Vous êtes sur la page 1sur 190

Revision and Proposed Improvement

for p-y Curves Design Criteria Used in


Offshore Wind Turbine Foundations

by

Julián Lajara Camacho


Exp. no: 325013

A thesis submitted in partial fulfilment


of the requirements for the degree of
Master of Science
in
Civil Engineering
(s.p. Structural Design and Construction)

Supervised by
Dr. Susana López-Querol (UCL)
Dr. Carmen Castillo Sánchez (UCLM)

E.T.S. de Ingenieros de Caminos, Canales y Puertos de Ciudad Real


Universidad de Castilla-La Mancha
University College London

Revision and Proposed Improvement for


p-y Curves Design Criteria Used in
Offshore Wind Turbine Foundations

by

Julián Lajara Camacho


Exp. no: 325013

A thesis submitted in partial fulfilment


of the requirements for the degree of

Master of Science
in
Civil Engineering
(s.p. Structural Design and Construction)

Supervised by
Dr. Susana López-Querol (UCL)
Dr. Carmen Castillo Sánchez (UCLM)

E.T.S. de Ingenieros de Caminos, Canales y Puertos de Ciudad Real

June 20th 2016


Personal information:

Julián Lajara Camacho


Exp. no: 325013

julian.lajara@gmail.com
julian.lajara@hotmail.com
+34 654 940 920
+44 (0)7821 482847

Supervisors:

Susana López-Querol, (UCL)


s.lopez-querol@ucl.ac.uk
Carmen Castillo Sánchez, (UCLM)
mariacarmen.castillo@uclm.es

General information

ETSI de Ingenieros de Caminos,


Canales y Puertos de Ciudad Real.
Universidad de Castilla-La Mancha
(UCLM), Avda. Camilo José Cela, s/n
13071 Ciudad Real (España)
Tel: +34 902 204 100

Department of Civil, Environmental


and Geomatic Engineering.
University College London (UCL),
Chadwick Building, Gower Street
London WC1E 6BT (United Kingdom)
Tel: +44 (0)20 7679 7224
“Civilization is civil works,
and in so far as these deteriorate so does society”

David P. Billington
Acknowledgements

This Thesis is the final piece of a challenging puzzle required to complete the +6 years
studies of Civil Engineering in Spain at the E.T.S.I. de Ingenieros de Caminos, Canales
y Puertos de Ciudad Real, University of Castile-La Mancha (UCLM). This work has
been carried out at the University College London (UCL) in the United Kingdom, as
part of the CEGE (Civil, Environmental and Geomatic Engineering department). After
all this time, my academic training has come to an end and hopefully I will be able to
put my knowledge and skills to good use in the very near future. But for now, in words
of a great artist, it is time to lay down this weary tune (and carry on).

During this work I have received advice from many people and I have learnt from topics
I had never been aware of before. I am grateful to Susana, my supervisor, for giving me
the opportunity to come to London without hesitation in this international internship
and learn from a different university than my alma mater. Aside from the technical
advice and deep knowledge of the topic, sometimes she provided me with inspiration
an insight of different topics such as professional engineering, which helped me to get
through this piece of work with an extra motivation.

Professors, classmates and acquaintances that have helped me all over this tough years or
we have gone through similar situations, thanks for your valuable addition to my persona.

And last, but certainly not least, many thanks to my parents in particular, and my family
in general (those who are here and those who are not) for supporting me unconditionally,
regardless of my (multiple) personal insecurities or the situation I am in.
Abstract

Offshore wind turbines farms are deemed to be the one of the most suitable solutions to
tackle future energy demands and emission restrictions in the European Union because
of the high efficiency of the system in terms of energy output per turbine compared to
onshore energy. Therefore, major Member States of the EU are now fostering the use
of such infrastructures, but the installation costs are very high compared to onshore
structures. Approximately, 25 − 30% of the cost of an offshore wind farm corresponds to
the foundations. Usually, for shallow waters, the preferred solution involves driving large
diameter hollow piles (monopiles) into the seabed, which are designed using standards
that reportedly lead to inaccurate designs. As a consequence of that, engineers are facing
an optimisation problem in the design of modern wind turbine foundations, especially
nowadays when newer models are getting bigger and will require a better performance of
the overall structure. The present report is focused on the study of monopile foundations
and the interaction between the soil and the structure. By means of Finite Element
Analysis (FEA), several comparisons have been carried out between a numerical model
and the so called p-y method, aiming at providing some recommendations for the design
of this particular foundations. This work deals with a more realistic approach to the
problem of modelling a laterally loaded monopile, as it includes dynamic conditions and
cyclic loads in a common but poorly characterised sandy soil, representing a step forward
with respect to most of the studies in this matter, in which the conditions simulated in
this document are only recreated in laboratory tests and never numerically. The analyses
include different loading scenarios and parametric studies of the interfaces between soil
and structure and plugging effects.

v
Resumen

Los parques eólicos offshore están destinados a ser una de las soluciones más factibles para
responder a la futura demanda energética y polı́tica de emisiones en la Unión Europea
debido a su alta eficiencia energética por turbina en comparación con la energı́a onshore.
Es por esto que la mayorı́a de Estados Miembros de la UE promocionan el uso de este tipo
de infraestructuras frente a las respectivas onshore. Aproximadamente, el 25 − 30% del
coste de un parque eólico offshore corresponde a las cimentaciones. Normalmente, para
profundidades reducidas, la solución más común consiste en la hinca de pilotes metálicos
huecos de gran diámetro (monopilotes) en el fondo marino, aunque estos elementos se
dimensionan en base a una normativa que lleva a resultados imprecisos. Es por esto que
los ingenieros se enfrentan al reto de optimizar el diseño de estos pilotes, y más ahora que
la tendencia que sigue la industria es la de aumentar el tamaño de las turbinas eólicas,
sometiendo a toda la estructura a un mayor nivel de solicitación. Este informe se centra
en el estudio de las cimentaciones de aerogeneradores con monopilotes y en el problema
de interación entre el suelo y la estructura. Usando el Método de los Elementos Finitos
(MEF), múltiples comparaciones se has llevado a cabo para contrastar los resultados
numéricos obtenidos con los ofrecidos por la normativa a través del llamado método de las
curvas p-y, tratando de ofrecer diversas recomendaciones para el proceso de diseño. Este
trabajo supone una apuesta más realista en la modelización de monopilotes sometidos
a carga lateral, puesto que incluye efectos dinámicos y cı́clicos en un suelo arenoso
mal caracterizado, suponiendo un paso adelante respecto a los trabajos que se están
desarrollando en la actualidad al respecto, en los que las condiciones aquı́ expuestas
son únicamente recreadas en el entorno de pruebas de laboratorio y nunca mediante
herramientas numéricas. Además, los análisis expuestos en esta tésis fin de máster
incluyen diferentes escenarios de carga ası́ como estudios paramétricos de los contactos
en la interfaz suelo-estructura y los efectos de la hinca.
Contents

Acknowledgements iii

Abstract vii

1 Introduction 1
1.1 Offshore wind energy background . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Offshore Wind Turbines description . . . . . . . . . . . . . . . . . . . . . 3
1.3 Foundations and monopiles . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.4 Problem description . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.5 Scope of the Thesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.6 Limitations and hypotheses . . . . . . . . . . . . . . . . . . . . . . . . . . 10
1.7 Outline of the Thesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12

2 Revision of the State of the Art 15


2.1 Offshore monopile foundations design . . . . . . . . . . . . . . . . . . . . . 15
2.1.1 P-y method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.1.2 Standards and codes . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.2 Laterally loaded monopiles: field tests . . . . . . . . . . . . . . . . . . . . 17
2.2.1 Dynamic behaviour . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.2.2 Cyclic behaviour . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.2.3 Centrifuge tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
2.3 Laterally loaded monopiles: Numerical models . . . . . . . . . . . . . . . 22
2.3.1 FEM . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
2.3.2 DEM . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
2.4 Installation of monopile foundations . . . . . . . . . . . . . . . . . . . . . 25
2.4.1 Open-ended and close-ended monopiles . . . . . . . . . . . . . . . 25
2.4.2 Driving and vibration . . . . . . . . . . . . . . . . . . . . . . . . . 26

ix
x CONTENTS

2.4.3 Plugging and cavity expansion . . . . . . . . . . . . . . . . . . . . 26


2.4.4 Friction and fatigue . . . . . . . . . . . . . . . . . . . . . . . . . . 27
2.4.5 Stress field alteration . . . . . . . . . . . . . . . . . . . . . . . . . . 28

3 Finite Element Model with ABAQUS 31


3.1 Brief remarks about ABAQUS . . . . . . . . . . . . . . . . . . . . . . . . 32
3.2 The Modules . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
3.3 Geometry and Boundary conditions . . . . . . . . . . . . . . . . . . . . . 34
3.4 Mesh . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
3.4.1 Element types . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
3.4.2 Meshing algorithm . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
3.5 Soil constitutive models . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
3.5.1 Revision of theory of plasticity in soils . . . . . . . . . . . . . . . . 42
3.5.2 Cyclic behaviour of sandy soils . . . . . . . . . . . . . . . . . . . . 43
3.6 Material properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
3.7 Contacts and interfaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
3.8 Loads . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
3.9 Types of analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
3.10 Time step . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
3.11 Summary of properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57

4 Validation of the Model 59


4.1 Point load . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
4.2 Shallow foundation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
4.3 Laterally loaded piles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
4.4 Geostatic step . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
4.5 Elastic models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
4.6 Modal analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
4.7 Soil column . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
4.8 Absorbing boundaries . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
4.9 Horizontal equilibrium . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77

5 P-y Method for Monopile Design 79


5.1 General background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
5.1.1 Elastic Approach: Beam on elastic foundation (BEF) . . . . . . . 79
5.1.2 P-y curves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
5.1.3 Capacity Approach: Brom’s method . . . . . . . . . . . . . . . . . 93
CONTENTS xi

5.1.4 Numerical Approach . . . . . . . . . . . . . . . . . . . . . . . . . . 95


5.2 Codes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
5.3 Monopile design according to the DNV-API . . . . . . . . . . . . . . . . . 97
5.4 Discussion and potential limitations . . . . . . . . . . . . . . . . . . . . . 100
5.4.1 Original testing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100
5.4.2 Initial slope Epy . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100
5.4.3 Layered soil . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
5.4.4 Soil-structure interaction . . . . . . . . . . . . . . . . . . . . . . . 101
5.4.5 Soil resistance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
5.4.6 Initial stress field . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
5.4.7 Sensitivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
5.4.8 Static and dynamic p-y curves . . . . . . . . . . . . . . . . . . . . 102
5.4.9 Concluding remarks . . . . . . . . . . . . . . . . . . . . . . . . . . 103

6 Comparison of the p-y Design Procedure and the FE Model 105


6.1 Interfaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
6.2 Soil plugging . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
6.3 2D v. 3D . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112
6.3.1 Absorbing boundaries . . . . . . . . . . . . . . . . . . . . . . . . . 115
6.4 Comparison FEM-API (DNV) . . . . . . . . . . . . . . . . . . . . . . . . 118
6.5 Peak loads . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120
6.6 Static-Dynamic response . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122
6.7 Influence of the cyclic loading . . . . . . . . . . . . . . . . . . . . . . . . . 129
6.8 p-y dynamic . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131

7 Conclusions and Recommendations 133


7.1 FEM modelling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133
7.2 Validation of the FEM model . . . . . . . . . . . . . . . . . . . . . . . . . 134
7.3 p-y method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136
7.4 Comparison p-y method and FEM . . . . . . . . . . . . . . . . . . . . . . 137
7.5 Comparison static-dynamic (cyclic) . . . . . . . . . . . . . . . . . . . . . . 138
7.6 Extra recommendations for designers . . . . . . . . . . . . . . . . . . . . . 139
7.7 Recommendations for the improvement of the codes . . . . . . . . . . . . 140

8 Conclusiones y Recomendaciones 141


8.1 Modelo con Elementos Finitos . . . . . . . . . . . . . . . . . . . . . . . . . 141
8.2 Validación del modelo de elementos finitos . . . . . . . . . . . . . . . . . . 143
8.3 Método p-y . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 144
8.4 Comparación método p-y - elementos finitos . . . . . . . . . . . . . . . . . 145
8.5 Comparación estático-dinámico (cı́clico) . . . . . . . . . . . . . . . . . . . 146
8.6 Recomendaciones para proyectos . . . . . . . . . . . . . . . . . . . . . . . 147
8.7 Recomendaciones para la mejora de las normativas . . . . . . . . . . . . . 148

9 Further Research 149

10 Appendices 161
10.1 Code for Infinite Elements . . . . . . . . . . . . . . . . . . . . . . . . . . . 161
List of Figures

1.1 a) Common terminology and b) general external actions . . . . . . . . . . 4


1.2 a) Ice on the platform (axial load increment) and b) marine growth (drag
increment). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.3 Schematic representation of the most commonly used foundation types.
a) Monopile, b) gravity foundation, c) suction caisson and d) tripod. . . . 6
1.4 Real examples of offshore wind turbine foundations. a) Tripod, b) jacket,
c) suction caisson, d) gravity foundation and e) monopile. . . . . . . . . . 7
1.5 Comparison between the London Eye and a 6 M W wind turbine rotor. . . 9

2.1 Summary of the research on field testing (Pile type: O = Open-ended,


C = Closed-ended. Installation method: D = Driven, J = Jacked, W =
Wished (pre-installed)), from de Blaeij (2013). . . . . . . . . . . . . . . . 18
2.2 150 · g centrifuge device. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
2.3 Degradation stiffness approach for the secant modulus of the soil, from
Achmus et al. (2009) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
2.4 Sample for triaxial DEM analysis, from Hu et al. (2010) . . . . . . . . . . 25
2.5 Soil arching and granular drag, from Liu et al. (2016). . . . . . . . . . . . 26
2.6 Soil flow and schematic horizontal stresses, from White et al. (2005) . . . 27
2.7 Experimental set-up and calibration chamber for ICP-05 testing, from
Jardine et al. (2013), Jardine et al. (2009a) . . . . . . . . . . . . . . . . . 29
2.8 Moving and Stationary radial stresses after different jacking strokes, from
Jardine et al. (2009b) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
2.9 Photoelastic effect in crushed glass near the pile toe, from Dijkstra and
Broere (2009) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30

xiii
xiv LIST OF FIGURES

3.1 Final geometry of the soil, after Achmus et al. (2009). The relative dis-
tances were taken wo the influence of the boundaries is minimised in the
surroundings of the monopile. . . . . . . . . . . . . . . . . . . . . . . . . . 35
3.2 Geometry of the pile . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
3.3 Mesh of the soil and the monopile . . . . . . . . . . . . . . . . . . . . . . 37
3.4 a) 20-node three-dimensional finite element C3D20R and b) integration
points per face . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
3.5 Seeding algorithm. The elements are reduced from the external to the
internal region by a constant 10%. The external blocks have a lenght of
7.2 metres, and the closer to the monopile 1.0 metres. . . . . . . . . . . . 41
3.6 Non-associative flow rule . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
3.7 Cyclic load for dynamic analyses . . . . . . . . . . . . . . . . . . . . . . . 53
3.8 Ramping function for static, quasi-static analyses and p-y derivation in
ABAQUS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
3.9 Equivalent stress for the pile cap . . . . . . . . . . . . . . . . . . . . . . . 54

4.1 Point load domain . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60


4.2 Shallow foundation problem dimensions . . . . . . . . . . . . . . . . . . . 61
4.3 2D Model of the strip foundation . . . . . . . . . . . . . . . . . . . . . . . 62
4.4 Equivalent plastic strain. Note the Prandtl mechanism formed due to an
imposed rough interaction between de coupled nodes of the foundation
and the soil. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
4.5 t-z curve for the foundation . . . . . . . . . . . . . . . . . . . . . . . . . . 64
4.6 Pile model for displacement comparison with Brom’s abaqus . . . . . . . . 65
4.7 Comparison between Brom’s method and the FEM model . . . . . . . . . 65
4.8 Geostatic step check. Although the red range appears to be negative, in
an adequate plot, the stresses at the top of the column are null, and it is
just an error of the plotting range (in [Pa]). . . . . . . . . . . . . . . . . . 66
4.9 Geostatic step in the 3D model (in [Pa]). . . . . . . . . . . . . . . . . . . . 67
4.10 Geostatic step in the 2D model (in [Pa]). . . . . . . . . . . . . . . . . . . . 67
4.11 Elastic model geometry . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
4.12 a) Permanent push load and b) triangular load . . . . . . . . . . . . . . . 69
4.13 Sinusoidal load . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
4.14 Deflection for a slender and non-slender pile . . . . . . . . . . . . . . . . . 70
4.15 Natural frequencies of the soil model . . . . . . . . . . . . . . . . . . . . . 71
LIST OF FIGURES xv

4.16 Analytic solution for an elastic, incompressible and isotropic soil column
subjected on its base to a sinusoidal acceleration field . . . . . . . . . . . 73
4.17 FEM solution for an elastic, incompressible and isotropic soil column sub-
jected on its base to a sinusoidal acceleration field. . . . . . . . . . . . . . 74
4.18 a) Column model in ABAQUS and b) horizontal displacement field (plot-
ted in 4.17). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
4.19 FEM model of the soil column with infinite elements . . . . . . . . . . . . 76
4.20 Solution to the infinite column problem, from López-Querol et al. (2014) . 77
4.21 FEM solution with ABAQUS . . . . . . . . . . . . . . . . . . . . . . . . . 77
4.22 Reaction in the soil on a middle plane of the monopile . . . . . . . . . . . 78

5.1 Beam on Winkler foundation . . . . . . . . . . . . . . . . . . . . . . . . . 80


5.2 Laterally loaded pile . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
5.3 Infinitesimal element analysis and geometric relations . . . . . . . . . . . 82
5.4 Representation of the soil reaction in a) after installation, and b) when
loaded. Notice that the lateral interaction (side friction) is neglected . . . 84
5.5 Concept of beam on non-linear elastic foundation . . . . . . . . . . . . . . 86
5.6 a) Elements of a p-y curve and b) variation of the subgrade reaction modulus 87
5.7 In-situ testing of laterally loaded piles, from Pando (2013). . . . . . . . . 87
5.8 Experimental set-up for laterally loaded piles in the Mustang Islands, by
Reese et al. (1974). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
5.9 Elements of a p-y curve, according to Reese et al. (1974). . . . . . . . . . 89
5.10 Failure modes for a laterally loaded pile, with a) for shallow depth and b)
for deep depth, by Reese et al. (1974). . . . . . . . . . . . . . . . . . . . . 89
5.11 Variation of the parameter A, by Reese et al. (1974) . . . . . . . . . . . . 90
5.12 Variation of Pc with depth, pile diameter and friction angle. Notice that
the bigger the diameter, the lower the transition depth (XR ) marked
with a circle, and most controlled the failure by the shallow mode, after
BrØdbæk et al. (2009). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
5.13 Variation of the parameter B, by Reese et al. (1974) . . . . . . . . . . . . 91
5.14 Estimation of kh as a function of the internal angle of friction, and the
relative density of the soil [API (1993)]. . . . . . . . . . . . . . . . . . . . 92
5.15 Behaviour of laterally loaded piles, by Helwany (2007). . . . . . . . . . . . 93
5.16 Brom’s abaqus for piles embedded in cohesive soil, with a) for short piles
and b) for long piles, from Helwany (2007) . . . . . . . . . . . . . . . . . . 94
xvi LIST OF FIGURES

5.17 Brom’s abaqus for piles embedded in cohesionless soil, with a) for short
piles and b) for long piles, from Helwany (2007) . . . . . . . . . . . . . . . 94
5.18 Variation of C1 ,C2 , and C3 with the internal angle of friction, from API
(1993). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96
5.19 Variation of Pc with depth x, with D = 6 m and φtr = 35o , according to
Figures 3.1 and 3.2 in Chapter 3 . . . . . . . . . . . . . . . . . . . . . . . 97
5.20 P-y curves method A . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
5.21 P-y curves method B . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
5.22 Method comparison 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
5.23 Method comparison 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99

6.1 Opening soil-monopile for µ = 0.3 (in [m]) . . . . . . . . . . . . . . . . . . 107


6.2 Opening soil-monopile for µ = 0.4 (in [m]) . . . . . . . . . . . . . . . . . . 107
6.3 Vertical slip of the toe of the monopile for µ = 0.3 and µ = 0.4 (in [m]) . . 108
6.4 Normal and shear forces in the contacts for µ = 0.3 (in [Pa]) . . . . . . . . 109
6.5 Normal and shear forces in the contacts for µ = 0.4 (in [Pa]) . . . . . . . . 109
6.6 Displacements at the tip of the monopile for different friction coefficients. 110
6.7 Different levels of plugging in the monopile . . . . . . . . . . . . . . . . . 111
6.8 Example of interaction between plugg and monopile (in [Pa]) . . . . . . . 111
6.9 Horizontal displacement field for the 2D model (in [m]) . . . . . . . . . . . 113
6.10 Horizontal displacement field for the 3D model (in [m]) . . . . . . . . . . . 113
6.11 Plastic equivalent strains in the 2D model . . . . . . . . . . . . . . . . . . 114
6.12 Plastic equivalent strains in the 3D model . . . . . . . . . . . . . . . . . . 115
6.13 Wave propagation problem in the 2D geometry . . . . . . . . . . . . . . . 116
6.14 Influence of the infinite elements in the stress field . . . . . . . . . . . . . 117
6.15 Example of the influence of the absorbing boundaries . . . . . . . . . . . . 117
6.16 Comparison of a mid plane of the monopile (p-y 0.5 m) . . . . . . . . . . 118
6.17 Comparison of a mid plane of the monopile (p-y 12 m) . . . . . . . . . . . 118
6.18 Comparison of a mid plane of the monopile (p-y 30 m) . . . . . . . . . . . 119
6.19 Horizontal subgrade reaction modulus variation with monopile depth . . . 120
6.20 FEM fitting (p-y 0.5 m) . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121
6.21 FEM fitting (p-y 12 m) . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122
6.22 FEM fitting (p-y 30 m) . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122
6.23 Horizontal displacement of the monopile . . . . . . . . . . . . . . . . . . . 123
6.24 Horizontal displacement of the mudline . . . . . . . . . . . . . . . . . . . 124
6.25 Horizontal displacement for n = 40 . . . . . . . . . . . . . . . . . . . . . . 125
6.26 Vertical displacement of the monopile . . . . . . . . . . . . . . . . . . . . 126
6.27 Vertical displacement of the mudline . . . . . . . . . . . . . . . . . . . . . 126
6.28 Vertical field of displacements for n = 40 . . . . . . . . . . . . . . . . . . . 127
6.29 Damage in the soil. The whiter the colour, the more accumulation of
plastic strains. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127
6.30 Crack opening at mudline level . . . . . . . . . . . . . . . . . . . . . . . . 128
6.31 Integration of the shear skin on the monopile surface . . . . . . . . . . . . 129
6.32 Integration of the base shear stresses . . . . . . . . . . . . . . . . . . . . . 129
6.33 Stress field (load peak) (in [Pa]) . . . . . . . . . . . . . . . . . . . . . . . . 130
6.34 Maximum horizontal stresses during peak load . . . . . . . . . . . . . . . 130
6.35 Horizontal displacement (load peak) (in [m]) . . . . . . . . . . . . . . . . 131
6.36 P-y static FEM vs. p-y dynamic FEM vs. p-y according to the codes . . . 132
List of Tables

3.1 Soil parameters for general soil model (Mohr-Coulomb) . . . . . . . . . . 45


3.2 Steel for monopiles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
3.3 Steel equivalent parameters for “solid” monopiles . . . . . . . . . . . . . . 46
3.4 Typical quasi-static loads for a 5 − 6 M W at mudline . . . . . . . . . . . 52
3.5 Environmental conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . 52

4.1 Soil parameters for the strip foundation problem . . . . . . . . . . . . . . 62


4.2 Elastic soil parameters for the reduced soil model (linear elastic) . . . . . 68
4.3 Soil parameters for the soil column problem . . . . . . . . . . . . . . . . . 72
4.4 Soil parameters for the infinite soil column problem . . . . . . . . . . . . . 75

5.1 BEF Relations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84


5.2 Recommended criteria for p-y curves [Reese and Van Impe (2001)]. . . . . 92
Chapter 1

Introduction

1.1 Offshore wind energy background

In the last two decades, the economic and social development of the European Union has
led to an increase of the energy demand. Industry and transportation take the biggest
share of the total power output, and when electric transportation becomes a reality,
the pressure on electric energy will increase. Thus, it is very important to do a good
planning of the foreseeable energy demands of the Member States.

Nevertheless, the main problem that society is facing is not only economical, but
also environmental and social. The dependency on fossil-based energy is strictly related
with the petroleum market, usually controlled by countries in political and social in-
stability, and the depletion of the fossil resources is expected to be a limiting factor
to the economy from the second half of this century. Aside from that, public aware-
ness and environmental responsibility are realities in today’s social opinion, conceiving
a more sustainable model of growth and demanding policies in accordance. It is clear
that renewable energies are a clear and valid alternative for achieving a better society,
a sustainable growth model and security of supply.

In terms of production, renewable energies have seen an increasing trend since 1990,
growing at an annually average rate of 4.4% in the European Union, and the share of
energy from renewable sources in the gross final energy consumption reached a 12.5% in
2010 [Eurostat (2012)].

In a multinational scenario, Europe is the world leader in offshore power generation

1
Chapter 1. Introduction 2

[EWEA (2011)]. The promotion of renewable, clean and indigenous energy is one of
the most important strategies of development in the European energy policies, and in-
sofar, maritime power is a relevant alternative. There are multiple lines of research and
development in this field, such as wind, tidal, wave or current energy, but in terms of
production, wind power is clearly the one with more potential [EWEA (2010)]. Since
the last two decades, this has been a major line of investment, expecting by 2020 to
supply the 20% of the energy demand of the European Union [EUP (2009)].

Nowadays, the power produced by offshore wind farms increases annually at gy-
gawatts levels, contributing to lower the emissions and helping to achieve a non-dependency
on fossil-based energies. The European Wind Energy Association (EWEA) estimates
that between 20 GW and 40 GW of offshore wind energy capacity will be operating
in the European Union by 2020 [EWEA (2011)]. This development responds to the
well established European Energy Roadmap, aiming to reduce by 2050 the greenhouse
gas emissions by 80 to 90% [EC (2011)]. But only with a high scale investment and a
global strategy, this target will be accomplished. Consequently, the two major technical
problems that will have to be faced are first how to build in a cost-effective way this in-
frastructure, and second, how to interconnect different wind farms to achieve a common
European grid and provide a global power system, reliable, safe and integrated with the
existing supply network for the European Union.

Recently, some countries have decided to foster offshore generation to the detriment of
onshore energy. As an example of this, the United Kingdom. Compared to onshore wind
power generation, the offshore generation has several benefits, such as a higher energy
output per unit, stronger and more sustained winds and low visual impact. Further-
more, wind farms can even contribute to the protection of certain maritime ecosystems.
The current systems present efficiencies up to a 70% extraction of the kinetic energy of
the wind, according to Betz’ law [EWEA (2015)], and the small size of this structures,
compared to their relatives maritime structures and platforms, allow for flexibility in
distribution and installation.

On the opposite, there are uncertainties regarding wind power in a marine envir-
onment. The competition of this source against the gas an oil industries in terms of
funding, equipment and expertise may lead to companies encountering difficulties when
trying to develop the offshore generation. The absence of a general transmission system
at the sea may create problems when distributing the energy on land or in inter-regional

M.Sc. Civil Engineering 1.1. Offshore wind energy background


3 Chapter 1. Introduction

networks. And all the protected areas in the marine environment have not been desig-
nated yet by the European Environmental Agency (EEA), consequently making difficult
to delimit the boundaries of the wind farms. Another major concern of the sector are
the expenses, as the sector is constantly bringing down the costs to achieve a more com-
petitive model. Those, which represent the setbacks of offshore wind power, represent
as well the main challenges engineering must address.

With such scenario, the number of wind farms and civil works associated to it are
expected to increase, constituting a good opportunity for the development of the renew-
able sector and a definite split from the traditional model of oil-dependent energy in
Europe.

