC results are shown in Figure 3 and indicate
vertical & lateral variability can be evaluated. Also,
OCRs decrease from 8 at the mudline to about 2 at 30 m
continuous readings are recorded with depth and the
depth in the BCF. They are in good agreement with the
results can be ascertained quickly while still onsite. In
reference oedometer values and the calculated profile
this regard, it is of interest to evaluate the profile of pre-
based on simple unloading of p  400 kPa.
consolidation stress from the in-situ test measurements.
For the piezocone results, first order expressions
3.4 Direct simple shear testing for the effective preconsolidation stress are available
from both the net cone resistance and excess porewa-
A series of five specimens was subjected to direct ter pressure readings (Kulhawy & Mayne 1990):
simple shear testing (DSS) and prepared using
SHANSEP procedures (Ladd 1991). Both the triaxial (4)
series and DSS data showed the expected trend of
a power law relation between su/vo  and OCR, as
shown in Figure 4. Critical state soil mechanics can be (5)

953

Copyright 2005 Taylor & Francis Group plc, London, UK


Preconsolidation Stress, p' (kPa) Preconsolidation Stress, p' (kPa)
0 100 200 300 400 500 600 700 800 0 100 200 300 400 500 600 700 800
10 10

Depth Below MLLW (meters)


VST Relation
Depth Below MLLW (meters)

15 Vs Trend
15
Oedometer
20 svo'
20
25
25
30
30 35
0.33(qtnet)
35 0.53(u2-uo) 40
Oedometer
45
40 svo'
Figure 6. Preconsolidation estimates from VST and DHT.
45

Figure 5. Preconsolidation stresses from piezocone C28. extends back into the fill to derive the soil frictional
resistance, similar to earth reinforcing tie strips or
Using the data from Figure 1 (CPTU C28), the derived grids that are normally laid horizontally.
p profiles appear in good agreement with the refer- Long-term strengths for the overconsolidated clay
ence oedometer values as presented in Figure 5. Of are well represented by c  20 kPa and   27
course, the occasional drops in the porewater-related (Fig. 2). With the open cell gravity structure as high
values reflect sand lenses or stringers within the sound- as 25 m, portions of the clay can become normally-
ing and should be filtered from the processed results. consolidated, with corresponding parameters:
An evaluation of p is available from the vane shear c  0 kPa and   27.
test data and plasticity index (Mayne & Mitchell 1988): For short-term loading during construction and/or
seismic design cases, the undrained shear strength
(6) characteristics of BCF clay may be applicable. A full
hierarchy of strength modes can be considered within
which shows reasonable agreement with the oedome- a critical-state framework (Kulhawy & Mayne 1990)
ter values in Figure 5. that can account for variations in OCR, strain rate,
A final first-order estimate can be made from the anisotropy, and loading directions (plane strain vs. tri-
shear wave measurements (Mayne et al. 1998): axial, compression and extension).
Derived profiles of undrained shear strength from the
(7) lab and in-situ tests are presented in Figure 6. Here, the
CPTu data provided the corresponding OCR profiles in
which is units dependent for p (kPa) and Vs input in the Bootlegger clay and eqns (2) and (3) used to obtain
meters/sec. It is noted that the field report (Volume 2 su values for CIUC and DSS modes. These compare
of Pearce & Hale 2004) noted some difficulties in the well with the lab series based on the oedometric OCRs
downhole testing program with source generation, fil- to drive the recompression-type SHANSEP series of
tering, and postprocessing of Vs profiles. The derived triaxial and simple shear strengths (e.g. DeGroot 2001).
preconsolidation stresses here compare on the lower The vane data shown here in Figure 6 are raw suv values
bound of the trends, as evidenced by Figure 5. and not corrected for strain rate and anisotropy effects
(e.g. Chandler 1988), however would not be consid-
ered crucial given the low plasticity characteristics of
4.2 Undrained strength profiles
BCF. Moreover, it is not common practice to correct
As two primary options are pursued for the POA dock vane suv in offshore applications (Randolph 2004).
construction (piled wharf or open cell sheet pile Alternative to the aforementioned, an empirical
structure), interpreted shear strength profiles within assessment of undrained strength modes can be made
the BCF clay strata were needed for both designs. The (e.g. Ladd 1991, DeGroot 2001).
current wharf facility is an open deck on driven 0.6-m
diameter open pipe piles. In the Cook Inlet region
near Anchorage, a reinforced earth fill design with 5 CONCLUSIONS
driven sheeting as the fill template has been success-
fully used for several private, commercial, and indus- A nearshore geotechnical site investigation of the
trial waterfronts. The vertical face of the sheeting Bootlegger Cove Formation clay has been made for a

954

Copyright 2005 Taylor & Francis Group plc, London, UK


Undrained Shear Strength, su (kPa) within a critical-state framework that uses stress his-
0 20 40 60 80 100 120 140 160 180 200 tory as a common reference basis, which in turn, gen-
10 erates undrained shear strength profiles.
Mudline Lab DSS Data
15 Lab Triaxial Data
Soft Vane Shear
Mu
dline ACKNOWLEDGMENTS
Depth (meters)

20 Silt TC from CPTu


DSS from CPTu
25 The authors extend appreciation to Bill Humphries at
Upper
Bootlegger ICRC, Diana Brake, William Sheffield, and Cheryl at
30
Cove Clay POA, Dave Woeller at ConeTec, and Barney Hale at
35 Titan Atlantic for their efforts and contributions to
Lower
Bootlegger
this study.
40
Cove Clay
45
REFERENCES
Figure 7. Undrained shear strength profiles in BCF clay.
Chandler, R.J. 1988. The in-situ measurement of the
undrained shear strength of clays using the field vane.
Vane Shear Strength Testing in Soils, STP 1014, ASTM,
West Conshohocken, PA: 1344.
DeGroot, D.J. 2001. Laboratory measurement and interpre-
tation of soft clay mechanical behavior. Soil Behavior
and Soft Ground Construction, GSP 119, ASCE: 167200.
Kulhawy, F.H. and Mayne, P.W. 1990. Manual for Estimating
Soil Properties for Foundation Design, Report EL-6800,
Electric Power Res. Inst., Palo Alto, CA, 306 p.
Ladd, C.C. 1991. Stability evaluation during staged con-
struction. Journal of Geot. Engrg. 117 (4): 540615.
Mayne, P.W. 1980. Cam-clay predictions of undrained
strength. Journal of Geot. Engrg 106 (11): 12191242.
Mayne, P.W. 1988. Determining OCR in clays from labora-
tory strength. Journal of Geot. Engrg 114 (1): 7692.
Mayne, P.W. and Mitchell, J.K. 1988. Profiling OCR in clays
Figure 8. The SeaCore field testing operations at Anchorage. by field vane. Canadian Geot. J. 25 (1): 150157.
Mayne, P.W., Robertson, P.K. and Lunne, T. 1998. Clay
stress history evaluated from SCPTU. Geotechnical Site
large dock expansion at the Port of Anchorage, includ- Characterization (2), Balkema, Rotterdam: 11131118.
ing in-situ and lab testing. Field tests was performed Pearce, R.A. and Hale, B.C. 2004. Preliminary Report for
Port of Anchorage Marine Geotechnical Services.
using a jackup platform (Figure 7) and consisted of
Terracon Report Proj. No. 70030532 to Koniag Services,
piezocone testing, soil borings with sampling and Anchorage, Alaska, Two Volumes.
SPTs, vane shear, and downhole shear wave velocity Randolph, M.F. 2004. Characterization of soft sediments for
measurements. Lab testing included index, consolida- offshore applications. Geotechnical & Geophysical Site
tion, triaxial, and simple shear. The laboratory and Characterization, Vol. 1, (Proc. ISC-2, Porto), Millpress,
in-situ field results are found compatible and consistent Rotterdam: 209232.

955

Copyright 2005 Taylor & Francis Group plc, London, UK


The geotechnical diving bell equipment used in Brazil to perform
nearshore and offshore geotechnical investigations

Francis Bogossian, Arnaldo Muxfeldt & Angle B. Dutra


Geomecnica S.A., Rio de Janeiro, Brazil

ABSTRACT: The expansion of the Brazilian pipeline network has resulted in new technical and environ-
mental challenges for shore approach studies. The geotechnical diving bell, due to its flexibility of use and cost
effectiveness is well suited to perform geotechnical investigations for such studies, even in the surf and breaker
zones and unsheltered areas where conventional methods such as jack-up platforms are usually ineffective.
Conceived and developed in Brazil to perform offshore geotechnical investigations in water depths up to 55 m,
the geotechnical diving bell is carried to the location and positioned by a vessel-of-opportunity, which also func-
tions as a support boat carrying all necessary equipment and personnel. Divers work inside the dry chamber,
previously filled with gas, fitted with all the necessary equipment for the execution of soil investigations. The
method has been extensively and successfully used along the Brazilian coast for numerous geotechnical site
investigations performed for offshore foundations studies, dredging as well as port installation studies.

1 INTRODUCTION million barrels/day in 2004, mainly due to offshore


discoveries in water depths ranging from shallow to
The development of offshore geotechnics in Brazil is ultra deep (over 3000 meters).
directly linked to the significant oil and gas findings The need for special structures for exploitation and
along the Brazilian coast starting in the 70s. production has required specialist geotechnical tools,
In the last 50 years, Brazilian oil production has such as the geotechnical diving bell (Fig.1), to
raised from the initial 2700 barrels/day to over 1.5 respond to these challenges (Bogossian, F. 2004).
This equipment, which is easily deployed from a
vessel-of-opportunity, consists of a diving chamber,
equipped to perform shallow underwater geotechnical
investigations and was developed in Brazil to respond
to the specific needs of platform foundation studies
along the Brazilian coast where it has been exten-
sively and successfully used in water depths up to
55 meters (Bogossian, F, e Spatz, F. 1978).
With Brazilian oil concentrated in its vast majority
offshore, the pipeline network has been widely
expanded in the later years to ensure a national oil
supply, thus bringing about new technical and envir-
onmental requirements for shore approach studies.
Due to its flexibility and cost-effectiveness, the
geotechnical diving bell is able to respond to these
new challenges.

2 TECHNICAL CHARACTERISTICS OF THE


EQUIPMENT

2.1 The geotechnical diving bell


The geotechnical diving bell allows the execution of
Figure 1. The geotechnical diving bell in operation. offshore geotechnical investigations with the use of

957

Copyright 2005 Taylor & Francis Group plc, London, UK


standard onshore equipment, thus avoiding all the
well known issues usually introduced when jack up
rigs and other conventional offshore methods are used,
such as air gap, wave and current movements etc.
(Bogossian, F. e Machado, C. F. Dias. 1981). Two dif-
ferent sizes of geotechnical diving bells are available,
a small one for current sampling campaigns and a larger
one for more complex testing programs. The small one
is 3.0 m high, with base dimensions 2.2 m  2.2 m and
weighs 4 tons. The large one is 4.5 meters high, with
base dimensions of 2.3  2.5 meters, weighing 6 tons.
Its three shoes are hydraulically regulated. The
geotechnical diving bell is equipped with a transpon- Figure 2. The geotechnical diving bell being carried to the
der for positioning. The chamber is electrically lighted investigation site.
and communication includes internal and external TV
and independent voice systems.
All the equipment, including conventional drilling
material, such as rods, barrels and etc., is adapted to
the walls of the geotechnical diving bell which is
ergonomically designed to ensure more efficient pro-
duction with less strain on the operators.
A hydraulic lightweight drill is used to perform the
soil borings and rock coring. Thin-walled tube sampling
and in situ tests as well as split spoon sampling and
standard penetration tests (SPT) can also be performed.
In situ tests, such as CPT, vane, dilatometer and
pressiometer tests can also be easily included.
Energy, communication and air supply are transmit-
ted to the bell chamber by means of umbilical cables
connected to the support boat on the surface. It is also
equipped with floaters that enable shorter displace-
ments to be made without the need of hoisting the
geotechnical diving bell to the support boat.

2.2 The support vessel requirements


The diving bell can be operated from a number of dif-
ferent vessels, the sole requirement being a stern
A-frame of 8 tons capacity. This flexibility enables Figure 3. Divers operating inside the geotechnical diving
vessels like supply boats, barges and survey vessels to bell.
be used subject to the specific conditions of the loca-
tion. Access to the port and support, water depth and
sea conditions are among the elements that will positioning system (DGPS) and anchored at the bor-
define the type of vessel to be used among the ones ing location, the diving bell is lowered to the seabed.
mentioned above. Once on the seabed, the geotechnical diving bell is
Additional equipment required is basically winches positioned with the transponder, ballasted, the three
for hoisting the bell, an air compressor, a diesel pump shoes leveled and the gas mixture injected through
and an electric generator to supply power to the bell. the umbilical, thus expelling all the water from inside
A decompression chamber must also be available to be the chamber until a water seal of 0.5 m is created
used according to the standard decompression tables. which is kept throughout the operation to give access
to divers. The gas mixture that is injected inside the
bell is also carefully monitored.
3 EQUIPMENT OPERATION The divers work in pairs inside the chamber per-
forming the drilling, sampling and in situ tests (Fig. 3).
The vessel with a stern A-frame carries the geotech- Their work is supervised by a geotechnical engineer
nical diving bell to the investigation site (Fig. 2). With or geologist who remains onboard the support vessel
the deployment vessel positioned using digital global together with the other members of the crew, namely

958

Copyright 2005 Taylor & Francis Group plc, London, UK


a laboratory technician, two diving supervisors, a After arriving at the sampling site, there was a
mechanic, for equipment operation and maintenance stand-by period of two days before positioning condi-
and a diving safety technician. tions could be attained. There were also two other
stand-by days during displacement between locations.
During the performance of the services there were
4 OPERATIONAL CONDITIONS only two days missed with stand-by between borings.
The six boreholes totaled 90 m and were performed in
Sea conditions for positioning the bell should not be 23 days, including mod/demob and installation phases.
stronger than those for force 3 on the Beaufort scale,
with winds of 7 to 10 knots of speed and waves up to
60 cm high. During bell operation force 4 is accepted,
which imply in winds of 11 to 16 knots and waves 6 CONCLUSIONS
of 1.5 m.
In breaker zones these values change to force 2 for The geotechnical diving bell has many characteristics
positioning and force 3 for operational procedures. that make it extremely valuable for underwater geo-
In case of rougher sea conditions, the geotechnical technical investigations in waters up to 55 m deep,
bell can be abandoned with no risk for the divers and among which the following can be highlighted:
crew until operational weather conditions are restored
It allows the performance of nearshore and offshore
and work can be resumed.
geotechnical investigations with the use of stand-
The maximum water depth for operation should
ard onshore drilling and testing equipment, with-
not be higher than 55 m whereas the minimum oper-
out introducing, any new variables to the results.
ational water depth, being a function of the vessels
A long history of successful results, with data used
minimum draught, should be no smaller than 2.9 m.
in foundation studies of innumerable existing pro-
Borehole depths of 80 m have been achieved.
duction oil platforms.
Production rate is a function of the water and bore-
A cost-effective solution for geotechnical soil bor-
hole depth, thus for the limit operational water depth
ings performed in the surf and breaker zones, in
of 55 m and a borehole depth of 30 to 40 m a produc-
water depths of 2.9 m, where conventional jack up
tion rate of 5 m per shift can be attained while for shal-
platforms might not be so efficient, which makes it
low water depths of around 10 m, a rate between 10 to
especially suitable for pipeline shore approach
15 m per shift can be expected for the same borehole.
studies.
Diving planning and routines follow safe diving
standards and international procedures. All diving
crew have extensive training on environmental pro-
tection and occupational health and safety together REFERENCES
with the onboard crew, besides being trained to exe-
cute the soil investigation procedures. Bogossian, F. 2004. A geotecnia offshore, no Brasil,
aplicada indstria do petrleo. In Geotecnia Revista
da Sociedade Portuguese de Geotecnia 2329. in
5 A RECENT OPERATION Portuguese.
Bogossian, F. e Machado, C. F. Dias. 1981. Energy dissipa-
The geotechnical diving bell was recently used for tion on the SPT rods. In Proc. Of tenth International
geotechnical investigations for the submarine sewage Conference on Soil Mechanics and Foundations Engi-
discharge pipeline in Ipanema Beach, Rio de Janeiro neering; Stockolm, 449450. Rotterdam: Balkema.
(Mello, Jayme R. C. and Gaulgoul, N. S. 2004). The Bogossian, F, e Spatz, F. 1978. Underwater soil and rock
scope of work was the execution of six percussion soil geotechnical borings for engineering purposes. In
Offshore Brazil; Proc. OB 78.06.5 12, Rio de Janeiro.
borings with SPT tests and split spoon or thin-walled Mello, Jayme, R. C. and Gaulgoul, N. S. 2004. Foundation
sampling subject to the varying soil characteristics in Reinforcement of the Ipanema Beach Offshore Sewage
the breaker zone. The maximum water depth at the Pipeline 1.83; Proc. Fifth International Conference on
location was 27 m. The vessel used was a barge with Case Histories en Geotechnical Engineering; New York,
stern A-frame. NY, April 1317, 2004.