1.2 Offshore Wind Turbines description


Obviously, any type of work in a marine environment represents an additional challenge
for engineers. This means that the cost of offshore infrastructure is way higher than
equivalent onshore (in terms of energy output), therefore it is paramount to find the
formula for achieving efficiency offshore. Engineers play a significant role in every stage
of this type of projects, from inception, to design and construction. Some of the extra
challenges engineers are presented with in offshore wind farm projects are, basically,
construction, maintenance and connection to the existing grid and power supply sys-
tem. On the contrary, dealing with typical problems of onshore projects such as space
availability, turbulences or inconstant winds, noise and a more restrictive environmental
awareness are kept in the background.

Within offshore farms projects, multiple disciplines of engineering are involved. Some
of the most interesting problems that arise during initial stages are turbine distribution
(in plan), wind disturbance minimisation, power hubs localisation, grid and cable lay-
outs, cable length optimisation, turbine design, superstructure and foundation design,
and so on.

Particularly, this document deals with the structure of the wind turbine itself, focus-
ing on the foundation. In order to set and describe the terminology of this document in
regards to the wind turbines, Figure 1.1 shows a generally accepted nomenclature.

1.2. Offshore Wind Turbines description M.Sc. Civil Engineering


Chapter 1. Introduction 4

Figure 1.1: a) Common terminology and b) general external actions

From top to bottom, the nacelle is the cover casing that houses all of the generating
components in a wind turbine, including the generator, drive train, and brake assembly.
The blades are the aerodynamic elements designed to transfer the kinetic energy from
the wind to the power converter or generator. The hub is the piece where the blades
are bolted, transmitting the rotary movement from the blades to the generator inside
the nacelle. These three elements together form the rotor. The rotor is mounted on

M.Sc. Civil Engineering 1.2. Offshore Wind Turbines description


5 Chapter 1. Introduction

the tower, which is made of steel. The transition piece is the element that connects the
tower to the foundation. The J-tube (J for the shape) is the protective cover of the
cables coming from the nacelle to the power grid. The rest of the terms are easier to
understand. In offshore foundations, the seabed is often referred as mudline, and later
on in this document will be described and studied in more detail.

Without considering the self weight of the structure, the main actions acting on the
structure are wind (most relevant), waves, currents, ice (all along the structure) and
the dynamic action of the rotating blades. The reaction of the soil (which is different
depending on the foundation type) provides stability to the tower. Other actions include
sand migration and marine growth, or accidental actions, such as impacts coming from
ships or ice blocks.

Figure 1.2: a) Ice on the platform (axial load increment) and b) marine growth (drag
increment).

1.3 Foundations and monopiles

The foundations for offshore structures are conditioned by the water depth at the loca-
tion. For shallow waters (up to 30 m), typically suction caissons (also known as suction
buckets), gravity foundations or monopiles are used. For transitional depths (30−60 m),
monopiles, tripods and jacket structures are common as well. For deep waters (> 60 m),
the previous solutions are not feasible, and designers have to recur to floating structures,
using tensioned guy wires to stabilise the structure or anchors to the sea bottom. Com-
binations of these solutions are also used (e.g. monopile with guy wires), but are very
rare.

1.3. Foundations and monopiles M.Sc. Civil Engineering


Chapter 1. Introduction 6

Figure 1.3: Schematic representation of the most commonly used foundation types. a)
Monopile, b) gravity foundation, c) suction caisson and d) tripod.

From all the types of foundations, monopiles are by far the most common solution.
The structure is very simple, as it is a steel pipe connected to the tower of the wind tur-
bine by a transition piece, which serves different functions, like protection from impacts
or mooring and landing of boats. Depending on the seabed conditions, the monopile can
be driven or nailed until the toe of the monopile reaches the design depth. Compared to
other foundations, the environmental impact is minimum. So it has been the easiness
of manufacturing and installation, and the low cost and impact what have made the
monopiles the most popular solution adopted by the sector.

Typically, the monopiles are floated to the site, or transported in vessels, and then
driven into the soil using hydraulic hammers in vessels with special equipment for this
tasks. The handling of the monopiles is delicate, and requires the use of special cranes
and clamps mounted on the above mentioned vessels. Installation times are relatively
short, for instance, individual monopiles constructed in 2004 as part of the Scroby Sands
wind farm in Norfolk, United Kingdom, required less than 24 hours to install [Malhotra
(2011)].

Since wind is the dominant action, the monopiles are subjected to non-traditional
working conditions. Any type of pile is generally employed in situations were the bearing
capacity of the soil is not enough to withstand vertical loading. In offshore wind farms,

M.Sc. Civil Engineering 1.3. Foundations and monopiles


7 Chapter 1. Introduction

Figure 1.4: Real examples of offshore wind turbine foundations. a) Tripod, b) jacket,
c) suction caisson, d) gravity foundation and e) monopile.

the vertical actions are relatively small compared to the horizontal ones, which are as
well very variable. This is also differential from the traditional oil and gas structures,
where wave and current loading control the design process.

All these particularities make interesting the study of the interaction of this structures
with the soil, under lateral loading conditions. Specially, monopiles, because they are
the less stable from the available options.

1.4 Problem description

Traditionally, the non-linear behaviour of the soil in problems of laterally loaded piles
is modelled by means of the so-called p-y curve method, in which the soil is divided
in independent non-linear springs. However, this soil-structure interaction problem was

1.4. Problem description M.Sc. Civil Engineering


Chapter 1. Introduction 8

based on field experimentation, using piles with small diameters, and later verified with
full scale testing (with 2 m diameter piles). But this method, which is currently admitted
by different standards, presents important limitations which lead to over-conservative
designs. If the companies want to optimise the design of the monopile foundation and
cut the cost of this foundations, it is imperative to find more efficient ways to size the
monopiles. Furthermore, lateral loading in a maritime environment are purely dynamic
and stochastic, which is not accounted for directly in the design codes.

With regards to the geometry, monopiles of around 4−5 m are common in the latests
wind farms. However, the aim of achieving a higher level of efficiency has pulled the
limits to monopiles of 6 − 7 m in diameter, and bigger ones are under study, contem-
plating the possibility of reaching 8 m and above. The wider the monopile, the better,
as more soil reaction is achieved with less penetration in the soil, which affects the size
of the driving machines, the efficiency of the process, the amount of steel in the struc-
ture, and eventually, the total cost of installation. Also the more it would behave as
a rigid pile and not as a flexible one. The limiting factor is the rotation at the head
of the monopile, as the wind turbines have a trouble-free range of operation in terms
of displacement and angular deviation. Nonetheless, wider monopiles are more likely
to present a softer dynamic behaviour, closing the gap between the natural frequen-
cies of their working range and the limiting ones [Cuéllar (2011)], and furthermore, the
driving process might become even infeasible. So the actual trends pointing to smaller
in length and wider monopiles [Achmus et al. (2009)]. This has been borne in mind
in this document as this particular case of monopiles will be studied. The geometry
and all the relevant data for this thesis have been extracted from a pre-design study of a
6 MW wind turbine [DOWEC (2003)] carried out by the Dutch company Ballast Nedam.

Aside from the increasing on the monopile diameter, the size of the turbines is in-
creasing notably as well. The bigger the turbine, the more energy can be produced.
Figure 1.5 shows a real comparison between the size of the rotor of a 6 MW wind tur-
bine rotor and the London Eye, a well-known structure without scaling, to appreciate
the size of the turbines and the general structure.

Thus, the question that has to be answered and the question this master thesis will
try to answer is: are the current methods and design standards representing accur-
ately enough the real interaction between actions, structure and foundation, leading to
cost-efficient infrastructure, or in fact, are the codes and national standards misleading

M.Sc. Civil Engineering 1.4. Problem description


9 Chapter 1. Introduction

designers to over-sizing on purpose, creating a gap in knowledge between what is gen-


erally accepted and what can be really done, comprising a problem that has not been
taken care of yet, with a lot of opportunities for improvement within the actual methods?.

Figure 1.5: Comparison between the London Eye and a 6 M W wind turbine rotor.

1.5 Scope of the Thesis


The overall objective of this document is to understand how the p-y method works, and
to compare it with a numerical model of the same characteristics with Finite Elements,
in an attempt to quantify analytically what the likely routes to achieve better designs
are. This thesis is aimed to provide a critical analysis of the p-y curves design method for
offshore wind turbines foundations (monopiles) compared to a modern numerical model
with finite elements, and to understand the interaction effects between foundation and
soil in terms of stresses, displacements, etc. Regarding this, the main tasks carried out
have been:

• To revise the state-of-the-art in regards to offshore wind turbines foundations and


other transversal topics, such as empirical experimentation, installation or con-
struction of this type of structures.

• To provide a critical review of the traditional design approach of offshore wind


turbine foundations.

• To analyse the influence of changing different parameters of the soil and the struc-
ture in terms of stresses, displacements, etc. in the numerical models.

1.5. Scope of the Thesis M.Sc. Civil Engineering


Chapter 1. Introduction 10

• To analyse and understand the behaviour of laterally loaded monopiles under dif-
ferent load scenarios, and under static, pseudo-static and dynamic analysis.

• To understand the limitations of the numerical models (soil plugging, constitutive


laws, contact elements, etc.), and the problems that arise when using finite elements
(convergence, computational time, numerical errors, etc.).

• To analyse and compare monotonic and cyclic loading scenarios in offshore envir-
onments.

• To understand the differences and simplifications of 1D, 2D and 3D models, and


to compare the results with the traditional design methods.

• To understand the position of the energy industry and the researchers in different
hot topics.

• To quantify qualitatively effects which are not accounted for in the p-y method,
but can be determined in a finite element model.

• To propose some guidelines and recommendations based on the analysis to improve


the existing methodology, gaining a better understanding and control over future
designs.

1.6 Limitations and hypotheses


Traditionally, the biggest offshore projects are privately owned by oil and gas companies,
so the existing information in terms of ground testing and general know-how is owned
by those. Besides, the majority of the information is available only for those geographic
areas where research and testing have been carried out, areas with fossil resources, etc.
The recently development of the offshore energy implies that data from soil testing is still
very scarce. Energy companies invest big amounts of money and are reluctant to provide
such information or make it public. This, for the academic realm, represents a challenge,
because of the lack of information in general terms. In this thesis, the soil parameters
and other information of this kind has been extracted from a public-private collaboration
between the Technical University of Delft in joint venture with a few energy companies
interested in studying the introduction of bigger turbines, such as Ballast Nedam. And
even the full document is difficult to find, so the information from the draft has been
used. This was the only information available to develop the models for study, and it is
based in a campaign of soil testing in the North Sea, in three different locations. The

M.Sc. Civil Engineering 1.6. Limitations and hypotheses


11 Chapter 1. Introduction

chosen location was characterised by its sandy soils, which is a fair comparison with the
conditions found in most of the UK shore. But still, soil testing in the sea is very diffi-
cult, and the harsh conditions made the company responsible for the bore holes reaching
only a depth of 12 m beneath the mudline. With that data, it is difficult to extrapolate
the values of the different parameters to higher depths, specially when it is expected to
reach 30 − 40 m under the mudline. This of course is a limitation of the results presen-
ted in this document, but no other data has been found or provided. In fact, this is the
same limitation that underlies in almost every technical publication and research papers.

Another important remark is that the analysis with FEM have not included the influ-
ence of the water. Part of the structure is under the water level, but the hydrodynamics
effects have not been taken into account (currents, bed drag and similar effects). An-
other consequence is that the soil under saturated conditions. Although this included
in the parameters of the soil, the absence of water in the model may have influence the
dynamic analysis (the mass of the water is missing in the equation) and no variation
in pore water pressure is estimated. Other effects such as scour around the pile in the
mudline, subsidence or settlements have also been neglected.

Likewise, the second major limitation of this dissertation has been the computation
capacity. All the models and calculation have been done using a dual core standard
processor in a laptop. This means that the number of elements in the numerical models
have been reduced as much as possible so that the calculations were ready in a reas-
onable time (which in some cases was three or four days, with many simplifications).
The numerical complexity will be a topic discussed when the finite element model is
described, but the three main variables influencing the overall time was the number of
elements, the plastic analysis and the dynamic analysis. In an attempt to overcome this
problem, different simplifications were introduced, and a simplified 2D model was also
studied.

All those limitations however are extended throughout all the publications found in
the literature because of the complexity of the numerical simulations in consideration
here.

1.6. Limitations and hypotheses M.Sc. Civil Engineering


Chapter 1. Introduction 12

1.7 Outline of the Thesis

This thesis is divided in 8 chapters. Chapter 1 is pretty much an introduction to the


topic of offshore energy, offshore wind farms and marine structures and foundations. In
Chapter 2, an overview of the research concerning the monopile foundations is presen-
ted, including topics such as the dynamic and cyclic behaviour of offshore wind turbines,
laboratory testing, ground conditions, numerical modelling of soils, installation effects
of monopiles, stress alteration and plugging and fatigue effects.

In Chapter 3 the Finite Element model developed and used for the numerical calcu-
lations in this thesis is presented and explained, paying special attention to the different
details that make it complex to understand. Some remarks about the modelling soft-
ware (ABAQUS) are commented as well. Some topics discussed are the geometry of the
model, boundary conditions and loads applied, the finite element types used, constitutive
models available in ABAQUS, contacts, absorbing boundaries, size of the model, type
of analysis or time step considered in dynamic simulations.

As there is no site date or lab testing backing up this thesis, Chapter 4 is extremely
important as it takes the most conflictive aspects of the numerical model and get them
throughout an extensive validation process, providing confidence in the subsequently
presented results in the following chapters of the thesis.

Chapter 5 is devoted to the p-y method. Along the various theoretical methods for
designing laterally loaded piles, the accepted p-y method is explained and discussed in
detail on the generally accepted p-y method. After analysing the p-y curves method, this
curves are obtained following different procedures to get a closer insight of the design
using this approach and applying this methodology to the problem being studied.

In Chapter 6 the overall results obtained by means of numerical analysis are exposed
and discussed, contrasting the traditional p-y method with the numerical dynamic and
static simulations. Besides that, the influence of important aspects of the numerical
models (interfaces, soil plugging, displacements, stress fields, 2D versus 3D comparis-
ons) have also been studied and discussed.

Chapter 7 is a summary of the findings of this work, addressing most of the topics
in the scope of this report, terminating the analysis process in Chapter 8 with some ref-

M.Sc. Civil Engineering 1.7. Outline of the Thesis


13 Chapter 1. Introduction

erences to those topics and aspects that have not been fully understood due to practical
reasons and need further addressing and maturing.

An last after Chapter 8, the Bibliography and references are listed. An Appendix
with relevant information about a code for infinite elements is also included.

1.7. Outline of the Thesis M.Sc. Civil Engineering


Chapter 2

Revision of the State of the Art

In this chapter, a review of different relevant technical papers will be carried out in
regards to the main subject of this thesis. The objective of this state-of-the-art is to
understand the research done up to date and to review results and methods developed
by other more experienced researchers that might result useful to produce outcomes in
accordance to what has been done and is supposed to be scientifically correct and right-
eous. Although many topics have been covered during the production of this document,
the most important ones are those intimately related to numerical simulation of soil-
structure interactions in piles and sandy soils, constitutive models of soils and monopile
foundations under lateral cyclic loading conditions. In many cases, the revision of the
available bibliography helped the author to build up his own numerical models taking
into consideration the existing ones and making possible to check the similarities of the
results with those already published.

2.1 Offshore monopile foundations design

As stated in many recent reports, offshore wind energy is a solid candidate to achieve
the future renewable energy production requirements set by the EU owed to the low
environmental impact and high efficiency of the system [EWEA (2011), EWEA (2010),
EWEA (2015)]. But the further the wind turbines are from the shoreline, the less feas-
ible is their installation because of the harsh weather conditions and the increasing water
depth. The compromise between energy generation and cost effectiveness of wind farms
projects has lead to the monopile foundations to become the preferred alternative by
the industry when it comes to shallow depths (up to 30 m), representing between the 30
and the 40% of the cost of a complete wind turbine, assuming an average of 1.5 million

15
Chapter 2. Revision of the State of the Art 16

Euros per megawatt installed [Byrne and Houlsby (2003)]. Those figures represent the
direct cost of a wind turbine, and they are strictly related to the design stage. In the
case of a monopile foundation, it depends on the type of steel (density and strength),
the amount of steel (embedded length, total length, thickness and diameter) and the
installation method (hammered, driven, drilled, etc.). Another big advantage of the use
of monopiles is the easiness of installation and manufacturing.

The advantages that piles present with regards to axial loading have been known
and studied during centuries, and it is not the scope of this thesis to analyse axial effects
in piles but to focus on the lateral behaviour and capacity of these elements, which is
a field currently undergoing revision mainly because the high applicability in offshore
wind farms, and the lack of knowledge in some aspects, like the response under cyclic
loading conditions.

2.1.1 P-y method

The analysis of the deformational behaviour of a soil started later in the XIX century
with the studies of Wrinkler and the development of Bernouilli’s beam theory for found-
ations and railway engineering [Winkler (1897), Hetenyi (1946)]. However, the idea of
a subgrade reaction modulus, a constitutive parameter of the soil linking force and de-
formation in piles was first introduced by Terzhagi (1955). Ideally, the displacements
could be computed on the assumption that the pressure acts on an elastic layer with
thickness equal to three times the diameter of the pile, and a subgrade reaction modulus
varying linearly with the depth of the pile. Unfortunately, no experimental or analytical
was provided to verify those assumptions. Later on, Reese and Matlock (1956) continued
with the research in the matter.

The p-y curve method can be attributed to McClelland and Focht (1958) and their
research based on consolidated undrained triaxial tests in deformational behaviour, but it
was not until Reese et al. (1974) and the Mustang Islands tests that the p-y curves started
to adopt its current form. This was a further development in the theory of the subgrade
reaction modulus, which was translated into non-linear springs as a continuation or
improvement of the analysis of a beam on elastic foundation. This initial experimental
and analytical p-y curves were intended for specific soils (those existing in the Mustang
Islands), so an adaptation for other soils and a generalisation was required, task carried
out by Murchison and O’Neill (1983), methodology which was later including in the API

M.Sc. Civil Engineering 2.1. Offshore monopile foundations design


17 Chapter 2. Revision of the State of the Art

(1993). For further references and an extensive derivation of this theories, please see
Chapter 5 of this master’s thesis.

2.1.2 Standards and codes

The basic document for the design of laterally loaded monopiles is the API (1993),
or the 2007 updated version, by the American Petroleum Institute. It is based on
the findings of McClelland and Focht (1958) and Murchison and O’Neill (1983), and
incorporates different guidelines, rules for design and recommendations, although most
of them are not quantified. The rest of the standards refer to or are based on this
American code. The other codes, DNV (1992), GL (2005), or national recommendations
such as M.E.L.T (1993) (France) or P.H.R.I (1980) (Japan) introduce additional rules
or local recommendation for good practices in design, serving as national annexes for
offshore wind turbine foundation design.

2.2 Laterally loaded monopiles: field tests


As said before, the simplicity of a monopile foundation and it potential use in wind
farms have made necessary a revamp of the studies and an update of the existing codes
and rules. The two existing tools are experimentation and numerical modelling, and
a combination of both. Most of the existing studies are experimental tests [de Blaeij
(2013)]. It is not the intention of this section to give account of the vast amount of
studies, but just a mere summary of the most significant. Please refer to Table 2.1 for
summed up information on this. Also, some of the terms used in the Table 2.1 will be
mentioned in subsequent sections.

2.2.1 Dynamic behaviour

As opposed to a traditional static analysis, the nature of the loads in a maritime envir-
onment is highly dynamic. Therefore, static approaches seem not to be the best way to
address the design of foundations using the p-y curves.. According to Cuéllar (2011),
there are two main trends in the way to treat the dynamic problem in the offshore wind
turbine literature.

On the one hand, a unspecific cyclic design approach, aiming to provide a lower
bound to the cyclic behaviour. Such approaches are characterised by a static design
affected by some constant factor, which is the same for all forms, types and magnitudes

2.2. Laterally loaded monopiles: field tests M.Sc. Civil Engineering


Chapter 2. Revision of the State of the Art 18

Figure 2.1: Summary of the research on field testing (Pile type: O = Open-ended,
C = Closed-ended. Installation method: D = Driven, J = Jacked, W = Wished (pre-
installed)), from de Blaeij (2013).

of cyclic loading (as can be seen in the API with the introduction of a 0.9 coefficient
degrading the overall p-y curve in all depths).

On the other hand, models and formulas focusing on certain aspects of cyclic loading
(for instance the number of applied cycles or the load level), in order to predict more
accurately the evolution of the pile behaviour. Now, the first method is used in most
of the design rules, as a lower bound means design by the principle of caution, but give
over-conservative results. The second method is usually based on experimentation, and
its difficult to extrapolate to any given design situation. Plus, the use of scale models

M.Sc. Civil Engineering 2.2. Laterally loaded monopiles: field tests


19 Chapter 2. Revision of the State of the Art

does not help sometimes when dealing with full scale foundations. Achmus et al. (2009)
criticises the fact that the existing correction factor in the codes are based on test with
less than 200 cycles, and for numbers above that, the correction factor might become
uncertain.

The influence of the soil in the dynamic response of the structure is remarkable, and
worth studying [Cuéllar (2011)]. There are two mainly effects: the first one is the direct
influence of the soil in the dynamic behaviour of the wind turbine. The second one is the
influence that the movement of the turbine has in the soil properties. This two situation
are not independent, but they will be treated as so in this section.

Generally speaking, the stiffness of the soil contributes to the natural frequencies
of the whole structure [Van der Temple (2006)]. A dynamic analysis is very important
in this type of structures, as the vibration of the rotor in working conditions cannot
coincide with the natural frequencies of the structure, as a necessary condition for the
serviceability state. The natural frequency in design of an offshore turbine should be such
that is does not coincide with the so-called 1P and 3P frequency intervals for variable
rotors and wave frequencies. Wave frequencies are generally lower than the rotational
frequency of the rotor. 1P and 3P are the frequencies respectively belonging to the
corresponding peak loads of the rotation frequency of the rotor and the frequency of all
the blades passing by the tower. According to Van der Temple (2006), these frequencies
divide the frequency range into three intervals (soft-soft, soft-stiff and stiff-stiff) suitable
for designing the natural frequency of the wind turbine. The softest stiffness, e.g. lowest
natural frequency, is considered to be the best from an investment point of view. A
lower stiffness will result in lower overall natural frequency which may be compensated
by increase in monopile diameter [Van der Temple (2006)]. According to these results,
obtaining the right stiffness of the soil is paramount in design stages, as the structural
stability of the turbine relies on that, and may depend on the geometrical properties of
the foundation [Achmus et al. (2009)], whereas the API (1993), for instance, considers
the stiffness of the soil independent of the geometry parameters.

2.2.2 Cyclic behaviour

The cyclic behaviour is a particular situation of dynamic loading, which accounts for re-
peatability of load cycles. It seems likely that the more cycles the foundation is subjected
to, the more displacement and degradation of the soil is expected, as the traditional fa-

2.2. Laterally loaded monopiles: field tests M.Sc. Civil Engineering


Chapter 2. Revision of the State of the Art 20

tigue equations, like the Miner’s rule, state. In the literature, such effects are explained
with two terms [Cuéllar (2011)]: the first, shakedown (the loading level is high enough
to produce plastic deformations in the soil at some depths, but after a finite number of
cycles the pile response stabilizes and from then on only elastic deformations occur) and
the second, degradation (i.e. a material degradation of the soil leading to a decrease
of its strength and stiffness). However, both the rate of accumulation of permanent
deformations and the displacement cyclic amplitude tend normally to diminish, which
implies that the reloading stiffness has to be higher than the virgin stiffness, and also
grow with every cycle [Cuéllar (2011)], in other words, there is a certain amount of cyc-
lic stiffening occurring in the soil when it is subjected to cyclic loading. Some of this
effects (obviously depending on the type of soil) might be explained with the idea of a
mechanical reorganisation of the soil particles in the surroundings of the pile, allowing
for a densification of the material. But, due to the saturated conditions of the seabed,
with conditions close to undrained, the increased compacting tendency of symmetrical
loading could eventually lead to a faster pore water pressure accumulation and therefore
to an increased liquefaction potential, with the subsequent reduction of the stiffness of
the soil [Cuéllar (2011)]. In Hu et al. (2010), it was proved that low-level cyclic loading
can improve pile capacity, whereas high-level cyclic loading can be highly detrimental to
shaft capacity, although this was only involving axial loading, and the lateral influence
was not mentioned.

Furthermore, the analysis of offshore piles poses the additional problem of the defini-
tion of the number cycles itself, given the irregular and stochastic nature inherent to the
offshore loading, not only in value, but also in direction. In fact, there are very few cyclic
tests on laterally loaded piles, and all of them unidirectional. And, in most of them, the
number of cycles is very limited (20 − 1000 cycles in elevated stress conditions) and not
enough to infer a trend in the behaviour of the structure in the long term of a life cycle
(≈ 25) years, with more than 109 cycles expected [Cuéllar (2011)].

2.2.3 Centrifuge tests

The two main experimental test are the so-called 1 · g tests, and the centrifuge test.

1 · g testing is the traditional method, and is a good approach to determine the influ-
ence of the different factors governing the behaviour of the piles, or to do a parametric
analysis of sensitivity. A good example of this is the analysis carried out by Cuéllar

M.Sc. Civil Engineering 2.2. Laterally loaded monopiles: field tests


21 Chapter 2. Revision of the State of the Art

(2011), which was able to demonstrate the influence of the cyclic loading in the densific-
ation of the soil in the surroundings of the pile, or the increment in pore water pressure.
The problems of this type of models are, first, the scale (soil particle against structure
unless in full scale tests) and secondly that absence of realistic effective stresses in the
soil. To overcome the problem of the effective stresses, N · g tests or centrifuge tests are
often used, when the equipment is available.

The underlying concept of the centrifuge is to subjects a scale model to a rotational


centrifugal force to simulate real stresses in the soil by an increase in the gravitational
force, so that more representative results can be achieved. However, the apparatus is
expensive, and requires experimented and well trained professionals.

Figure 2.2: 150 · g centrifuge device.

On of the world’s top institutions in this sort of testing is the Cambridge University,
under the supervision of Prof. Andrew N. Schopfield, one of the earliest in using centri-
fuge analysis in soil mechanics.

Regarding monopile and p-y analysis, in 2012 six centrifuge tests have been carried
out under 100 · g conditions in saturated clay showing that although the use of the DNV
(1992) and design p-y curves is conservative for ultimate limit state design, its usage
may be inappropriate for cyclic loading design as the initial soil-foundation stiffness may
not be accurately estimated, resulting in inaccurate calculations of the change in loads
applied to the system [Lau et al. (2014)].

Three dimensional Finite Element Analysis (FEA) was undertaken using the com-
mercial software ABAQUS to complement the understanding of the lateral response of
the monopiles obtained through centrifuge testing. Centrifuge testing and fully-coupled

2.2. Laterally loaded monopiles: field tests M.Sc. Civil Engineering


Chapter 2. Revision of the State of the Art 22

3D finite element analysis was undertaken to investigate the lateral load-deformation


response of monopiles in soft clays. The study was carried out on a monopile with an
outer diameter of 3.8 m, total length of 50 m and embedding depth of 20 m. Three types
of soft clay with varying undrained shear strength profiles were used. Axial load and
lateral displacement were applied to the monopile to represent turbine self-weight and
wind and wave loads respectively. Compatible results from both these methods verified
the hypothesis that a large diameter monopile behaved as a short rigid pile by deforming
via rotation about a pivot point with its toe undergoing negative displacement. It was
also determined that the depth of this pivot point increased with an increase in the
applied lateral displacement [Haiderali et al. (2014)].

But dynamic centrifuge testing has also been undertaken to investigate the effects
of axial load on the lateral response of pile groups and earthquake influence. Instability
effects lead to amplification of any existing lateral deflections which may be present as
a result of construction-induced imperfections, historical lateral displacements or owing
to earth pressures from laterally spreading soil. Apart from that, the relative pile–soil
flexibility was found to have a strong influence on the increasing lateral deflection oc-
curring in the tests [Knappett and Madabhushi (2009)].