959

Copyright 2005 Taylor & Francis Group plc, London, UK


Conquering new frontiers in underwater cone penetration testing

Karin van den Berg, Axel Walta & Titus de Wolff


A.P. van den Berg, Heerenveen, The Netherlands

ABSTRACT: The future poses a challenge with the need of accurate information about the bearing strength and
shear strength of very soft seabed at ever greater water depths. The static cone penetration testing (CPT) method
is a technique that has proven to be very suitable for obtaining reliable data in the harsh environment of the deep
sea. Special requirements apply for the two main parts that make up the CPT unit: the pushing device and the
measuring device. This paper deals with the aspects of designing these parts in such a way that they meet the chal-
lenges posed by the deep sea environment at 5000 m water depth and anticipates further future developments.

1 INTRODUCTION to settlements and differential settlements. This devel-


opment required a better and more accurate under-
What are extreme environments? They are merely cir- standing of the behaviour of their foundations and that
cumstances that we are not used to. We define a depth of the subsoil. Especially after 1945 the development
of 5000 m below sea level as an extreme environment, of the CPT method was remarkable. The method of
but the fish that are living there would probably say, if interpretation of the reading from CPTs in order to
they could talk, that we are the ones living under predict the bearing capacity of foundation piles
extreme circumstances, that is, under an extremely low replaced the pile driving formulae entirely, as well as
pressure. So trying to work under extreme environ- the driving of test-piles. By means of a very large num-
ments means nothing more than trying to broaden your ber of load-tests it was proven that by considering the
horizon and to conquer new, so far unknown territory, CPT as a model test-pile, a sufficiently accurate pre-
which is what mankind has always endeavoured. diction of its bearing capacity could be obtained.
Nowadays it is generally considered unacceptable to
1.1 Cone penetration testing design a pile-foundation in the Netherlands without
having available sufficient CPT data.
Cone Penetration Testing is all about conquering new
territory. It has been used for over 60 years in the 1.2 Advantages of the CPT method
Netherlands for the prediction of pile behaviour under
loading conditions of the approximately 1 million Characteristic advantages of the CPT method over
foundation piles yearly put into practice. Without other types of soil investigation are:
these piles it would be virtually impossible to build 1 A continuous stream of information covering the
(lasting) structures in the Netherlands. full depth is obtained so that the layers of different
The geological conditions in the Netherlands make composition or behaviour can be shown;
it necessary, already since several hundreds of years 2 Measurement of the parameters is done in-situ, so
ago, to apply pile-foundations on a large scale. At first that the actual stress conditions are influenced as
timber piles were used exclusively, while the driving little as possible, this in contrast with tests per-
depth was determined during the process of pile- formed just below the bottom of a borehole and
driving. Only much later the procedure was improved especially when it concerns so-called undisturbed
in such a way that simple pile driving formulae, based samples for testing in the laboratory;
on empirical data, came into use, in order to assess a 3 The results are very reliable because the human
relation between driving depth and bearing capacity. influence on registration of the results can be
Between 1930 and 1935 the CPT method came into entirely excluded;
use. It enabled the pile-length and the bearing capacity 4 The method provides results, which can be inter-
to be determined prior to pile installation. The process preted on the spot;
was also accelerated by the general trend in the building 5 The costs involved are infinitely lower than those
trade to erect larger structures, which are more sensitive of other kinds of soil investigations.

961

Copyright 2005 Taylor & Francis Group plc, London, UK


2 HISTORY OF OFFSHORE SOIL
INVESTIGATION EQUIPMENT

2.1 Wireline CPT: WISON-APB


Offshore soil investigation using the static Dutch
cone penetration testing method was started in 1970
with wire-line down-the-hole CPT equipment,
called WISON (stands for WIreline SONdeerapparaat),
built by A.P. van den Berg for Fugro. Since that time
numerous units were manufactured by several different
parties and many different technical developments
have taken place. Nowadays the wireline equipment
made by A.P. van den Berg is called WISON-APB.
The CPT unit is lowered into a pre-drilled bore-hole
to measure cone tip resistance, local friction and
pore-water pressure over 3 m intervals up to maxi-
mum 50 kN, 1 m intervals up to maximum 100 kN
and 1.6 m intervals up to maximum 150 kN. Figure 1. Mixture of groups.

2.2 Seabed CPT: ROSON Table 1. Considerations regarding group: handling.


In the eighties the ROSON (stands for ROtating
Sub-group Possible solutions
SONdeerapparaat) was put into use for the first time:
a CPT unit that could be lowered onto the seabed for Weather conditions Light weight, small
continuous penetration testing at capacities ranging Sea conditions Light weight, small
from 10 kN to 200 kN. The driving force was supplied Skills of the crew Easy to operate
by electrically driven friction wheels, ensuring the Long way up and down Hydrodynamic shapes
continuous movement into the soil.

2.3 Recent offshore developments


for both main parts that make up a CPT system,
A selection of products developed (or in develop-
namely the pushing device and the measuring system.
ment) by the R&D team in the last three years is:
These technical challenges arise from:
Seabed sampler (development phase; since 2003)
6 The ambient water pressure of 50 N/mm2;
Wireline system with a depth rating of 2000 m
7 Salt water;
(development phase; since 2003)
8 The very soft seabed;
Cones with a depth rating of 4000 m (custom made
9 The accuracy of the cone to measure the soil with
for the investigation of the oil carrier Prestige)
a nominal bearing strength of 0.05 N/mm2
Depth encoder with a depth rating of 7000 m (taken
(50 kPa) with respect to the ambient water pres-
into production since 2002; 2 systems are commer-
sure of 50 N/mm2 (50 MPa);
cially in use)
10 The long way down and up, to and from the bot-
Acoustic data transmission for the Roson system
tom of the sea.
with a depth rating of 2000 m (taken into produc-
tion since 2000; 3 systems are commercially in use).
3.1 Considerations regarding the pushing device
3 CONQUERING NEW FRONTIERS IN Designing offshore CPT equipment is completely dif-
UNDERWATER CONE PENETRATION ferent compared to the designing of onshore CPT
TESTING equipment. Figure 1 gives a rough idea about the con-
ditions to be considered when designing CPT equip-
The need to perform offshore soil investigations at a ment for offshore deployment. Tables 1 thru 4 show
water depth of 5000 m requires a whole new design of the typical groups, divided into subgroups. Below
the CPT equipment in order not only to survive the every subgroup, possible solutions are mentioned to
extreme environment at 5000 m water depth but also cope with the issues posed by that specific subgroup.
to obtain data with the required accuracy and reliabil- The conclusion should be easy: design a machine that
ity. The challenges encountered need to be addressed contains all these subgroups solutions.

962

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 2. Considerations regarding the group: environment. To be still able to penetrate to 40 m with less
weight and a smaller frame, the following modifica-
Sub-group Possible solutions tions are needed:
Seawater Protecting coating; seawater resistant 11 the cross sections of the cone and the rods are
materials reduced (  F/A);
Ambient pressure Strong housing or fluid filled with 12 a cartridge system with CPT rods and a manipula-
compensator tor to connect and disconnect the rods.
A computer with special software running on deck
Table 3. Considerations regarding group: seabed conditions. communicates with the submerged multiplexer on the
unit and collects the CPT data. A window on the
Sub-group Possible solutions screen shows the movements of the manipulator.
After a sounding depth of 40 m the manipulator dis-
Slope Adjustable support plates connects the rods and put them back in the cartridge.
Soft Large surface of the support plates to avoid When lifting the equipment from the seabed the
sinking in winch on deck will have a hard time to pull the machine
Hard Heavy reaction weight upwards against the massive resistance. Hydrodynamic
smooth topside will help to ease the lifting.
All in all the entire unit is made smaller, lighter,
Table 4. Considerations regarding group: conditions for safer to handle, even in bad weather, and still able to
measurements.
push 40 m.
Sub-group Possible solutions

As little as possible Large surface of the support


3.1.2 Environment
disturbance of the seabed plates to avoid sinking in 3.1.2.1 Water pressure
No undesirable Large stable unit to lead the At 5000 m water depth the ambient water pressure
movements of the reaction forces to the seabed will be a staggering 50 N/mm2. While lowering the
equipment
machine, a pressure compensator will be necessary to
deal with the pressure difference of 50 MPa between
the inside and outside of the housings.
But the solution for one can undo the solution for Some chambers and housings in the unit, which
the other. And this is the challenging part for the need an atmospheric pressure of 0.1 MPa, like elec-
Research & Development team. tronic chambers, have to be sealed off from the high
The Research & Development team has to get the pressure. This affects the seals and the geometry of
right balance between the solutions of the subgroups. the design. Some of the chambers or housings have to
A technical background, a good sense of imagination be pressure compensated otherwise the unit will be
and an extensive working experience makes all the too heavy. The compensating system has to react on
difference in creating a good working design. every change in ambient pressure otherwise the dif-
ferential pressure can damage the seals and seawater
3.1.1 Handling will have access to important parts of the unit.
In order to be able to perform soil tests up to 40 m The ambient water pressure especially has effect
depth, the pushing device needs a heavy weight to on the cone. If a normal onshore cone is used it will
counter the pushing thrust and a tall frame to protect the be impossible to measure a value accurately because
long CPT string. A large seabed contact base for stabil- the measured value will be less than 0.1% of the
ity and minimum disturbance of the bottom adds to the ambient water pressure. A special cone design will be
heavy weight. However a sudden weather change, like necessary to make accurate measurements.
hard wind and high waves will turn a heavy machine
like that into an uncontrolled piece of equipment. 3.1.2.2 Salt water
Weight reduction will help to control the machine. It is Salt water can be very aggressive to certain materials,
clear that a critical balance has to be found for the and may therefore cause failures. Especially saltwater
opposing pre-conditions with respect to weight. in combination with electricity will give corrosion
While lowering the equipment to the seabed the problems. Next to the underwater problems, a prob-
speed cannot be high otherwise it will tumble over lem can also occur on deck. When the system is used
and the CPT rod protection part will collide with the and put back on deck, the water will evaporate and the
lifting cable and the umbilical with serious risk of salt crystals will stay behind. These salt crystals stick
damage. With a less tall frame and a smooth base the to grease and can affect the function of for example a
lowering procedure will be less risky and quicker. bearing or a seal.

963

Copyright 2005 Taylor & Francis Group plc, London, UK


3.1.3 Seabed conditions
The very soft seabed consists of very small particles
suspended in water. The challenge is to detect the bor-
der between water and seabed. When the system is
lowered with a certain speed and hits the seabed, dis-
turbance of the soil will occur. The weight of the sys-
tem is also a problem if the base plate is too small.
The unit will sink into the seabed. If the pushing sys-
tem sinks into the soil the consequence is that the
cone or vane test apparatus starts below the seabed
surface. This will preload the cone during zero read-
ing. A preloaded cone or vane test apparatus will give
data biased with the preload value. Distance to seabed
measurement during landing will prevent this.
Before testing measurement of the horizontality
and possible correction of horizontality is necessary.
Three moveable legs with levelling plates with suf-
ficient area answer to these issues. Before lifting the
equipment after performance of the test, the legs are
retracted.

3.2 Considerations regarding the measuring Figure 2. Standard A P van den Berg cone dimensions.
system
One of the most important challenges is an accurate
measurement of the soil strength parameters in deep In the case of A.P. van den Berg 10 cm2 cones this
water. The main issues to be addressed with measur- is 7.5 cm2 (  0.75). Per 1000 m water depth this
ing soil strength in deep water are: results in a force in the tip load cell of approximately
13 Sensitivity versus water pressure, especially for 750 kN. This equals 7.5 MPa tip load. When measur-
friction measurements. The friction of the O ing soft clays the preloading of the cone exceeds a
ring seals to the other parts increases with the measured value of the clay by far (a typical value
water pressure and limits the free movement of measured in soft clays is 50 kPa).
for instance the friction sleeve
14 Water pressure consuming the measuring range of 3.2.1 Possible solutions
the cone tip resistance qc and pore water pressure U.
3.2.1.1 Hydrostatically compensated cone.
The force acting on the cone resulting from the In a hydrostatically compensated cone the soil seals
penetration speed is passed on to the pushing rod in and seals for water tightness can be avoided when a
several ways. Only the way through the load cell is subtraction design is chosen. The strain gauges can
measured by the data acquisition system. Other pos- then be fitted on the inside of the two load cells while
sible ways are for instance through O rings, soil the outside is exposed to the surrounding water. This
seals and dirt or contamination. eliminates the problem mentioned under 3.2, sub 13
O rings for instance deform under high pressure but adds the problems associated with subtraction
and exert a higher force on its metal surrounding. cones like inaccuracy of the local friction.
This causes a greater part of the generated force to In the hydrostatically compensated cone the pre-
pass through the O rings, resulting in a lower meas- load caused by the water pressure due to area affects
ured parameter. is grossly avoided. However the problem of calibra-
The area on which the water pressure U2 acts pre- tion is introduced.
loads the cone load cell. The surrounding water pressure is now present on
Fh  ghzA all sides of the load cell in the cone leading to a three
dimensional deformation of the material of the load-
Where; cell and the strain gauges. This deformation of the
Fh is the force on the load cell material of the cone differs from the deformation
is the unit weight of (sea) water caused by the force from the soil. The IKS institute of
g is the earth gravity at the location the University of Siegen, Germany found a deviation
hz is the water column of 1% in span when testing torque load cells under
A is the area on which the water pressure acts. 40 MPa pressure.

964

Copyright 2005 Taylor & Francis Group plc, London, UK


Load resulting from a 0.002MPa weight

Figure 3. Relation between water pressure and measured


load.

A hydrostatically compensated cone needs to be


calibrated under the same pressure as it is used under,
which makes the calibration procedure technically
very challenging and only valid at the calibrated
Figure 4. Calibration trend lines.
pressure.

3.2.1.2 Specially designed normal cone The repeatability or precision of the load cell itself
The ambient water pressure now only acts on the out- is much better than 0.1%. When only a small part of
side of the cone. The deformation of the material on the full measuring range is used the calibration pro-
the outside of the cone can be calculated and taken cedure must be adjusted to this. Modern software
into consideration in the cone design. packages are well capable of comparing accurate
The load cell is at atmospheric pressure and nor- 100-point calibration data with acquired soil data
mal calibration data applies. eliminating non-linearity effects of the load cell.
The inaccuracy due to the problems mentioned in In a calibration procedure the cone load cell is com-
3.2, sub 13 and 3.2, sub 14 need to be addressed to see pared to a standard. Two points (zero and full load) are
if the problems can either be avoided or be known so used to determine a straight line. This trend line is
accurately that the effects can be adjusted. used to calculate the actual load from the loadcell out-
put. Accuracy is the deviation of the actual load on the
loadcell from the calculated load (see Fig. 4).
3.2.1.3 At 3.2, sub 13
A different situation occurs when only a small part
Measuring the friction of the O rings under pressure
of the measuring range is used and the zero reading is
is determined by applying a small accurately known
recorded at the start of the test and used as the refer-
force to the friction sleeve of a normal cone at various
ence for testing.
pressures. For this purpose AP van den Berg designed
The trend line differs from reality in two ways: an
a high-pressure vessel that is hinged in the middle. By
absolute distance between the lines and a deviation in
turning the vessel upside down the known weight of
angle.
the friction sleeve can be applied to the load cell or
Taking zero readings close to the measured value
not. Mounting a friction sleeve with defined weight
eliminates the inaccuracy due to the absolute distance
onto the cone and turning the vessel enables measure-
between the lines. The inaccuracy due to the differ-
ment of the friction of the O rings.
ence in angle however is exaggerated.
A.P. van den Berg used this method to test her cones
By determining the real correlation for instance at
for deep water. Having a reliable measuring method
100 points divided over the range the correlation
resulted in small design changes that minimizes the
between load and load cell output is accurately known
effects of O ring friction as is shown in Figure 3.
and the software can calculate the load on the loadcell
accurately from the loadcell output.
3.2.1.4 At 3.2, sub 14 Now repeatability has become accuracy. This
With only a small part of the range used for measuring brings deepwater CPT data well within the required
the soil data special care needs to be taken with cone accuracy with an atmospheric cone. This led Thales-
calibration. The cone is usually specified as 0.1% Geosolutions to choose A.P. van den Berg as manu-
accurate. The overall accuracy depends further on the facturer of the cones to CPT around the Prestige at
data acquisition system, the cabling in between and 3800 m water depth with excellent results. Figure 5
force bypass effects as mentioned before. shows one of the cones used in the Prestige project.