2.3 Laterally loaded monopiles: Numerical models

2.3.1 FEM

This subsection of the thesis is specially important, because most of the decision-making
process on certain aspects of the model from Chapter 3 has been based in the recom-
mendations or the decisions made by these authors in their numerical models and FEA
know-how.

Probably one of the first studies using FEA was carried out by Trochanis et al. (1991).
In this paper, the nonlinear response of pile foundations to axial and lateral loads, both
monotonic and cyclic.The piles and the soil were modelled by solid three-dimensional
quadratic elements (with ABAQUS) including contact elements. Only square piles were
modelled. The pile elements were assumed to remain elastic at all times, while the soil
was idealized as either a linear elastic material or a Drucker-Prager elastic-plastic ma-
terial.

M.Sc. Civil Engineering 2.3. Laterally loaded monopiles: Numerical models


23 Chapter 2. Revision of the State of the Art

With time, models were growing to introduce a greater number of elements, but the
basics remained constant. One of the most referenced experts in numerical simulation
of monopiles is Professor Martin Achmus. In Achmus et al. (2009), he introduced one of
the most interesting improvements in the models. Basically, he decided to avoid the use
of complex and sophisticated material laws. He discovered that, under cyclic conditions,
if the rate of displacement accumulation is low, this may lead to the accumulation of
numerical errors. To prevent this from happening, he proposes the so-called Degradation
Models Stiffness, in which a variation of the soil properties are correlated with the rate
of deformation and the number of load cycles by an exponential law. A typical form of
this laws are as follows: a
EsN ∼ εcp,N =1 b2
= a = N −b1 (X) (2.1)
Es1 εcp,N
The degradation stiffness approach to account for cyclic loading effects is presented in
Figure 2.3. In a cyclic triaxial test, an increase of the plastic axial strain can be observed.
Assuming the elastic strain to be negligible, the degradation rate of secant stiffness after
first cycle Es1 and N th cycle EsN can be presented by the plastic axial strains after first
cycle εacp,N =1 and after N th cycle εacp,N according to the equation. Here N is the number
of load cycles and X is the cyclic stress ratio for cohesionless material as follows:

σ1,cyc
X= (2.2)
σsf

where σsf is the major principal stress at static failure state and σ1,cyc is the major
principal stress for the cyclic stress state under consideration. b1 and b2 are calibration
parameters. The cyclic stress ratio is thus dependent on the confining pressure and on
the cyclic stress level.

Nevertheless, perhaps the big concern with this method is the use of an elastic-plastic
law, bearing in mind that a Mohr-Coulomb is not precisely a constitutive model for a
soil, but a failure criteria. The influence of the diameter of the monopiles was another
topic of study using a Mohr-Coulomb failure criterion and 3D FEA. It was found again
that the use of the p-y method with large diameter monopiles led to conservative results
in terms of displacements and rotations of the pile head [Achmus et al. (2008)], finding
this significant for monopiles with D > 3 m.

In order to account for a more complex numerical model instead of a Mohr-Coulomb,

2.3. Laterally loaded monopiles: Numerical models M.Sc. Civil Engineering


Chapter 2. Revision of the State of the Art 24

.
Figure 2.3: Degradation stiffness approach for the secant modulus of the soil, from
Achmus et al. (2009)

Damgaard et al. (2014) and Bourgeois et al. (2010) worked in 3D FEA with cyclic loading,
but incorporating more complex constitutive laws. Damgaard et al. (2014) considered
the use of a isotropic and poroelastic material, simulating the soild phase of the soil and
the liquid phase separately and then combining them. Bourgeois et al. (2010) introduced
a specific law for the cyclic behaviour of the soil using its own software (CESAR-PLC).

2.3.2 DEM

In this section it is worth mentioning as well other numerical models which are being
developed as a tool for understanding the behaviour of the particulate matter that form
the soil. This methods are put together as Discrete Elements Methods (DEM) and
represents a field with a promising future, as it can represent the interaction between
particles, treating the soil medium as a compendium of microscopic elements and not as a
continuum. DEM tests have been proved useful for the study of stress-induced anisotropy
in sand under cyclic loading [Hu et al. (2010)], and for calibrating soil parameters using
analytical triaxial tests, which later were used for foundations subjected to cyclic loading
analysis [Sim et al. (2013)]. In Figure 2.4 an example of a DEM model of a triaxial
test for deriving soil properties and soil investigation can be seen. Ideally, the DEM
method would be applied more and more to complex soil problems, and specifically to
the monopile problem. Even though due to a scale factor a monopile interaction problem
is easier to model with FEA, the DEM can complement very well in topics were a FEM
model can lack applicability, for instance, the installation problems, in which penetration
and crushing of particles are highly influential in the results.

M.Sc. Civil Engineering 2.3. Laterally loaded monopiles: Numerical models


25 Chapter 2. Revision of the State of the Art

.
Figure 2.4: Sample for triaxial DEM analysis, from Hu et al. (2010)

2.4 Installation of monopile foundations

Several authors have investigated the influence of the installation of a pile in the overall
behaviour of the structure. As it is a clear soil-structure interaction problem, the initial
stress conditions imposed by the installation method will have consequences in terms
of displacements and stresses, an this is critical to understand the short term and long
term response of the pile.

2.4.1 Open-ended and close-ended monopiles

Monopiles are a particular type of piles. Due to its hollow structure, during the install-
ation process, a soil column is likely to penetrate up to a certain height. During the
installation of these open steel pipe piles a plug can form inside the pile. Such plug
formed during the installation has an influence on the installation process of the steel
pipe pile as well as on the bearing behaviour and the pile resistance. This extra res-
istance comes from soil arching, which ultimately creates an increase in the horizontal
stresses, and therefore an increase of the skin friction in the pile, both in the outer an
inner surfaces [Paikowsky (1990)] (see Figure 2.5). Dijkstra and Broere (2009) however
attributes this to a formation of a cone of soil (like a failure mechanism) at the toe of
the open-ended monopile. The forming of the plug depends on different parameters, e.g.
the pile diameter, the soil conditions and the installation method [Lüking and Kempfert
(2013)], but the phenomenon of the plug inside open-ended steel monopiles is not fully
investigated yet [Labenski and Moormann (2015)]. Some of this hypothesis were corrob-
orated in experimental tests scaled tests in sands by Liu et al. (2016), stating that the
amplitude of the cyclic loading can influence the macro-scale material response as well
as the particle scale response in the soil.

2.4. Installation of monopile foundations M.Sc. Civil Engineering


Chapter 2. Revision of the State of the Art 26

Figure 2.5: Soil arching and granular drag, from Liu et al. (2016).

2.4.2 Driving and vibration


Driving and vibration are the two most common installation methods for offshore mono-
pile foundations [Labenski and Moormann (2015)]. The main difference between both
is the equipment required and the duration of the process. While vibration may take
minutes, driving may take hours. For the driving process, a hydraulic hammer is re-
quired, and a noise-mitigating shield has to be allocated on top of the monopile or in
the hammer. There are several analytical approaches as well as national standards to
determine the bearing capacity of driven pipe piles, no matter if open or close-ended.
This is an important parameter when selecting the hammering equipment. Nevertheless,
there are just a few analytical approaches and so far no national standards to determine
the bearing capacity of vibrated pipe piles [Labenski and Moormann (2015)].

The existing methods for driving (not valid for vibrating) are the ICP-05 method
[Jardine et al. (2005)], UWA-05 method [Lehane et al. (2008)], UCD-11 method [Igoe
et al. (2011)], and the HKU-12 [Yu and Yang (2011)] all based in the cone penetrometer
test.

2.4.3 Plugging and cavity expansion


The plugging of an open-ended monopile can be characterised by the so-called Final
Filling Ratio (FFR) and the Incremental Filling Ratio (IFR). The IFR is a measure
of soil displacement near the pile tip and depends on the inner pile diameter, pile wall
thickness, plug densification or dilation and installation method [Lehane et al. (2008)].
As the IFR approaches zero, the behaviour of the pile is the same as that of a fully
plugged pile. If the IFR and FFR approaches to 1, the pile behaves fully cored, which

M.Sc. Civil Engineering 2.4. Installation of monopile foundations


27 Chapter 2. Revision of the State of the Art

is approximately equivalent to a bored pile [Lehane et al. (2008)]. The densification


of the soil, and the relation between the FFR and IFR is summarised in the cavity
expansion theory [Xu et al. (2005)], which takes into consideration the flow of soil passing
through during the installation process and the variation of soil properties because of
the alteration of the soil. The level of displacement in the soil can be related to the
effective area ratio A∗rs . The relation between plugging theory and cavity expansion was
studied by Xu et al. (2005) and White et al. (2005).

D0.2
 
F F R ≈ min 1, i (2.3)
1.5
 2
Di
A∗rs = 1 − IF R · (2.4)
Do
Where Di and Do refer to the inner and outer diameters respectively.

Figure 2.6: Soil flow and schematic horizontal stresses, from White et al. (2005)

2.4.4 Friction and fatigue


Those two concepts arise again from the installation process of the monopiles (although
it is applicable for any type of pile), and remarks again the massive influence of the

2.4. Installation of monopile foundations M.Sc. Civil Engineering


Chapter 2. Revision of the State of the Art 28

construction process in the eventual behaviour and response of the wind turbine in
general and the foundation in particular. Different authors have stated that the soil
particles around the shaft of the monopile are actually broken soil particles, which lead
to a contraction of the interface layer of soil and a progressive densification [White et al.
(2005)]. According to these tests a reduction of lateral stress and local shaft friction
as a function of pile displacement is visible from the measurements. This reduction of
lateral stress varies over the length of the pile and depends on the magnitude and cycles
imposed by the installation method [White and Lehane (2004), White et al. (2005)].
White and Lehane (2004) also showed that the friction fatigue in higher for stiff soils,
demonstrating that the reduction of lateral stresses is stronger for driven than for jacked
piles (one-way installation cycles). It is also demonstrated that the number of loading
cycles has a greater influence on the contraction of the interface layer. The influence
of this papers is relevant when studying contact elements and how to model interfaces
between piles and soils, although to the best of my knowledge, no research has been done
with this respect, trying to link the results from the frictional fatigue to a constitutive
law governing the behaviour of the contacts.

2.4.5 Stress field alteration

As the monopile is driven into the seabed, the stresses around the pile toe and the pile
body are altered, as it can be seen in Figure 2.6. White et al. (2005) tried to work
around the IFR and FFR to parametrise the influence of the plugging in relation to
the type of the pile (open-ended, close-ended and so on). Probably, the most extensive
and successful analyses in this topic are presented by Richard Jardine, resulting in the
recommendations of the ICP-05 method [Jardine et al. (2005)], but most of the research
existing in this topic is pure laboratory based with scale models and for close-ended piles
[Jardine et al. (2009a)].

Richard J. Jardine reviewed the potential effects of cyclic loading on offshore piles,
and considered how these may be addressed in practical design. They outlined the in-
dicative ranges for cyclic loading components that might apply to the range of multiple
pile structures, noting that the loads vary with platform weight, water depth, ocean
environment, and structural forms [Rimoy et al. (2013)].

During the standardised process ICP-05 (ICP stands for Imperial College Pile), it
was shown that the installation imposes a two-way cyclic failure, with both contractant

M.Sc. Civil Engineering 2.4. Installation of monopile foundations


29 Chapter 2. Revision of the State of the Art

and dilative phases of interface shear developing during each stroke of the installation
[Jardine et al. (2009b)] and the stress development at points away from the shaft, at the
end of installation, appear to be only weakly dependent on the total number of jacking
cycles [Jardine et al. (2009b)], but the stresses developed at any point depend principally
on the local tip resistance and the spatial position relative to the pile tip [Jardine et al.
(2009b)].

.
Figure 2.7: Experimental set-up and calibration chamber for ICP-05 testing, from
Jardine et al. (2013), Jardine et al. (2009a)

However, The number of jacking cycles had little influence on the soil mass stresses,
and no strong time effects were evident. These parameters are thought to be more im-
portant at the pile-soil interface, during dynamic installation and over extended periods
of time [Jardine et al. (2013)]. See Figure 2.8.

Different plots showed intense stress concentrations emanating from the tip, with soil
located within 10 times de radii of this moving focus experiencing a relatively high-level
stress cycle during each installation jack stroke [Jardine et al. (2013)]. This is corrobor-
ated by Dijkstra and Broere (2009), who carried out photo-elastic tests to understand
the effects of the pile toe (see Figure 2.9 ). Particle crushing and shear band formation
processes have a key bearing on the phenomena observed [Jardine et al. (2009b)].

2.4. Installation of monopile foundations M.Sc. Civil Engineering


Chapter 2. Revision of the State of the Art 30

Figure 2.8: Moving and Stationary radial stresses after different jacking strokes, from
Jardine et al. (2009b)

Figure 2.9: Photoelastic effect in crushed glass near the pile toe, from Dijkstra and
Broere (2009)

The analysis of these results is complicated. Interpretation of the dip in radial stresses
developed above the tip as being owed to the strain paths developed around the pile tip,
as well as the effects of cyclic loading, particle breakage, creep and stress relaxation is a
feasible option. These processes were found to persist as penetration continues, and are
thought to contribute to the steady decline [Jardine et al. (2013)].

M.Sc. Civil Engineering 2.4. Installation of monopile foundations


Chapter 3

Finite Element Model with


ABAQUS

A numerical model has various advantages worth recalling. Compared to a physical


model, they are much more cheaper and require less time to produce. However, even-
tually, a physical model, or data from the field will be required in order to calibrate
and validate the numerical model. Once this has been done, the model can be used and
reused for multiple purposes, such as design, parametrisation, inverse analysis and so
on. Although nowadays almost every phenomena in engineering can be modelled (with
better or worse results), one of the more limiting factors is the computational cost of
the problem as it has an effect on the total computational time. So the feasibility and
accuracy of the modelling process is sometimes subdue to it, as it was discussed in the
introduction to this thesis when writing about the limitations of the model.

In this chapter, the basics of numerical simulation with ABAQUS will be discussed
(in regards to this project). The following sections are aimed to describe those key points
in the process of simulating the behaviour of laterally loaded monopiles that needs to be
understood in order to extract the most of this document. Some consideration during
the modelling processes are underpinned in this chapter as well. Another objective of
this chapter it to set a common list of vocabulary an terms, so when further in the thesis
different aspects of the modelling process are referred to or named, those can be easily
understood reading the remarks hereby stated. Some of the issues and decisions made
based on the troubles that were popping up when progressing in the simulation are also
briefly discussed in this chapter.

31
Chapter 3. Finite Element Model with ABAQUS 32

3.1 Brief remarks about ABAQUS


ABAQUS is a multi-purpose finite element code which is composed by the CAE (Com-
plete ABAQUS Environment or Computer-Aided Engineering), the Viewer (pre and
post-processors), and the Standard/Explicit processors (solvers).

The CAE is a complete interface that provides a simple base for creating, submitting,
monitoring, and evaluating results from ABAQUS/Standard and ABAQUS/Explicit
simulations. ABAQUS/CAE is divided into modules, where each module defines a
logical aspect of the modelling process. For instance, defining the geometry, defining
material properties, and generating a mesh. Moving from module to module, you build
the model from which ABAQUS/CAE generates an input file that you submit to the
ABAQUS/Standard or ABAQUS/Explicit analysis product. The analysis product per-
forms the analysis, sends information to ABAQUS/CAE to allow you to monitor the
progress of the job, and generates an output database. Finally, you use the Visualization
module or (Viewer) of ABAQUS/CAE to read the output database and view the results
of your analysis.

The most experienced users of this software can use a text editor for all the modelling
stages, creating a .INP file with the commands in ABAQUS. For the sake of this thesis,
the user interface has been used most of the time, because for beginners it is the most
efficient way to get acquainted with the software. However, in some cases, the .INP file
had to be used, as several options and commands of analysis are not incorporated to the
user interface and has to be written in a text file.

3.2 The Modules


In order to define the physical properties of the model, ABAQUS uses the so called
Modules. The process of defining the model must follow a cascade scheme dictated by
the modules. The list of Modules and its functions are specified here:

1. Part module: It creates a feature, captures your design intent and contains geo-
metry information as well as a set of rules that govern the behaviour of the geo-
metry. Essentially, the part contains the information that defines the geometry,
systems of coordinates, planes of reference, etc.

2. Property module: In this module, the materials, section profiles, assign sections,

M.Sc. Civil Engineering 3.1. Brief remarks about ABAQUS


33 Chapter 3. Finite Element Model with ABAQUS

orientations, normals, and tangents to the parts are defined.

3. Assembly module: When you create a part, it exists in its own coordinate system,
independent of other parts in the model. In contrast, you use the Assembly module
to create instances of your parts and to position the instances relative to each
other in a global coordinate system, thus creating the assembly. You position part
instances by sequentially applying position constraints that align selected faces,
edges, or vertices or by applying simple translations and rotations. This Module
has a paramount importance in this thesis, as it allows to joint the monopile
foundation part with the soil part, each of those with different properties.

4. Step module: It creates the analysis steps, specifies the output requests or the
adaptive meshing (feature not used in this work) and specifies analysis controls,
such as exit gates for iterations. Within a model you define a sequence of one
or more analysis steps. The step sequence provides a convenient way to capture
changes in the loading and boundary conditions of the model, changes in the way
parts of the model interact with each other, the removal or addition of parts, and
any other changes that may occur in the model during the course of the analysis.
In addition, steps allow you to change the analysis procedure, the data output, and
various controls. You can also use steps to define a linear perturbation analysis
about non-linear base states. You can use the ”replace“ function to change the
analysis procedure of an existing step. The steps are very important in the dynamic
analysis. Another noteworthy feature of the step module is the initialisation of the
geostatic conditions in the soil.

5. Interaction module: It defines the mechanical and thermal interactions between


regions of a model or between a region of a model and its surroundings. It has
been used to define the contact interactions between the monopile an the soil, and
the use of damping elements.

6. Load module: It sets load conditions, boundary conditions, load cases and pre-
defined fields (such as stress fields). Prescribed conditions are step-dependent
objects, which means that you must specify the analysis steps in which they are
active. You can use the load, boundary condition, and predefined field managers
to view and manipulate the stepwise history of prescribed conditions. The Load
module and the Step module are very related one another.

7. Mesh module: It contains the tools and algorithms to mesh the geometry defined
in the part sections. Meshes can be associative or independent (depending on

3.2. The Modules M.Sc. Civil Engineering


Chapter 3. Finite Element Model with ABAQUS 34

our model). It also features seeds, mesh techniques, and element types. The
seed feature was very useful for generating variable meshes, which will be later
explained.

8. Optimization module: This module was not used in any of the calculations, al-
though it can be used to change the topology of the model.

9. Job module: the Job module was used to create and manage analysis jobs and
to view a basic plot of the analysis results, also to create and manage adaptivity
analyses and co-executions (paralellisation or use of various processors and GPU
enhancement).

3.3 Geometry and Boundary conditions


The geometry is one of the most important aspects of a model. It represents the math-
ematical domain used to build the stiffness matrix of the FE model, but it also has
influence in the mesh properties, such as the meshing algorithm (swept, free, etc.). As a
representation of the domain, the boundaries are also represented, and they can play an
important role in the model, as they can generate noise or disturbances when reading
the results from the calculations. This assuming a static analysis. During a dynamic
analysis, the boundaries represent hard points in which the energy from the loads can
reflect and, again, distort the results.

Apparently, it seems obvious that in order to avoid distortion from the contours, a
bigger model is required. But a bigger model implies that, in order to maintain accur-
acy, more elements are needed, which means more calculations. If we decide to keep the
number of elements to a minimum, we may find that there are geometrical distortion
issues with the meshing, and again the accuracy of the method is clearly affected.

Now, within the different elements present in the software, the estimation of the
field of displacements can be done using linear or quadratic formulation, and different
number of nodes per element, which is translated into more equations when solving
the stiffness matrix of the model. Therefore, it is important to notice that geometry,
mesh and element type are related in this particular research. This has been conflict-
ive during the development of the numerical models, as the computation capacity was
small. As an idea, the models were ran in a laptop equipped with a dual core pro-
cessor and 8 Gb of RAM through a virtual machine (as the OS could not handle the

M.Sc. Civil Engineering 3.3. Geometry and Boundary conditions


35 Chapter 3. Finite Element Model with ABAQUS

version of ABAQUS). As seen in papers and tech reports, similar problems were solved
using parallel computation (up to 16 high-performance processors) with 256 Gb of RAM.

In most of the cases, and particularly in the final model, the domain is a semi-
cylinder. Although there is no physical correlation, the use of a semi-cylinder is the
preferred option for most of the people doing research on pile foundations [Achmus et al.
(2009)], but a rectangular domain can also be used [Bourgeois et al. (2010)]. Some of the
advantages of the use of a cylindrical domain in ABAQUS is the use of sweeping meshing
techniques, in which de mesh is applied following a certain parametric curve or sweeping
path. Besides, as the shape of the pile is cylindrical, it seems logic to think in radial
coordinates instead of rectangular coordinates. Another benefit is that the distribution
of the nodes in the mesh can be denser in the centre close to the pile than in the outer
regions, creating elements of similar sizes at the same time that the radii is increasing.
That avoided distortion problems in adjacent elements.

Figure 3.1: Final geometry of the soil, after Achmus et al. (2009). The relative distances
were taken wo the influence of the boundaries is minimised in the surroundings of the
monopile.

Special attention has been paid to the geometry of the monopile and the soil inside
the monopile. The monopile has been considered as a solid body (including soil and

3.3. Geometry and Boundary conditions M.Sc. Civil Engineering


Chapter 3. Finite Element Model with ABAQUS 36

Figure 3.2: Geometry of the pile

steel) and not as two separate interfaces. It has been assumed that the monopile is fully
filled with soil and it has a tie constrain with the inner skin of the monopile. As the
behaviour of the soil inside the monopile is out of the scope of this thesis, its effect has
been neglected (in terms of shearing stresses) albeit it has an added value to the bending
resistance and mass of the model, and a consideration on its weight in terms of density
of the pile. The hypothesis of modelling the monopile as a solid element was proved
right by Achmus et al. (2009), and it has been introduced to this work. The material
properties assigned to this element were obviously arranged to be equivalent to the case
we were modelled. As this thesis is more focused on the soil behaviour, the outputs in
terms of stresses for the monopile were disabled, so computer resources were saved.

With respect to the boundary conditions, the bottom is fixed, the symmetry plane
has a symmetry boundary condition and the external face has the vertical displacement
unconstrained to allow for vertical deflection or settlement.

3.4 Mesh

Once those previous considerations about geometry has been considered, the rest of
the models shared similar characteristics in terms of finite element types, meshing and

M.Sc. Civil Engineering 3.4. Mesh


37 Chapter 3. Finite Element Model with ABAQUS

meshing algorithms. During the post-process phase of ABAQUS, it carries out several
checks on mesh properties. For instance, it establishes an automatic mesh-distortion
criteria, so if it founds that some of the elements do not satisfy these requirements,
those are marked and can be replaced or the mesh can be redefined to avoid element
distortion (angular distortion). When computing the solution, it was checked that the
model complied with the distortion criteria.

Figure 3.3: Mesh of the soil and the monopile

In the soil, the total number of nodes is 8803 and the total number of elements is
1824, 970 nodes and 924 elements for the soil with a total volume of 137722.89 m3 (soil)
and 543.90 m3 (pile). This is only valid for the final model, which has been used for the
most complex static and dynamic calculations and has been object of validation. Other
models have been used, but in the end were not relevant for the final objective of the
document.

3.4.1 Element types

ABAQUS contains a library of solid elements for two-dimensional and three-dimensional


applications. The two-dimensional elements allow the modelling of planes, axisymmet-
ric problems and include extensions to generalized plane strain (when the model exists

3.4. Mesh M.Sc. Civil Engineering


Chapter 3. Finite Element Model with ABAQUS 38

between two planes that may move with respect to each other, providing thickness dir-
ection strain that may vary with position in the plane of the model but is constant
with respect to thickness position). The material description of three-dimensional solid
elements may include several layers of different materials, in different orientations, for
the analysis of laminated composite solids. A set of nonlinear elements for asymmetric
loading of axisymmetric models is also available, and linear infinite elements in two and
three dimensions can be used to model unbounded domains.

The solid element library includes isoparametric elements: quadrilaterals in two di-
mensions and “brick” (hexahedra) in three dimensions. These isoparametric elements
are generally preferred for most cases because they are usually the more cost-effective
of the elements that are provided in ABAQUS. For practical reasons it is sometimes
not possible to use isoparametric elements throughout a model; for example, some mesh
generators use automatic meshing techniques that rely on triangulation to fill arbitrarily
shaped regions. Because of these needs ABAQUS includes triangular, tetrahedron, and
wedge elements.

Solid elements are provided with first-order (linear) and second-order (quadratic)
interpolation, and the user must decide which approach is more appropriate for the ap-
plication. Some of the guidelines are as follows. Standard first-order elements are essen-
tially constant strain elements: the isoparametric forms can provide more than constant
strain response, but the higher-order content of the solutions they give is generally not
accurate and, thus, of little value. The second-order elements are capable of represent-
ing all possible linear strain fields. Thus, in the case of elliptic problems—problems for
which the governing partial differential equations are elliptic in character, such as elasti-
city, heat conduction, acoustics, in which smoothness of the solution is assured (much
higher solution accuracy per degree of freedom is usually available with the higher-order
elements). Therefore, it is generally recommended that the highest-order elements avail-
able be used for such cases: in ABAQUS this means second-order elements. Although
further approximations exist, little is gained by going beyond the second-order elements,
so ABAQUS does not offer any higher-order forms.

Another application of the use of quadratic elements is the fact that for non-linear
analysis, the strains are calculated as the integral of the rate of deformation, instead of
directly from the deformation gradient matrix (which is a first order approximation and
assumes small strains). According to the Theory Guide [Simulia (2010)] of ABAQUS,

M.Sc. Civil Engineering 3.4. Mesh


39 Chapter 3. Finite Element Model with ABAQUS

second-order elements provide higher accuracy in ABAQUS/Standard than first-order


elements for “smooth” problems that do not involve severe element distortions. They
capture stress concentrations more effectively and are better for modelling geometric
features as they can model a curved surfaces with fewer elements.

Another feature of ABAQUS is the use of reduced integration. Second-order reduced-


integration elements in ABAQUS/Standard generally yield more accurate results than
the corresponding fully integrated elements. Reduced integration usually means that an
integration scheme one order less than the full scheme is used to integrate the element’s
internal forces and stiffness. Superficially this appears to be a poor approximation, but
it has proved to offer significant advantages. For second-order elements in which the
isoparametric coordinate lines remain orthogonal in the physical space, the reduced-
integration points have the Barlow point property (the strains are calculated from the
interpolation functions with higher accuracy at these points than anywhere else in the
element). Furthermore, reduced integration decreases the number of constraints intro-
duced by an element when there are internal constraints in the continuum theory being
modelled, Finally, reduced integration lowers the cost of forming an element, i.e. a fully
integrated, second-order, 20-node three-dimensional element requires integration at 27
points, while the reduced-integration version of the same element only uses 8 points and,
therefore, costs less than 30% of the fully integrated version, according to the Theory
Guide [Simulia (2010)]. This cost savings are specially significant in cases where the
element formation costs dominate the overall costs, such as problems with a relatively
small wave front and problems in which the constitutive models require lengthy calcu-
lations, which happens to be the case of the simulations carried out in this thesis.

After all these considerations, and bearing in mind all the recommendations by the
Theory Guide of ABAQUS [Simulia (2010)], the element type chosen from the library
was the C3D20R, a stress/displacement element for standard calculation, 20-node quad-
ratic brick, with reduced integration and hourglassing control (check to avoid null-stress
integration points). 2D models used the equivalent bidimensional element.

Pore water pressure has been neglected. In case of consideration, those element type
would be no longer of use, as there are special elements for poro-elastic problems.