965

Copyright 2005 Taylor & Francis Group plc, London, UK


meeting these other demands. Challenges to construc-
tion and technical performance posed by the environ-
ment of deep seas have been successfully met up to a
water depth of about 4000 m and further designs for
investigations of the soil at 5000 m water depth are
very real, with an anticipated 0.5% accuracy of the
measuring results. The challenges we are facing are
leading us to a higher level of technology which may
eventually even help us conquer the threshold of
extraterrestrial soil investigation.
Figure 5. 2.6 cm2 cone, depth rated 4000 m.

REFERENCES
4 CONCLUSIONS
Martin Knight, Hugh Burbidge & Brian Hitchcock, 2001.
What does the future hold in store for us? The ever- Pennys Bay reclamation A case study demonstrating
growing demand of energy and living space leads to the successful application of an advanced seabed PCPT
system.
the exploration of areas in ever more extreme envir- Lunne, T., Robertson, P.K. & Powell, J.J.M. 1997. Cone
onments. At the same time higher demands are set to Penetration Testing in Geotechnical Practice.
environmental concerns, reliability, safety and costs. ASTM 1995. ASTM Standard Test Method.
The CPT method has proven to be the best solution Dutch Standard Test Method NEN5140.
so far in the reliable and accurate collection of soil ISSMFE 1089. International Reference Test Procedure
data under extreme environments, at the same time (IRTP).

966

Copyright 2005 Taylor & Francis Group plc, London, UK


Cyclic friction piezocone tests for offshore applications

G.L. Hebeler & J.D. Frost


Georgia Institute of Technology, Atlanta, United States

J.A. Schneider & B.M. Lehane


University of Western Australia, Perth, Australia

ABSTRACT: Soil strength degradation due to cyclic loading may have significant implications on foundation
performance, particularly for offshore structures. For soils types that may incur significant disturbance during
sampling, such as very soft clays, sands, or structured materials, an in situ measurement of cyclic degradation
characteristics may be more representative of field performance. The addition of cycling into an in situ testing
programme adds little cost, and typically requires no design modifications to most common in-situ penetration
devices and push equipment. However, the added data relating to large strain cyclic degradation can be effect-
ively used to assess issues such as soil sensitivity and post installation pile shaft friction. Preliminary studies of
cyclic piezocone penetration tests, with emphasis on sleeve friction response, are discussed herein. The current
study utilizes a traditional 15 cm2 CPTU device outfitted with four additional friction and five additional piezo
sensors to obtain a more complete representation of cyclic shaft response.

1 INTRODUCTION

Waves, wind, earthquakes, or construction activities


apply various types of cyclic loading to foundation
elements. This cyclic loading may lead to generation
of pore pressures, degradation of strength, and settle-
ment. The significance of cyclic loading effects is a
function of amplitude, number, and other load cycle
characteristics. Typically the assessment of cyclic
degradation is addressed through laboratory test studies
(e.g. Andersen et al. 1980, Jewell & Randolph 1988,
Airey et al. 1992), although, some recent site investi-
gations have included a component of cyclic penetra-
tion testing (Randolph 2004, Peuchen et al. 2005).
Evaluation of cyclic degradation through in situ test-
ing may be more appropriate for soils types that may
incur significant disturbance during sampling, such
as very soft clays, silts, sands, or structured materials.
Most previous cyclic penetration studies have
focused solely on tip resistance, including cone, ball,
and T-bar geometries. The current cyclic soundings
were performed using a conventional 15 cm2 piezo-
cone penetration testing (CPTU) unit configured with
a multi piezo friction attachment (MPFA), Figure 1
(Hebeler 2005). The MPFA includes 4 additional fric-
tion and 5 additional pore pressure sensors along the Figure 1. Multisleeve piezo-friction penetrometer attachment
shaft of the penetrometer. These sensors are located configured with conventional CPTU Module. (a) schematic
more than 10 cone diameters behind the tip, outside brackets indicate sensor offset from tip in meters, (b) design
of the complex stress regime created near the cone tip detail, and (c) piezo friction sleeve mandrel design detail.

967

Copyright 2005 Taylor & Francis Group plc, London, UK


during penetration. The current study investigates the The friction sleeves of the MPFA have been pre-
use of such sensors to evaluate the rate and magnitude loaded by approximately 10 to 20 kPa using a rubber
of cyclic interface friction degradation. Data from o-ring, and thus have better low load resolution, as
two onshore sites in Western Australia, a soft clay site well as the ability to measure a limited range of nega-
(Burswood) and an uncemented calcareous sand site tive sleeve friction values. Thus, during cycling the
(Ledge Point), are used to investigate the viability of multi sleeve sensors can measure limited negative val-
the testing methods and device performance. ues during extraction (10 to 20 kPa), and very low
residual values which are often below the resolution
of conventional sleeve friction designs.
2 THE MULTI PIEZO FRICTION
ATTACHMENT (MPFA) 3 CYCLIC TESTING RESULTS,
INTERPRETATION, AND DISCUSSION
2.1 Device specifications
The MPFA device is the second generation of multi 3.1 Overview of test sites
friction devices developed at the Georgia Institute of Two onshore test sites at Burswood and Ledge Point,
Technology (GIT). The precursor, developed by DeJong representing problematic soil conditions often found
& Frost (2002), is the multi-friction sleeve attachment offshore, were used in the current study.
(MFSA), and is comprised of four independent fric- The Burswood site is located along the Swan
tion mandrels accepting replaceable sleeves of varying River, three kilometers upstream from the centre of
texture. The MFAs are designed for use behind con- Perth, Western Australia. Burswood consists of soft,
ventional 15 cm2 CPT devices to allow for simultan- high plasticity clay (LL  80, PI  40 to 50) with
eous measurement of CPT parameters (e.g. q c, fs, u2) in shell fragments and silt lenses. The undrained shear
conjunction with additional interface shear measure- strength profile gradually increases from approxi-
ments, and in the case of the MPFA, the correspond- mately 16 kPa at 4 m depth to 35 kPa at 17 m depth,
ing pore pressure response. When configured with a with an apparent OCR due to ageing of 1.5 to 2. The
conventional digital CPT module (0.61 m in length) sensitivity based on field vane data shows values
the MPFA has a total assembled length of 2.49 m. between 4 and 9 above 7 m, and between 2 and 4 below
7 m. Additional information and test results from the
2.2 Friction sleeve design Burswood site are contained in Randolph (2004) and
Chung & Randolph (2004).
From a conceptual perspective, evaluation of pile shaft The Ledge Point site is located approximately
friction based on an in situ measurement of interface 100 km north of Perth, Western Australia, along the
friction should provide a more accurate and reliable coast of the Indian Ocean. Ledge Point consists of an
prediction than a parameter controlled by a different uncemented calcareous uniform fine sand (SP). The
failure mechanism, such as cone tip resistance. How- D50 grain size is 0.24 mm, the sand has less than 5
ever, the CPT friction sleeve (fs) measurement obtained percent fines, and the maximum void ratio ranges
in the standard configuration directly behind the tip is from approximately 1.2 to 1.4.
considered highly variable and less reliable than the
CPT tip (q t) and pore pressure (u 2) measurements.
3.2 Cyclic results in soft clay (Burswood)
The high variability of traditional fs measurements
has been shown to be primarily due to the effect of the A CPT-MPFA sounding was performed at the Burs-
sleeve being located within a zone of high stress gra- wood test site, with ten full length (1 m) 2-way cycles
dient that develops behind the cone tip (Campanella carried out at 20 mm/s between tip depths of 5.8 and
& Robertson 1981), the sleeve roughness changing 6.8 m. A dissipation test was performed for approxi-
within and between soundings (DeJong et al. 2000), mately 2 hours after the initial virgin penetration to
and the use of load cells with varying resolution and 6.8 m, before the initial upward half cycle. Figure 2
configurations (Lunne et al. 1997). Given these undesir- shows the cyclic behavior of the corrected tip stress (qt),
able influences, the CPT fs measurement in its current u2 pore pressure, a representative multi sleeve (MS)
configuration does not lend itself to accurate predic- stress (fs2), and a representative multi piezo (MP) pres-
tion of average interface strength well behind sure (u s2), respectively. The q t and u 2 measurements
a pile tip. The MFAs overcome the above shortcom- show hysteresis loops that are affected by the relative
ings by accepting sleeves of specified roughness, position of the sensors within the cycle, as a result of the
providing four independent measures of interface large influence zone around the tip. However, the multi
friction spaced 260 mm apart within the stable shearing sleeve and piezo sensors located further up the shaft
region along the shaft, and by the use of isolated non- measure the behavior over a smaller influence zone
subtraction load cells. and only show variability in the cycle due to layering.

968

Copyright 2005 Taylor & Francis Group plc, London, UK


qt tip stress (kPa) To quantify the effect of cycling on the various
-300 -100 100 300 sensor results, the average value for each full stroke
5.4
(or half cycle) is shown as a function of cycle number
(a)
in Figure 3. This framework is comparable to that dis-
5.8 cussed in Matlock & Foo (1980), but it is noted that
the cumulative strain is as important as number of
Depth (m)

6.1 stress reversals (cycles) when quantifying cyclic degrad-


ation (Poulos 1981). The cycles discussed in this paper
6.4
(1 m in length) are relatively large and create a stable
cyclic response.
Average degradation of qt with large scale cycling
6.8
is presented in Figure 3a. The average (q t) tip resistance
decreases with each cycle, with the initial portion of
u2 pore pressure (kPa)
each downward cycle exhibiting negligible resistance.
-50 0 50 100 150 200 250
The initial extraction resistance is relatively high due
5.4 to the aforementioned dissipation test, and subsequent
(b) cycles show higher resistance during insertion as com-
5.8
pared to extraction. The pore pressure measurements
in Figure 3b mirror the cone tip resistance, since u2
is more strongly influenced by the reduction in octa-
Depth (m)

6.1
hedral pore pressures due to degradation in soil strength
than the increase in u2 caused by shear induced pore
6.4 pressures from soil destructuring.
Figure 3c and 3d show the average cyclic degrad-
6.8 Hyd Pressure ation of sleeve friction and pore pressure for a repre-
Pore pressure sentative MS and MP sensors centered 1070 and
1140 mm (24.6 and 26.2 diameters) behind the cone
fs2 sleeve stress (kPa)
tip, respectively. Reduction in pore pressures along
-20 -15 -10 -5 0 5 10 15 20
4.4
the shaft due to the dissipation test performed after
virgin penetration results in minimal initial shaft fric-
(c)
tion reduction (compare the degradation rate in down-
4.7
ward cycles 0 to 1, to that of cycles 1 to 2). The shaft
response degrades to approximately 40 to 60 percent
Depth (m)

5.0
of the end of dissipation values as a result of 10 large
scale cycles. During this test series, the ratio of tension
5.4
to compression sleeve friction reduces from approxi-
mately 90 percent during initial loading after dissipation
5.7 to 50 percent at 10 cycles. Since the soil is essentially
sliding along the friction sleeve with no additional
6.0 radial soil displacement, the pore pressures measured
us2 pore pressure (kPa) further back along the shaft primarily represent the
0 50 100 150 interface shear behaviour. There is an initial drop in
4.3
pore pressures due to the 2 hour dissipation, followed
(d)
by positive us2 pore pressures relating to degradation
4.6
at fs2.
Depth (m)

5.0
3.3 Cyclic results in calcareous sand (Ledge Point)
5.3 A CPT-MPFA sounding was performed at a site in
Ledge Point Western Australia with ten full length
5.6 (1 m) 2-way cycles performed at 20 mm/s between tip
Hyd Pressure
depths of 6.7 and 7.7 m. No dissipation test was per-
MP Pressure
60 formed for this data set, but dissipation will have little
influence at a sand site as compared to the Burswood
Figure 2. Cyclic CPT-MPFA results between tip depths of clay site. Figure 4 shows the cyclic qt, u2, and represen-
5.8 and 6.8 m at the Burswood soft clay site: (a) qt tip stress, tative MS and MP (located at the same distance behind
(b) u2 pore pressure, (c) fs2 stress, and (d) us2 pressure. the tip as in Figures 2 and 3) measurements, respectively.

969

Copyright 2005 Taylor & Francis Group plc, London, UK


Cycle # qT tip stress (kPa)

0 2 4 6 8 10 -5000 0 5000 10000 15000 20000 25000


300 6.4
(a) (a)
250
6.7
Average qt per cycle (kPa)

200
150

Depth (m)
100 7.1

50
0 7.4

-5
- 100 7.7
Downward strokes
- 150
Upward strokes
- 200
u2 pore pressure (kPa)
Cycle # -50 0 50 100 150
0 2 4 6 8 10 6.4
160 Pore pressure
(b)
(b)
140 Hyd Pressure
6.7
Average u2 per cycle (kPa)

120

Depth (m)
100 7.1

80
7.4
60

40 7.7
Downward strokes
20 Upward strokes
Hyd Pressure
0
fs2 sleeve stress (kPa)
Cycle #
-30 -20 -10 0 10 20 30
0 2 4 6 8 10 5.3
15
(c)
(c)
10 5.7
Average fs2 per cycle (kPa)

5
Depth (m)

6.0

0
6.3
-5

6.6
-10 Downward strokes
Upward strokes
-15
Cycle # us2 pore pressure (kPa)
0 2 4 6 8 10 0 20 40 60 80 100
120 5.3
(d) (d) MP response

100 Hyd Pressure


Average us2 per cycle (kPa)

5.6

80
Depth (m)

5.9
60

6.2
40
Downward strokes
20 6.6
Upward strokes
Hyd Pressure
0

Figure 3. Average cyclic degradation results from tip depths Figure 4. Cyclic CPT-MPFA results between tip depths of
of 5.8 to 6.8 m depth at the Burswood soft clay site, (a) qt tip 6.7 and 7.7 m at the Ledge Point calcareous sand site: (a) qt tip
stress, (b) u2 pore pressure, (c) fs2 stress, and (d) us2 pressure. stress, (b) u2 pore pressure, (c) fs2 stress, and (d) us2 pressure.