3.4. Mesh M.Sc. Civil Engineering


Chapter 3. Finite Element Model with ABAQUS 40

Figure 3.4: a) 20-node three-dimensional finite element C3D20R and b) integration


points per face

3.4.2 Meshing algorithm

The mesh was created by density specification, creating seeds along the edges of the
model to indicate where the corner nodes of the elements should be located. By doing
this, a finer mesh is allocated close to the pile (the main area of interest) and coarser
close to the boundaries of the domain. The algorithm followed a top-down generation,
in which the 3D model is created in first place, and then the geometry is meshed. In
order to control the meshing algorithm, two preferred options (aside from the use of
seeds and partitions) were used: structured and swept mesh. Structured meshing is
the top-down technique that gives you the most control over your mesh because it ap-
plies pre-established mesh patterns to particular model topologies. The monopiles were
meshed with this generation technique. Swept meshes are generated by internal creation
of the mesh on an edge or face and then sweeping that mesh along a sweep path. The
result can be either a two-dimensional mesh created from an edge or a three-dimensional
mesh created from a face. This technique resulted very useful when meshing the semi-
cylinder of the soil domain, as the sweeping path given was conformed by the curvature
of the external circumferences.

Another important part was the mesh transition between the monopile and the soil.
According to the ABAQUS guide, it is recommended to mesh each of the instances
separately using independent meshes. Independent meshes consume more computer re-
sources, but are necessary when there is no continuity in the material properties of the
different parts of the model.

In Figure 3.5 the importance of the seeds can be seen. The seed were variable in

M.Sc. Civil Engineering 3.4. Mesh


41 Chapter 3. Finite Element Model with ABAQUS

each of the 3 components to make a thinner mesh around the monopile. The density
of elements in the monopile has to be sufficient as well. As the contact elements use
the monopile surface as the master surface, enough number of elements are required to
compute the interaction forces accurately.

Figure 3.5: Seeding algorithm. The elements are reduced from the external to the
internal region by a constant 10%. The external blocks have a lenght of 7.2 metres, and
the closer to the monopile 1.0 metres.

3.5 Soil constitutive models

Selecting the constitutive model in ABAQUS was a task that required a lot of time
for this document, as it was important to revise the advantages and disadvantages that
each of those presented. Another problem was the lack of soil parameters. Soil testing
in marine environments is very complex, and the data available are scarce. This means
that, albeit a complex model is required, if the characteristic parameters of a soil for a
certain ground condition are unknown, the constitutive law cannot be used.

3.5. Soil constitutive models M.Sc. Civil Engineering


Chapter 3. Finite Element Model with ABAQUS 42

3.5.1 Revision of theory of plasticity in soils

The soils present in the North Sea are characterised as loose sands, which are equival-
ent to normally consolidated clays (drained case). Due to a rearrangement of the soil
particles, they tend to show hardening under higher deviatoric stresses. Traditionally,
the fundamentals of plasticity in soils have been supported by 3 concepts: a yielding
surface, a flow rule and a hardening law.

• Yielding surface: a strength criteria for a certain organisation of the particles of


the soil, establishing a limit to admissible stress states. In the p0 − q space it is
defined by a convex function satisfying

F (p0 , q, X) ≤ 0 (3.1)

• Flow rule: it is a measure of the reorganisation of the particles in the soil. It


establishes the flowing character of the deformation rate.

dε̄¯P = dλ · →

n (3.2)

where dλ represents the plastic multiplier and →



n = ∂G
¯
∂ σ̄ being G the plastic potential
¯
function and σ̄ the second-order stress tensor.

• Hardening law: it represents the change in the parameters that define the reorgan-
isation of the particles.
dX = f (dε̄¯P ) (3.3)
∂X
with dX = ∂ ε̄¯P
∂ ε̄¯P , being X the hardening function.

The traditional critical state models present isotropic hardening with associative flow
rule. This is equivalent to say that the plastic multiplier (dλ) is a scalar. However, loosing
sands as the soils in the North Sea from where the samples have been investigated, are
better represented by a non-associative flow rule. This means that the gradient of the
plastic potential and the gradient of the yielding surface are not coincident.
Another concept that had to be understood in order to use the best possible con-
stitutive model for the soil was the Consistency Condition introduced by Prager. In a
plastic process, the actual value of the plastic multiplier is obtained by prescribing that,
upon plastic loading conditions, the stress state must remain on the yielding surface.
This is important when the plastic multiplier is not an scalar, and must be defined as:

M.Sc. Civil Engineering 3.5. Soil constitutive models


43 Chapter 3. Finite Element Model with ABAQUS

Figure 3.6: Non-associative flow rule

 
1 ∂F ¯¯ e
dλ = · C̄ · dε̄¯ (3.4)
kbp ¯
∂ σ̄

with
∂F ¯¯ e ∂G ∂F
kbp = · C̄ · − · dX (3.5)
¯
∂ σ̄ ¯
∂ σ̄ ∂q
¯
where C̄¯ e represents the constitutive tensor of fourth order of the soil.

The failure criterion imposes limits on stress states that can be reached or not.
The plastic potential function produce relative magnitudes of plastic strain increments
(its functionality is similar to Poisson’s ration in elastic strain). The yielding criteria
determines when plastic increments occur, and the hardening/softening function defines
the magnitudes of plastic strain increments.

3.5.2 Cyclic behaviour of sandy soils

With those previous remarks of plasticity, and bearing in mind mind that in the end,
a dynamic process should be modelled, choosing the most appropriate constitutive law
for the soil became an important objective of this thesis, so the available options in
ABAQUS were explored. The keywords for the search were dynamic hardening, non-
associative flow rule, isotropic material, sands, cyclic loading and rate dependency. In
the library of plastic models in ABAQUS we can find models for metals and structures
and “other” plasticity models, in which the soil models are included. The summary of
the different models included in ABAQUS is presented below:

3.5. Soil constitutive models M.Sc. Civil Engineering


Chapter 3. Finite Element Model with ABAQUS 44

• Extended Drucker-Prager: Used to model frictional materials, presents pressure-


dependent yield, allows isotropic hardening, can be used with progressive damage
and failure criterion and it is intended for monotonic loading conditions.

• Modified Drucker-Prager/Cap Model: Used in cohesive soils, presents pressure-


dependent yield, allows inelastic hardening and it is intended for monotonic loading
conditions. It adds a “cap” to the extended Drucker-Prager.

• Mohr-Coulomb Plasticity: For frictional materials, allows isotropic hardening and


it is intended for monotonic loading conditions.

• Critical State (Clay) Plasticity Model: For clays, not sands.

• Crushable Foam Plasticity Model: Intended for energy absorption problems and
not for soils.

What could be extracted from the study of the different constitutive models is that
none of them represents the behaviour we wanted to analyse. Plus, the Mohr-Coulomb
and the Drucker-Prager model are a failure criterion which has been adapted to a con-
stitutive model. Although this might not be correct from the point of view of soil
mechanics, it is the sort of common trend for complex simulation [Achmus et al. (2009),
Damgaard et al. (2014), Bourgeois et al. (2010)], i.e. essentially, in theory those are
not constitutive models. Apart from this, most of these models depended on multiple
parameters which are not available in traditional soil testing. For instance, the crushable
foams model rely on 6 different parameters, or the Cap models, which let you choose what
kind of cap it is needed to form a linear, hyperbolic or exponential yield surface function.

3.6 Material properties


In ABAQUS, the materials are defined by assigning properties and parameters, e.g. for
the monopile, a isotropic linear elastic behaviour was chosen, so the input parameters
are the elastic modulus and the Poisson’s ration. So basically ABAQUS has a library
of constitutive models which can be selected. As said, the material behaviour of the
steel of the monopile was straightforward, albeit the soil constitutive behaviour required
more research. If a constitutive law is not found in the ABAQUS library, it can be
implemented using UMATs (User MATerials), which are text documents containing the
information for ABAQUS to compute. Although this was considered, the use of UMATs

M.Sc. Civil Engineering 3.6. Material properties


45 Chapter 3. Finite Element Model with ABAQUS

presents huge handicaps, being extremely inadvisable the use of them.

With this perspective, an extensive research on how the different experts in this mat-
ter modelled the soil behaviour in monopile foundation has been carried out. The fact
that, up to date, no very much has been done in laterally cyclic loaded monopiles, was
not helping. The approach that was adopted was taken from professor Martin Achmus’
recommendation, whom decide to adopt a simple constitutive model (Mohr-Coulomb)
combined with a degradation of the stiffness of the soil, because parameters were avail-
able, and the fact that using complex constitutive model may become unsuccessful due
to the fact that numerical errors are likely to accumulate and distort the results. Us-
ing a Mohr-Coulomb method also allowed to compare results in terms of stresses and
displacements with those obtained by professor Achmus to make sure that the order of
magnitude of the values was correct and in range.

The soil parameters of the model were as follows:

Table 3.1: Soil parameters for general soil model (Mohr-Coulomb)

E 40000 kP a

ν 0.25

c0 1 kP a

φ 35o

δ 5o

γ 20.00 kN/m3

The steel of the monopile has the parameters of Table 3.2:


However, following Achmus’s hypothesis, the monopile was modelled as a rigid body
with equivalent properties. The equivalences represent the change from a hollow cylinder
to a solid cylinder. The two properties that needed to be changed were the mass (m) and
the bending stiffness of the monopile ((EI)m ), as in this thesis the only effects analysed
are the lateral ones. First, the mass (mm ) of a monopile of length L, outer diameter Do
and thickness t plus the mass of the soil inside has to be equal to the mass (ms ) of the
solid body of length L and diameter D:

3.6. Material properties M.Sc. Civil Engineering


Chapter 3. Finite Element Model with ABAQUS 46

Table 3.2: Steel for monopiles

E 210000000 kP a

ν 0.3

γ 78.50 kN/m3

mm + msoil = ms
π(Do2 − (Do − t)2 ) π(Do − t)2 πD2
Lγsteel + Lγsoil = Lγeq
4 4 4
D2 − (Do − t)2 (Do − t)2
γeq = o γ steel + γsoil
D2 D2

Getting γeq = 21.16415 kN/m3 . The same procedure applies for the bending stiffness
and the Young’s modulus:

(EI)m = (Eeq I)s


(EI)m
Eeq =
Is
[Do4 − (Do − t)4 ]
Eeq = Esteel
D4

Having Eeq = 8274838 kP a. This is just a mere procedure to make sure that the
behaviour and weight of the solid body is the same than the monopile. The Poisson’s
ratio was conserved without alteration.

Table 3.3: Steel equivalent parameters for “solid” monopiles

E 8274840 kP a

ν 0.3

γ 21.17 kN/m3

M.Sc. Civil Engineering 3.6. Material properties


47 Chapter 3. Finite Element Model with ABAQUS

3.7 Contacts and interfaces

When dealing with foundations, the contact between surfaces need a special character-
isation. There is an important interaction between the monopile and the adjacent soil
which has to be considered in order to achieve the most accurate and representative
solution of the physical phenomena. In this model, the focus has to be put on the con-
tact surface (external skin) between the monopile and soil. It is remarkable that for all
cases, the soil inside the monopile is assumed to have a tie constrain to the monopile.
This means that it counts for mass and bending resistance, but not for shearing stresses.
The monopile is in fact a solid block made of soil and steel, with equivalent properties.
Because of the loads acting on the monopile, the steel is subjected to high stresses, and
those need to be transmitted to the mesh of the soil. It is therefore inevitable that, when
two very different material are in contact, they behave quite differently. So the question
is now how to accommodate the displacements and stresses of the monopile with the
displacements and stresses of the soil. The numerical analysis of the interfaces of the
model require a very careful treatment. The critical surface were interaction properties
have been applied are critical in terms of convergence criteria, giving problems most of
the time, mainly because this contact elements are intended for explicit analysis and not
for implicit analyses.

ABAQUS allows two types of solutions: contact elements and node coupling, and a
combined one.

• Contact Elements: it requires the duplication of the nodes of the contact surface. It
introduces a fictitious layer (null thickness) between the elements of the monopile
and the soil. This layer has normal stiffness and a shear stiffness and allows the
separation of the nodes of the pile and soil, i.e., no physical contact, something
that has been demonstrated experimentally. This solves the problem of finding
tensile stresses in the soil continuum. On the other hand, if a compressive stress is
acting, the tangent shear moduli allow for the development of shear stresses. This
friction is relevant in the problem, not only in terms of stresses, but also in terms
of energy dissipation when dealing with dynamic analysis. However, this algorithm
increases considerably the computational cost of the problem and incorporate new
parameters to the model (the frictional law of the interfaces), which would require
special characterisation and a special treatment. For simplicity, a conventional
tangential frictional law has been used, as recommended by various authors.

3.7. Contacts and interfaces M.Sc. Civil Engineering


Chapter 3. Finite Element Model with ABAQUS 48

• Coupling Nodes: it requires the duplication of the nodes in contact, one of the
nodes is part of the upper layer and the other one is part of the bottom layer.
Between opposite nodes there is a kinematic constraint. This condition is applied
to the “master” node, and the other, the “slave” node faces the condition imposed
by the master. The normal movement with respect to the contact surface is free
whereas the relative displacement between surfaces is restrained. This algorithm
provides a solution for the discontinuity of stresses and displacements between the
monopile and the soil without increasing the computational time. However, it
does not take into account the friction between the elements and the associated
dissipation of energy.

There is a possibility of combining both methods, and it is defining surfaces instead


of nodes (surfaces are associations of nodes in ABAQUS, but defined previously to the
mesh). This means that contact elements can be used in combination with master and
slaves surfaces. According to the Theory Guide, the master surface has to be the one
with more toughness in comparison to the slave surface. Therefore, the skin of the mono-
pile constitutes the master surface and the adjacent soil the slave surface. In between,
null thickness interface elements with a frictional constitutive law. However, the increase
in computational time may become restrictive in some cases, the convergence is in most
cases not guaranteed and the determination of a frictional law between the steel and the
soil is not obvious. As a first approach, the friction can be estimated using geotechnical
criteria assuming perfect interaction between concrete and granular materials. This is
assumed as 2/3 of the friction angle of the soil.

For an initial approach, this could be valid, but the contacts were subject of study
in this thesis. The model selected for the interaction of the soil and the monopile was
defined as purely frictional. This is what the authors in the literature reviewed go for.
The reason why is because it is the simplest of the models, as it only requires two para-
meters (friction coefficient and limit shear stress). However, there is a simplified version,
which is the one that have been used to model all the contact behaviours in the FEM
model with ABAQUS, which is the frictional law with small sliding. The “small sliding”
just means that there is no need to impose a limiting shear stress (mainly because of the
lack of experimental test or field data in this matter). The analysis on this thesis focused
on determining how much the monopile separates from the soil and the influence of the
variation of the friction coefficient in terms of energy dissipation and shearing forces.

M.Sc. Civil Engineering 3.7. Contacts and interfaces


49 Chapter 3. Finite Element Model with ABAQUS

As µ = 0.3 gave more conservative results, this will be the value of the parameter
for the general soil model.

3.8 Loads
In general, the main load components to be taken into account for the structural assess-
ment of the wind turbine’s foundation are the turbine’s own weight and those produced
by the wind and wave action. Further loads, as for instance those originating from sea
currents, sea ice, earthquake motion, operational loads or a boat impact, may play a sig-
nificant role depending on the specific site location, and can be estimated on the basis of
the usual offshore standards. Regarding the monopiles, their behaviour is mainly influ-
enced by the fact that the horizontal loads in the turbine are much more significant than
the vertical loads, when in traditionally pile foundations an opposite scenario is found.
The loads, as described in the codes, can be divided in two groups: hydrodynamics and
aerodynamics.

• Hydrodynamic loading: In addition to the drag and inertia forces exerted by the
passing waves on the offshore structure, further loads, such as those caused by
the sea currents, by the breaking waves or by the wave-slamming (impact) effects,
need to be considered. They can all be loosely termed as hydrodynamic loads, and
depend mainly on the kinematics of the waves, the density and depth of the water
and on the shape of the support structure.

The calculations of the hydrodynamic loads usually follow the next sequence:

1. Consideration of the local site statistics of the sea state parameters (wave
height and period), the maximum wave height and corresponding period are
determined for both the extreme (50 year return period) and service (1 year
return period) storm conditions. A first approximation may be obtained
from charts or by means of empirical formulas relating the wind velocity to
the significant wave height and period, such as the JONSWAP formulation.
2. Then, the “design wave” kinematics (wave length and water velocities and
accelerations) are defined at every depth by means of a suitable wave theory
(for instance, linear wave theory)
3. The distributed drag and inertia forces per unit length of submerged structure
can be then estimated by introducing the water particle velocities and accel-
erations into the Morison equation. This equation also requires the provision

3.8. Loads M.Sc. Civil Engineering


Chapter 3. Finite Element Model with ABAQUS 50

of the hydrodynamic coefficients, which may be obtained from the literature


and mainly depend on the Reynolds and Keulegan-Carpenter numbers as well
as on the surface roughness of the structure. This values can be found in the
codes.
4. The resulting lateral loads can be computed then by integration of the dis-
tributed forces along the length of the supporting structure that is exposed
to the action of the waves.

• Aerodynamic loading: The flow of air and its interaction with the wind turbine
produces in general both quasi-static and dynamic loads on the structure. They are
conditioned by the design parameters (e.g. the aerodynamic shape of the blades,
the rotational speed of the rotor, the turbulence generation, etc.) as well as by the
wind characteristics (average wind speed at hub height, speed distribution with
height, air density, etc.), which are stochastic by nature. The aerodynamic loads
acting on the rotor itself (thrust, drag and lift forces) are usually computed com-
bining the so-called Blade Element Momentum Theory (BEM) and vortex theory
for small perturbations, usually with CFD software or with a commercial software
called FAST, which include fatigue analysis. Additionally, there are a number of
factors that may have a relevant influence on the wind loads, as for instance the
wind field perturbations caused by the structure itself (wake and upwind effects),
vorticity, stall and aeroelastic effects, possible flow asymmetries or the dynamic
response when the turbine is at standstill conditions under strong wind, and, there-
fore, their importance should also be assessed during design.

The calculations of the aerodynamics loads usually follow the next sequence:

1. Determination of the temporal statistics for the wind speed at the wind-farm
site and fitting to a certain distribution (Weibull generally).
2. Then, the wind speed at a given height (usually 10 m) is calculated for the
extreme event, for instance the strongest wind gust (3 seconds in duration)
with a recurrence period of 50 years.
3. The maximum wind speed for service conditions is defined by the rotor design
(the cut-off speed above which the blades are rotated in order to offer min-
imum resistance to wind flow).
4. Calculation of the wind-speed distribution with height, for instance by means
of the Hellmann distribution.

M.Sc. Civil Engineering 3.8. Loads


51 Chapter 3. Finite Element Model with ABAQUS

5. The quasi-static pressure of the flowing wind upon the structure at every
height can be then estimated through the Bernoulli equation.
6. The integration of the pressure over the structure, considering the exposed
surfaces, the angle of attack and the aerodynamic form coefficients, provides
finally the total horizontal loads.

Nevertheless, The consideration of a single wave or wind gust cannot possibly rep-
resent the loading conditions of the offshore wind turbine, since the sea states and wind
conditions are irregular and stochastic in nature, and the structure will be subjected to
billions of load cycles of different magnitude during its lifetime. For the assessment of the
turbine’s response to extreme short-time events, a transient calculation can be performed
by adopting a suitable load history time signal of a certain length (for instance, 600s).
Such time signal can in principle be derived by means of an inverse Fourier transform
of a wave spectrum or be directly based on measured data of wind speed and sea water
level. However, in order to assess the performance of the structure during its design life,
it is necessary to extrapolate the short-time data in some suitable way that provides
the magnitude and occurrence of both the large cycles in low numbers (i.e. the extreme
events) as well as the lower cycles in large numbers (relevant for fatigue analysis where
appropriate).

Neither the complete monopile nor the whole turbine have been modelled because of
the limited computational resources available. The analysis only considers the interac-
tion between the monopile and the soil from the mudline and downwards. This implies
that the representative load must be translated from the whole structure to only the
seabed line. Luckily, the values of the loads a the mudline level were available and could
be used. The design loads at mudline for the North Sea have been obtained from Lesny
and Wiemann (2005) which are the ones which the DOWEC (2003) are based on. This
is the report from which conditions, dimensions and actions have been taken. To the
best of the author’s knowledge, blade and aero-elastic effects are not considered in the
actions gathered in Table 3.4.

To give an order of magnitude, the environmental conditions in which this loads act
are presented in Table 3.5.

This thesis only considers the horizontal action at mudline. As the model uses
symmetry, the load has to be divided by 2. One of the objectives of this thesis is to find

3.8. Loads M.Sc. Civil Engineering


Chapter 3. Finite Element Model with ABAQUS 52

Table 3.4: Typical quasi-static loads for a 5 − 6 M W at mudline

Horizontal Load 15 M N

Vertical Load 35 M N

Bending Moment 562 M N m

Torsional Moment 4 MNm

Table 3.5: Environmental conditions

50-year extreme 10min wind @ hub-height 50 m/s

50-year extreme 5s wind @ hub-height 60 m/s

Mean water depth 30 m

50-year extreme water depth 41 m

50-year maximum wave heigh Hmax 22.3 m

Related wave period 14.5 s

50-year tidal current surface velocity 1.71 m/s

50-year storm surge current surface velocity 0.43 m/s

the p-y curves of the soil and compare results with the p-y obtained in a more realistic
loading condition (cyclic loading conditions). The static p-y curves were found using a
ramping function with a certain load. But the dynamic case was different, as also field
conditions wanted to be understood as well. As only monotonic loads are admitted by
the constitutive model in ABAQUS, the cyclic loading considers a sinusoidal function of
t = 1 s modelled as a Fourier series with the following parameters:

N
X
P (t) = P0 + An (cos nω(t − t0 )) + Bn (sin nω(t − t0 )) (3.6)
n=1

where P0 = 0, the cosine terms were neglected, Bn = 1, t0 = 0 and ω = 2π. Later


on, the function was multiplied by a constant (7.5 M N ) to give physical significance, as

M.Sc. Civil Engineering 3.8. Loads


53 Chapter 3. Finite Element Model with ABAQUS

pictured in Figure 3.7.

Figure 3.7: Cyclic load for dynamic analyses

Figure 3.8: Ramping function for static, quasi-static analyses and p-y derivation in
ABAQUS

The applied load is not modeled as a point load, because in a 3D model this would
cause a concentration of stresses in one node of the pile. To avoid this, ABAQUS uses
rigid bodies. There are two rigid and non-deformable bodies available: rigid cap and
rigid cone. The cone allows to use a point load which will be transferred to the pile.

3.8. Loads M.Sc. Civil Engineering


Chapter 3. Finite Element Model with ABAQUS 54

The rigid cap allows the use of stress instead of point loads. The later one was used
because of the symmetry plane. When using symmetry boundary conditions is better
to work in terms of stresses (otherwise, the point load should be divided and applied in
nodes). The total horizontal load is calculated multiplying the stress time the area of
the monopile (not the pile cap). This equivalent load takes into consideration the effect
of the wind, the sea-swell and the currents.

Figure 3.9: Equivalent stress for the pile cap

Regarding the load conditions, the soil was preset or preloaded with a geostatic load
step, in order to account for the gravity and introduce the influence of the K0 earth
pressure coefficient, as the initial stress field is very important in the calculation of final
stresses. For instance, the shear stresses in the interaction between monopile and soil
depend on the normal stresses, which are higher if the weight of the soil is considered.
This could also be achieved by inducing a gravity field in the model, but the interesting
point of using the geostatic step is the fact that te software automatically put the dis-
placement field to null.

3.9 Types of analysis


The methodology has followed a natural evolution of the model, building up from a
simple shallow strip foundation to the full model of the monopile and the soil. First,
static analyses were carried out, without accounting for dynamic effects or using a dy-
namic amplification factor for the static loading conditions. After that, a quasi-static
model involving the discretisation of a time dependent load, but without accounting for
dynamic effects (mass and inertia). In this case, for each time step the software needs to
solve a different load step according to the variation of the load within time. But that
load has not been conceded as a transient load. That was the next step to perform. A
fully dynamic model in which the load varies within time. This will complete the cycle of

M.Sc. Civil Engineering 3.9. Types of analysis


55 Chapter 3. Finite Element Model with ABAQUS

simulations in order to take into consideration all the effects of a real case situation. But
there are also many more issues to address. For example, the influence of the dynamic
vertical stiffness in the overall behaviour of the structure. Another step would be the
incorporation complex elastic-plastic laws for the soils while maintaining the dynamic
analysis.

The fact that implicit elements were selected due to its applicability to non-linear
problems, implied that the use of implicit analyses were a constant in all the models.
In the implicit analysis, the stiffness matrix of the whole structure has to be inverted in
order to solve for the acceleration field iteratively. This has a huge computational cost
when the model becomes greater in size or for the use of contact elements. Although the
explicit method was tried, because it is faster and it does not need the inversion of the
stiffness matrix, the Courant number was a limiting factor, as it had to be addressed
independently for each material and ultimately lead to time steps too small which actu-
ally led to a higher computational time.

For the calculation, the Direct Integration Method (DIM) has been used. Modal
superposition would require a very large number of modes since the spatial distribution
of the largely sparse and the mode shapes have to be computed at every step. For the
DIM, the general equation of the structure has to be solved at every integration time
step:
M (t) ü + C (t) u̇ + K (t) u = P (t) (3.7)

Where M is the mass matrix, C is the damping matrix, K is the stiffness matrix and
P is the load vector, variable with time. In our case, damping was neglected. Notice
that the equation above is linear, but the matrices are time dependent.

The time integration method implemented is the Hilbert-Hughes-Taylor (HHT) method,


which is like the Newmark method, but improved. This is a method that allows for
second order accuracy (which is not possible with the regular Newmark method). De-
pending on choices of input parameters, the method can be unconditionally stable. The
only difference with Newmark is the inclusion of the α parameter. α = 1 corresponds
to the Newmark method. α should be between 0.67 and 1.0. γ and β have been set
to 0.5 and 0.25 respectively (Rayleigh’s damping) in order to avoid a solution which is
affected by the so called “numerical damping”.The smaller the α, the smaller the numer-
ical damping. This default values ensure that the method is second order accurate and

3.9. Types of analysis M.Sc. Civil Engineering


Chapter 3. Finite Element Model with ABAQUS 56

unconditionally stable [Arias-Trujillo et al. (2012)]. α has been set to 0.67. The final
proposed values were give by default by ABAQUS after choosing the so-called option
“moderate dissipation”, a particular case recommended by the manual when solving
problems involving contacts. In order to account for big deformations, another options
was checked in the options of the analysis (NLG ON). This means non-linear geometry
an considers large strains in the model [Simulia (2010)].

3.10 Time step


The time step is also an important parameter when deciding how to represent the loading
scenarios in dynamic analysis, but it also has a big influence when the solver is iterating
to find a solution. Most of the convergence errors come from the fact that the time
interval is too big to find a certain present accuracy. It also represents the frequency the
software is getting data to perform the calculations. There are many recommendations
for the time step in the literature, but the two most recommended are the one proposed
by Chopra (δt/T < 0.1) where T is the period of the highest frequency and the one
related to the wave celerity, which can be derive from δt < V /(A · c), where V is the
volume of a finite element is, A is the area of the largest side of the finite element and c
is the highest wave celerity in the soil.

The minimum time step for the integration stage was selected so that the time step
that the software takes to integrate the general equation is significantly shorter than
the nth fundamental period of the system. Therefore, every single mode shape will be
recorded in case it is produced. During the elastic calculations, the smallest period is
T = 0.01 s. To rely on the side of caution during calculation, and due to the simplicity
of the problem, the integration time step employed has been i = 0.001s. Sometimes,
when convergence errors were difficult to deal with, ABAQUS had the option to use an
auto-interval for the time step. This feature increased notably de computational time,
as it has to compute first the time step required for the next iteration and then sole the
system.

Another problem arising from using small time steps is the amount of data that has
to be stored in the computer. Many times, the time step had to be set to a very small
value but it has to be specified that the software should only store data every 10 steps.
Many times the memory of the computer was filled or the output files were so big that

M.Sc. Civil Engineering 3.10. Time step


57 Chapter 3. Finite Element Model with ABAQUS

could not be edited. Sometimes this was a matter of balance between the sampling
theorem (Nyquist) and the data representation, in order to avoid the aliasing of the
output “signals”.

3.11 Summary of properties


To sum up this chapter, the parameters and the considerations of the 3D model that
later will be validated and used to perform all the complex calculations are presented.

• Geometry: According to a 6 M W offshore wind turbine monopile described in


technical reports [DOWEC (2003)]. Lenght of the monopile (only underground):
37 m; thickness of the wall 60 mm; diameter 6 m. The dimensions of the soil model
were related to the diameter of the monopile, making sure that the boundaries
would not affect the solutions and taking as a reference other models described
in the revision of the state-of-the-art of this thesis. The radial geometry is used
by multiple authors, and allow radial meshing. The graphic representation can be
seen in Figures 3.1 and 3.2. Studies with a 2D geometry were also carried and
discarded and infinite “acoustic” elements were used to avoid wave reflection in
the boundaries.

• Mesh: 8803 nodes and 1824 elements (soil) and 970 nodes and 924 elements (pile).
9773 nodes and 2748 elements in total. A volume of 138266.79 m3 of model. The
meshing in the pile and the soil is variable, from elements with a length of ∼ 7 m to
∼ 1 m in the soil and a length of ∼ 1 m to ∼ 0.5 m in the pile. The radial geometry
helped to implement sweep meshing techniques, instead of random meshing. This
controlled element angular distortion.

• Element types: C3D20R Stress/displacement solid elements, 20-node quadratic


brick with reduced integration and hourglassing control, recommended for non-
linear problems. The acoustic elements were directional solids CIN3D8.

• Material properties: For the soil, an elastic-plastic behaviour has been con-
sidered. The parameters were extracted from ground testing in the southern part
of the North Sea, with bore holes at 10 m depth [DOWEC (2003)], as described in
technical reports. The elastic regime is linear, homogeneous and isotropic elastic
model, and the plastic regime follow a traditional Mohr-Coulomb law. The metal-
lic monopile follow a linear, isotropic an homogeneous elastic law with equivalent

3.11. Summary of properties M.Sc. Civil Engineering


Chapter 3. Finite Element Model with ABAQUS 58

parameters to model a hollow pile as a solid body. The material properties can be
checked in Tables 3.1 and 3.3.

• Interfaces: Contact elements with null thickness. The master surface is defined
in the outer skin of the monopile and the slave surface is the equivalent surface in
contact with the soil. Frictional behaviour (µ = 0.3) and small sliding formulation.
The bottom of the monopile follow the same law, allowing for detaching.

• Loads: Sinusoidal load for cyclic effects. Ramping load for p-y and capacity
determination. Load applied as a stress on top of the monopile to avoid the
pile distortion and stress concentration. The peak value has been considered the
maximum quasi-static load for a 6 M W wind turbine at mudline level for conditions
in the North Sea according to Lesny and Wiemann (2005) and DOWEC (2003)
i.e., 15 M N for the whole pile, but because of the symmetry conditions, half of
the load has been considered.

• Boundary conditions: Fixed conditions on the bottom, vertical displacement


allowed in the surroundings of the soil. Symmetry condition in the respective
plane, see Figure 3.1.

• Type of analyses: Static (lateral static load), quasi-static (ramping loads vari-
able in time but without mass and inertial effects) and dynamics. Implicit analysis
conditions thorough all the simulations. Direct integration using a stable HHT
method for time integration, allowing the software to control the parameter based
on the recommendations for problems with contacts (moderate dissipation) and
non-linear geometry options activated to account for large strains. When perform-
ing dynamic calculations, the time interval was chosen so that no aliasing was
allowed. Sometimes, an automatic interval was picked by the software in order to
guarantee convergence of the system. Because of the high amount of data handled
by the computer, the outputs (mainly in history analysis) had to be carefully
chosen. The initial step of every analysis incorporated a geostatic stress field.

• Other considerations: Soil plugging effects, installation effects, densification of


the soil around of the monopile, pore water pressure or the degradation of the
stiffness of the soil have not been possible to model in the general scheme. Some
of these effects were studied, and the results can be seen in Chapter 6 but should
be subject of further research.

M.Sc. Civil Engineering 3.11. Summary of properties


Chapter 4

Validation of the Model

Numerical models are carried out to determine the specific behaviour or response of
the phenomena being studied without using scale physical models. By using software,
money and time can be saved. However, the results from the model will always be an
approximation with a certain amount of uncertainties. Thus, it is needed to reduce the
amount of error sources within the model. The process of validating some properties of
the numerical model will be explained in detail. In the previous chapter, the process of
building up a model was described. In this chapter, the scope is to validate different
aspects of the model in order to achieve the better solution possible when modelling a
soil-structure interaction with FE. This also helps to get a better understanding of the
dominant parameters.

First, a property of the model was chosen e.g. the constitutive model. Then, a
simple model with a close analytical solution was used to compare the results. Once
the analytical and numerical solution lied within tolerance, it was considered accurate
enough to be implemented later on in the complete model. The reason of this procedure
is the lack of information from other projects or reports in this matter. But, there is
a second motivation, which is to being able to control the software, and to learn the
expected outcomes from it.

4.1 Point load


In order to understand how ABAQUS integrates the stress and displacement fields, and
how to obtain results at single spatial points of the domain (later this was going to be
used to find stresses and displacements for the p-y method) a single elastic problem was

59
Chapter 4. Validation of the Model 60

used. A point load was applied in a Boussinesq semispace.

Figure 4.1: Point load domain

For z = 0.2 m and z = R, and using the elastic relationship for points loads Poulos
and Davis (1974) we have that:

3Pz3
σz = = 119.4 kP a (4.1)
2πR5
Plotting the stress field in ABAQUS and the vertical component (S33), we find that
σz = 121.9 kP a. There is a slight difference between analytical and numerical results,
but the error is minimum and probably with a finer mesh both values would have been
identical. With this simple exercise how to work out the stresses and how to obtain
them at a particular location in the model was learnt. How to use symmetry conditions
with the software was also practised. The influence of the boundary conditions in the
whole domain was also understood.

4.2 Shallow foundation


The first an easier problem intended to understand how the software works was a simple
rigid strip shallow foundation problem on clay. The most interesting part was to learn
about coupling nodes and DOF’s constraints to allow for a rigid behaviour of the strip

M.Sc. Civil Engineering 4.2. Shallow foundation


61 Chapter 4. Validation of the Model

footing and a rough contact between soil and structure. Aside from that, it could be
understood how the software treated plasticity in soils, as a Mohr-Coulomb constitutive
law was used. Using this simple model, an strategy to obtain the p-y curves was derived
as well. In this case, instead of p − y, a t − z curve was found (where t is the vertical
reaction of the soil and z the vertical displacement). The idea of this simple model is
to compare the numerical results with those proposed by Terzaghi in 1943. First, the
general shear mode failure was checked followed by the bearing capacity of the strip
foundation in this type of soil. The shear failure is expected to follow the Prandtl cri-
teria. In this particular example, instead of applying a load to the strip foundation, a
displacement of −0.5 m was imposed to the foundation. As the concrete of the founda-
tion was not part of the model, a set of coupled nodes was created in order to simulate
a rough contact between soil and foundation. Again, symmetry conditions were applied.
To account for the settlement of the foundation, the vertical boundaries were assumed
as rollers, whereas the bottom was fixed. The surcharge of the soil was assumed as a
distributed load. Shell elements were used for the model for plane strain conditions.

A schematic representation of the problem can be seen in Figure 4.2. This problem
was taken from Helwany (2007), so the results from the book and the results from the
model could be compared as well.

Figure 4.2: Shallow foundation problem dimensions

The soil properties used for this were:

4.2. Shallow foundation M.Sc. Civil Engineering


Chapter 4. Validation of the Model 62

Table 4.1: Soil parameters for the strip foundation problem

E 20000 kP a

ν 0.33

c0 100 kP a

φ 0

γ 18.14

As explained in the previous Chapter, in ABAQUS it is important to keep the model


as small as possible. This will later save computational time and will make easier to work
with the outputs files and the post-processing. Obviously for a simple model as this it
will not make a difference, but when dealing when complex constitutive laws, loads and
big domains it will set the difference between getting any results or not. Thus, in this
simple problem was important to understand how symmetry is used by the software and
how to use it to the benefit of the simulation process.

Figure 4.3: 2D Model of the strip foundation

Using Terzhagi’s equation for the bearing capacity, and assuming the parameter of

M.Sc. Civil Engineering 4.2. Shallow foundation


63 Chapter 4. Validation of the Model

Table 4.1:
qu = c0 Nc + qNq + 0.5γBNy = 570 + 6.89 = 576.9 kP a (4.2)

where Nc = 5.70, Nq = 1 and Ny = 0 are the Terzaghi’s coefficients. Now, we need to


compare this value, which is the ultimate bearing capacity of the soil, with the FEM
model. To do so, we need to find again the bearing capacity in terms of stresses. As
explained before, instead of using a load on the foundation, a displacement was imposed,
to make sure that the soil started to fail. This can be checked in the next plot of the
equivalent plastic strain:

Figure 4.4: Equivalent plastic strain. Note the Prandtl mechanism formed due to an
imposed rough interaction between de coupled nodes of the foundation and the soil.

Once the failure criteria was found, the maximum reaction force has to be found
under the coupling nodes that simulate the strip foundation. Plotting this reaction
against the displacement under the foundation we can find the so-called t − z curve of a
foundation. For this case, the t − z plot can be seen in Figure 4.5.

It can be seen that the maximum bearing force (asymptotic value in the graph 4.5)
is 175.017 kN , marking the point where the soil starts to flow plastically. Taking that
reaction force, and dividing by the area of the foundation we have:

qu = 175.017 kN/0.3 m2 = 583 kP a (4.3)

which is almost the same value found with Terzaghi’s equation.

4.2. Shallow foundation M.Sc. Civil Engineering


Chapter 4. Validation of the Model 64

Figure 4.5: t-z curve for the foundation

4.3 Laterally loaded piles

Once sufficient confidence was gained in the model with ABAQUS and how to interpret
results and stresses, the next step was to be familiar with numerical models of piles. In
this case, the soil had the properties presented in Table 4.1, and the main objective was
to compare the displacements obtained with the FEM model and Brom’s method (it can
be found in Chapter 5, in Figure 5.16). This exercise is presented in Helwany (2007), so
the results from the book and the results from the model could be compared this time
as well. Still the model was not definitive, and the geometry was still rectangular and
not radial, as eventually was. This exercise was also important to understand the parti-
tions in the numerical model, i.e., using a finer mesh the closer the elements get to the
monopile, and coarser close to the boundaries of the domain. Some of the parameters
that later would be used in the final model were also taken from recommendations from
Helwany (2007) in the chapter of the book devoted to pile foundations (Chapter 8). For
this problem, the same procedure followed for the shallow pad was followed, imposing a
displacement instead of a load, so the resisting capacity could be found.

In this case, as later in this document will be discussed, the capacity of the pile
according to the numerical model is higher than the traditional Brom’s method for
laterally loaded piles. In this case, FEM’s ultimate capacity is 526 kN and Brom’s

M.Sc. Civil Engineering 4.3. Laterally loaded piles


65 Chapter 4. Validation of the Model

Figure 4.6: Pile model for displacement comparison with Brom’s abaqus

abaqus give 265 kN . One is almost twofold the other. The traditional method is
overconservative (it is in the side of caution), i.e., it assumes less capacity than the
expected with a FEM model. It is also true that no contacts were allowed in this
particular case. Considering the interface between the soil and the monopile probably
would have brought down the FEM curve in Figure 4.7, giving a solution closer to the
one proposed by Brom (for a same load, a model with contacts will result in bigger
displacements on top of the monopile and therefore less bearing capacity).

Figure 4.7: Comparison between Brom’s method and the FEM model

4.3. Laterally loaded piles M.Sc. Civil Engineering


Chapter 4. Validation of the Model 66

4.4 Geostatic step


Generally, a geostatic step should be used to establish in-situ stresses in the soil profile
concurrent with no deformation which is the case for most of the geotechnical prob-
lems. ABAQUS has a predetermined option to do so, and initiate the stresses. This
is important as, for instance, the at-rest earth pressure coefficient is taken into account
automatically by using this option. The exact geostatic stresses are calculated using the
gravity force (which is what is happening) in an elastic media. Another option could
be approximate K0 with Jaky’s equation (1 − sin φ0 ), but this would be indeed a worse
approximation. In ABAQUS this is introduced as an initial step (previous to the loads).
The way it acts is equilibrating external actions with gravity actions. This implies than
internal iterations have to be carried out by the software, and there is always a certain
error. The validation of the geostatic step was also important to gain confidence in the
stresses of the most advanced models.

Initially, a soil column of 1 m2 and 10 m length was defined with a density of


γ = 20 kN/m2 simulating water, so the expected vertical stresses were obviously γ · z.
The deformations were 0 as well, which means that the geostatic step was performed
correctly.

Figure 4.8: Geostatic step check. Although the red range appears to be negative, in
an adequate plot, the stresses at the top of the column are null, and it is just an error
of the plotting range (in [Pa]).

Then, the geostatic step was applied to all the models. One of the important things

M.Sc. Civil Engineering 4.4. Geostatic step


67 Chapter 4. Validation of the Model

of applying geostatic conditions come from the contact between the monopile and the
soil. Those contact elements follow a frictional law in which the tangent stresses depend
on the normal stresses. If the geostatic field is not taken into account, the normal stresses
in the contacts would be reduced, and therefore, the shear forces will be lowered and the
displacements of the monopile increased.

Figure 4.9: Geostatic step in the 3D model (in [Pa]).

Figure 4.10: Geostatic step in the 2D model (in [Pa]).

4.4. Geostatic step M.Sc. Civil Engineering


Chapter 4. Validation of the Model 68

4.5 Elastic models


Previous to the creation of the final domain, a small domain was used with two object-
ives. One, the validation of elastic solution in quasi-static and dynamic regime under
different load conditions. Second, the influence of the pile diameter in the deflected
shape under a certain load. In this cases, the geostatic step was not performed, as this
was added later on the modelling process.

Figure 4.11: Elastic model geometry

Table 4.2: Elastic soil parameters for the reduced soil model (linear elastic)

E 20000 kP a

ν 0.33

γ 20.00 kN/m3

3 different load scenarios were applied to the top of the pile. This are plotted in
Figure 4.12 and 4.13.

With the base of this model and this loads, after the quasi-static elastic analysis,
an elastic-plastic quasi-static analysis was performed, and eventually an elastic-plastic

M.Sc. Civil Engineering 4.5. Elastic models


69 Chapter 4. Validation of the Model

Figure 4.12: a) Permanent push load and b) triangular load

Figure 4.13: Sinusoidal load

dynamic acting as the precursor of the final full model explained in Chapter 3 of this
document. In every case, it was find that the stress field gave similar results when
P = 500 kN was reached.

It was also important to see the influence of the pile diameter, as the slenderness
(length over diameter of the pile ) of the structure influences the overall response of
the soil. For slender structures, in which bending stresses are governing the behaviour.
Wider piles, as the commonly used in recent years tend to deflect as a rigid body. For
this checks, the push load in elastic conditions (Figure 4.12) was used.

This results would later become a subject of discussion in Chapter 5. The tests in

4.5. Elastic models M.Sc. Civil Engineering


Chapter 4. Validation of the Model 70

Figure 4.14: Deflection for a slender and non-slender pile

which the p-y method are based considered only slender piles, therefore the deflection is
closer to than to. This implies that the displacements fields may not be correct. There-
fore, this issue might be a source of errors when comparing the FEM and the traditional
method. The rotation centre is changed from one type of pı̀le to the other as well, finding
that for wider monopiles it is found around 1/3 from the pile toe. The pile has 2 points
with null deflection below the sea level. The deflected shape corresponds to the dis-
placements of a flexible monopile. In fact, the expected shape would be closer to a rigid
monopile, but the fact that the monopile is modelled as an equivalent structure might
have a certain influence. Neither shear stress diagrams nor bending moment diagrams
have been plotted or analysed. This is because the way the pile was modelled. In order
to avoid further interactions between the soil inside the monopile and the inner skin of
the monopile, the structure was assumed to be a whole, taking advantages of equivalent
properties such as mass and stiffness. To mention a setback of this, it is precisely the
fact that the stress diagrams are not representative, as the stiffness of the structure is
equivalent and includes the soil and not only the steel element.

It is fair to say that the order of magnitude of the displacements was found to be
around centimetres (0.06 m), and the normal stresses to the pile in the soil usually
ranged between 120 and 290 kP a in the elastic regime.

M.Sc. Civil Engineering 4.5. Elastic models


71 Chapter 4. Validation of the Model

4.6 Modal analysis

Apart from the information about the type of analysis given in Section 3.9, a modal
analysis was also carried out in order to find the first five natural frequencies of the
system. It is known experimentally that the frequency corresponding to the first mode
n = 1 is about 1 Hz.

Figure 4.15: Natural frequencies of the soil model

4.6. Modal analysis M.Sc. Civil Engineering


Chapter 4. Validation of the Model 72

The numerical model also gives for the first mode frequencies close to 1. There are
two main reasons to perform a modal analysis of the system. The first one is because the
turbine manufacturer need to know the approximate value of the structure+soil system
for the operating range of the turbine. The second one is because of the nature of the
loads. Adding the effect of currents, wind and waves, the frequency of the load is again
close to 1 Hz so for practical purposes, the natural vibration of the whole system need
to be checked when performing design works.

4.7 Soil column

This validation of the model was very important for dynamic problems. To do so, the
response of a vertical column of elastic soil under a sinusoidal acceleration field in its
base has been studied. This validation was carried out previous to the initial dynamic
loading cases of the monopile. The height of the soil column is H = 30 m,the width is
1 m, and the base is assumed to be rigid. Shell elements have been used for the model,
assuming plane strains. The soil properties, modelled as an elastic, incompressible and
isotropic media have the following values:

Table 4.3: Soil parameters for the soil column problem

E 58060 kP a

ν 0.4516

γ 19.20 kN/m3

The acceleration field has the following equation ah = 3 · sin(2πt) (amplitude Ac =


3 m/s2 and angular velocity ω = 2π)and its applied to the base as a dynamic bound-
ary condition. The total time of simulation is t = 10 s with a discretisation each
∆t = 0.005 s.

For this particular problem, Vicente Cuéllar derived in 1974 an analytic solution for
the horizontal displacement field with time:

M.Sc. Civil Engineering 4.7. Soil column


73 Chapter 4. Validation of the Model

4Aω 2
X (2n−1)π (2n − 1)πy
ux (y, t) = − νs2 (2n−1)2 π 2
·( )·
− ω2 2H
4H 2
2Hω νs (2n − 1)πt
· (sin(ωt) − · sin( )) (4.4)
νs (2n − 1)π 2H

p
Where νs = G/ρ is the velocity of the shear waves (G = E/2(1+ν)) and A = Ac /ω 2
is the amplitude of the displacement of the base. The plot of the analytic solution can
be seen in Figure 4.16 and the solution for the displacement obtained with ABAQUS in
Figure 4.17.

Figure 4.16: Analytic solution for an elastic, incompressible and isotropic soil column
subjected on its base to a sinusoidal acceleration field

In ABAQUS, the incompressibility was guaranteed constricting the vertical displace-


ment of the column. The analytical and numerical solutions are almost equivalent,
except for some values around t = 7 s, so the validation of the model under dynamic
conditions was successful.

4.7. Soil column M.Sc. Civil Engineering


Chapter 4. Validation of the Model 74

Figure 4.17: FEM solution for an elastic, incompressible and isotropic soil column
subjected on its base to a sinusoidal acceleration field.

Figure 4.18: a) Column model in ABAQUS and b) horizontal displacement field (plot-
ted in 4.17).

M.Sc. Civil Engineering 4.7. Soil column


75 Chapter 4. Validation of the Model

4.8 Absorbing boundaries

The absorbing boundaries are very important in elastic and non-elastic problems when
wave reflection in dynamic simulations can disturb the results. If the computational cost
is not a problem, the bigger the domain the more amount of time may take an elastic
wave to reappear. If the computational cost is a problem, absorbing boundaries are
required. These boundaries were used in the 2D elastic and plastic models, where the
alteration induced by wave reflection was considerably large and had to be addressed.
In the 3D model, however no big influence was found for the studied amount of cycles
and loads.

The way to overcome wave disruption in 3D was by means of infinite elements.


ABAQUS has a library of “acoustic” elements that satisfy the conditions of infinite ele-
ments, but they are not available from the user interface, so a small amount of coding
is required in the .INP root file of the model. The personal code for the implementation
of the infinite elements in an .INP file of ABAQUS can be found in the Appendix 10 of
this thesis.

To validate the result and gain confidence on the acoustic elements, an infinite soil
column was proposed. On top of the column, an instantaneous pressure (p = 3 kN/m2 )
is applied. For practical reasons, the model of the column is 5 m long, and the dis-
placements have been measured 1 m below the point of application of the pressure. The
column is 1 m wide and the hypothesis of plane strain have been assumed as well. The
model is bidimensional, with shell elements.

Table 4.4: Soil parameters for the infinite soil column problem

E 98000 kP a

ν 0

γ 21 kN/m3

In 2005, Arnold Verruijt proposed an equation for the vertical displacements:

4.8. Absorbing boundaries M.Sc. Civil Engineering


Chapter 4. Validation of the Model 76

p · νc (t − z/νc )
w(z, t) = · H(t − z/νc ) (4.5)
E

where w denotes the vertical displacement, z is the depth from top of the column,
p
H(...) represents the Heaviside function, νc = D/ρ is the velocity of the compression
waves and t represents the time. If the absorbing boundary is properly dismissing the
elastic wave, numerical with FEM and analytical solutions have to concur. The FEM
solution with ABAQUS was compared with the one obtained in López-Querol et al.
(2014). The agreement between the solutions is perfect, so the infinite elements (or
acoustic elements) were understood and proved to be fully functional.

Figure 4.19: FEM model of the soil column with infinite elements

M.Sc. Civil Engineering 4.8. Absorbing boundaries


77 Chapter 4. Validation of the Model

Figure 4.20: Solution to the infinite column problem, from López-Querol et al. (2014)

Figure 4.21: FEM solution with ABAQUS

4.9 Horizontal equilibrium


In order to check te equilibrium of the system, the load applied at the level of the mudline
to the monopile must be equal to the soil resistance. The forces in the model are integ-

4.9. Horizontal equilibrium M.Sc. Civil Engineering


Chapter 4. Validation of the Model 78

rated as explained before. For this, a horizontal load Q = 200 kN has been applied on
top of the monopile using a rigid cap. This load, although not representative, will be
used to check equilibrium conditions and to get an idea of the behaviour of the monopile.

Integrating the horizontal forces along the monopile, we find that ptotal = −189.1 kN .
Considering that the applied load was Q = 200 kN it is clear that a relative error of
5.5% which is within a reasonable limit of tolerance. Therefore, it can be said that
equilibrium is achieved.

It can be see in Figure 4.22 that from depth −20 m and on, the collaboration of the
soil with the overall capacity of the monopile is negligible, which means that a shorter
monopile could have been used for such a wider monopile, provided that the load is the
static Q mentioned before.

Figure 4.22: Reaction in the soil on a middle plane of the monopile

M.Sc. Civil Engineering 4.9. Horizontal equilibrium


Chapter 5

P-y Method for Monopile Design

In this chapter, an empirical analysis using the p-y design methodology will be pro-
posed. By far, the p-y method is the most commonly used in industry to design monopile
foundations. In this method, the soil is ideally represented by non-linear springs which
reproduce the reaction applied by the soil to a laterally loaded monopile, as discrete
forces. Although the p-y curves were devised by the oil industry due to site investiga-
tion and design of large and deep foundations in the sea, they have been widely used in
many other civil engineering works, such as offshore wind turbine foundations (our case)
or bridge foundations. Both case studies sharing a common situation, which is heavy
lateral loading scenarios, extreme winds or braking forces respectively. Bur reportedly,
the use of p-y curves in the design of monopile foundations leads to inaccurate designs,
resulting in oversized elements.

5.1 General background


If we take a look back at the existing background in design of piles subjected to lateral
loading, we may find three basic approaches: elastic approach, capacity approach and
numerical approach (from older to the more recent).

5.1.1 Elastic Approach: Beam on elastic foundation (BEF)


The elastic approach was used to estimate displacements in the pile due to lateral loading
assuming that the soil behaves as a perfectly elastic material. After 1897, the method
was also known as Beam on Winkler Foundation (BWF). This method came from the
existing knowledge of design of footings (pads, stripes, rafts...) mainly used by structural

79
Chapter 5. P-y Method for Monopile Design 80

engineers. Basically, the soil was modelled as independent springs acting on compres-
sion, and the vertical resistance of the subgrade was then proportional to the vertical
displacement of the soil underneath the above-mentioned structures. In the form of a
equation:
q = K · ∆y (5.1)

where q is the applied contact pressure, ∆y is the settlement and K is the so-called
modulus of subgrade reaction.

Figure 5.1: Beam on Winkler foundation

However, among the geotechnical engineers, the moduli of subgrade reaction is not
very representative, as it is not a characteristic property of the soil. Let’s do a very
quick example. If we go to a book of elastic solutions for soils (see Poulos and Davis
(1974)), we can find that, for a circular footing, the settlement is given by the following
equation:
qB(1 − ν 2 )
∆y = I · (5.2)
E
where I is an influence factor, B is the diameter of the foundation, ν is the poisson’s
ratio and E is the elastic modulus of the soil. As the contact pressure is known, we can
define the subgrade reaction modulus as:

E
K = q/∆y = (5.3)
IB(1 − ν 2 )

But, ν may vary from drained conditions (0.2 or 0.3) to undrained conditions (0.5),
and the same applies for E, E undrained is not the same than E drained. Therefore,
the elastic properties are variable with time and soil conditions, not fundamental. And
besides, there are many other factors that influence the displacement of a footing, such
as size, time, soil type, shape, roughness of the interaction soil-structure and relative
stiffness of the structure with respect to the soil (stiffer footings transfer the loads in a
different way than flexible footings). Plus, the units of the subgrade reaction modulus
are poorly conditioned ([F/L3 ]) for its application. Nevertheless, this approach is not
completely wrong, it may give a good initial guess, specially if we predict that the dis-

M.Sc. Civil Engineering 5.1. General background


81 Chapter 5. P-y Method for Monopile Design

placements of the foundations are going to be very very small.

So, if we think of a stripe foundation instead of a circular footing, and we think of


a vertical structure, instead of a horizontal structure, this was the early way to solve
displacements in piles subjected to lateral loads. In fact, this initial approach is very
easy to implement using the stiffness method in structures (assuming elastic behaviour
of the soil).

The next step of this process was to add the interaction of the pile with the soil by
means of the Euler-Bernoulli beam theory. In this case, a fourth order linear differential
equation have to be derived and solved. A laterally loaded single pile presents a soil-
structure interaction problem.

Figure 5.2: Laterally loaded pile

The soil reaction is dependent on the pile movement, and the pile movement is de-
pendent on the soil reaction. The solution must satisfy a non-linear differential equation
as well as equilibrium and compatibility conditions. The solution usually requires several
iterations and there is no straightforward analytical solution (only for simple boundary
conditions and non-realistic soil reaction distribution).

5.1. General background M.Sc. Civil Engineering


Chapter 5. P-y Method for Monopile Design 82

The first derivation of the equation was proposed by Hetenyi (1946).