970

Copyright 2005 Taylor & Francis Group plc, London, UK


Figure 5 shows the average cyclic response for both the Cycle #
downward and upward half cycles for each sensor 0 2 4 6 8 10
10000
over ten 2-way cycles. (a) Downward strokes
Figure 5a, shows the average downward tip response Upward strokes

Average qT per cycle (kPa)


8000
increasing from the 4th cycle onward. This is not
6000
indicative of an increase in soil strength, but rather
denotes an increase in soil density caused by collapse
4000
of the hole and displacement of infilling soil. This effect
is quantified by the qt stress at the bottom of penetration 2000
increasing from a virgin value of approximately 11 MPa
to a value of 23 MPa after 10 full cycles. The tip behav- 0
ior shows negligible resistance for the upper half of each
downward stroke, and shows no resistance during the -2000

upward strokes. Cycle #


0 2 4 6 8 10
The u2 pore pressure shows a slight difference 65
between upward and downward cycling, with slightly (b)
Downward strokes
positive pore pressures during downward penetration Upward strokes

Average u2 per cycle (kPa)


60 Hyd Pressure
and slightly negative pore pressures during upward
penetration. The pore pressures along the shaft, meas-
ured by the MP sensor also show minimal change
55
with large scale cycling.
Cyclic sleeve friction behaviour is shown on Figures
4c and 5c. The MS friction response shows a large 50
degradation due to cycling. The average reduction in
shaft friction after the first cycle is 60 percent, with
further degradation to only 13 percent of the initial 45
resistance (or 7.5 times lower) after three 2-way cycles. Cycle #
0 2 4 6 8 10
As discussed for the tests at Burswood, the MPFA can 25
provide an indication of the ratio of tension to com- (c) Downward strokes
20
pression capacity. The ratio of tension to compression
Average fs2 per cycle (kPa)

Upward strokes
MS friction reduces from 75 percent at initial loading 15
to 60 percent at 10 cycles. 10
Low pile shaft friction in calcareous soils is well
5
known, and typically attributed to cyclic loading and
contraction at the interface during installation and load- 0
ing. Shaft friction evaluated from model tests and con- -5
ductor pullout tests at the North Rankin A platform
-10
off the Northwest shelf, Australia, reduced from mono-
tonically installed values of approximately 40 kPa to -15
cyclically installed (driven) values of 5 to 20 kPa (Abbs Cycle #
et al. 1988, Poulos et al. 1988). Interface friction at large 0 2 4 6 8 10
55
displacements reduces by approximately an additional Downward strokes
(d)
50 percent (Poulos et al. 1988). For conductor load tests
Average us2 per cycle (kPa)

Upward strokes
offshore South Africa, Ebelhar et al. (1988) show CPT 50 Hyd Pressure
sleeve friction values of 20 to 40 kPa, and conductor
pullout resistances of 10 to 20 kPa.
Field results at North Rankin A and offshore South 45
Africa are in general agreement with the MPFA obser-
vations at Ledge Point. As such, cyclic data collected
40
during site characterization can provide insight into
sands with a high potential for cyclic contraction.
However, the rate of degradation for cyclic MPFA 35
tests should be much higher than that of large diameters
piles, since the rate of degradation will be influenced Figure 5. Average cyclic degradation results from tip depths
by the normal stiffness condition, and thus pile diameter of 6.7 to 7.7 m depth at the Ledge Point calcareous sand site:
(kCNS  4 G/D; see, Boulon & Fouray 1986, Lehane (a) qt tip stress, (b) u2 pore pressure, (c) fs2 stress, and (d) us2
et al. 1993, White 2005). Additionally, the minimum pressure.

971

Copyright 2005 Taylor & Francis Group plc, London, UK


achievable shaft friction is probably not solely a func- REFERENCES
tion of active earth pressure, but also arching behav-
iour around the pile or penetrometer (Randolph 2003). Abbs, A.F., Waterton, C.A., Bell, R.A., Khorshid, M.S. &
These effects, among others, need to be quantified in Jewell, R.J. 1988. Evaluation of pile friction from con-
a theoretical framework before cyclic MPFA results can ductor load test, Calcareous sediments; Proc. int. conf.,
2, Perth: 607616. Rotterdam: Balkema.
be applied quantitatively to large diameter pile design. Airey, D.W., Al-Douri, R.H. & Poulos, H.G. 1992. Estimation
of pile friction degradation from shearbox tests, ASTM
Geotechnical Testing Journal, 15 (4): 388392.
4 CONCLUSIONS AND RECOMMENDATIONS Andersen, K.H., Pool, J.H., Brown, S.F. & Rosenbrand, W.F.
1980. Cyclic and static laboratory tests on Drammen
Results of cyclic piezocone tests with high quality clay, Journal of the Geotechnical Engineering Division,
sleeve friction readings using the MPFA have illus- ASCE, 106 (GT5): 499529.
trated the significance of cyclic interface degradation Boulon, M. & Fouray, P. 1986. Physical and numerical simu-
at soft clay and calcareous sand sites. The testing meth- lation of lateral shaft friction along offshore piles in
sand, Numerical methods in offshore piling; Proc. 3rd int.
ods require little modification to currently available
conf., Nantes: 127147.
in situ testing equipment, as cyclic testing has previ- Chung, S.F. & Randolph, M.F. 2004. Penetration resistance
ously been shown successful in an offshore envir- in soft clay for different shaped penetrometers, Int. site
onment (e.g. Peuchen et al. 2005). characterization, ISC
2, Proc. 2nd int. conf., Porto:
Cyclic MPFA or MFSA measurements show prom- 671677. Rotterdam: Millpress.
ise as a method for identifying soils with the potential DeJong, J.T. & Frost, J.D. 2002. A Multi-Friction Sleeve
for significant interface contraction and a resulting Attachment for the Cone Penetrometer, ASTM Geotechnical
loss of friction during cyclic loading. The main design Testing Journal, 25(2): 111127.
applications are related to pile design and drivability Ebelhar, R.J. Young, A.G. & Stieben, G.S. 1988. Cone
penetration and conductor pullout tests in carbonate soils
studies. Currently the assessment of cyclic MPFA and
offshore Africa, Calcareous sediments; Proc. int. conf.,
MFSA soundings are qualitative, but results agree with 1, Perth: 155163. Rotterdam: Balkema.
frameworks for assessing cyclic loading of piles, as Hebeler, G.L. 2005. PhD Dissertation, Georgia Institute of
outlined by Poulos (1981) and Matlock & Foo (1980). Technology.
The large 1 m cycles used in the current preliminary Jewell, R.J. & Randolph, M.F. 1988. Cyclic rod shear tests
study seem excessive for friction degradation studies, in calcareous sediments, Calcareous sediments; Proc.
and increase the potential for gapping. Cycling over a int. conf., 1, Perth: 215222. Rotterdam: Balkema.
length of 2 to 3 diameters (approximately 1 friction Lehane, B.M., Jardine, R.J., Bond, A.J. & Frank, R. 1993.
sleeve length) may provide results that are still within Mechanisms of shaft friction in sand from instrumented
pile tests, Journal of Geotechnical Engineering, 119(1),
the relatively constant large displacement range (Poulos
ASCE: 1935.
1981, Poulos et al. 1988), while decreasing the testing Lunne, T., Robertson, P.K. & Powell, J.J.M. 1997. Cone
time and gapping concerns. However, differences in Penetration Testing in Geotechnical Practice, Blackie
cyclic degradation response as a function of amplitude Academic & Professional, New York, 312 pp.
may be observed, warranting additional investigations Matlock, H. & Foo, S.H.C. 1980. Axial analysis of piles
to optimize cyclic MPFA procedures and interpretation. using a hysteretic and degrading soil model, Numerical
methods for offshore piling, 127133. London: ICE.
Peuchen, J., Adrichem, J. & Hefer, P.A. 2005. Practice notes
ACKNOWLEDGEMENTS on push-in penetrometers for offshore geotechnical
investigation, Frontiers in offshore geotechnics, ISFOG,
Proc. int. symp., Perth.
Support from the ARC and the US NSF (OISE- Poulos, H.G. 1981. Some aspects of skin friction of piles in clay
0412895) of the first author and through an IPRS under cyclic loading, Geotechnical Engineering, 12: 117.
scholarship for the third author are gratefully acknowl- Poulos, H.G., Randolph, M.F. & Semple, R.M. 1988.
edged. Additional thanks are extended to Prof. Mark Evaluation of pile friction from conductor tests,
Randolph, Dr. Dave White, and others at COFS for ini- Calcareous sediments; Proc. int. conf., 2, Perth:
tial discussions regarding cyclic penetration testing, 599605. Rotterdam: Balkema.
cyclic installation and loading of piles, and on applica- Randolph, M.F. 2003. Science and empiricism in pile foun-
tions of the MPFA. Probedrill, Pty, of Western Australia dation design, Gotechnique, 53(10): 847875.
Randolph, M.F. 2004. Characterisation of soft sediments
is acknowledged for patient use of their CPT rig at the for offshore applications, Int. site characterization, ISC
2,
Ledge Point site, and COFS is thanked for the use of Proc. 2nd int. conf., Porto: 209232, Rotterdam: Millpress.
the UWA CPT rig at the Burswood site. The anony- White, D.J. 2005. A general framework for shaft resistance
mous reviewers are thanked for helping to clarify the on displacement piles in sand, Frontiers in offshore geot-
manuscript. echnics, ISFOG, Proc. int. symp, Perth.

972

Copyright 2005 Taylor & Francis Group plc, London, UK


Practice notes on push-in penetrometers for offshore geotechnical
investigation

J. Peuchen & J. Adrichem


Fugro Engineers B.V., Leidschendam, The Netherlands

P.A. Hefer
Advanced Geomechanics, Perth, Western Australia

ABSTRACT: Push-in penetrometers continue to see further development in interpretation theory and practice.
The principal driver is offshore development of increasingly deeper water and/or more complex sites. This
paper presents recent achievements and limitations taken from offshore geotechnical practice.
Particularly, examples of site-specific comparisons are presented for the cone (CPT), T-bar (TBT) and ball
penetrometers (BPT). The comparisons include TBT and BPT pore pressure monitoring.
Offshore Mini Cone Tests (MCT) use penetrometers having a cross sectional area of typically 1 cm2 or 2 cm2.
Their ease of deployment is a great advantage. This must be offset against a requirement for cautious interpret-
ation of test results, as illustrated by recent research results and practice.
Conventional practice includes a seabed frame providing seafloor support and/or reaction to a push-in device.
This is a practical yet wanting deployment solution, particularly for accurate characterisation of extremely soft
seabed soils. A discussion of practical considerations is presented, including assessment of measurement accuracy
for the upper few metres below seafloor.

1 INTRODUCTION A real world demands concessions in investing in


technology and limiting operational cost. The following
Geophysical reconnaissance dominates offshore sections explore achievements and limitations taken
ground characterisation, particularly in deeper from offshore practice.
waters. Technological progress in this area is signifi-
cant, yet ground truthing remains necessary for the
time being. Deployment of push-in penetrometer 2 CPT, TBT AND BPT
devices is often the preferred practice for ground
truthing, in comparison with sampling and laboratory The CPT has been used commercially for many years.
testing. This preference for penetrometers can largely The TBT made its commercial debut in Australia
be attributed to cost-time-benefit issues. during 1996 (Randolph et al. 1998) and has subse-
The drive for the ideal offshore penetrometer is quently gained significant offshore industry recogni-
and should be alive. This ideal measuring instrument: tion as an improved profiling tool, specifically for
deepwater soft clays. TBT geometry suits seabed-
gives unambiguous soil type identification type deployment. This limits the total depth of pene-
provides unlimited measuring resolution and tration to less than approximately 50 m below seafloor.
accuracy A miniature TBT for downhole deployment (i.e.
has a closed-form theoretical interpretation model through the drill string) was developed and used dur-
directly linked to fundamental soil mechanics ing 1998. It was found to lack the robustness required
penetrates at high speed to any desirable depth for the typically onerous offshore operating condi-
below seafloor tions. The BPT on the other hand has more suitable
allows cheap operation from a small AUV geometry for deployment down a borehole. The first
(autonomous underwater vehicle) launched at a commercial BPT was developed and trialed by Fugro
convenient coastal port. during 2003.

973

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 1. Penetrometer characteristics. Liquid limits range from about 30% to 90%, with
plasticity indices between 5% and 60%.
Tool Ap As/Ap Pore pressure filter(s) Laboratory triaxial and simple shear tests were
undertaken on undisturbed samples recovered within
CPT 15 cm2 0.59 At cone shoulder
the development area. Assuming the simple shear tests
TBT 100 cm2 0.07 At side-centre and
at side-end of T-bar axis to represent an average laboratory strength and using
BPT 48 cm2 0.05 At bottom face of sphere all the CPT results from the investigation, a cone
and at push rod correlation factor, Nk of 12 was found to provide the
immediately above best correlation to the undrained shear strength, cu
sphere (qn/Nk). Net cone resistance qn represents cone
resistance adjusted for pore pressure and total vertical
As: smallest cross-sectional area of shaft immediately stress. Since a CPT, BPT and TBT were undertaken
below load sensor; Ap: projected cross sectional area of within 17 m from each other, it is possible to derive
penetrometer.
similar correlation factors, Nt and Nb for the TBT (bar
factor) and BPT (ball factor) respectively. Undrained
shear strength profiles derived from these three tests
are shown on Figure 1. It is noted that Nt  11 and
Nb  10 provide the best fit to the CPT profile. The
bar factor falls within the bar factor range of 9.1 to
11.9 derived from a theoretical plasticity solution
(Randolph & Houlsby 1984). The ball factor falls just
outside a similarly derived range of 11 to 15.5 for a
sphere (Randolph et al. 2000). However, current and
ongoing research on penetrometers in soft soils indi-
cates that the magnitude of soil sensitivity and vis-
cous rate effects in the soil can affect the correlation
factors, sometimes resulting in these factors falling
outside the theoretical ranges. This issue will need to
be resolved as more real field data is recovered from
various geographical regions.
Pore pressure measurements in both the TBT and
BPT provide supplementary data. The pore pressure
ratio, Bq  u2/qn used for the CPT data provides a
convenient form of comparison between the three
tools. Here, u2 refers to excess pore pressure meas-
ured at the cone shoulder. In the case of the TBT and
Figure 1. Undrained shear strength.
BPT, the measured penetration resistance is used in
place of qn to arrive at equivalent Bt and Bb ratios
The BPT trial followed a detailed geotechnical respectively. However, as for the CPT, the location of
investigation of the Chinguetti Field located on the the pore pressure filter affects the manner in which
continental shelf of Mauritania in water depths of 640 m the pore pressure ratio is calculated. The TBT has
to 820 m. The geotechnical investigation included two filters mounted along the axis of the bar, one at
extensive CPT and TBT data acquisition. Tool details the centre and one at the edge. The centre filter was
are summarised in Table 1. used for calculation of Bt. Figure 2 shows that Bq and
The continental shelf off Mauritania comprises Bt compare favourably. The BPT has one filter mounted
complex geohazards, but the development area is on the bottom face of the sphere (centered) and a
located within a relatively stable zone. Geophysical second in the shaft just above the sphere. The former
results indicate parallel bedded layers. A number of was used for calculation of Bb. As for pore pressure
thin marker layers comprising coarser and denser sandy measurements at the cone face in the CPT, an adjust-
silt interspersed throughout the profile correlate with ment factor K needs to be applied to the pore pressure
the geophysical results (Fig. 1). measured at the bottom face of the sphere to allow
Soils at this site are laterally quite homogenous and comparison with a CPT filter located at the cone
predominantly comprise a mixture of calcareous shoulder. A factor of 0.3 has been used for Bb on
clayey silt and silty clay down to a depth of at least Figure 2. This value is lower than comparable values
100 m below seafloor. Particle size fractions on average for relating CPT excess pore pressure at the cone face
are 10% sand, 65% silt and 25% clay. Calcium carbon- (u1) to u2. K is believed to be a function of the soil
ate contents of between 7% and 18% were measured. characteristics such as fabric, overconsolidation ratio,

974

Copyright 2005 Taylor & Francis Group plc, London, UK


Conventional interpretation of MCT results is gen-
erally by a direct, uncorrected link to a comprehensive
database available for CPT interpretation (Lunne et al.
1997). This link may have been promoted by laboratory
research and development activities with mini cone
penetrometers, suggesting equivalence. In fact, many
of these laboratory studies used genuine offshore
penetrometers. Tentative conclusions are as follows:
Values for MCT cone resistance in clay strata are
typically within about 10% of CPT cone resistance
values (Titi et al. 2000, De Lima & Tumay 1990).
Variations in very soft clays will be greater, due to the
less accurate nature of the MCT system, as explained
further below.