Figure 5.3: Infinitesimal element analysis and geometric relations

Considering an infinitesimal element of the beam of length dx, the equilibrium of


internal moments yields the following relation:

X
M = (M + dM ) − M − V dx + dQ · (dy/2) − p dx · (dx/2) = 0 (5.4)

Neglecting the quadratic terms and differentiating twice, we find:

d2 M dV d2 y
− + Q =0 (5.5)
dx2 dx dx2

The term of the axial load can be neglected (Q) as we are not taking into consideration
axial effects in this analysis. The moment acting on the infinitesimal section of the beam
can be worked out integrating the normal stresses (σz ) within the normal cross area (A):
Z
M= σz · z dA (5.6)
A

Considering Bernouilli’s hypothesis (plane sections remain plane) we can find geometrical
relations to link stresses and strains. Assuming symmetry (neutral axis at mid cross
section) and the relation given by the rotation of the section, the following relations can
be obtained:
dy
u(x, y) = θz = z (5.7)
dx
du d2 y
ε(z) = = 2z = κz (5.8)
dx dx

M.Sc. Civil Engineering 5.1. General background


83 Chapter 5. P-y Method for Monopile Design

σ(z) = Ep · ε(z) = Ep κ z (5.9)

Where u(x, y) is the deflection of the pile along the longitudinal axis, ε(z) is the strain,
and z is the distance to the neutral plane and Ep is the Young’s modulus of the material.
Substituting (5.9) in (5.6): Z
M= (Ep κ z) · z dA (5.10)
A

Assuming that the material of the pile is linear elastic and isotropic, we would have:

d2 y
Z
M = Ep κ z 2 dA = Ep Ip κ = Ep Ip (5.11)
A dx2

Substituting (5.11) into (5.5) without considering axial loads:

d4 y dV
Ep Ip − =0 (5.12)
dx4 dx

Now, from horizontal equilibrium of forces in the infinitesimal element:


X
Fh = p(x) dx + dV = 0 (5.13)

Which leads to the governing equating for a laterally loaded pile:

d4 y
Ep Ip = p(x) (5.14)
dx4

where Ep is the elastic modulus of the pile, Ip is the second moment of area with respect
to the neutral axis of the structure, y is the horizontal component, z is the vertical
component. The variable p(x) corresponds to the resultant soil resistance force per unit
length of pile that occurs when the unit length of pile is displaced a lateral distance y
into the soil. Therefore, it is paramount to represent properly the value of p(x), or in
other words, the soil reaction (see Figure 5.4).

Finding a close-form analytical solution of this equation is only possible for simple
functions of p(x) [F/L]. However, the relationship between the soil reaction (p(x)) and
the pile deflection y [L] is non-linear and also varies along the pile depth. In practise,
the only way to solve this equation is recurring to numerical methods. Using the Finite
Difference Method (FDM) is the most extended numerical solution, and most software
devoted to pile analysis use the solutions given by this method. As it requires the division
of the pile in finite elements, the soil reaction p(x) might as well be finitely determined,
allowing for the use of the p-y curves.

5.1. General background M.Sc. Civil Engineering


Chapter 5. P-y Method for Monopile Design 84

Figure 5.4: Representation of the soil reaction in a) after installation, and b) when
loaded. Notice that the lateral interaction (side friction) is neglected

As a result of the analysis, the beam deflection, bending moment and shear force
along at any given point can be determined, allowing for a better understanding of the
stresses of the pile and making it easier to design. The relations given by successive
integration of the general equation are written in the following table:

Table 5.1: BEF Relations

Name Equation Units

Rotation θ = dy/dx [Dimensionless]

Curvature κ = θ2 [Rad/L]

d y2
Bending moment M (x) = Ep Ip dx2 [FL]
3
d y
Shear force V (x) = Ep Ip dx 3 [F]
4
d y
Soil reaction p(x) = Ep Ip dx4 [F/L]

It is important to highlight here that in foundation springs, the spring constant mul-
tiplied by the spring deflection (which is the same as the beam deflection) produces the
resistive force of the foundation (soil) per unit beam length. The modulus of subgrade
reaction multiplied by the width of the beam gives the foundation spring constant. In
fact, the spring constants are often estimated by determining the soil subgrade reaction
modulus (the modulus can be determined experimentally, e.g., by performing a plate
load test). Note that this idea requires that the coefficient of subgrade reaction at vari-
ous depths be known, and this is difficult to measure in the field.

M.Sc. Civil Engineering 5.1. General background


85 Chapter 5. P-y Method for Monopile Design

Up to this point, the theory of a beam on elastic foundation is rather general. In fact,
these very same equations represent the behaviour of pile sheet walls, flexible cantilever
walls, and so on. The main difference is the way of modelling the soil reaction, or the
function p(x). In flexible walls, a very extended method is to use the Rankine Thrust
theory, assuming elasto-plastic springs in which the relation between the deformation of
the soil and the structure is related through a function taking into account the relation
between displacement and active and passive thrust coefficients.

This, however, is only valid for plane strain problems such sufficiently long retaining
flexible walls. In piles, another approach had to be used. Is here where the p-y curves
started to play an important role. From this point, the beam on elastic (or Winkler)
foundation has passed to be called beam on elasto-plastic foundation.

5.1.2 P-y curves

The P-y curves is the most extensive method to design pile foundations subjected to lat-
eral loads. It was developed around the seventies [Matlock (1970), Reese et al. (1972),
Reese et al. (1974)], and it gives an estimation of the non-linear behaviour of the soil
reaction by means of a function that correlates lateral load with deflection at a given
depth. Depending on the discretisation level of the pile, so does the number of non-linear
springs that can be modelled with a p-y curve.

At a given depth, the prediction of the soil resistance is the most critical factor.
After the installation of the pile, at a certain depth, we can assume that the stress dis-
tribution can be constant and equal around the pile. Once the pile is laterally loaded,
the pile is deflected and the soil induces a stabilising stress around the pile. It is im-
portant to point out that some of the stresses will not be perpendicular to the pile wall
due to development of shear stresses at the interface between the pile and the soil. The
net soil reaction p(x) is obtained by integrating the stresses around the pile cross section.

In general, p-y curves are non-linear and they are a function of depth, soil type, and
pile dimensions and properties. Ashour and Norris (2000) used the strain wedge model
to study analytically the influence of some of these factors on p-y curves. They found
that for uniform sand deposits, a stiffer pile results in stiffer p-y curves. They also found
that if two piles have the same width, but one has a circular cross section and the other
has a square cross section, the resulting p-y curves will be different. The square pile in

5.1. General background M.Sc. Civil Engineering


Chapter 5. P-y Method for Monopile Design 86

Figure 5.5: Concept of beam on non-linear elastic foundation

sand showed a soil-pile resistance higher than the circular pile. They even proposed a
chart with corrections depending on the shape of the pile, but the findings from Ashour
and Norris are based on analytical studies, and to the best of the author’s knowledge no
full-scale experiments have been reported to confirm their findings.

Before starting defining the p-y curves, it is necessary to define several variables, be-
cause the terminology might be confusing in the literature. kh is defined as the horizontal
modulus of subgrade reaction. It relates lateral pressure with displacement (qh = kh · y)
and its dimensions are [F L−3 ]. If we multiply the horizontal modulus of subgrade re-
action by the width of the pile, we obtain the subgrade reaction modulus of the soil
(K), thus p = K · y, with K = kh · B, with dimensions [F L−2 ]. Once this has been
clarified, the code defines the coefficient of subgrade reaction (nh ), which relates the rate
of increase of the subgrade reaction modulus (K) with depth (z), being Ki = nh · zi .
From now and on, this will be the terminology used.

A typical p-y curve is composed by the following terms:

• Initial slope Epy−max = kh · B = K

• Secant slope Epy−sec

M.Sc. Civil Engineering 5.1. General background


87 Chapter 5. P-y Method for Monopile Design

• Asymptotic value Pult

• Transition curve

Figure 5.6: a) Elements of a p-y curve and b) variation of the subgrade reaction
modulus

Bearing this in mind, with this elements, different authors have used different meth-
ods to estimate the p-y behaviour of the soil. Basically the parameters represented in
Figure 5.6 are obtained experimentally. Other authors (see Reese et al. (1974)) propose
how to calculate the p-y curve of a sandy soil based on multiple experimental results.
Here, the so-called Method A will be explained.

The common way to find the p-y curve is through experimentation. In fact, at
inception of the p-y curves, they were a result of pure experimentation in piles in the
Mustang Islands (Texas) by Reese et al. (1974).

Figure 5.7: In-situ testing of laterally loaded piles, from Pando (2013).

5.1. General background M.Sc. Civil Engineering


Chapter 5. P-y Method for Monopile Design 88

A basic test consisted of a pile loaded with sensors, such as inclinometers. With
such devices the deflection and curvature of the pile can be plotted. Having found that
function by fitting, using the governing equation, a solution can be integrated accurately.
The method used in the on-site experimentation is shown below in Figure 5.8:

Figure 5.8: Experimental set-up for laterally loaded piles in the Mustang Islands, by
Reese et al. (1974).

As a result of that testing campaign, the Method A for p-y curves was developed.
The proposed formulation consists of three curves: an initial straight line, 1, a parabola,
2, and a straight line, 3, all assembled as a continuous piecewise curve, fully differenti-
able (see Figure 5.9). This methodology proposed to build the respective curves giving
different values to create a parametric curve using the pairs (Pult , yult ); (Pm , ym ) and
(Pk , yk ) plus the functions mentioned above. yult , ym , yk refer to a certain experimental
displacement dependent on the pile width.

M.Sc. Civil Engineering 5.1. General background


89 Chapter 5. P-y Method for Monopile Design

Figure 5.9: Elements of a p-y curve, according to Reese et al. (1974).

Pult is related to the maximum lateral resistance, with

Pult = A Pc (5.15)

where A is a correction factor to account for the scale of the test (a poor agreement
between scaled test and full-scale tests was found to be relevant), and Pc is the ultimate
resistance. This value is computed analytically by means of either statically or kinemat-
ically admissible failure modes (limit analysis). Depending on the depth, there are two
failure modes that provide different solutions for the ultimate soil resistance, having Pcs
for shallow depths and Pcd for deep depth.

Figure 5.10: Failure modes for a laterally loaded pile, with a) for shallow depth and
b) for deep depth, by Reese et al. (1974).

If we write down the equilibrium equations, we find the following relations:

5.1. General background M.Sc. Civil Engineering


Chapter 5. P-y Method for Monopile Design 90

K0 x tan φtr sin β tan β (D − x tan β tan α)


Pcs = γ 0 x [ +
tan(β − φtr ) cos α tan(β − φtr )
+ K0 x tan φtr (tan φtr sin β − tan α) − Ka D] (5.16)

Pcd = Ka D γ 0 x (tan8 β − 1) + K0 D γ 0 x tan φtr tan4 β (5.17)

where φtr is the internal angle of friction based on triaxial tests,γ 0 is the effective
unit weight, α = φtr /2 and β = 45o + φtr /2, and Ka and K0 are the active and at
rest Rankine coefficients. As the pile does not behave the same way in shallow depth
(Pcs ) than in deep depths(Pcd ), there is a transition zone between them, given by the
intersection of the functions.

The next plots show the variation of A (Figure 5.11) and Pc (Figure 5.12) with depth
for a generic pile. As shown, the transition depth increases with diameter and angle of
internal friction. Hence, for piles with a low slenderness ratio the transition depth might
appear far beneath the pile toe.

Figure 5.11: Variation of the parameter A, by Reese et al. (1974)

Pm is related to the maximum lateral resistance again, with

Pm = B Pc (5.18)

Where B is a correction factor depending on the dimensionless factor x/D (see Figure
5.13).

M.Sc. Civil Engineering 5.1. General background


91 Chapter 5. P-y Method for Monopile Design

Figure 5.12: Variation of Pc with depth, pile diameter and friction angle. Notice that
the bigger the diameter, the lower the transition depth (XR ) marked with a circle, and
most controlled the failure by the shallow mode, after BrØdbæk et al. (2009).

Figure 5.13: Variation of the parameter B, by Reese et al. (1974)

The initial straight line 1 (see Figure 5.9) depends on Epy−max which relies on a good
estimation of the subgrade reaction modulus. kh varies with depth along the pile, and
this is a purely experimental parameter based on soil testing. Most codes [API (1993),
DNV (1992)] offer a graph for the estimation of kh (see Figure 5.14).
The parabola 2 (see Figure 5.9) can be easily estimated once the rest of the points
have been already determined with the start and end point, and imposing the continuity
of derivatives in the transition points.

5.1. General background M.Sc. Civil Engineering


Chapter 5. P-y Method for Monopile Design 92

Figure 5.14: Estimation of kh as a function of the internal angle of friction, and the
relative density of the soil [API (1993)].

But instead of recurring to the method A every time, the most common references
for already-built p-y curves are summarised in the next table:

Table 5.2: Recommended criteria for p-y curves [Reese and Van Impe (2001)].

Soil type and condition Reference


Soft clay below the water table Matlock (1970)
Stiff clay below the water table Reese et al. (1975)
Soft clay above the water table Welch and Reese (1972, 1975)
Sands API (1993) & Reese et al. (1974)
Soils with cohesion and friction Evans and Duncan (1982)
Strong rock Reese (1997)
Weak rock Nyman (1980)

M.Sc. Civil Engineering 5.1. General background


93 Chapter 5. P-y Method for Monopile Design

5.1.3 Capacity Approach: Brom’s method

The ultimate load approach embodied in Broms’ method is suitable for short and long
piles, for restrained- and free-headed piles, and for cohesive and cohesionless soils [Hel-
wany (2007)]. It assumes that a short pile will rotate as one unit when it is subjected
to lateral loads for short piles. The soil in contact with the short pile is assumed to fail
in shear when the ultimate lateral load is reached. On the other hand, a long pile is
assumed to fail due to the bending moments caused by the ultimate lateral load; that
is, the shaft of the pile will fail at the point of maximum bending moment, forming a
plastic hinge.

Figure 5.15: Behaviour of laterally loaded piles, by Helwany (2007).

Also, the term restrained-headed pile indicates that the head of the pile is connected
to a rigid cap that prevents the head of the pile from rotation. Brom proposed different
abaqus for design (See Chapter 8 from Helwany (2007)). As we can see in some of the
graph, Brom include Rankine passive earth pressure coefficient, assuming a limit equi-
librium approximation and failure criteria.

5.1. General background M.Sc. Civil Engineering


Chapter 5. P-y Method for Monopile Design 94

Figure 5.16: Brom’s abaqus for piles embedded in cohesive soil, with a) for short piles
and b) for long piles, from Helwany (2007)

Figure 5.17: Brom’s abaqus for piles embedded in cohesionless soil, with a) for short
piles and b) for long piles, from Helwany (2007)

M.Sc. Civil Engineering 5.1. General background


95 Chapter 5. P-y Method for Monopile Design

5.1.4 Numerical Approach

The numerical approach is basically solving the governing equation for the pile using a
numerical methods (generally finite differences) implemented in a code. Typically the
commercial software has already a library of p-y curves, making easier for the engineer
to work with it. Some of these are COM624, LPILE, or FB-Pier among others.

An interesting feature of some of this packages is the possibility to account for the
development of a plastic hinge in the pile. This is achieved by a more complex formu-
lation of the p-y curves, but a likely situation when the maximum capacity wants to be
found or when dealing with slender piles.

But there are other ways to solve the problem. Finite Element analysis has been
proven to be valid as well, although it requires a better understanding of the constitutive
models and the input parameters.

5.2 Codes
If Reese introduced the Method A (see next paragraph to Figure 5.8), the API (1993)
proposes a complementary method called Method B, although both methods share its
inception. The American Petroleum Institute will be consider the reference, as the
European design codes DNV (1992) and GL (2005) refer to the API rules. What the
code provides is a way to choose and build the p-y curves for its application in the sub-
sequent analysis. Next equations describe Method B.

The modification comes from the fact of using simplified equations and parameters to
define the p-y curves. The ultimate resistance is estimated from the following equation:

 Pus = (C1 x + C2 D)xγ 0

Pult = min
0
ud = (C3 D)xγ

 P

C1 , C2 , C3 can be selected as well from the proposed plot as a function of the internal
angle of friction, or be computed with the following relations:
   
tan β cot(β − φtr )
C1 = tan α tan β − K0 + K0 sin β tan φtr + tan β
tan(β − φtr ) cos α
(5.19)

5.2. Codes M.Sc. Civil Engineering


Chapter 5. P-y Method for Monopile Design 96

tan β
C2 = − Ka (5.20)
tan β − φtr

C3 = Ka (tan8 β − 1) + K0 tan φtr tan4 β (5.21)

Which can be related with the analysis with limit equilibrium carried before.

Figure 5.18: Variation of C1 ,C2 , and C3 with the internal angle of friction, from API
(1993).

Then, the following equation is proposed to describe the p-y curve:


 
kx
p(y) = A Pu tanh y (5.22)
A Pu

With A being still the same parameter from method A, or using the next formula:
 x
A = 3 − 0.8 ≥ 0.9 (5.23)
D
A = 0.9 will only be used for cyclic loading conditions. The initial slope of both
methods is identical, and Pult is similar.

M.Sc. Civil Engineering 5.2. Codes


97 Chapter 5. P-y Method for Monopile Design

5.3 Monopile design according to the DNV-API


This section is devoted to get the p-y curves given by the codes which lately will be
compared with the ones found in the numerical model with FE to see the if there is any
difference. In order to do so, both Method A and Method B (previously explained) have
been used to compare the likelihood of the results. For doing so, a personal spreadsheet
has been used. The parameters involved in the definition of the soil are the ones used in
for the Mohr-Coulomb constitutive law of the FE model, which was already tested and
calibrated, provided by the DOWEC (2003). The geometry is the same that has been
already explained in previous chapters.

First, Pc was computed to see the influence of the failure modes. As the diameter of
the monopile is relatively big, the failure mode was expected to be determined by the
shallow failure criteria. Therefore, the soil strength variation with depth is given by the
following graph:

Figure 5.19: Variation of Pc with depth x, with D = 6 m and φtr = 35o , according to
Figures 3.1 and 3.2 in Chapter 3

Once Pc (x) is known, obtaining the p-y curves for our soil is a straightforward pro-
cess, following both Method A and B. In order to use the functions for the parameters
A and B, a Matlab routine was written to obtain an interpolation function. For the rest
of the cases, the functions given were used, instead of the plots. An example of some
p-y curves are presented below:

5.3. Monopile design according to the DNV-API M.Sc. Civil Engineering


Chapter 5. P-y Method for Monopile Design 98

Figure 5.20: P-y curves method A

Figure 5.21: P-y curves method B

In the old codes, it was a common practice to get the first p-y curve at 0.5 m depth,
and the rest at each meter, however, this recommendation lacks sense with modern com-
puters. From depths of x = 36 m and more, the p-y curves are almost identical. This is

M.Sc. Civil Engineering 5.3. Monopile design according to the DNV-API


99 Chapter 5. P-y Method for Monopile Design

related with A and B parameters and the determination of the dimensionless coefficient
x/D.

It is also interesting to see if the results obtained with each of the methods are
comparable. It seems like there is a good agreement in the initial slope, but there is a
certain variation in the ultimate strength Pu . For greater depths, Method B appears to
give a greater ultimate resistance, whereas for shallow depths, it gives a smaller value of
Pu compared to Method A.

Figure 5.22: Method comparison 1

Figure 5.23: Method comparison 2

5.3. Monopile design according to the DNV-API M.Sc. Civil Engineering


Chapter 5. P-y Method for Monopile Design 100

5.4 Discussion and potential limitations


Originally, the p-y method was intended for small piles (diameters around 1 m), while
current monopiles are much larger than that (5 − 6 m). The designs are based on codes
with more than 20 years, with very few revisions. But while it seems obvious that an
update of the methods would be advisable, the vision of the industry does not share the
same direction. While academia is trying to improve the design of monopile foundations
in an attempt to save in total capital expenditure of projects (assuming that almost the
30% of the cost of a wind turbine comes from the foundations), the industry remains
reluctant to adopt the proposed changes. Besides, there is no official method for design,
as the existing codes are more recommendations than regulations.

5.4.1 Original testing

As stated in the previous section, the p-y curve method stems from the analysis carried
out in the Mustang Islands. It is important to bear in mind that, in experimentation,
the outcomes are completely influenced by the conditions of testing.

Those tests were carried out for only one type of pile (circular, slender and metallic),
with one only diameter, and in only one type of soil (sand), subjected to seven load
cases. Given the complexity of soil-structure interaction problems, it seems like bold
idea to extrapolate the results to any other case. As a matter of fact, further research
on full scale piles uncovered a huge bias in the results, in comparison with the initial
estimations. Subsequently, correction factors (A and B) had to be introduced to account
for that.

5.4.2 Initial slope Epy

In the formulation of the beam on elastic-plastic foundation, when deriving the bending
stiffness, the only term used is the bending stiffness of the pile (Ep ). However, in the
p-y curves, great part of the influence of the curve is given by the parameter Epy or the
equivalent to the subgrade reaction modulus. Nonetheless, there is no relation between
Ep and Epy whatsoever in the formulation. If the reaction of the soil depends on the
deflection of the structure through the curves, it can be questioned whether those two
parameters might relate one another or not.

Another point to address would be if the shear deformation of the pile is relevant to

M.Sc. Civil Engineering 5.4. Discussion and potential limitations


101 Chapter 5. P-y Method for Monopile Design

determine the soil reaction. Assuming the hypothesis that plane sections remain plane
eliminates the term of deformation in the cross section due to shear stresses.

And, the fact that k is given in the codes as a function of the internal angle of friction
and the relative density of the soil as characteristic parameters of the soil, when actually
they are not such thing.

5.4.3 Layered soil

By using springs the discretisation of the soil is divided in “layers” with no interaction
between each other. If the soil is supposed to act as a continuum, the methods are not
considering the interaction between springs. The shear modulus of resistance of the soil
G is not considered at all in the formulation.

5.4.4 Soil-structure interaction

The “skin” friction is a term that has not been considered for the determination of the
ultimate resistance (with the failure modes) neither in Method A nor in Method B.
Therefore, a term, depending on the diameter and shape of the pile, is missing from the
analysis. It is , in fact, assumed a perfectly smooth pile, as the Rankine criteria has been
used. Some authors have proposed alternative methods including in the net ultimate
lateral resistance and the friction resistance.

The soil plugging at the pile toe and the shear resistance at that point is not con-
sidered in the p-y formulation.

5.4.5 Soil resistance

The soil resistance is intimately related to the failure mode, which is somehow depend-
ent on the flexibility of the pile. The slenderness parameter of the piles in the Mustang
Islands was L/D = 34.4, whereas the monopile studied in this dissertation has a slender-
ness ratio of L/D ≈ 6. For modern monopiles, a slender ratio < 10 is common. Notice
the massive difference, moreover when the bending stiffness is related to the slenderness
and the soil reaction is influenced as well by the flexibility of the structure.

The failure modes could also be questioned. As can be seen in Figs. 5.12 and 5.19,
the transition depth is far under the pile toe, specially for large diameters, which means

5.4. Discussion and potential limitations M.Sc. Civil Engineering


Chapter 5. P-y Method for Monopile Design 102

that the problem is governed by failure in shallow mode for most modern cases, even
though the pile is more likely to behave as a rigid body than as a flexible structure.

5.4.6 Initial stress field

It has been demonstrated experimentally that the behaviour of the piles are influenced by
the initial stress field developed after installation [Jardine et al. (2013)]. The behaviour
of the pile will change effectively depending on whether the monopile has been drilled
or nailed.This methodology does not take into account the initial stress field of the soil
(altered during the installation process) nor the possible influence of the vertical load of
the structure.

5.4.7 Sensitivity

Using a spreadsheet gives a lot of possibilities to change values and parameters. It has
been found that the most influential parameter are k when finding the initial slope of
the p-y curves and D for the ultimate resistance of the soil. While D is given by design,
is a chosen value, k is assumed a property of the soil dependent on the relative density.

5.4.8 Static and dynamic p-y curves

Up to now, not much has been said of the dynamic p-y curves, but in fact those were
considered in the Mustang Island original tests. The reason is that in oil platforms wind
and wave swell, which are the main lateral loads, have an inherent dynamic component.
During the Mustang Island tests, the piles were in total subjected to seven horizontal
load cases consisting of two static and five cyclic. And this indeed is taking into account
in the parameters A and B. Later, method B included a correction factor for dynamic
loading cases again. However, it is important to notice that the nature of the tests was
clearly monotonic and not cyclic.

In fact, there are very few cyclic tests on laterally loaded piles [Pando (2013)]. And,
in most of them, the number of cycles is very limited (20 − 500 cycles) and not enough to
infer a trend in the behaviour of the structure in the long term of a life cycle (≈ 25) years,
with more than 109 cycles expected [Cuéllar (2011)]. The codes merely recommend a
degradation of a 10% in comparison to the static p-y curves. It is indeed surprising that
a problem with such a dynamic nature is treated in such a trivial way in the codes. But
the experimental data is scarce, and the results are certainly biased.

M.Sc. Civil Engineering 5.4. Discussion and potential limitations


103 Chapter 5. P-y Method for Monopile Design

5.4.9 Concluding remarks


All the simplifications made during the process is what has made researchers to question
the validity of this method. Good news is that the simplification leads to very conser-
vative results. If a model is capable of adding the skin friction, the influence of the shear
modulus of the soil, the effect of the shape of the pile, the initial stress field influence and
the shearing resistance at the toe, the final p-y curve would be more accurate. Besides,
the method was intended to be used in sands, but the nature of the soil is very variable.
There are no special remarks for monopile foundations in bed rock or in multi-layered
soil.

On the side of industry’s opinion, it is clear that using a conservative method will
lead to safer structures, and the saving in detailing and steel during the manufacturing
process of the foundation is not worth bothering. Reportedly, no wind turbine has col-
lapsed up to date. But the information is very scarce and companies does not provide
such information.

It can be said that the p-y method is an experimental analysis backed by years of
experience in the matter without any significant failures that would push designers to
devise more accurate proceedings.

5.4. Discussion and potential limitations M.Sc. Civil Engineering


Chapter 6

Comparison of the p-y Design


Procedure and the FE Model

The first condition that the model needs to satisfy is the equilibrium condition between
actions on the turbine and soil reaction, which is achieved with sufficient pile length.
Using ABAQUS, the p-y curves of the soil can be found working with the finite elements.
The horizontal displacement of the monopile (y) can be computed in working planes,
taking the value from the horizontal field of deformation of the interface soil-structure.
The soil resistance (p) is found in the same working planes that the displacements. The
interface elements are not used when determining the p-y curves, as they have been
assigned only tangential behaviour and not normal behaviour. This however does not
mean that they do not play an important role in the stress resultant.

But, figuring out the p-y curves is not immediate with a FEM model. According to
the solid elements in ABAQUS, the stress field of the soil present 6 independent com-
ponents. In the Mohr-Coulomb formulation, those can be written for general states of
stress in terms of three stress invariants. These invariants are the equivalent pressure
stress in terms of the maximum and minimum principal stresses, the Mises equivalent
stress, and the third invariant or deviatoric stress. It is important then to select the
horizontal component of these stresses (S11 in general). Now, ABAQUS gives the values
of the stresses at the integration points, meanwhile the stresses at the nodes are linearly
interpolated from the stresses in the stress points. The averaged stresses of the element
cannot be calculated directly from the stress points. By using the information on the
nodal points, we miss information in the middle of the element. When the elements
are sufficiently small this effect can be neglected, but with relatively big elements it is

105
Chapter 6. Comparison of the p-y Design Procedure and the FE Model 106

necessary to take this effect into account. To do so, the stresses in the nodes have to
be multiplied by a weight factor, depending on the local coordinates inside the element.
Thus, the averaged stresses are calculated by multiplying the stresses on the interface
of the element by the weighting factors. Then, the total stress has to be divided by
4 (the total sum of the weight factors, as an averaged weight). Then, the calculated
average stress (taking into account the different components of the stress tensor), in
the face are multiplied by the area of the interface element. The stresses (in kN/m2 )
are multiplied by the area (in m2 ) to obtain a force. This process is carried automat-
ically by ABAQUS in order to obtain the equivalent forces in the selected working planes.

According to the codes, the reaction of the soil has to be determined first at level
−0.5 m and then every meter along the monopile. However, for practical reasons, just
a few of the results will be presented in this chapter. The reason for this is that from a
certain depth and on, the displacements of the monopile are very close to zero no matter
what real load is applied, so for a real case (as it is obvious) the soil is very far from
yielding and in the very first stages of an elastic behaviour, due to the small amount of
strain. On the contrary, the closer to the seabed, the yielding of the soil is practically
imminent.

6.1 Interfaces
The next models only feature the contact elements. There is no agreement between
authors in which coefficient should be taken. For traditional concrete piles, the values
range between µ ≈ 0.3 − 0.5. In this thesis values between 0.3 − 0.4 have been chosen,
as the roughness of the steel is usually lower than the roughness of the concrete. The
model is elastic, and subjected to a cyclic sinusoidal (T = 1s) horizontal load of 100 kN
on top of the monopile. The influence of the friction coefficient will be analysed hereby.
The base of the monopile has the same frictional formulation than the skin.

The first thing to notice is that the monopile and the soil are detached under cyclic
loading conditions. This has been checked in-situ and, in fact, it is a situation that
occurs.

Comparing both result, the maximum crack opening on the tip at the monopile is
around 5.8 mm, so the influence of the friction on the opening is nt significant. On the
contrary, the opening in the pile toe seem to differ in ≈ 2 mm. There is no physical

M.Sc. Civil Engineering 6.1. Interfaces


107 Chapter 6. Comparison of the p-y Design Procedure and the FE Model

Figure 6.1: Opening soil-monopile for µ = 0.3 (in [m])

Figure 6.2: Opening soil-monopile for µ = 0.4 (in [m])

evidence of detaching in the toe area of the monopile.

The toe detachment was studied by means of the vertical slip of the monopile under
cyclic conditions. Because of the particular geometry, a rigid monopile is likely to be

6.1. Interfaces M.Sc. Civil Engineering


Chapter 6. Comparison of the p-y Design Procedure and the FE Model 108

affected by vertical displacements at the toe, as the bending deformation is relatively


small. This effect, although seen in the numerical model, should be studied experiment-
ally, because the monopiles are usually nailed to the ground, and specially that particular
area of the foundation.

Figure 6.3: Vertical slip of the toe of the monopile for µ = 0.3 and µ = 0.4 (in [m])

Another topic of discussion was the plane of maximum normal stresses. Using the
contact elements, this can be easily plotted. Apparently, the middle plane should be
subjected to the higher normal stresses. This has been shown with the finite element
model. Albeit, it is very interesting the fact that the maximum shear stresses does
not occur in that middle plane. These facts remark the influence of a 3D geometry in
the modelling of pile problems, making difficult to change to a bidimensional geometry,
because that transversal influence of the shearing forces would be neglected. This phe-
nomena is pictured in Figures 6.4 and 6.5, contrasting left and right soil models.

Eventually, it was also obtained the displacement at the tip of the monopile. It was
then interesting to see whether the different friction value could influence the displace-
ments of the monopile. It was demonstrated that, although the influence is very small
(tenths of millimetres), there is a slight difference in terms of horizontal displacement of
the tip. This can be appreciated in Figure 6.6. If the horizontal displacement of the pile
is influenced, then the horizontal displacement of the soil is slightly influenced as well,
affecting likely to the p-y curves.

Another important aspect apart from the shearing forces in the skin of the monopile
is the shearing forces at the base o the monopile. Since the behaviour of the monopile

M.Sc. Civil Engineering 6.1. Interfaces


109 Chapter 6. Comparison of the p-y Design Procedure and the FE Model

Figure 6.4: Normal and shear forces in the contacts for µ = 0.3 (in [Pa])

Figure 6.5: Normal and shear forces in the contacts for µ = 0.4 (in [Pa])

is closer to a rigid body, the shear at the base will create a counterbalancing moment in
the structure that will help reduce the deflection, and therefore the p-y curves will be
affected. From the FE analysis, it can be concluded that shear at the base and at the skin
in non-slender monopile has a very big influence in the overall behaviour compared to
flexible monopiles, and might be a reason for discrepancies between traditional methods
and FE methods. By plotting the p-y curves of the soil with and without contacts in
the model, the influence of the interface can be measured.

6.2 Soil plugging

This was an attempt to understand the soil plugging and its influence on the overall
behaviour on the monopile and during the installation phase. To do so, a more accurate

6.2. Soil plugging M.Sc. Civil Engineering


Chapter 6. Comparison of the p-y Design Procedure and the FE Model 110

Figure 6.6: Displacements at the tip of the monopile for different friction coefficients.

model of the monopile was created. In it, the monopile was modelled as a pipe with a
thickness of t = 60 mm (which is a common value), and not as a solid body. Therefore,
thins shell elements were used. The soil inside the body of the monopile was made up
using the same elements used in the soil of the general model. This was the general
scheme that allowed to derive the equivalent properties for the solid body monopile.
The first idea was to check whether the soil inside the monopile due to installation has a
certain influence in the overall bending behaviour of the monopile. The second objective
was to compare the numerical results with a close solution of the soil plugging. It was
also checked out a hypothesis stated by Professor Achmus, in which he claimed that
there was no significant difference in using a full solid body monopile with equivalent
properties than modelling the whole monopile and the soil inside. As the soil on the
inside of the monopile is not fixed to the skin, the soil inside the pipe appears to shear
and move according to the displacement of the steel structure. The interaction between
the inner skin of the monopile and the pile was the same than for the outer skin and the
soil surrounding the pile. The base of the monopile was fixed, and an horizontal load of
1 kN was applied. In essence, the simplified model behaved as a cantilever beam stuffed
with soil.

What was observed is that the pressure over the inner skin of the monopile was in-
creased, and therefore the friction. However, this did not have a great influence on the

M.Sc. Civil Engineering 6.2. Soil plugging


111 Chapter 6. Comparison of the p-y Design Procedure and the FE Model

Figure 6.7: Different levels of plugging in the monopile

overall behaviour of the structure, as it is the steel element the one that carries more of
the stresses. In fact, the difference between the whole pile plugged and the pile without
plugging was of the order of millimetres. This seems to back up Achmus’ theory. Non-
etheless, it is true that the installations effects have not been considered and the soil
inside the monopile might not follow traditional laws.

Figure 6.8: Example of interaction between plugg and monopile (in [Pa])

6.2. Soil plugging M.Sc. Civil Engineering


Chapter 6. Comparison of the p-y Design Procedure and the FE Model 112

6.3 2D v. 3D

One way to save time and calculations is by reducing the dimensions of the model. The
majority of the models described in this document has taken advantage of the use of
symmetry boundary conditions. By using this, a model can be reduced to half of the
size of the initial one. Then the solutions are obtained in a semi-domain which is equi-
valent to the whole domain. The symmetry conditions need to be used with care, but
essentially creates a mirror model which later can be restored and interpreted in full
scale. In the case of most of the monopile models revised, this technique has been used.
By doing this, the geometry is reduced, allowing the author to have a finer mesh in the
semi-domain, instead of a coarser one in the whole domain.

In order to save more time and resources, a 2D model was studied and proposed. It
was used to compare between 3D and 2D and see how close in terms of displacements
and stresses could it get one another. The hypothesis was that if the pile is wide enough
(which certainly is), the critical band (centre of the pile-soil, mid plane) can be simulated
using a 2D model of 1 m wide, instead of a full 3D model (even with symmetry bound-
aries). This possibility was explored, and the results are described in the following lines,
but it did not work out as expected, so the full 3D models were necessary and adopted
eventually. Another problem with the 2D model is the fact that no shell elements could
be used, solid elements had to be used, and later the width was set in 1 m. Such thing
was done in order to account for the contact elements of the interface. The formulation
of the shell elements does not include the interaction with other shell elements, so a
contact cannot be modelled with shell elements in ABAQUS.

The 2D model was derived from the 3D model, “cutting” the middle plane of the
monopile, as it is theoretically the most representative of the behaviour of the system.
After the initial checks with a purely elastic model, a Mohr-Coulomb law was defined
with the same values of the 3D model. Following the same principle than in the full
model, first a distributed load was applied on top of the monopile. It is better to work
in terms of stresses and not in terms of the absolute load to make easier the comparison
between the models. Plotting the displacements fields the first impressions can be ob-
tained.

After applying a similar stress on top, the horizontal field is the expected one for a
retaining wall and different from the 3D mechanism (note that the direction of the load

M.Sc. Civil Engineering 6.3. 2D v. 3D


113 Chapter 6. Comparison of the p-y Design Procedure and the FE Model

Figure 6.9: Horizontal displacement field for the 2D model (in [m])

Figure 6.10: Horizontal displacement field for the 3D model (in [m])

is changed in the figures). The maximum horizontal displacement in the soil in 2D is


ux = 0.6125 m and the respective in 3D ux = 0.05397 m, almost 10 times less. Another

6.3. 2D v. 3D M.Sc. Civil Engineering


Chapter 6. Comparison of the p-y Design Procedure and the FE Model 114

observation is the location of the centre of rotation of the pile. While in the 2D model
the centre is almost at the toe of the pile (completely rigid pile) in the 3D model is found
at 1/3 of the height of the monopile. This has to be studied separately, as the bending
resistance of the monopile is the main factor influencing this condition.

The are two main reasons for it. Using a 2D model we are assuming the hypothesis
of plane strain. However, a pile problem is likely to be controlled by a 3D geometry.
Recalling Figure 6.4 and subsequent ones, the maximum shear stresses are not find in
the middle plane of the monopile. Obviously, with a 2D model these effects are not
taken into consideration. The other might be the use of wrong boundary conditions in
the 2D model, as no symmetry could be defined in the z axis of the 2D model.

Going back to the 3D model, another proof of the 3D influence can be seen when
plotting the points in the soil which are yielding.

Figure 6.11: Plastic equivalent strains in the 2D model

Again, a typical active-passive schemes can be seen in the 2D model, whereas in the
3D model (although similar to an active-passive outlines) is not that close. The plastic
flow occurs in much more higher areas (close to the mudline) an with a tridimensional
aureole.

M.Sc. Civil Engineering 6.3. 2D v. 3D


115 Chapter 6. Comparison of the p-y Design Procedure and the FE Model

Figure 6.12: Plastic equivalent strains in the 3D model

6.3.1 Absorbing boundaries

Although the 2D option was discarded, a dynamic analysis was performed on it. A
problem with elastic wave reflection was expected in the soil domain. The problem of
the reflection of the waves in the boundaries can be overcome using damping elements.
When modelling geomechanical systems, an infinite domain is often required to repres-
ent the ground or other larger bodies. For this type of situations, using non-reflecting
elements become useful. These conditions, which the software applies automatically to
the nodes, will prevent the artificial wave reflections generated at the boundary from re-
entering the model and contaminating the results. When those conditions are included,
the algorithm computes an impedance matching functions for all the boundary segments
based on an assumption of linear material behaviour. The problem can also be addressed
using numerical absorbing boundary conditions based on a damped wave equations.

Figure 6.13 shows the propagation of the elastic wave generated by a push load in
dynamic conditions.

ABAQUS can treat this types of problems using dash-pots. Nonetheless, dash-pots
need to be attached to a rigid element and need to be defined with a direction and an
impedance. Dash-pots become useful when, for example, isolators in earthquake engin-
eering are modelled, situation in which a building is isolated from the soil.

6.3. 2D v. 3D M.Sc. Civil Engineering


Chapter 6. Comparison of the p-y Design Procedure and the FE Model 116

Figure 6.13: Wave propagation problem in the 2D geometry

The other option, which has been the preferred one for the monopile foundations, is
the use of the so-called infinite elements, or acoustic elements as ABAQUS calls them.
This type of elements, with lots of applications in geotechnical engineering absorb any
type of wave (P, S, Love...), but they have a certain computational cost and are difficult
to implement in the model, as they are not available in the user interface and they re-
quire a particular line of code in the .INP file to be accessible and usable in ABAQUS.
One of the disadvantages of using infinite elements is the fact that they do not only
“absorb” the waves, but also the stress and displacements fields, which means that the
model need to be bigger in order to use the full potential capabilities of these special
elements.

To present a comparison, one point of the domain was selected to compare the
model with and without absorbing boundaries. While the absorbing boundary keeps
the solution stable in time, the elastic rebound creates and important distortion in the
displacement fields (and consequently in the stress fields). The damping and absorbing
effect can be seen in Figure 6.15.

After performing a similar analysis with the 3D model, it was found that wave
reflection was not a problem, so the use of infinite elements was avoided initially in the
full model. Wave affects may become visible if the elastic strain tensor is plotted in

M.Sc. Civil Engineering 6.3. 2D v. 3D


117 Chapter 6. Comparison of the p-y Design Procedure and the FE Model

Figure 6.14: Influence of the infinite elements in the stress field

Figure 6.15: Example of the influence of the absorbing boundaries

ABAQUS, but it is not significant as the geostatic stresses and the plastic tensor are
much more important. This happens as well because the low velocity of application of
the cyclic load. As a safety measure, lately the absorbing boundaries were also included
in the 3D model, taking advantage of the effects studied in the 2D model.

6.3. 2D v. 3D M.Sc. Civil Engineering


Chapter 6. Comparison of the p-y Design Procedure and the FE Model 118

6.4 Comparison FEM-API (DNV)


The comparison between the FEM model and the standards for offshore foundations
design is shown in Figures 6.16, 6.17 and 6.18.

Figure 6.16: Comparison of a mid plane of the monopile (p-y 0.5 m)

Figure 6.17: Comparison of a mid plane of the monopile (p-y 12 m)

The results between FEM and standards are very different. At first sight, the FEM
method looks like it proposes a stiffer behaviour of the soil. This means that the soil
has much more capacity than it is believed using the codes. The codes therefore are in
the side of caution by a great deal. This however is not constant along the monopile.

M.Sc. Civil Engineering 6.4. Comparison FEM-API (DNV)


119 Chapter 6. Comparison of the p-y Design Procedure and the FE Model

Figure 6.18: Comparison of a mid plane of the monopile (p-y 30 m)

The closer we are to the mudline, the closer the FEM and the p-y method are. The
upper parts of the monopile are subjected to high stresses and the amount of soil is
small, so it is the weaker part of the soil. The deeper it goes, the more pronounced is the
difference. From a certain point (around −15 m) there is no point in plotting more p-y
curves comparison graphs, as the displacement in the FEM model is almost zero for the
service loading conditions, and the soil behaves in a elastic range. Another remarkable
point on the graphs is the fact that the initial slope in the FEM and the p-y curves are
also different (the initial stiffness of the soil changes between models). And the finite
element software, when reaching the plastic zone or the yielding points in the soil, start
to loose continuity and the curve is not very clear some times (the parabolic shape is
somewhat missing). This probably is caused by the automatic integration steps of the
plastic strains in the Mohr-Coulomb constitutive model.

Therefore, the p-y method is overestimating displacements in the soil. Regarding the
stiffness of the soil, the horizontal subgrade modulus can be easily calculated with the
FEM results an be compared to the ones used in the p-y method. Plotting the results
in Figure 6.19 kh can be easily understood.

What can be seen is that with depth, kh in the FEM model gets bigger with respect
to the linear approximation of the p-y method. Result after 15 m depth could not be
plotted, as the displacements are very close to zero. It can be seen as well a relation
between the initial slopes in the p-y curves with this plot, where in the first meters of the

6.4. Comparison FEM-API (DNV) M.Sc. Civil Engineering


Chapter 6. Comparison of the p-y Design Procedure and the FE Model 120

Figure 6.19: Horizontal subgrade reaction modulus variation with monopile depth

monopile the stiffness between models is somehow similar and the deeper the monopile
goes, the steeper the FEM initial stiffness gets.

Other reasons for the difference between methods may stem from the limitations
of the p-y method commented in previous chapters. First, the p-y method is missing
different effects that the FEM (although with a poor constitutive model) accounts for,
for instance, the lateral skin friction and pile tip influence in large diameter monopiles
is very big, as it generates additional forces at the pile, meaning less displacement in
the soil. Compared to small diameter monopiles (those used in the p-y method) large
diameter monopiles has more rotation than deflection, which means that additional shear
stresses are developed resulting in a bigger resistance. The fact that the walls of the
monopiles are more separated at the toe of the pile creates more resistance (larger arm
if we think on moments an couples of forces). Ultimately, large diameter monopiles
mobilise more soil than small diameter monopiles which means more energy dissipation
and bearing capacity.

6.5 Peak loads


The difference in the peak load given by the different methods can be worked out fitting
and adjusting function to the results from the FEM model. It can be seen that the close
they get to the seabed, the closer the p-y method gets to the FEM analysis. The further
in depth the more differences we find. What seems to be clear is that the capacity of

M.Sc. Civil Engineering 6.5. Peak loads


121 Chapter 6. Comparison of the p-y Design Procedure and the FE Model

the soil according to the FEM model is greater than the one found with the p-y method.
For the first meters the methods are relatively accurate. For high depths, the difference
is more pronounced. Specially if we compare the initial stiffness or slope, which is lower
in the p-y than in the FEM model. For intermediate depths, the FEM capacity seems to
exceed the p-y capacity in a 50% more. It is important to remember that the p-y does
not take into account pile toe resistance (base shear) or skin friction (lateral shear), and
the correspondent base moment and skin moments which add extra capacity the FEM
model.

Figure 6.20: FEM fitting (p-y 0.5 m)

For mid depths, an elastic trend can be seen in the plots, and the when yielding, the
points start to scatter a bit, not following a clean path in the plastic regime.

For greater depths it makes no sense to force the soil to reach yielding, mainly be-
cause the results for upper parts of the monopile would be unrealistic, but it is clear
that the initial stiffness is improved with depth. This can also be checked out in Figure
6.19.

6.5. Peak loads M.Sc. Civil Engineering


Chapter 6. Comparison of the p-y Design Procedure and the FE Model 122

Figure 6.21: FEM fitting (p-y 12 m)

Figure 6.22: FEM fitting (p-y 30 m)

6.6 Static-Dynamic response


In the next section, some of the most interesting findings during cyclic and dynamic
simulations will be discussed. In Figure 6.23 the horizontal displacement of the monop-
ile under the load represented in Figure 3.7 is plotted versus time. First thing to notice
is the big influence of the first cycle. In fact, the peak displacement is found for the
first push. After that, the peaks are very similar and the horizontal displacement is
stabilised and decreased slightly after 40 s of simulations. Another important issue with
the first cycle is the directionality of the action of the load. This affect the whole life of

M.Sc. Civil Engineering 6.6. Static-Dynamic response


123 Chapter 6. Comparison of the p-y Design Procedure and the FE Model

the monopile. Let’s assume that in a real situation with real environmental conditions,
there is a preferential wind or wave direction. Automatically, the soil will have “memory”
and a preferred direction of oscillation and tilting of the monopile will be determined.
Furthermore, is has been reported that this effect affects also to the subsidence and
scour in the surroundings of the monopile. When the monopile deflects and a crack is
formed between the steel and the soil (as it was demonstrated in previous sections of
this Chapter), assuming that the soil is saturated with water, during the next cycle that
water in the crack will be “pumped” upwards and in the same direction of the monopile
movement, creating erosion and making that the history of displacements increase during
the lifespan of the wind turbine. With respect to the plot and the directionality, it can
be seen that the range of displacements goes from +0.047 m to −0.025 m approximately.
This range is consequence of the initial push and the directional effect mentioned above.
In any case, it can be seen that the range of displacements is around 0.06 m which is not
a very high value for the magnitude of the structure above the monopile (and without
considering bending moments and vertical stresses).

Figure 6.23: Horizontal displacement of the monopile

Consequently with the displacement of the monopile head, it is important to see


how the soil is behaving under that cyclic loading. Again the influence of the first
load peak is found. Perhaps, the results should ignore that initial effect, but for the
sake of this research purposes, the graphs have been left untreated. In Figure 6.24
it can be seen that the effects of the first push are more important in the soil. As
a plastic law have been used, the mudline is subjected to the biggest displacements

6.6. Static-Dynamic response M.Sc. Civil Engineering


Chapter 6. Comparison of the p-y Design Procedure and the FE Model 124

during the firsts seconds of the simulation, therefore getting permanent (plastic) strains,
indicating that the maximum capacity has been reached at that level. After that, the
displacement history tends to be maintained and starts to decrease slightly after 20
seconds of simulation.

Figure 6.24: Horizontal displacement of the mudline

These plots also remark the importance of the stress and strain history in problems
involving cyclic conditions. Another approach for this could be treating the problem of
monopiles subjected to lateral cyclic loading as a problem of fatigue in the soil. This
may become useful for simulating the whole life of the structure (< 109 ) cycles and not
only 40 (or 40 seconds). Doing a rough number, an assuming a cycle per second, ≈ 109
corresponds with 31 years, and probably nobody will be willing to run a simulation
like that. Therefore, the long term knowledge of the monopiles is topic where extensive
research is needed.

The influence in terms of horizontal displacements in the vicinity of the monopile was
also plotted. It was found that after the simulation, and assuming that after that time
an asymptotic degradation has been reached, it can be concluded that the soil affected
by the monopile In Figure 6.25 can be seen this trend. It is also important to notice
the unsymmetrical influence (again, because of the directionality of the problem). It is
expected that the horizontal influence reaches a distance of ≈ 2D.

After checking the horizontal displacement history, the vertical one was also analysed.

M.Sc. Civil Engineering 6.6. Static-Dynamic response


125 Chapter 6. Comparison of the p-y Design Procedure and the FE Model

Figure 6.25: Horizontal displacement for n = 40

Although apparently under horizontal load conditions no vertical influence should be de-
tected, there is a certain downwards movement of the monopile. Figure 6.26 represents
the vertical movement of the monopile during the simulation. A settlement of around
0.006 m is found. After speaking with monopile designers, they have found that the set-
tlement of the monopiles is an effect that they have notice that happen within time, but
they are not very sure how to control or predict it. In the numerical analysis the vertical
movement of the foundation was also noticed when studying the contact elements, in
which a vertical slip variable was found. It is also important to notice that the vertical
slip is influenced by the friction of the contacts and by the way the pile deflects, finding
more interaction the more rigid the pile is.

So the lateral action influences also the vertical behaviour of the monopile. With
respect to the soil, it is found that due to cyclic loading, subsidence around the monopile
is produced. The model does not reproduce scour or similar alterations in the mudline.
The consequence of finding a significant subsidence of around 0.2 m close to the mudline
is a reduction of the soil reaction in that area, influencing the interaction between the
soil and the pile. In fact in Figure 4.22 can be seen how the point of maximum reaction
is not at mudline, but a few meters below that level.

Having analysed that, the next figures try to give more evidences to support this
findings. Plotting the vertical displacement field, it can be appreciated that close to
the monopile the soil has moved downwards and there is a certain heave in the vicinity
around that area. This is also linked to the shear failure on the surface of the soil. Notice

6.6. Static-Dynamic response M.Sc. Civil Engineering


Chapter 6. Comparison of the p-y Design Procedure and the FE Model 126

Figure 6.26: Vertical displacement of the monopile

Figure 6.27: Vertical displacement of the mudline

that in the plot only one direction is represented. In a and fully 3D real case, what could
be seen is an areola around the monopile and not only in two extremes of the monopile.

A combined representation of the equivalent plastic strains can also give important
information. In Figure 6.29 the progression of the “damage” in the soil motivated by
the cyclic action can be seen. It is also found that the shear mechanism is not fully
developed (as it is expected otherwise) as the soil is not close to its full bearing capacity,

M.Sc. Civil Engineering 6.6. Static-Dynamic response


127 Chapter 6. Comparison of the p-y Design Procedure and the FE Model

Figure 6.28: Vertical field of displacements for n = 40

although some plastic strains can be seen in the upper part of the monopile influence
zone.

Figure 6.29: Damage in the soil. The whiter the colour, the more accumulation of
plastic strains.

In order to continue with the analysis of the contacts, the opening of the monopile at
mudline level has also been obtained in the FEM model. It appears to remain constant
throughout the simulation. The relevance of the opening is however not sufficiently
understood with this study. The continuous opening and closing cycles may induce a

6.6. Static-Dynamic response M.Sc. Civil Engineering


Chapter 6. Comparison of the p-y Design Procedure and the FE Model 128

concentrated erosion phenomena close to the mudline, but this require a much more
detailed analysis, incorporating a problem of flow in saturated media.

Figure 6.30: Crack opening at mudline level

In the Mustang Island tests, and for the derivation of the equation in the codes, the
interaction between the monopile and the soil is assumed to be smooth. This implies
that technically, the resistance in terms of shearing forces developed around the mono-
pile and in the base are neglected. This is not the case of the FEM model, in which
the contacts have been modelled after different implications. After seeing the existing
differences between the FEM model and the p-y curves from the standards it is import-
ant to question where that deviation come from. In order to quantify these effects, the
integration of shear stresses on the body of the monopile an at the toe have been plotted
it graphs 6.31 and 6.32.

It is difficult to know how much of that skin shear goes into the increase of the
capacity according to the p-y method. To do so, shear forces per monopile meter should
have been necessary so that the influence on the p-y curves could have been determined
numerically.

For the case of the base there is a certain upwards tendency. The more cycles the more
base shear stress. This might be related to the vertical displacement of the monopile.
In fact, the value of these stresses at the base are similar to the load applied on top of
the monopile.

M.Sc. Civil Engineering 6.6. Static-Dynamic response


129 Chapter 6. Comparison of the p-y Design Procedure and the FE Model

Figure 6.31: Integration of the shear skin on the monopile surface

Figure 6.32: Integration of the base shear stresses

6.7 Influence of the cyclic loading

At a peak in the load sinusoidal signal, the maximum stresses found in the soil are
roughly 400 kP a peak (negative for compression).

But with the geostatic step finding the maximum horizontal stresses can be cumber-
some. In Figure 6.34 the points of maximum reaction are pictured. Those two points
corresponds with “passive” earth pressure areas. One is right in the base of the mono-
pile, and the other is 1/3 of the pile length below the mudline. This tool may become

6.7. Influence of the cyclic loading M.Sc. Civil Engineering


Chapter 6. Comparison of the p-y Design Procedure and the FE Model 130

Figure 6.33: Stress field (load peak) (in [Pa])

useful as well for designers trying to optimise te embedding length.

Figure 6.34: Maximum horizontal stresses during peak load

M.Sc. Civil Engineering 6.7. Influence of the cyclic loading


131 Chapter 6. Comparison of the p-y Design Procedure and the FE Model

The horizontal field of displacements show movements of 0.04 m peak. No shear


mechanism, or active-passive shapes as the soil is far from a limit state.

Figure 6.35: Horizontal displacement (load peak) (in [m])

6.8 p-y dynamic

In order to compare adequately, the first peaks were omitted of the analysis (because of
the great influence of the first push). The the procedure was similar to the one used for
the static p-y curves. Instead of using a ramping static load, different sinusoidal loads
were added to get the variation of the load. Owed to the complexity of this calculations,
and the amount of time used by the computer in doing the computations, only 3 points
have been represented in Figure 6.36. Initially it is intuitive to see that the initial slope
of the p-y has changed. the second implication is that the more cycles of simulation, the
lower the capacity of the soil gets. This is a typical property of materials, so the soil
presents a certain amount of fatigue. What this means is that perhaps, the static p-y
curve determination should be substituted by something in the fashion of the well-known
S-N curves in fatigue analysis of structures. If the soil presents degradation regarding the
cycles involved, then this might represent an important advance in the understanding of

6.8. p-y dynamic M.Sc. Civil Engineering


Chapter 6. Comparison of the p-y Design Procedure and the FE Model 132

the behaviour of the soil.

Figure 6.36: P-y static FEM vs. p-y dynamic FEM vs. p-y according to the codes

M.Sc. Civil Engineering 6.8. p-y dynamic


Chapter 7

Conclusions and
Recommendations

The aim of this work was to understand the behaviour of laterally loaded monopiles and
to understand the limitations of the current p-y method for design, comparing station-
ary and transient cases. For that purpose, the document was clearly structured in a
first part in which a numerical model of the problem was developed (as a tool for later
analyses) and a second part in which the actual p-y methodology was understood and
compared with the numerical calculations. Subsequently, some implications for practical
design have been explored:

7.1 FEM modelling


• It has been concluded that it is possible to model a full monopile-soil system using
large scale FEA, provided that the designer is aware of the necessary simplifications
of the model.

• The size of the model is mainly influenced by the boundaries, i.e., the boundary
conditions need to be sufficiently further away from the loaded element in order to
avoid the distortion of the stress field by the boundary conditions. In many cases,
for the models in this thesis, this was the main constrain, and not the size to avoid
the wave reflection. With the dimensions shown in this thesis, the accuracy was
proven sufficient.

• The models are very sensitive to the constitutive model and the contact elements

133
Chapter 7. Conclusions and Recommendations 134

employed in the interface between soil and monopile. The selected constitutive
model, although it is very simple, continues with the trend followed by the main
researchers in monopile design due to the complexity of the overall system. Fol-
lowing the same principle, the frictional behaviour of the contacts was chosen from
previous reports, leading to results within range and order of magnitude.