It is much more difficult to make comparisons


between cone penetrometers in sands (Diepstraten
2003). This is because of the natural variability of
sands over very short distances. In general, build-
up of MCT cone resistance is faster than that for
CPT cone resistance. Experience indicates that the
MCT results allow differentiation between loose
sands and dense sands, but it is not possible to say
Figure 2. Pore pressure ratios.
how they correlate with CPT results. Scaling
effects such as proposed by De Beer (1963) may
compressibility and crushability. Kelleher & Randolph not be universal (Meave Silva 1999, Diepstraten
(2005) show a ball penetrometer with pore pressure 2003). Recent experiences could be interpreted to
measured around the horizontal circumference of the show that in strongly dilatant sands, the MCT pro-
sphere. Their results show this geometry producing a vides a significantly higher cone resistance than a
Bb profile comparable to the corresponding CPT Bq CPT and that in sands with contractive behaviour
profile. This correlation and similar comparisons may the MCT will give lower cone resistance than a
not be universal. CPT. However, insufficient research quality data
All three tools allow measurements both during the are available to be definitive. Figure 3 presents a
push-in and pull-out of the penetrometers, when one-off example for an apparently uniform sand
deployed using a continuous push-pull system. Cyclic stratum at about 4 m depth. The differences between
measurements and retraction measurements provide MCT and CPT cone resistance can partially be
additional data on soil behaviour, such as soil sensi- explained by elastic theory according to Vreugdenhil
tivity to de-structuring and large shear. et al. (1994). Here, the Vreugdenhil model predicts
MCT cone resistance being up to about 25% higher
than CPT cone resistance.
MCT signature may differ from CPT signature
3 MINI CONE TEST MCT
in ground with an effective particle size D50
exceeding about 10% of the diameter of the MCT
3.1 MCT CPT comparison
penetrometer. Individual particles rather than the
Offshore Mini Cone Tests (MCT) have been around soil mass may contribute to the measurements.
for more than 10 years (Power and Geise 1995). The loading response of a smaller probe to soil layer-
Ease of MCT deployment is a great advantage, ing is more rapid than that of a larger cone pen-
promoting use. etrometer. This relates to soil failure mechanisms
The measuring probe for a MCT is a miniature ver- in layered soils. Depending on ground conditions,
sion of the reference cone penetrometer used for the smaller probe may show higher peak qc values
Cone Penetration Tests (CPT). MCT penetrometers and lower trough values.
have a cone base area of typically 1 cm2 or 2 cm2. Soil structure may cause failure along zones of
Recent CPT standardisation documents exclude mini- weakness. A larger cone penetrometer affects a
ature versions of the CPT because of accuracy limita- larger mass of soil. There is a greater potential for
tions. Specifically, cone penetrometers with a cone soil structure effects. Thus, the larger cone pen-
base area of less than 5 cm2 (500 mm2) are excluded, etrometer may exhibit a lower cone resistance in
e.g. NNI (1996), ISSMGE (1999) and CEN (2005). structured soils.

975

Copyright 2005 Taylor & Francis Group plc, London, UK


Figure 4a. Zero drift CPT-MCT in sand.

Figure 4b. Zero drift CPT-MCT in clay.

of the zero readings of a measuring system between


Figure 3. Comparison CPT-MCT, suggesting shift in MCT
cone resistance due to heat flux. the start and completion of a cone penetration test.
The comparison is directly from commercial off-
shore practice (7 MCT projects and 3 CPT projects,
The MCT rate of penetration used by Fugro is all Fugro seabed systems), with no particular data
approximately 40 mm/s. This is twice the rate of a clean-up. Apparent data quality for the 5 cm2 cone
standard CPT and may be compared with a non- penetrometers is much better than that for the 1 cm2
dimensional parameter V where V  vd/cv, v is the penetrometers.
penetration rate, d is the probe diameter and cv is the The following issues are believed to be important
coefficient of consolidation of the soil (Randolph but tentative. The authors are not aware of specific
2004). The diameter ratio for the 10 cm2 to 1 cm2 pen- verification studies.
etrometers is about 3. The selected penetration rate
1 The MCT probe makes use of miniaturisation tech-
thus represents a partial adjustment for probe size.
nology. This implies some trade-offs for operational
This rate adjustment helps to approximate the CPT
robustness. Geometrical tolerances will not be fully
transition zone between drained and undrained
scaled down. For example, the influence of O-rings
ground behaviour for a MCT. The higher penetration
on sensor hysteresis will be larger for a miniature
rate will affect the cone resistance measurements, but
penetrometer.
with a limit of a few percent (e.g. Randolph 2004).
2 Harsh offshore conditions will induce eccentric
loading of a cone penetrometer in natural soil.
Theoretically, load sensors of well-built penetro-
3.2 MCT accuracy
meters are compensated for this effect. Practice
The estimated minimum accuracy for MCT cone may be different. In this regard it is noted that the
resistance is the larger value of qc  400 kPa or 5% ratio of bending stiffness (EI) across the friction
of the measured value. This estimate is equivalent sleeve and load sensors is about 12:1 (10 cm2 cone
to the requirements of Accuracy Class 3 for CPTs penetrometer: 1 cm2 penetrometer) for a force (F)
(ISSMGE 1999). MCT sleeve friction and hence fric- normalised to cone base area: EI10/F10  12 EI1/F1.
tion ratio are generally more variable. This is for subtraction-type penetrometers. Bending
Figures 4a and 4b provide a CPT-MCT accuracy stiffness will generally be worse for penetrometers
comparison for tests in sand and clay respectively, with concentric load cells.
based on zero drift as key performance indicator 3 Heat flux phenomena for a miniature cone pen-
(ISSMGE 1999). Zero drift is the absolute difference etrometer are possibly more severe. Figure 3 presents

976

Copyright 2005 Taylor & Francis Group plc, London, UK


some provisional evidence. Particularly, the cone Table 2. Suggested coefficients for depth accuracy
resistance for the MCT shows a significant tempor- assessment.
ary drop after penetration of the competent sand
stratum below about 4.5 m depth. Heat flux is Depth Accuracy z
reported to give an apparent shift in cone resist-
Geotechnical system a b c
ance (Post & Nebbeling 1995). For example, fric-
tion in dense sand causes a 10 cm2 cone to heat by Downhole favourable 0.4 m 0.003 0.003
about 1C/MPa cone resistance. Resulting heat Downhole adverse 1.0 m 0.005 0.004
flux decreases cone resistance by an apparent shift Seabed favourable 0.2 m 0 0.01
in the order of 100 kPa to 200 kPa for a penetrating Seabed adverse 0.8 m 0 0.02
10 cm2 probe going from dense sand into clay. This
is a temporary decrease lasting about 5 minutes. Note: Estimated resolution: 50% of accuracy.
Ambient temperature compensation systems fail to
avoid heat flux effects. Depth Accuracy z [m]
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6
0

4 SEAFLOOR SUPPORT 50

Total Depth d+z [m]


4.1 Use of seabed frame 100

Geise & Kolk (1983) point toward a futuristic ROV


(remotely operated vehicle) scenario: no requirement 150

for temporary seafloor support just let ROV or AUV


thrusters provide active reaction to penetrometer 200

push-in. This is ideal for accurate characterisation of


extremely soft seabed soils. Unfortunately, current 250

practice still differs. It includes a robust seabed frame Water Depth: Favourable 50m Favourable 200m Adverse 50m Adverse 200m

providing seafloor support and/or self-weight reac-


tion to the push-in device. Figure 5. Example depth accuracy estimate for seabed
The following two points frequently attract prac- investigations.
tice discussions:
How much does the seabed frame sink into the Table 3. Premise to estimated depth accuracy seabed
seabed or, more generally, what is the depth accur- system.
acy of encountered stratum boundaries including
seafloor? Offshore setting seabed system
What is the effect of a seabed frame on CPT data
Characteristics Favourable Adverse
measured in the upper few metres below seafloor?
Vessel Variation within Variation
horizontal 5 m of target within 5 m
4.2 Depth accuracy
position of target
The following expression for depth accuracy assess- Vessel 1 m at hook 3 m at hook
ment is proposed: heave point point
Tidal 1.5 m 3m
variation
Seafloor Firm and level Very soft seabed
soils or very
rugged seafloor
where: a  constant depth error, i.e. the sum of all Penetration Vertical at start, Inclined at
constant errors in metres; b  error dependent on verticality with correction average 5 from
water depth, i.e. the sum of all water-depth dependent for measured vertical from
errors; c  error dependent on test depth, i.e. the sum inclination seafloor to test
of all test depth dependent errors; d  water depth in depth z
metres; z  test depth in metres relative to seafloor;
z  test depth accuracy in metres (say 95% confi-
dence level). seabed testing in water depths of 50 m and 200 m.
Table 2 presents suggested coefficients for conven- Table 3 presents accompanying assumptions. Accuracy
tional downhole and seabed based testing and sampling improvements should be feasible by incorporating
systems. The seabed case is illustrated in Figure 5 for additional and/or more accurate system sensors.

977

Copyright 2005 Taylor & Francis Group plc, London, UK


Offshore definition of the seafloor (ground surface) 5 ACKNOWLEDGEMENTS
is difficult for extremely soft ground. Penetration of
the reaction equipment into a near-fluid zone of the The authors are grateful to the Joint Venture Partners
seabed may take place unnoticed. Such settlement for the Chinguetti Development for permission to use
affects measured/assumed penetration depth z. Also, certain data in this publication. The JVPs in the
settlement may continue during testing. Chinguetti development are:
Seabed frame settlement is likely to be governed Woodside Mauritania Pty Ltd, Hardman Chinguetti
by the following factors: Production Pty Ltd, Roc Oil Mautania Company,
(1) descent velocity and penetration into seabed, Premier Oil Mauritania B Limited and Mauritania
(2) non-centric loading during touchdown and testing Holdings BV.
and (3) tensioning and hysteresis forces in a heave
compensation system (4) consolidation of seabed
sediments. REFERENCES
Touchdown velocity of a seabed frame depends
on controlled winch speed and on heave of the winch Baldi, G., Bellotti, R., Ghionna, V., Jamiolkowski, M. &
on a vessel. Descent velocity implies kinetic energy Pasqualini, E. 1986. Interpretation of CPTs and CPTUs,
that can increase frame penetration into the seabed 2nd Part: Drained Penetration in Sands, Fourth Int. Geot.
in comparison with a wished-in-place scenario. In Seminar Field Instr. and In-Situ Measurements, Nanyang
this regard it may be noted that deployment of light- Technological Institute, Singapore, pp. 143156.
weight systems is often without heave compensation. Beer De, E.E. 1963. Scale Effects, Gotechnique, Vol. 13,
The larger offshore geotechnical systems use passive pp. 3975.
heave compensation, which becomes effective only CEN European Committee for Standardization 2005.
Geotechnical investigation and testing Field testing
after touchdown. The authors are not aware of use Part 1: Electrical cone and piezocone penetration tests,
of active heave compensation systems for geotech- draft European standard prEN ISO 22476-1.
nical investigation. Demands on robustness and Diepstraten, E. 2003. Comparison of Cone Resistance of
power consumption are likely factors in holding back Mini Cone and Fugro Standard Cone in Sand, MSc
development. Dissertation Technical University of Delft, pp. 113 (in
Non-centric loading during touchdown and testing Dutch).
can occur. This is usually a result of an uneven sea- Geise, J.M. & Kolk, H.J. 1983. The Use of Submersibles
floor, seawater drag on the frame and lowering for Geotechnical Investigations, Proceedings of Subtech
cable(s) caused by minor vessel movements and/ 83, The Design and Operation of Underwater Vehicles,
Paper 7.3, SUT Society for Underwater Technology,
or seawater currents. A system for frame tilt mea- London.
surement is usually available. This offers some ISSMGE International Society of Soil Mechanics and
opportunity for providing evidence about any non- Geotechnical Engineering 1999. International Reference
centric loading. Test Procedure for the Cone Penetration Test (CPT)
Tensioning of a (passive) heave compensation sys- and the Cone Penetration Test with Pore Pressure
tem can help to reduce seabed frame settlement. (CPTU), Report of the ISSMGE Technical Committee
Variations in tension due to hysteresis are typically in 16 on Ground Property Characterisation from In-situ
the order of 10% of the total load on the heave com- Testing, Proceedings of the Twelfth European Conference
pensation system. on Soil Mechanics and Geotechnical Engineering,
Amsterdam, Edited by Barends et al., Vol. 3, pp.
21952222.
Kelleher, P.J. & Randolph, M.F. 2005. Seabed Geotechnical
4.3 Soil strength Characterisation with the Portable Remotely Operated
Drill, Proceedings International Conference on Frontiers
Discussions about apparent strong overconsolidation in Offshore Geotechnics, ISFOG, Perth.
zones just below seabed, formed by consolidation of Lima De, D.C. & Tumay, M.T. 1990. Development of the
seabed sediments, arise from time to time. Puech et al. Louisiana State University Calibration Chamber System
(2005) and Puech & Colliat (2005) investigated such (LSU/CALCHAS), 4th Int. Seminar on Calibration
overconsolidation zones for extremely soft clays in Chamber Research, Grenoble France.
the Gulf of Guinea. They conclude that their origin Lunne, T., Robertson, P.K. & Powell, J.J.M. 1997. Cone
is not related to seabed frame loading. Site investiga- Penetration Testing in Geotechnical Practice, Blackie
tion measurements were found to be fully compatible Academic & Professional, London.
Meave Silva, O.R. 1999. Shallow Cone Penetration Test/
regardless of presence or absence of a frame. For Shallow Penetration Tests in Dense Saturated Sand, MSc
example, inferred CPT strengths showed excellent Thesis Delft: International Institute for Infrastructural,
correlation with laboratory strengths for undisturbed Hydraulic and Environmental Engineering.
box corer samples. The authors are aware of similar NNI Nederlands Normalisatie-Instituut 1996. Geotechnics.
findings elsewhere. Determination of the Cone Resistance and the Sleeve

978

Copyright 2005 Taylor & Francis Group plc, London, UK


Friction of Soil. Electric Cone Penetration Test, Dutch Randolph, M.F., Hefer, P.A., Geise, J.M. & Watson, P.G.
Standard NEN 5140. 1998. Improved Seabed Strength Profiling using T-bar
Post, M.L. & Nebbeling, H. 1995. Uncertainties in Cone Penetrometer, Proceedings International Conference
Penetration Testing, Proc. International Symposium on Offshore Site Investigation and Foundation Behaviour
Cone Penetration Testing CPT95, Linkping, Sweden, New Frontiers, Society for Underwater Technology,
Vol. 2, pp. 7378. London, pp. 221235.
Power, P.T. & Geise, J.M. 1995. Seascout Mini CPT System, Randolph, M.F., Martin, C.M. & Hu, Y. 2000. Limiting
International Symposium on Cone Penetration Testing, resistance of a spherical penetrometer in cohesive mater-
Linkping, Sweden, Vol. 2, pp. 7984. ial, Gotechnique, Vol. 50, No. 5, pp. 573582.
Puech, A., Borel, D., Dendani, H. & Colliat, J.L. 2005 Randolph, M.F. 2004. Characterisation of Soft Sediments
Deepwater Geotechnical Site Investigation in the Gulf of for Offshore Applications, Proceedings of the Second
Guinea: Present Practice, Proceedings International International Conference on Site Characterisation,
Conference on Frontiers in Offshore Geotechnics, Porto, Portugal, Vol. 1, pp. 209232.
ISFOG, Perth. Titi, H.H., Mohammad, L.N. & Tumay, M.T. 2000.
Puech, A. & Colliat, J.L. 2005. Some Geotechnical Miniature Cone Penetration Tests in Soft and Stiff Clays,
Specificities of Gulf of Guinea Deepwater Sediments, ASTM Geotechnical Testing Journal, Vol. 23, No. 4,
Proceedings International Conference on Frontiers in pp. 432443.
Offshore Geotechnics, ISFOG, Perth. Vreugdenhil, R., Davis, R. & Berrill, J. 1994. Interpretation
Randolph, M.F. & Houlsby, G.T. 1984. The limiting pressure of cone penetration Results in Multilayered Soils, Int. J.
on a circular pile loaded laterally in cohesive soil, for Numerical and Analytical Methods in Geomechanics,
Gotechnique, Vol. 34, No. 4, pp. 613623. Vol. 18, No. 9, pp. 585599.

979

Copyright 2005 Taylor & Francis Group plc, London, UK


Comparison of cone and T-bar factors in two onshore and one
offshore clay sediments

T. Lunne
Norwegian Geotechnical Institute, Oslo, Norway

M.F. Randolph & S.F. Chung


Centre for Offshore Foundation Systems, The University of Western Australia, Perth, Australia

K.H. Andersen & M. Sjursen


Norwegian Geotechnical Institute, Oslo, Norway

ABSTRACT: Field developments in deep waters with very soft soils have led to increased reliance on the use of
in situ tests to evaluate soil design parameters. Recently the T-bar has been introduced due to its potential for increas-
ing the reliability of interpreted undrained shear strength relative to the CPTU for soft clays in deep waters.
Empirical correlations based on field tests and laboratory tests on samples at one offshore and two onshore soft clay
sites indicate that T-bar factors are in a somewhat narrower range compared to cone factors. Recommendations are
given in terms of cone and T-bar factors to use, and when T-bar tests should be carried out in addition to CPTU.

1 INTRODUCTION resistance. The other advantage of the T-bar is that it is


a so called full-flow penetrometer, so that for compu-
Since the introduction of the cone penetration test in tation of the undrained shear strength the total vertical
offshore soil investigations in the North Sea in 1972, stress need not to be subtracted as for the CPTU cone
it has become standard practice to base soil parameters resistance. This paper presents some results of a pro-
for foundation design on CPT/CPTU results and labora- ject carried out jointly by the Norwegian Geotechnical
tory tests on obtained samples (Lunne 2001). Institute (NGI) and The Centre for Offshore Foundation
With the trend to develop fields in gradually Systems (COFS). CPTU and T-bar tests have been car-
deeper water with very soft soils, there are problems ried out at the onshore soft clay test sites in Onsy in
with obtaining accurate CPTU data due to the large Norway (NGI) and Burswood in Australia (COFS). At
readings of the sensors on the sea bottom due to hydro- both sites the results of the CPTUs and the T-bar tests
static water pressure, and the relatively small magni- have been used to compute cone and T-bar factors based
tude of the additional loads when penetrating soft soils. on correlations to undrained shear strength measured
There are also difficulties with obtaining high quality by triaxial and direct simple shear tests carried out on
undisturbed samples due to handling equipment in large high quality samples. Results from one site offshore
water depths and also due to the large stress relief when Australia have also been used in the correlations, in
samples are brought up to deck level. Damage to the addition to some laboratory model tests at NGI and
soil structure can be especially severe if there is gas centrifuge tests by COFS.
dissolved in the pore water, as the gas will come out of
solution and expand due to the large stress relief (Lunne
et al. 2001). Laboratory tests on soils that are disturbed
2 DESCRIPTION OF SITES AND REFERENCE
in this way will result in measured soil parameters that
SOIL PARAMETERS
are not representative of in situ conditions.
In situ tests that can give reliable soil parameters are
2.1 Onsy test site, Norway
therefore needed. Recently the T-bar has been intro-
duced to offshore soil investigations (Randolph et al. The onshore soft clay site at Onsy has been used by
1998, Randolph 2004). The T-bar has a larger area, NGI for more than 30 years. It is described in detail
giving higher resolution in the measured penetration by Lunne et al. (2003). Figure 1 shows a soil profile

981

Copyright 2005 Taylor & Francis Group plc, London, UK


Soil Water content (%) Unit weight Undrained St Water content (%) Unit weight Undrained St
Soil
3
(kN/m ) shear strength (kPa) fall (kN/m3) shear strength (kPa) field
Description Description
20 40 60 80 141618 20 40 60 80 cone 20 40 60 80 14 16 18 20 40 60 80 vane
0 0
Dry crust sufall cone su CAUC 4.8
suCAUC su DSS
5 5
CLAY su CAUE s field vane 4
u
4.6 3.1
suDSS

Depth (m)
Depth (m)

10 su field vane 10
5.9
4.8
4.5
15 15
Used Nkt = 10 Used Nkt = 10

20 20
Water content, w Water content, w
Plasticity index, Ip Plasticity index, Ip
2.8
3.8
25 25
NOTE : Undrained su and St are only based on test results from block samples NOTE : Undrained su and St are only based on test results from block samples

Figure 1. Soil profile Onsy test site. Figure 3. Soil profile at Laminaria site (to be completed).

Water content (%) Unit weight Undrained


Soil St
(kN/m3) shear strength (kPa)

0
Description
40 60 80 100 14 16 18 20 40 60 80 5 15 2.3 Laminaria site, offshore Australia
Crust
CLAY, The tests at this site in the Timor Sea offshore Australia
medium, plastic
5
some shell e c e
c
were carried out as part of a site investigation for field
fragments CAU (e or c)
development (Randolph et al. 1998). Figure 3 shows a
Depth (m)

and silt lenses DSS


10 UU
Vane
soil profile at Laminaria. The reference CAUC and
DSS tests were carried out on 76 mm diameter piston
15
e c
c
tube samples, whose quality will be less good than the
e

20
block samples at Onsy. CAUE tests were not carried
Water content, w
Plasticity index, Ip
out on the Laminaria samples.
25

Figure 2. Soil profile Burswood test site. 3 EQUIPMENT AND PROCEDURES

close to the location where the CPTUs and T-bar tests The cone penetration tests have been carried out with
described in this paper have been carried out. The ref- equipment and procedures according to the Inter-
erence undrained shear strength parameters included national Reference Test Procedure (IRTP) published
in Figure 1 have been obtained by CAU (anisotropic- by the International Society of Soil Mechanics and
ally consolidated undrained triaxial) tests sheared in Geomechanical Engineering (ISSMGE 1999). In par-
compression (CAUC) and in extension (CAUE) and ticular the tests at NGI have been done with the ENVI
DSS (direct simple shear) tests carried out on high memocone, the tests at Burswood with the Hogentogler
quality samples taken with the Canadian Sherbrooke cone penetrometer and the tests at the Laminaria site
block sampler (Lefebvre & Poulin 1979). Reconstituted with the Fugro cone penetrometer. Pore pressure has
Onsy clay has also been used for some laboratory been measured at the location just behind the cone,
model tests where small-scale T-bar tests have been car- the so called u2 position, for the Onsy and Burswood
ried out. This will be further discussed in Section 5. tests. For the Laminaria tests pore pressures were
measured on the cone tip, u1.
The measured cone resistance has been corrected for
2.2 Burswood test site, Australia
the effects of pore pressure as required in the IRTP. The
An onshore soft clay site at Burswood (Perth) is used results below are reported in terms of the measured sleeve
by the University of Western Australia for research pur- friction, fs, the net cone resistance, qnet and the excess
poses. Figure 2 shows a soil profile at the location pore pressure, u.
where the present CPTUs and T-bar tests were carried At present there is no international standard for the
out. The reference CAUC, CAUE and DSS tests shown T-bar test. One conclusion of the work in the JIP on
in Figure 2 have been carried out on tube samples (71 which this paper is based, is the recommendation to
and 104 mm dia.). It should be borne in mind that the use a T-bar 40 mm in diameter and 250 mm long, thus
quality of these samples is not as good as those from the giving a projected area 10 times that of the standard
block sampler used at Onsy. Reconstituted Burswood cone. The penetration and extraction rate should be
clay has also been used for centrifuge tests carried out the same as for the CPTU, i.e. 20 mm/s. Normally the
at by COFS, with model CPTs and T-bar tests carried cone part of the penetrometer is replaced with the
out on the centrifuge samples as discussed in Section 5. T-bar so that the same load cell as used to measure

982

Copyright 2005 Taylor & Francis Group plc, London, UK


qT-bar (kPa)
-400 -200 0 200 400 600
0

10

Depth (m)
15
Figure 4. Picture of cone penetrometer and T-bar used at
Onsy.

20
qnet & (kPa) Individual tests
Average
0 200 400 600 800
0
25

Figure 6. Results of T-bar tests carried out at Onsy.


5
4 RESULTS IN SITU TESTS
Individual tests
Average
4.1 Onsy
10 Figure 5 shows the results of the five ENVI memocone
Depth (m)

qnet CPTUs carried out at the Onsy site. The individual val-
ues of sleeve friction, fs, corrected cone resistance, qt,
and penetration pore pressure, u2, from each test are
15 shown in shaded form, while the average profiles are
shown in bold. Figure 6 shows the results of altogether
four T-bar profiles carried out using the ENVI memo-
cone. Again the individual test results are shown shaded
20 and the overall average in bold. Note that the resistance
during extraction of the T-bar has also been measured,
Friction although these values are not discussed further in this
paper.
25
0 20 40 60 80 4.2 Burswood
Friction (kPa) Figure 7 shows the results of two CPTU profiles in
terms of fs, qt and u2. CPTU 1 is 25 m from the sample
Figure 5. Results of CPTUs carried out at Onsy.
boring and the T-bar tests, while CPTU 2 is close to the
sample borehole. Therefore the results of CPTU 2 have
been used in the analyses described in the next section.
cone resistance for the CPTU is used to measure the Figure 8 shows the results of 4 T-bar tests in shaded
T-bar resistance (Fig. 4). form and the calculated average profile. Again the
The project has also given recommendations on extraction profiles have been included. The reductions
cyclic T-bar testing that can be used to get information in extraction resistance in two of the profiles, at depths
on the remoulded shear strength (see Randolph 2004). of 4, 9 and 14 m, are due to cyclic T-bar tests having
However, the present paper does not include results been carried out. These have been excluded when cal-
from this part of the programme. culating the average profile.

983

Copyright 2005 Taylor & Francis Group plc, London, UK


qnet & u (kPa) qnet & u (kPa)
0 200 400 600 800 0 200 400 600 800
0

Individual tests
Average
0 5
Cone 1
Cone 2
4
10

Depth (m)
u
Depth (m)

qnet
8 qnet u
15

12 Friction

20
16 Friction

20 25
0 20 40 60 80 0 20 40 60 80
Friction (kPa) Friction (kPa)

Figure 9. Results of CPTUs carried out at Laminaria.


Figure 7. Results of CPTUs carried out at Burswood.

qT-bar (kPa)
qT-bar (kPa)
-400 -200 0 200 400 600
0 -400 -200 0 200 400 600
0

5
5

10
10
Depth (m)

Depth (m)

15 15

20 20
Individual tests Individual tests
Average
Average
25 25

Figure 8. Results of T-bar tests carried out at Burswood. Figure 10. Results of T-bar tests carried out at Laminaria.

984

Copyright 2005 Taylor & Francis Group plc, London, UK


qT-bar /qnet u /qnet
0.0 0.5 1.0 1.5 0.0 0.5 1.0
0.0 0.0

Onsy
Burswood
Laminaria
5.0 5.0

10.0

Depth (m)
10.0
Depth (m)

Onsy 15.0
15.0 Burswood
Laminaria

20.0
20.0

25.0
25.0
Figure 13. u/qnet vs depth.
Figure 11. qT-bar/qnet vs depth.
4.4 Comparison of CPTU and T-bar results at the
three sites
qT-bar /u
0 1 2 3 Figures 11 to 13 give for all three sites the ratios of the
0.0 average values of qT-bar and qnet, qT-bar and u (from
CPTU), and finally u and qnet, where qnet 
qt
vo. It is interesting to note that qT-bar/qnet tends to
decrease with depth for Burswood and Laminaria,
5.0 whereas for Onsy this ratio is about constant with
depth. The pore pressure ratio, Bq  u/qnet, profiles
are more or less constant for all three sites below a
depth of about 7 m, and so the ratio qT-bar/u follows
10.0 the same trends as qT-bar/qnet. The ratios in Figures 11
Depth (m)

to 13 may potentially be used to evaluate differences in


soil properties such as: OCR, rigidity index and
strength anisotropy. This is being explored further in
15.0
a joint ongoing study by NGI and COFS.

5 CORRELATIONS
20.0
Laminaria
For the CPTU the undrained shear strength, su, can be
Burswood
Onsy computed either from the net cone resistance or from
the excess penetration pore pressure using the follow-
25.0
ing formulas:
Figure 12. qT-bar/u vs depth. (1)
where vo is the total vertical stress and Nkt is a cone
4.3 Laminaria factor
Figure 9 shows two CPTU fs, qt and u1 (measured on (2)
cone tip) profiles and the average of the two profiles.
Figure 10 shows the results of two individual T-bar where uo is the in situ static pore pressure and Nu is
profiles and the computed average. another cone factor.

985

Copyright 2005 Taylor & Francis Group plc, London, UK


Even though the above two expressions are theor- 3 The range, or scatter, in T-bar and cone factors based
etically based, there are so many simplifications and on su,av is generally slightly smaller than when based
assumptions involved in these theories that it is com- on su,CAUC.
monly accepted that the cone factors have to be deter-
Other work in the JIP project (NGI/COFS 2002)
mined empirically. Numerous correlation studies have
can also shed light on the T-bar factors. In connection
been carried out in the past giving quite a range in the
with some pipeline-seabed model tests at NGI, T-bar
cone factors (Lunne et al. 1997). One major issue is
tests were carried out using a small T-bar (20 mm in dia;
which value of su to use. In the past NGI has mostly
used the undrained shear strength determined by a
CAUC (anisotropically consolidated undrained) triaxial Nkt
test sheared in compression, suCAUC. In the following 0 5 10 15 20
the average undrained shear strength from a CAUC and 0
a CAUE (sheared in extension) suCAUE and a direct Block Onsy
Burswood
simple shear test, sDSS
u have also been used. The average Laminaria
su has been denoted su,av  (suCAUE  suCAUE  suDSS)/3.
5
As mentioned in the introduction the T-bar is a full-
flow penetrometer so that it is not necessary to subtract
the total vertical stress when computing su, thus the

Depth (m)
T-bar factor NT-bar becomes: 10

(3)
15
Theoretical work done in the present project has
shown that although theories exist for how to interpret
the T-bar in terms of undrained shear strength param-
eters, there are many assumptions that are required in 20
the analyses (Randolph & Andersen 2005).
As such empirical correlations for the NT-bar factor NKt1 = (qt - v0)/su, CAUC
NKt2 = (qt - v0)/su,av
are still required at present, just as for the cone pen- Laminaria results su, av = su, DSS
etrometer. However, the simplicity of the flow mech- 25
anism around the T-bar offers the potential for analytical Figure 14. Nkt vs depth.
solutions that take account of secondary soil charac-
teristics such as strain rate dependency, strain soften-
ing and so forth. The remarks about the su value to use Nkt
in the correlations are valid for the T-bar also, although 0 5 10 15 20
the symmetry of the failure mechanism makes it more 0
logical to correlate the T-bar resistance with the aver- Onsy Block
Laminaria
age strength, su,av.
Burswood
Figures 14, 15 and 16 show computed values of
Nkt, Nu and NT-bar vs depth for Onsy, Burswood and 5
Laminaria clays respectively. The open symbols refer
to suCAUC and the filled symbols to su,av.
Tables 1 and 2 summarize the cone and T-bar factors
10
shown in Figures 14 to16. The number of laboratory
Depth (m)

tests at each of the three sites are still quite limited, there-
fore the range of the factors have been given instead of
the standard deviation. Although the number of values is 15
relatively small there are some trends that are emerging:
1 There is significant scatter in both the Nkt, Nu and
NT-bar values for the individual data points both within
each site and between the sites. The range in average 20
values for the various sites is, however, smaller. Nu1 = (u2 - u0)/su,CAUC
2 The range in the average NT-bar factors for the vari- Nu2 = (u2 - u0)/su,av
ous sites is somewhat smaller than for the cone fac- Laminaria results su,av = su,DSS
25
tors Nkt and Nu, implying that NT-bar may vary less
for different sites than Nkt and Nu. Figure 15. Nu vs depth.

986

Copyright 2005 Taylor & Francis Group plc, London, UK


125 mm long). These tests have been done on reconsti- ature T-bar tests (5 mm in dia; 20 mm long).
tuted Onsy clay, and also on reconstituted clay from Reference shear strengths have been determined by
Watchet Harbour (UK). Reference shear strengths have laboratory DSS tests.
been evaluated from plate load tests (Onsy clay) and Table 4 summarises the NT-bar factors evaluated
UU triaxial tests (Watchet Harbour clay). from the NGI model tests and the COFS centrifuge
Classification parameters for the Watchet Harbour tests. For the tests on reconstituted Onsy clay the
and Onsy clays are compared in Table 3. shear strength has been corrected to be valid for su,av
In addition COFS has carried out centrifuge tests as defined above.
on reconstituted Burswood clay with in flight mini- The NT-bar factors listed in Table 4 are within the range
of values obtained from the field tests based on su,av
NT-bar (Table 2) of 9.5 to 14.3 and give some confidence that
0 4 8 12 16 the T-bar factors are also applicable for shallow soils.
0
Burswood
Onsy 6 RECOMMENDATIONS
Laminaria
The findings in the previous chapter have been used
5
to arrive at the recommended values given in Table 5.
For the CPTU it is recommended that both the net
cone resistance, qt
vo, and excess pore pressure
10
Table 3. Soil classification and index parameters for
Depth (m)

reconstituted Watchet Harbour (WH) and Onsy clays.

Parameter WH clay Onsy clay


15
Fines content 35% 40%
(% 0.002 mm)
Plasticity index 42% 30%
20 Water content 69 to 73% 60 to 65%
Sensitivity 2.5 3,4
NT-bar1 = qT-bar /su, CAUC (from fall cone)
NT-bar2 = qT-bar /su, av Shear strength at 3.2 kPa 1.5 kPa
Laminaria results su, av = su, DSS clay surface
25 Shear strength at 6.0 kPa 4 kPa
200 mm depth
Figure 16. NT-bar vs depth.

Table 1. Summary of cone and T-bar factors based on su,CAUC.

Nkt Nu NT-bar

Site Avg Range Avg Range Avg Range

Onsy 11.7 10.313.2 8.6 7.39.5 8.6 7.89.9


Burswood 8.6 8.09.7 5.3 5.05.7 9.2 7.211.7
Laminaria 11.3 10.012.5 N/A N/A 10.2 8.812.1
Overall 10.5 8.013.2 7.0 5.09.5 9.3 7.212.1

Table 2. Summary of cone and T-bar factors based on su,av.

Nkt Nu NT-bar

Site Avg Range Avg Range Avg Range

Onsy 16.4 15.617.5 11.9 11.012.6 12.0 11.013.1


Burswood 11.5 11.012.6 7.0 6.77.3 10.6 9.511.6
Laminaria 13.7 12.914.8 N/A N/A 12.4 11.614.3
Overall 13.9 11.017.5 9.5 6.712.6 11.7 9.514.3

987

Copyright 2005 Taylor & Francis Group plc, London, UK


should be used to compute either su,CAUC or su,av. sampling and laboratory tests, and local correlations
Local experience may show that one cone factor works be developed.
better for a given clay. Whenever possible the undrained A general comment for both the CPTU and T-bar
shear strength values should be supplemented with factors is that when it is conservative to have a low
value of su, like in bearing capacity calculations, then
Table 4. NT-bar factors from model tests. the upper values in the ranges in Table 5 should be
used. When it is conservative to have a high undrained
Case NT-bar avg NT-bar range shear strength, like in skirt penetration analyses, the
lower values given in Table 5 should be used.
NGI tests, 11.9 11.013.4 It should be emphasised that the recommended
Onsy clay
CPTU and T-bar factors given in Table 5 have been
NGI tests, WH clay 13.0
Centrifuge tests 10.2 9.710.7 based on experience from a rather limited range of
(Burswood) clays, with values of soil plasticity varying from 33
to 45%, and OCR 1.8 to 1.3. There is clearly a need to
expand the basis for the correlations, especially
Table 5. Recommended CPTU and cone factors. to cover high plasticity clays like those encountered
offshore Africa.
Undrained
In situ Empirical shear Recommended
The CPTU should be the basic in situ test for off-
test factor strength, kPa range shore soil investigations, because of its excellent pro-
filing capability and because of the large amount of
CPTU Nkt su,CAUC 913 experience for interpretation of soil type and a range
su,av 1217 of soil parameters.
Nu su,CAUC 69 However, when it is important to have as reliable
su,av 712.5 undrained shear strength values as possible it is rec-
T-bar NT-bar su,CAUC 811 ommended to also carry out T-bar tests in addition to
su,av 1013 CPTU as summarized in Table 6 below.

Table 6. Applicability of interpreted soil parameters.

Applicability1

Geotechnical problem Depth range (m) Soil parameters required CPTU T-bar

Backfilled trenches: 01 Soil profile 12 3


upheaval buckling2 Classification 2
Soil density 23
Undrained strength 23 12
Pipeline and riser 03 Soil profile 12 3
interaction with soil3 Classification 2
Undrained strength 2 12
Remoulded strength 4 124
Skirted foundations: 015 or 405 Soil profile 12 3
penetration, capacity Classification 2
Undrained strength 2 12
Remoulded strength 4 124
Seabed templates: 010 Soil profile 12 3
penetration, stability, Classification 2
settlements Undrained strength 2 12
Remoulded strength 5 124
Settlements (34)6 (
)6
Geohazards Slope 010 or 1005,7 Soil profile 12 3
stability Classification 2
Undrained strength 2 12
Remoulded strength 34 124

Notes:
1. Authors views on relative applicability of tool, with 4. Requires cyclic T-bar tests (not considered here).
scale: 1 High; 5 Very low; not applicable. 5. Current interpretation only covered to 30 m depth.
2. Extremely soft soil may be encountered. 6. Assessment of settlement parameters not covered here.
3. Very soft soil may be encountered. 7. T-bar may be limited to 40 m depth for seabed systems.

988

Copyright 2005 Taylor & Francis Group plc, London, UK


7 SUMMARY AND CONCLUSIONS Watchet Harbor clay tests were made available from
the STRIDE JIP managed by 2H Offshore.
CPTU and T-bar tests have been carried out at NGIs and
COFS onshore soft clay test sites in Onsy, Norway,
and Burswood, Australia. Triaxial and direct simple REFERENCES
shear tests have been carried out on high quality sam-
ples, with an average undrained shear strength calcu- ISSMGE 1999. ISSMGE Technical Committee TC16 Ground
lated as the arithmetic mean of triaxial compression, Property Characterisation from In-situ Testing (1999) Inter-
direct simple shear and triaxial extension strengths. national Reference Test Procedure (IRTP) for the Cone
Results of similar tests have also been made Penetration Test (CPT) and the Cone Penetration Test with
available from the Laminaria soft clay site offshore pore pressure (CPTU). Proc. XIIth ECSMGE Amsterdam.
Australia, except that CAUE tests were not carried out Balkema. pp. 21952222.
so that the average strength was taken as the simple Lefebvre, G. & Poulin, C. 1979. A new method of sampling
shear strength. in sensitive clay. Canadian Geotechnical Journal, Vol. 16.
Lunne, T. 2001. In situ testing in offshore geotechnical investi-
Correlations studies have been carried out whereby gations. Proc. Int. Conf. on In Situ Measurements of Soil
the cone factors Nkt and Nu and the T-bar factor Properties and Case Histories, Bali, pp. 6181.
NT-bar have been computed based on both su,CAUC and Lunne, T., Robertson, P.K. & Powell, J.J.M. 1997. Cone pene-
su,av. It was found that T-bar factors lay in a somewhat tration Testing in Geotechnical Practice. Spon Press.
narrower range compared to cone factors. Provisional London. 312 pages.
recommendations are given in terms of what cone and Lunne, T., Berre, T., Andersen, K.H. & Tjelta, T.I. 2001.
T-bar factors to use, and when T-bar tests should be Deepwater sample disturbance due to stress relief. Proc.
carried out in addition to CPTUs. OTRC Int. Conf. On Geotechnical, Geological and Geo-
It is also recommended that there is a need to develop physical Properties of Deepwater Sediments, OTRC,
pp. 6485.
similar correlations also for other soils, especially the Lunne, T., Long, M. & Forsberg, C.F. 2003. Characteristics and
high plasticity clays as are found offshore Africa. engineering properties of Onsy clay. Proc. Workshop on
Characterisation and Engineering Properties of Natural
Soils, November 2002, Vol. 1, pp. 395428.
Norwegian Geotechnical Institute/Centre for Offshore
ACKNOWLEDGEMENT Foundation Systems. 2002. Characterisation of Soft Soils
In Deep Waters by In Situ Tests. JIP Report 20011026-4,
The authors would like to thank the participants in the dated 18 December 2002.
joint industry project within which the work described Randolph, M.F. 2004. Characterisation of soft sediments for
herein has been carried out: Statoil, BP Exploration offshore applications. Keynote Lecture in Proc. 2nd Inter-
Operating Company, Woodside Engineering, Norsk national Conf. on Site Characteristion, Porto, Sept. 2004,
Hydro, Petroleo Brasileiro and ConocoPhillips. NGIs pp. 209232.
participation has also been supported by the Research Randolph, M.F. & Andersen, K.H. 2005. Numerical analyses
of T-bar penetration in soft clay. Paper in preparation.
Council of Norway, while COFS has been supported Randolph, M.F., Hefer, P.A., Geise, J.M. & Watson, P.G.
under the Australian Research Councils Research 1998. Improved seabed strength profiling using T-bar
Centres Program. The data from the model testing on penetrometer. Proc. Int. Conf. Offshore Site Invest. and
resedimented Onsy clay in NGIs laboratory was Foundation Behaviour- New Frontiers, Society for
made available through the CARISIMA projects. The Underwater Technology, London, pp. 221235.

989

Copyright 2005 Taylor & Francis Group plc, London, UK


Considerations in evaluating the remoulded undrained shear strength
from full flow penetrometer cycling

N.J. Yafrate & J.T. DeJong


University of Massachusetts, Amherst, Massachusetts, United States

ABSTRACT: In recent years full-flow probes, such as the Ball and T-bar penetrometers, have been inves-
tigated as alternative in situ tests to the peizocone (CPTU) and the vane shear test (VST). These penetrom-
eters do not require correction for overburden pressure and have potential to provide both undrained and
remoulded undrained shear strength measurements in a timely manner. This paper summarizes a portion of
a research program that is utilizing a number of highly characterized test sites to comprehensively evaluate
aspects of the full flow penetrometers. The results presented herein are used to consider practical aspects
related to evaluating the remoulded strength from full flow penetrometers. Results from three test sites, which
range in sensitivity from about 4 to 100, are used for the analysis. It is shown that the number of cycles
required to fully remould the soil is less than 10, decreasing as the sensitivity of the soil increases. The cycling
interval distance required to establish steady state full flow conditions around a penetrometer and to remold
the soil in a consistent manner is about 5 and 10 cm for the T-bar and Ball penetrometers, respectively. This
corresponds to a distance slightly larger than the probe diameter. Incorporating practical implementation con-
siderations, a 20 cm cycling interval would be preferable. Cycling at this interval for a limited number of
cycles (up to 10) would enable measurement of the remoulded strength with substantial saving in time and
money relative to the VST.

1 INTRODUCTION typical hourly rate for offshore site investigations, the


opportunity to evaluate the remoulded strength via
The undrained and remoulded strength profile of soft cycling of full-flow penetrometers is attractive given the
clay deposits is often required for the design of offshore high potential cost savings over the VST procedure.
foundation systems. Due to the expense of sampling As the practice moves toward the use of full flow
and laboratory testing various in situ methods are probes it is necessary to investigate their implementa-
employed to provide strength profiles of the deposit. tion to determine the most appropriate testing methods.
The piezocone (CPTU) is often employed to estimate Variations in cycling distance, penetrometer type and
the undrained shear strength profile while the vane penetration rate generate different results. This paper
shear test (VST) is used to obtain undrained and addresses aspects related to evaluation of the remoulded
remoulded strength values at discrete depths. Unfor- shear strength via penetrometer cycling through exam-
tunately both devices require large geometric and/or ining tests performed at three highly characterized
empirical corrections that are inherent to device design test sites.
and/or testing method and whose effect is magnified
when performing tests in deep water. In addition, the
downhole VST is expensive since about one hour is 2 FULL FLOW PENETROMETERS
required to obtain a remoulded strength measurement
and the rate for offshore site investigations is about Full full penetrometers provide a measure of the soils
US$1050 k/hr. resistance to flow around the penetrometer. This is akin
In recent years full-flow probes, such as the Ball and to a viscosity measurement. This mechanism differs
T-bar penetrometers, have been investigated as alterna- significantly from the CPTU where the soil is fully
tive in situ tests. They do not require correction for over- displaced. As a result, many of the correction factors
burden pressure and have potential to provide both for stresses resulting from overburden and pore water
undrained and remoulded undrained shear strength pressure associated with full-displacement penetro-
measurements in a timely manner. Considering the meters such as the CPTU are not necessary.

991

Copyright 2005 Taylor & Francis Group plc, London, UK


influence
zone
influence zone

influence zone

remoulded
zone

remoulded
zone
Figure 2. Image of Plate, T-bar, and Ball full-flow
(a) (b) penetrometers.

Figure 1. Schematic of (a) continuous penetration to deter- 10 Hz provided measurements of penetration resist-
mine undrained shear strength and (b) interval cycling to ance every 2 mm.
determine remoulded shear strength.

Consequently, the penetrometer resistance measured 3.2 Test sites


during initial monotonic penetration (Figure 1a) can Full flow T-bar and Ball penetrometers were used to
be related directly to the undrained shear strength via investigate the measurement of remoulded undrained
shear strength at three highly characterized soft clay
(1) sites: Onsoy, Norway, and Louiseville, Quebec and
Gloucester, Ottawa Canada. The test sites were chosen
where q is the penetration resistance measured during to include a large spectrum of undrained shear strengths
penetration and N is the factor designated for the full- and sensitivities. Onsoy clay is a highly uniform, nor-
flow penetrometer. (NT corresponds to the T-bar, etc.) mally to lightly overconsolidated high plasticity marine
The N values are currently calibrated based on labora- clay. VST strength is approximately 10 kPa near the
tory tests or VST data from a site although analytical surface and increases to 25 kPa at 15 m (Lunne et al.,
analyses indicate a constrained range of N values that 2003). Louiseville clay is a firm high plasticity clay
is less than Nkt values for the CPTU (Randolph 2004). with undrained shear strength increasing from appro-
The remoulded undrained shear strength, sur, can ximately 30 kPa near the surface to 60 at 14 m. The
be determined by cycling the penetrometer about the overconsolidation ratio decreases from approximately
desired depth (Figure 1b). The soil in the cycling zone is 5 at the surface to 2 at 14 m (Leroueil et al., 2003).
destructured as the penetrometer is cycled. Once the Gloucester clay is a soft highly sensitive clay with
penetration resistance reaches a steady state the soil undrained shear strength of approximately 10 kPa near
has been remoulded and the clay sensitivity may then the surface to 50 kPa at 18 m (Lo et al., 1976). The
be evaluated via sensitivity values which range from approximately
(2) 48 in Onsoy, 22 in Louiseville, and up to 100 in
Gloucester. As documented elsewhere, the sensitivity
value is dependent to some extent of the test method
3 FULL FLOW PENETROMETERS, TEST used. A summary of test site properties is provided
SITES & TESTING PROGRAM in Table 1. Additionally, most cyclic testing was per-
formed at depths where the natural water content is
3.1 Full flow penetrometers greater than the liquid limit.
The T-bar and Ball penetrometers used in this study
3.3 Testing program
each have a projected area of 100 cm2 and a ratio of
projected area to push rod projected area of 10 to 1 The full flow testing program focused on determining
(Figure 2). A penetration rate of 2 cm/s was maintained the effects of cycling distance and cycle number on
through out the testing sequence. Data acquisition at measured remoulded undrained shear strength.

992

Copyright 2005 Taylor & Francis Group plc, London, UK


Table 1. Summary of select test site properties. penetration and extraction data, the gradual smooth-
ing of features with cycling that reflects soil remould-
Onsoy, Louiseville, Gloucester, ing, and separately by features detected in pore pressure
Test site Norway Canada Canada profiles obtained in miniature piezoprobe soundings.
Although the diameter of full flow penetrometers is
Depth Interval 15.016.0 m 11.512.5 m 8.09.0 m
Wn (%) 62 64 85 larger than the CPTU, they are not insensitive to minor
WL (%) 60 65 59 stratigraphic changes.
IL 0.75 1 2 For the Onsoy clay cycling the resistance during
su(FVT) (kPa) 27 50 34 penetration and extraction in later cycles is nearly
St(FVT) 68 n/a 50100 constant during each stroke and the absolute resist-
St(fallcone) 46 22 88 ances are very similar. There is a slight increase in the
absolute resistance during a stroke and a slight bias in
absolute magnitude of resistance towards the penetra-
Cycling interval distance was evaluated by pen- tion stroke. Generally these biases are more pronounced
etrometer cycling over depth intervals of 5, 10, 20, in the Ball data. The Gloucester and Louiseville data
50, and 100 cm. Cycling was performed until the pen- exhibit similar and stronger trends. Although the pos-
etrometer resistance reached a steady value, reflect- itive bias is not always observed, these trends are
ing complete remoulding of the soil. relatively consistent. The cause and/or mechanisms
generating these differences is currently unknown and
under further examination. It is plausible that instru-
4 CHARACTERISTIC CYCLING RESPONSE mentation performance, changes in flow conditions,
changes in flow patterns resulting from the drill string,
Cycling was performed at two 1 m depth intervals at and/or soil properties may contribute to these changes.
each site with the Ball and T-bar penetrometers. Results Analytical and numerical analyses to date do not exhibit
from the lower cycling interval are presented here as or predict these biases.
the OCR between sites was similar at depth (about
1 to 2) although values and trends between the two
intervals were consistent. Figure 3a and b present the 5 INFLUENCE OF CYCLE NUMBER
penetration resistance during cycling for both pen-
etrometers at the Onsoy and Gloucester test sites, The number of cycles required to reach a fully
respectively, representing sensitivity extremes of 5 and remoulded soil condition was examined through deter-
88 (Table 1, based on fall cone). The Louiseville data mining the mean resistance at the mid-location of each
contained similar trends. It is noted that throughout penetration and extraction stroke. The penetration
the paper analysis is performed using the measured stroke was considered the first half of the cycle and
resistance to penetrometer penetration and extraction. the extraction stroke considered the second half of a
The conversion to undrained strength values is sensitive cycle. The absolute resistance of each half cycle for
to the benchmark data, and presentation in this format, all three test sites for the T-bar and Ball penetrometers
although possible, would only complicate addressing is presented in Figure 4. The data is also presented in
the objectives of the paper. Corresponding strength val- a normalized format by dividing the resistance by the
ues could be roughly approximated using an N value resistance measured during the initial penetration.
of about 10.5 although it may be somewhat site specific This enables comparison between sites and compari-
and may differ depending on the desired measurement. son at different depths within the same profile.
The initial penetration resistance is relatively con- It is important to recognize that the penetrometer
stant with minimal variability with depth for the resistance measured during the initial penetration is
Onsoy site (Figure 3a). In general, the penetration not a direct measure of undrained strength. Rather,
resistance is about 10% larger for the Ball than the it is an average measurement of the undrained resist-
T-bar during initial penetration, which would result in ance of the soil within the full flow zone. An element
a slightly higher N value for the Ball for estimating of soil undergoes undrained shear as it comes into the
undrained strength. The difference is 4% and 17% for full flow mechanism, after which the element under-
the Gloucester and Louiseville cycling, respectively, goes high strains that results in a strain-softened, par-
indicating consistently higher resistance to a varying tially remoulded strength state as the soil element
degree with the Ball. This is consistent with previous exits the flow mechanism above the penetrometer. The
observations (DeJong et al., 2004, Randolph 2004). strain an element of soil undergoes as it passes through
The penetration profile for Gloucester (Figure 3b) a full flow mechanisms varies depending on location,
appears much more erratic but actually reflects strati- but average 400% and 225% for the T-bar and
graphic variations formed during deposition. This is Ball penetrometers, respectively, according to analyt-
confirmed by the symmetry between features in the ical studies (Randolph 2004).

993

Copyright 2005 Taylor & Francis Group plc, London, UK


(a) Onsoy, Norway
1m cycle, 15-16 m depth
T-bar Ball
14.6 14.6

14.8 14.8

15.0 15.0
Depth (m)

Depth (m)
15.2 15.2

15.4 15.4

15.6 15.6

15.8 15.8
-300 -200 -100 0 100 200 300 400 -300 -200 -100 0 100 200 300 400
Penetration Resistance (kPa) Penetration Resistance (kPa)

(b) Gloucester, Canada


1m cycle, 8-9 m depth
T-bar Ball
7.6 7.6

7.8 7.8

8.0 8.0
Depth (m)

Depth (m)

8.2 8.2

8.4 8.4

8.6 8.6

8.8 8.8
-100 0 100 200 300 -100 0 100 200 300
Penetration Resistance (kPa) Penetration Resistance (kPa)
(c) Gloucester, Canada
0.1 m cycle, 8.25-8.35 m depth
T-bar Ball
6.26 6.24

6.28 6.26

6.30 6.28
Depth (m)

Depth (m)

6.32 6.30

6.34 6.32

6.36 6.34

6.38 6.36
-200 -100 0 100 200 300 -200 -100 0 100 200 300
Penetration Resistance (kPa) Penetration Resistance (kPa)

Figure 3. Example cycling data for T-bar and Ball penetrometers over various cycling distances at two test sites.

994

Copyright 2005 Taylor & Francis Group plc, London, UK


(a) T-bar, 1 m cycle
500 100

% of Initial Penetration Resistance


Penetrometer Resistance (kPa)

Onsoy (15-16m)
400 Louiseville (11.5-12.5m) 80
Gloucester (8-9m)
300 60

200 40

100 20

0 0
0 2 4 6 8 10 12 0 2 4 6 8 10 12
Cycle # Cycle #

(b) Ball, 1 m cycle


500 100
% of Initial Penetration Resistance
Penetrometer Resistance (kPa)

400 80

300 60

200 40

100 20

0 0
0 2 4 6 8 10 12 0 2 4 6 8 10 12
Cycle # Cycle #

Figure 4. Degradation of soil during cycling plotted as cycle # versus (a) absolute penetrometer resistance and (b) percent
of penetrometer resistance during initial penetration. Note that resistance during penetrometer penel is presented at 1/2 cycle
# values and resistance during penetrometer extraction presented as cycle #s.

Rapid degradation of penetrometer resistance is evi- cycles for the Onsoy and Louiseville clays, respect-
dent in the plots of penetrometer resistance and nor- ively. As highlighted by the sensitive Goucester clay,
malized resistance in Figure 4. For all sites rapid the cycles required to reach a remoulded state appears
degradation occurs within early cycles and the pen- to decrease with increasing sensitivity. However, the
etrometer resistance continues to gradually decrease number of cycles required is dependent on the rate of
in an asymptotic manner with cycling. Conceptually, decay from undrained to remoulded strength, a factor
the penetrometer resistance would stabilize, as do that is related to, but not equivalent to, sensitivity. Add-
laboratory fall cone trials, once full remoulding has itional research is underway to separate the influence
occurred. For this study it is assumed that a remoulded of these two factors.
state has been reached once the variability of cycling An estimate of the sensitivity that would be predicted
resistance between cycles becomes comparable with by the full flow penetrometers (assuming a constant N
the rate of degradation. value) can be obtained from the inverse of the percent
Degradation occurs within three and four cycles of initial penetration resistance at the cycle number
for the T-bar and Ball penetrometers, respectively, for where a remoulded condition is established. Using the
Gloucester clay, the most sensitive soil. Degradation cycle numbers identified above, the sensitivity values
to a remoulded condition occurs within about 7 and 9 estimated for the T-bar and Ball for the Gloucester,

995

Copyright 2005 Taylor & Francis Group plc, London, UK


Louiseville, and Onsoy test sites are about 14, 6, and 3, To investigate this, cycling tests were performed using
respectively. The sensitivity values agree reasonably the T-bar and Ball penetrometers at the Louiseville
well for the lower sensitivity soils, which primarily exist and Gloucester test sites. In one sounding cycle inter-
offshore. For high sensitivity soils it is probable that vals of 100, 50, 20, 10, and 5 cm were performed
the degree of softening that occurs during initial pene- sequentially with depth. Plots of 100 and 10 cm cycling
tration is substantial and this results in an average at Gloucester are in Figure 3b and c.
measured resistance during initial penetration that is The degradation curves from the T-bar and Ball
significantly lower than the undrained soil strength. cycling intervals performed in Gloucester clay are
presented in Figure 5. Evidence of a general increase
in penetration resistance, which corresponds to an
6 INFLUENCE OF CYCLING INTERVAL increase in undrained shear strength, with depth is
present. With cycling the penetrometer resistance
With the number of cycles required to establish generally converges, which would be expected for the
remoulded conditions known, it is necessary to remoulded strength, although the resistance appears
determine the cycling interval distance required to to be somewhat dependent on the stress level (resist-
establish full flow conditions within a cycling stroke. ance increasing with depth).

(a) T-bar
Gloucester, Canada
250 100
% of Initial Penetration Resistance
Penetrometer Resistance (kPa)

1.0 m
200 0.5 m 80
0.2 m
0.1 m
150 0.05 m 60

100 40

50 20

0 0
0 2 4 6 8 10 12 0 2 4 6 8 10 12
Cycle # Cycle #

(b) Ball
Gloucester, Canada
300 100
% of Initial Penetration Resistance
Penetrometer Resistance (kPa)

250
80
200
60
150

100 40

50
20
0
0
0 2 4 6 8 10 12 0 2 4 6 8 10 12
Cycle # Cycle #

Figure 5. Degradation of soil during cycling for various cycling intervals plotted as cycle # versus (a) absolute penetrometer
resistance and (b) percent of penetrometer resistance during initial penetration.

996

Copyright 2005 Taylor & Francis Group plc, London, UK


Examination of the degradation rate normalized as a partially remoulded strength behind the pen-
a percent of the initial penetration resistance for the etrometer. This average measurement is less than
T-bar reveals identical degradation behaviour for all the undrained strength and will result in a lower
cycling intervals between 5 and 100 cm. For the Ball, estimate of soil sensitivity. This also implies that N
the degradation behaviour is similar for cycling intervals values at sensitive sites may be lower.
of 10 cm and larger. Given T-bar and Ball penetrome- The cycling distance required to establish flow
ter diameters of 4.0 cm and 11.2 cm, this corresponds conditions is about equal to the diameter of the full-
to a cycling interval equal to about one penetrometer flow penetrometer. Considering practical implemen-
diameter. tation and data interpretation considerations a cycling
As evident in the 10 cm cycling plots (Figure 3c), distance of at least 20 cm is recommended.
there can be some variability in the displacement limits
when cycling is manually controlled with hydraulic
push systems. To minimize the influence of limit vari- ACKNOWLEDGEMENTS
ations a larger cycling interval in combination with
analyzing data at the center of the cycling interval The authors would like to Geoff Christoph, Don
may be preferable. Numerical analyses (Randolph 2004) DeGroot, Tom Lunne, Serge Leroueil, Mark Randolph,
of a perfectly placed penetrometer also indicate that the NGI, COFS-UWA, and UMass Amherst in the design,
full flow mechanism is fully created after displacements execution, and/or analysis:. Funding from the National
of about one to two penetrometer diameters. Con- Science Foundation (CMS# 0301448) is appreciated.
sidering these factors, a cycling interval of 20 cm could
be reliably implemented in practice and the full flow
mechanism would be fully mobilized at the center of
REFERENCES
the cycling interval.
DeJong, J.T., Yafrate, N.J., DeGroot, D.J., and Jakubowski, J.
2004. Evaluation of the Undrained Shear Strength Profile
7 CONCLUSIONS in Soft Layered Clay Using Full-Flow Probes, 2nd
International Site Characterization Conference, Porto,
Results from the field testing of T-bar and Ball pen- Portugal, Vol. 1, pp. 679686.
etrometers at three highly characterized test sites have Lo, K.Y., Bozozuk, M., and Law, K.T. 1976. Settlement
been presented to examine practical aspects related to analysis of the Gloucester test fill, Canadian
the measurement of remoulded strength via full-flow Geotechnical Journal, 13(4), 341354.
penetrometer cycling. The following observations have Randolph, M.F. 2004. Characterisation of soft sediments for
offshore applications. 2nd International Site Characteri-
been made: zation Conference, Porto, Portugal, Vol. 1, pp. 209232.
Soil remoulding typically occurs within 10 cycles. Leroueil, S., Hamouche, K., Ravenasi, F., Boudali, M.,
The number of cycles required to reach a steady Locat, J., and Virely, D. 2003. Geotechnical characteriza-
resistance was lower for the highly sensitive tion and properties of a sensitive clay from Quebec.
Characterization and Engineering Properties of Natural
Gloucester site due to a higher rate of decay. Soils, Tan et al. (eds.), Balkema, Vol. 1, pp. 363394.
It is important to recognize that the initial penetra- Lunne, T., Long, M., and Forsber, C.F. 2003. Characterization
tion of the penetrometer is an average measure- and engineering properties of Onsoy clay. Characterization
ment, with undrained shear being mobilized ahead and Engineering Properties of Natural Soils, Tan et al.
of the penetrometer and the soil softening towards (eds.), Balkema, Vol. 1, pp. 395427.

997

Copyright 2005 Taylor & Francis Group plc, London, UK


Well deformations at West Azeri, Caspian Sea

J.D. Allen
BP Exploration and Production Technology, Houston, Texas, USA

K. Hampson
BP Exploration and Production Technology, Sunbury upon Thames, Middlesex, UK

C.J.F. Clausen
Norwegian Geotechnical Institute, Oslo, Norway

C. Vermeijden
Fugro Engineers BV, The Netherlands

ABSTRACT: In the summer of 2003 the Dada Gorgud semisubmersible under contract to BP drilled one
well and completed parts of nine others in a twelve-slot template at the West Azeri platform site in the Caspian
Sea. Part way through the process the drillers determined that segments of some casings had moved laterally
more than a meter creating significant doglegs in the casings at depths of about 150 meters below mudline. The
template simultaneously experienced subsidence. A geotechnical investigation showed excess pore pressures
supporting 40 to 50% of the weight of the formation with correspondingly low clay strengths in some strata. The
low strength allowed the weight of the overburden to push soil into any open drilled hole when the hole was
filled only with seawater and cuttings. The soil movement carried adjacent casings along with it.

1 BACKGROUND 1.2 Drilling procedures


The semisubmersible drilling rig Dada Gorgud drilled
1.1 Project description
the wells through the seabed template. The first hole
The West Azeri offshore platform will be one of was drilled with sea water using a suite of tools that
several in the Azeri-Chirag-Gunashli (ACG) develop- would produce a 36 inch (0.91 m) diameter hole to a
ment in the Caspian Sea. The development area is depth of about 160 meters below mudline. The cuttings
about 45 km long and the West Azeri platform will sit flowed up the hole to the seafloor. Then the hole was
near the center of it about 130 km east-southeast from swept with a mixture containing guar gum followed
the city of Baku. Installation of the platform is sched- by more sea water and finally by prehydrated ben-
uled for completion during October of 2005. The plat- tonite mud with a specific gravity of 1.25. Then a 30(
form will be placed in 118.4 meters of water and will (0.76 m) casing full of sea water was lowered into the
contain positions for 48 wells. Production from the hole. The bottom end of the 30( casing, called the shoe,
platform is expected to peak at 342,000 barrels of oil stopped at about 154 meters below mudline and the
per day. weight of the 30( was supported on the seabed tem-
In order to accelerate production from the platform plate. Drillers cemented the annulus around the 30(
the project team planned to install 10 wells prior to using cement having a specific gravity of 1.9. Excess
installation of the platform. Contractors placed a steel quantities of cement would flow up in the annulus to
template on the sea bottom in the spring of 2003 to the seafloor.
serve as a guide for these first wells. The template was The next step was drilling a 26( (0.66 m) diameter
supported by three piles and also contained slots for straight hole through the 30( casing to a depth of 180 to
two docking piles that would serve as guides for set- 190 meters. The hole was swept with 50 barrels (7950
ting the platform. The template guided the wells into liters) of viscous liquid followed by further circulation
a 2.8 m by 2.6 m grid spacing. with seawater. Then drillers measured the orientation

999

Copyright 2005 Taylor & Francis Group plc, London, UK


of the hole along its full length with a gyroscope. This -14 -13 -12 -11 -10 -9 -8 -7 -6 -5 -4 -3 -2 -1
2

Vous aimerez peut-être aussi