• It is important to use variable meshes to account for particularities in the vicinity


of the monopile and improve the accuracy of the model. Finer meshes did not
improve accuracy significantly, however, if the interaction of the contacts want
to be represented within a small range of tolerance, the monopile needs sufficient
elements so that they can keep up with the subdivisions already made in the
soil. Furthermore, the monopile has a circular geometry, but the elements are
“straight” tetrahedrons, so sufficient discretisation needs to be provided. This is
also important to avoid angular distortion in the mesh. Owed to this distortion,
angular interfaces might lock, resulting in concentrating stresses at the sides of the
contact elements.

• Under lateral loads, it is possible to model the monopile as a solid body and use
equivalent properties for the monopile, avoiding extra interactions in the inner skin
of the hollow element.

• To avoid localised distortion under the application of loads in nodes, the use of
rigid elements, e.g. rigid caps, is advisable.

• It is important to find the right balance between number of elements and con-
stitutive law of the soil, as the computational time might be restrictive.

• Implicit analyses have been the preferred option, as explicit analyses required very
small time steps to guarantee convergence of the solution.

• The use of a 2D model did not produce the expected outcomes. Theoretically a
2D model might approximate the response of a middle plane of a wide monopile
accurately enough. However, this could not be proven, and the behaviour of a 2D
model was closer to the one of a flexible retaining wall than to the monopile’s.

7.2 Validation of the FEM model


• When there is no empirical data available or previous FEA to match results, an
extensive validation process has to be carried out to check that the model is giving

M.Sc. Civil Engineering 7.2. Validation of the FEM model


135 Chapter 7. Conclusions and Recommendations

the correct response.

• The initiation of the analysis with a geostatic step is very important to get the
real stresses in the soil for the subsequent steps of simulation.

• The absence of pore water pressure implies that the stresses obtained in the sim-
ulations are effective.

• From the calculations, the differences between a slender and a chunkier monopile
in terms of deflection were checked. The implications arising from this affect the
p-y methodology, as it is based in slender monopiles.

• It can be concluded that the use of damping elements is a requirement for dynamic
simulations.

• The dynamic response of the model was also checked using a soil column, obtaining
a numerical solution almost identical to the analytical solution.

• It can be concluded that the first natural mode shape of the soil-monopile system
is close to 1 Hz.

• The frequency of the load is very small compared to other dynamic problems (for
instance, a moving train passing over a rail track). This means that the velocity of
application of the loads in the turbine in relation with the celerity of the wave in the
soil media are very close, which means that the influence of a wave propagating in
the domain is negligible. Although this is true, it is also true that when cyclic loads
are applied in the domain for a continuous period of time, more and more energy
is put into the media, which eventually might suffer from energy accumulation and
wave reflection. Thus, it is recommended to use damping boundaries or acoustic
elements to reduce those effects when simulating cyclic loading scenarios for long
periods of time.

• Before finding the p-y curves of the soil, the condition of equilibrium in the system
needs to be found and satisfied.

• It is also interesting to study the influence of the stiffness in the localisation of


the centre of rotation of the monopile. For slender monopiles it can be found that
there are two points with null deflection. For wide monopiles, there is only one,
and it is usually located at 1/3 of the monopile length from the base.

7.2. Validation of the FEM model M.Sc. Civil Engineering


Chapter 7. Conclusions and Recommendations 136

7.3 p-y method


• From the comparison of Method A and B, it seems like there is a good agreement
in the initial slope, but there is a certain variation in the ultimate strength Pu . For
greater depths, Method B appears to give a greater ultimate resistance, whereas
for shallow depths, it gives a smaller value of Pu compared to Method A. This
appears to be feasible, as the upper zones in the soil are not subjected to the
confinement of the lower areas.

• The soil resistance is intimately related to the failure mode, which is somehow
dependent on the flexibility of the pile. The slenderness parameter of the piles
in the Mustang Islands was L/D = 34.4, whereas the monopile studied in this
dissertation has a slenderness ratio of L/D ≈ 6. For modern monopiles, a slender
ratio < 10 is common. Notice the massive difference, moreover when the bending
stiffness is related to the slenderness and the soil reaction is influenced as well by
the flexibility of the structure.

• This methodology was originally devised with slender piles. This implies that for
wide monopiles (or non-slender monopiles) the formulation need to be updated.
Otherwise the results are far from represent the real problem.

• The behaviour of the pile will change effectively depending on whether the monopile
has been drilled or nailed. This methodology does not take into account the initial
stress field of the soil (altered during the installation process) nor the possible
influence of the vertical load of the structure. Also the level of confinement inside
de monopile and the plugging effects are not accounted for in the p-y derivation.
The soil plugging at the pile toe and the shear resistance at that point is not
considered in the p-y formulation. The “skin” friction is a term that has not
been considered for the determination of the ultimate resistance (with the failure
modes) neither in Method A nor in Method B. Therefore, a term, depending on the
diameter and shape of the pile, is missing from the analysis. It is, in fact, assumed
a perfectly smooth pile, as the Rankine criteria has been used. Some authors have
proposed alternative methods including in the net ultimate lateral resistance the
frictional resistance.

• It has been found that the most influential parameter are kh when finding the initial
slope of the p-y curves and D for the ultimate resistance of the soil. While D is
given by design, is a chosen value, kh is assumed a property of the soil dependent

M.Sc. Civil Engineering 7.3. p-y method


137 Chapter 7. Conclusions and Recommendations

on the relative density.

• The codes merely recommend a degradation of a 10% in comparison to the static


p-y curves. It is indeed surprising that a problem with such a dynamic nature is
treated in such a trivial way in the codes. But the experimental data is scarce,
and the results are certainly biased.

• It can be said that the p-y method is an experimental analysis backed by years
of experience in the matter without any significant failures, but it needs to be
updated and adapted to recent research and current designs.

7.4 Comparison p-y method and FEM


• The matching in upper soil levels (−3 m) between FEM and API is relatively
good, however from that point, the API starts to give less capacity to the soil,
i.e. PuF EM > PuAP I . This leads to inaccurate design of the monopiles, as they are
oversized. Using the API means using a conservative method. It should be also
noticed that although the geotechnical design of the monopiles is controlled by the
p-y interaction, the structural design of the monopile is controlled by the Fatigue
Limit State (FLS) during installation, so in many instances the dimensions of the
monopiles are owed to the fatigue factor.

• From a depth of about −15 m the p-y curves are difficult to derive, as for normal
loading conditions the level of deformation in the soil is negligible.

• The main reason in the difference of the initial slope or stiffness is believed to come
from the fact that the p-y method takes the variation of kh linear with depth, when
in reality the FEM model provides a different value. It can be concluded that the
variation kh is therefore non-linear in depth.

• The main differences from the FEM and the API come from the API approach.
The p-y curve method stems from the analysis carried out in the Mustang Islands.
It is important to bear in mind that, in experimentation, the outcomes are com-
pletely influenced by the scale and the conditions of testing. Those tests were
carried out for only one type of pile (circular, slender and metallic), with one only
diameter, and in only one type of soil (sand), subjected to seven load cases. Given
the complexity of soil-structure interaction problems, it seems like a bold idea to
extrapolate the results to any other case. As a matter of fact, further research on

7.4. Comparison p-y method and FEM M.Sc. Civil Engineering


Chapter 7. Conclusions and Recommendations 138

full scale piles uncovered a huge bias in the results, in comparison with the initial
estimations. Subsequently, correction factors (A and B) had to be introduced to
account for that. Anyway, small diameter monopiles are subjected to less overall
rotation and horizontal displacement. Therefore more soil resistance and the shear
stresses at the base and due to skin friction are not considered in the API. From the
calculations with the FEM, these effects are estimated in a 30 − 40% of additional
forces in the FEM derivation of the p-y curves with respect to the API curves.

• The plane of maximum normal forces corresponds with the mid plane of the mono-
pile, but it does not coincide with the plane of maximum shear forces, speaking in
terms of resultants.

• The level of plugging inside the monopile influences the internal friction under
lateral loading conditions and therefore the deflected shape of the monopile. Albeit,
the stress state of the soil inside is difficult to model specially after the aggressive
installation procedures.

7.5 Comparison static-dynamic (cyclic)


• It could be checked the influence of the direction of application of the force in
the FEM model. Although this may seem trivial, in a real wind turbine with
currents and strong winds with a clear preferential direction, this fact can cause
a preferential direction of displacement of the wind turbine. Eventually, this may
cause problems to the nacelle, as the most restrictive condition for its operation
is the rotation. It can be concluded that the history of stresses can have a strong
influence in the overall behaviour in the long term analysis.

• With FEM could also be checked that under lateral conditions, there is a certain
influence on the vertical behaviour of the monopile, creating additional settlements
in the structure.

• Because of the plastic strains at mudline level, the opening crack between monopile
and soil increases in time. Although it cannot be analysed with FEA, in cyclic
conditions water can penetrate and be pumped out of that crack, creating scour
problems and erosion. Ultimately erosion can cause a reduction in the resistance
of the soil, as the level of the mudline around the monopile is lowered.

• The densification of the soil in the surroundings of the monopile could neither been
modelled nor studied.

M.Sc. Civil Engineering 7.5. Comparison static-dynamic (cyclic)


139 Chapter 7. Conclusions and Recommendations

• The influence at the mudline of the cyclic loading expands to 10 m around the
extremes of the monopile, appearing to remain constant after the first load cycles.

• In the p-y curves, the influence of a cycle load has two direct implications according
to the results. The first one is a reduction of the initial stiffness or slope (albeit this
is suspected to happen because of the initial push and influence of the direction of
the load). The second one is the more load cycles the less resistance in the soil is
found. Cycling conditions create a degradation in the soil stiffness.

• According to the simulation, shear stresses at the base and skin increase in time,
which may lead to think that the capacity of the cyclic p-y curve is improved
with time. However, although shear stresses are higher, the p-y curves present
lower capacity due to the accumulation of plastic deformations in the model. An
improvement in the capacity of the soil might be associated with the densification
of the soil around the monopile, but that phenomenon has not been explored.

• In cyclic conditions, in which both sides of the surrounding soil are being loaded
alternatively, a passive Rankine thrust can be slightly devised sometimes.

• A forty-second simulation (40 cycles) is not enough to find out general trends in
the behaviour of the monopile, but helps to understand qualitatively what may
happen.

• During the 109 cycles in a lifetime of a wind turbine (or 40 in the case of this
thesis) the plastic deformation in the soil is accumulative and it increases with
time. Also if the soil conditions are undrained, the excess of pore water pressure
due to cycling loading may affect the long term behaviour.

7.6 Extra recommendations for designers


• Aside from the use of national standards and the use of particular software for
monopile design, it is recommended to create a FEM model for a large diameter
monopiles and see the difference and the sensitivity of the different methods, and
the behaviour of the soil.

• It is very important to characterise the dynamic loads, and not only the static
ones.

• Pay special attention to the constitutive model and the associated failure criteria.

7.6. Extra recommendations for designers M.Sc. Civil Engineering


Chapter 7. Conclusions and Recommendations 140

• Always check the equilibrium of the structure in the FEM model.

• Carry long term analysis, studying the whole life of the structure and not only the
static case.

7.7 Recommendations for the improvement of the codes


• Carry out extensive revision of the p-y formulation, starting from the derivation
of the equations to include the effects of the stiffness and geometry of the piles.

• Redo the failure criteria in limit state of the monopiles, including the determination
of the influencing geometry parameters.

• Revise the influence of the parameter kh and its linear variation in depth.

• Include revision for wide (non-slender monopiles) and correction factors, e.g. shear
stresses at base and skin, plugging effects and installation effects.

• Include cyclic and dynamic effects with following a more technical approach and
not only with a constant degradation factor.

• Carry extensive analysis in the long term, all along the whole life cycle.

• Include liquefaction effects and influence of the pore water pressure.

• Provide further recommendations for multi-layered soil and not only for homogen-
eous soil. This include analysis when the seabed is a bed rock.

M.Sc. Civil Engineering 7.7. Recommendations for the improvement of the codes
Chapter 8

Conclusiones y Recomendaciones

El objetivo de este trabajo ha sido entender el comportamiento de monopilotes cargados


lateralmente y entender las limitaciones del método p-y, comparando casos estacionarios
y transitorios. Para ello, el documento ha sido claramente estructurado en una primera
parte en la que el modelo numérico fue desarrollado (como una herramienta para su pos-
terior uso) y una segunda parte en la que la metodologı́a p-y fue entendida y comparada
con los resultados numéricos. Consecuentemente, alguna implicaciones para el diseño
han sido exploradas:

8.1 Modelo con Elementos Finitos


• Se puede concluir que realizar un modelo completo del sistema suelo-monopilote
usando el MEF es factible, siempre que el proyectista sea consciente de las sim-
plificaciones del modelo.

• El tamaño del modelo viene condicionado principalmente por las condiciones de


contorno, i.e., los contornos deben estar lo suficientemente alejados del elemento
solicitado para evitar la distorsión del campo de tensiones. En muchos casos, para
los modelos expuestos en esta tésis, este ha sido el mayor condicionante, y no
el tamaño para evitar la reflexión de ondas. Con las dimensiones propuestas se
conseguı́a la suficiente precisión en el modelo.

• Los modelos son muy sensibles a los modelos constitutivos y a los elementos de
contacto usados en la interfaz suelo-monopilote. La ley constitutiva elegida, si bien
es simple, sigue con la tendencia seguida por muchos investigadores debido a la

141
Chapter 8. Conclusiones y Recomendaciones 142

complejidad general del modelo. Siguiendo el mismo criterio, el comportamiento


friccional de los contactos fue elegido acorde a estudios previos, coincidiendo los
resultados obtenidos con los esperados en rango y orden de magnitud.

• Ha quedado probada la importancia de usar mallas variables para considerar las


particularidades alrededor del monopilote y mejorar la precisión del modelo en esa
zona. A menudo, una malla más fina en general no mejoraba significativamente
la exactitud de los resultados, sin embargo, si la interacción de los elementos de
contacto quiere ser captada lo suficientemente bien, el monopilote debe ser discret-
izado de tal manera que se establezca una cierta continuidad entre las divisiones
en el suelo y en el pilote. Además, debido a que la geometrı́a semicircular em-
pleada fue mallada usando elementos tetrahédricos de aristas rectas, es necesario
un número suficiente de estos elementos. Esto es especialmente importante para
evitar la concentración de tensiones en los nodos laterales de los contactos debido
a la angularidad de los elementos y la distorsión angular.

• Bajo condiciones de carga lateral, es posible modelar el monopilote como un ele-


mento sólido y usar propiedades equivalentes, evitando interacciones extra en el
interior del elemento hueco.

• Para evitar efectos locales al aplicar la carga en los nodos superiores del monopilote,
es remendable el uso de elementos rı́gidos como rigid caps.

• Es importante encontrar el compromiso entre número de elementos y ley con-


stitutiva del modelo, ya que el tiempo computacional puede ser una limitación.

• El método implı́cito ha sido el preferido, ya que el explı́cito requiere pasos tem-


porales muy pequeños para garantizar la convergencia de la solución en análisis
dinámico.

• El uso de un modelo 2D no produjo los resultados esperados. Teóricamente un


modelo 2D deberı́a poder usarse para aproximar la respuesta en un plano medio
de un monopilote lo suficientemente ancho. Sin embargo, esto no ha posido ser
demostrado, resultando en que el comportamieto del modelo 2D era más próximo
al de un muro de contención flexible que al de un monopilote.

M.Sc. Civil Engineering 8.1. Modelo con Elementos Finitos


143 Chapter 8. Conclusiones y Recomendaciones

8.2 Validación del modelo de elementos finitos


• Cuando no hay datos empı́ricos o investigaciones previas en la materia, un proceso
extensivo de validación es necesario para garantizar la confianza en los resultados
obtenidos.

• Es importante comenzar los análisis con un paso geoestático en el programa para


inicializar el campo tensional real en el suelo, siendo este propagado en los siguientes
pasos de cálculo.

• La ausencia de presión de poro y agua en el modelo garantiza que las tensiones


obtenidas durante las simulaciones son efectivas.

• De los cálculos, las diferencias entre monopilotes esbeltos y no esbeltos fueron


comprobadas. Estas diferencias son muy importantes, ya que afectan directamente
a la metodologı́a p-y, ya que está basada en pilotes esbeltos.

• La importancia de los elementos absorbentes en los contornos es muy importante


en el cálculo dinámico.

• La respuesta dinámica del modelo fue validada a su vez con una columna de
suelo, obteniendo una solución numérica prácticamente coincidente con la solución
analı́tica.

• Se puede decir que la primera frecuencia natural del sistema suelo-monopilote


ronda 1 Hz.

• La velocidad de aplicación de la fuerza es pequeña comparado con otros problemas


dinámicos (como por ejemplo, un tren pasando sobre raı́les). Esto implica que la
velocidad de aplicación de la fuerza y la velocidad de propagación de la onda en el
medio elástico son cercanas, por lo que los efectos derivados de este fenómeno son
despreciables. Aunque esto sea ası́, es cierto que en problemas en los que el modelo
se somete a una carga cı́clica, la energı́a del sistema crece y crece pudiendo llegar a
un punto en el que la influencia sea notable si no se ponen elementos absorbentes.
Por esto, se recomienda el uso de contornos absorbentes para reducir estos efectos
al simular carga cı́clica en el modelo.

• Es necesario satisfacer la condición de equilibrio en el modelo de elementos finitos


antes de proceder a la obtención de las curvas p-y.

8.2. Validación del modelo de elementos finitos M.Sc. Civil Engineering


Chapter 8. Conclusiones y Recomendaciones 144

• Es también interesante estudiar la influencia de la rigidez del monopilote en la


situación del centro de rotación de la estructura. Para monopilotes esbeltos se
encontró que hay dos puntos con desplazamiento nulo, mientras que solo hay uno
en el caso de monopilotes rı́gidos, normalmente situado a 1/3 de la longitud del
pilote desde la base del mismo.

8.3 Método p-y


• De la comparación del Método A y B parece que existe una buena correlación
entre las pendientes iniciales, aunque existe una cierta variabilidad en el valor de
la resistencia última Pu . A mayor profundidad, el Método B parece que da una
mayor carga última, mientras que para zonas superficiales da un menor valor que
el Método A. Esto parece razonable dado que las zonas superiores estas sometidas
a un menor nivel de confinamiento del suelo.

• La resistencia del suelo está intimamente relacionada con el modo de fallo, que en
este caso es dependiente de la flexibilidad del monopilote. La esbeltez en los pilotes
empleados para los ensayos en las Islas Mustang era de L/D = 34.4, mientras que
la del monopilote estudiado en esta tesina de máster es L/D ≈ 6. Para monopilotes
modernos, la esbeltez suele ser < 10. Es importate destacar la gran influencia de
este paraámetro, y la gran diferencia existente entre los pilotes empleados para la
derivación de los métodos A y B y los monopilotes actuales, especialmente cuando
la rigidez flexional del monopilote influencia la reacción en el suelo.

• El comportamiento del monopilote cambia notablemente dependiendo de si el


método de instalación ha sido la hinca o la perforación. Ni la metodologı́a p-
y ni el modelo de EF tiene en cuenta la historia tensional en el suelo debido al
proceso constructivo, ni la posible influencia de las acciones verticales en la es-
tructura. Del mismo modo, los efectos del taponamiento del monopilote no son
considerados. A su vez, la fricción en la pared exterior e interior del monopilote
no ha sido considerada a la hora de determinar los modos de fallo del suelo.

• Se ha comprobado que el parámetro kh es de los más influyentes a la hora de


calcular la pendiente inicial de las curvas p-y, y D a la hora de encontrar la carga
última del monopilote. Mientras que el diámetro es un parámetro que viene del
diseño, kh es asumido como una propiedad del suelo dependiente de la densidad
relativa.

M.Sc. Civil Engineering 8.3. Método p-y


145 Chapter 8. Conclusiones y Recomendaciones

• La normativa recomienda una degradación del 10% de las curvas p-y dinámicas
en comparación con las estáticas. Resulta sorprendente que los efectos dinámicos
sean tratados de manera tan trivial, teniendo en cuenta la naturaleza dinámica y
cı́clica de este problema.

• Se puede decir que el método p-y viene respaldado por numerosos años de experi-
encia y diseños, pero que necesita una actualización significativa para adaptarlo a
las nuevas circunstacias e investigaciones.

8.4 Comparación método p-y - elementos finitos


• Las curvas p-y son casi coincidentes en los niveles superiores del suelo (hasta −3 m)
entre el MEF y la API, aunque a partir de ese punto, la API empieza a dar menos
capacidad al suelo, i.e. PuM EF > PuAP I . Esto lleva al sobredimensionamiento de los
monopilotes, ya que usar la API significa usar un método conservativo. También
cabe destacar que el diseño de los monopilotes no viene ifluenciado por aspectos
geotécnicos, sino más bien por el estado lı́mite último de fatiga durante el proceso
de hinca del mismo.

• A partir de una profundidad de aproximadamente −15 m la obtención de las


curas p-y reales del suelo es complicado, ya que bajo las condiciones de solicitación
standard no se llega al punto de que el suelo plastifique y alcance su capacidad
última, como es normal por otro lado.

• La principal diferencia entre las pendientes iniciales entre los distinto métodos
puede deberse al hecho de que el método de las curvas p-y asumen que kh varı́a
linealmente con la profundidad, cuando ha sido comprobado con el MEF que esto
no sucede ası́ en realidad.

• Las principales diferencias entre los resultados del MEF y al API provienen de los
cálculos en los que se basa la normativa. Hay que recordar que la metodologı́a
p-y nace de los ensayos experimentales a escala realizados en las Islas Mustang
con pilotes circulares, metálicos y flexibles en un único tipo de suelo y con un
único diámetro. Por tanto, los resultados vienen totalmente condicionados por esta
situación de partida, y resultarı́a difı́cil extrapolar las conclusiones de esos ensayos
al resto de casos posibles de monopilotes en distintos tipos de suelos. Es por esto
que diversos parámetros correctores fueron incorporados, pero aun ası́, el método
es inapropiado cuando se comparan resultados reales con los ofrecidos por la API.

8.4. Comparación método p-y - elementos finitos M.Sc. Civil Engineering


Chapter 8. Conclusiones y Recomendaciones 146

Aparte de esto, el compartamiento de los monopilotes con diametros reducidos es


distinto al de los pilotes rı́gidos, ya que estos últimos están sometidos a ua rotación
y un desplazamiento horizontal generales superiores a los primeros, lo cual lleva
a la generación de mayores esfuerzos cortantes en base y fuste del monopilote.
De los cálculos en elementos finitos, se estima que estos efectos pueden constituir
entre el 30 − 40% de fuerzas adicionales en el modelo numérico con respecto a la
formulación de la API.

• El plano de máximas fuerzas normales corresponde con el plano medio del suelo-
monopilote, pero no coindice con el plano de máximas fuerzas de corte, en términos
de resultante de fuerzas.

• El nivel que alcanza el suelo dentro del monopilote influye en su comportamiento


a flexión ya que aumenta la fricción interior. Sin embargo es un fenómeno muy
difı́cil de caracterizar, especialmente tras un agresivo proceso constructivo en el
suelo.

8.5 Comparación estático-dinámico (cı́clico)


• Pudo ser comprobada la influencia de la direccionalidad de la carga en el modelo.
Aunque esto pueda parecer trivial, en realidad lo que se crea es una dirección
preferencial de desplazamientos, algo que puede suceder cuando la dirección del
viento o de las olas es muy marcada. En última instancia, este hecho puede crear
problemas a la turbina, ya que el principal rango operativo viene definido por la
rotación en cabeza de la estructura. Por tanto, la historia tensional en el suelo
tiene una influencia muy importante en el comportamiento a largo plazo de la
estructura.

• Con el MEF pudo comprobarse como las condiciones de carga lateral pueden con-
dicionar también el comportamiento vertical del monopilote, generando asientos.

• Debido a las deformaciones plásticas en el fondo marino, la separación entre el


monopilote y el suelo aumenta con el tiempo. Aunque esto no pudo ser estudiado,
se cree que el agua penetra en la apertura y es empujada hacia arriba y hacia el
suelo, creando problemas de erosión que pueden desencadenar una reducción en la
resistencia local en esa zona debido a que el nivel del suelo ha descendido.

• La densificación del suelo en los alrededores del monopilote no ha podido ser estu-
diada o comprobada.

M.Sc. Civil Engineering 8.5. Comparación estático-dinámico (cı́clico)


147 Chapter 8. Conclusiones y Recomendaciones

• La influencia en el fondo marino de los efectos cı́clicos se expanden hasta 10 m


alrededor de los extremos del monopilote, manteniéndose constante tras los primeros
ciclos de carga.

• Según los resultados obtenidos, la influencia de la carga cı́clica sobre las curvas
p-y tiene dos lecturas principales. La primera es una reducción de la rigidez inicial
del suelo o pendiente inicial en el rango elástico (aunque esto se puede deber al
empuje inicial de la carga y a los efectos direccionales de la misma). La segunda
tiene que ver con la degradación producida por los ciclos de carga. A más ciclos,
menor resistencia del suelo. Por lo tanto, se puede concluir que bajo condiciones
cı́clicas se produce una degradación de la resistencia del suelo.

• Según las simulaciones, los esfuerzos cortantes en base del monopilote aumentan
con el tiempo, algo que puede llevar a pensar que la capacidad de las curvas p-y
puede incrementarse con el tiempo. Sin embargo, aunque los esfuerzos cortantes
aumenten, la capacidad de las curvas p-y decrece debido a la acumulación de
deformaciones permanentes en el suelo. La única mejora de las curvas p-y puede
deberse a una densificación del suelo, hecho que no ha sido contrastado o estudiado
en este documento.

• Una simulación de 40 segundos (40 ciclos de carga) no es suficiente para encon-


trar tendencias generales en el comportamiento de la cimentación, pero ayuda a
comprender cualitativamente lo que sucederı́a.

• Durante los 109 ciclos de carga en la vida útil del monopilote (o 40 en el caso
de este documento) la deformación plástica del suelo es acumulativa e incrementa
con el tiempo, aunque tiende a estabilizarse. Igualmente, si las condiciones son
no drenadas y la presión de poro debido a las condiciones cı́clicas aumenta, puede
afectar al compartamiento a largo plazo del suelo y de la estructura.

8.6 Recomendaciones para proyectos

• Además de usar las distintas normativas y software para obtener los esfuerzos
en los monopilotes, es recomendable crear un modelo en EF para comprobar las
diferencias y caracterizar el comportamiento del suelo.

• Caracterizar correctamente las cargas dinámicas, y no solo las estáticas.

8.6. Recomendaciones para proyectos M.Sc. Civil Engineering


Chapter 8. Conclusiones y Recomendaciones 148

• Prestar especial atención al modelo constitutivo que represente el comportamiento


tenso-deformacional del suelo.

• Siempre comprobar la condición de equilibrio en el modelo de elementos finitos.

• Llevar a cabo un análisis a largo plazo, estudiando la vida útil de la estructura y


no únicamente centrarse en el caso estático.

8.7 Recomendaciones para la mejora de las normativas


• Llevar a cabo una actualización general de la formulación p-y, comenzando por la
obtención de las ecuaciones de gobierno del problema para incluir los efectos de la
rigidez y geometrı́a del pilote.

• Investigar los criterios de fallo en análisis lı́mite para monopilotes de gran diámetro.

• Revisar la influencia del parámetro kh en la rigidez del suelo y su variación lineal


con la profundidad.

• Incluir una revisión para monopilotes rı́gidos y factores de corrección para incluir
los efectos del cortante en base y piel del monopilote, efectos del taponamiento y
la historia tensional según proceso contructivo.

• Incluir lo efectos ciı́clicos y dinámicos de una manera más técnica o experimental,


y no solo incluyendo un coeficiente de degradación.

• Llevar a cabo análisis a largo plazo, teninedo en cuenta toda la vida útil del aero-
generador.

• Incluir los posibles efectos de la licuefación e influencia de la presión de poro en


condiciones de carga cı́clica.

• Incluir recomendaciones para suelos no homogéneos y con varios estratos, incluy-


endo los casos en los que haya que instalar la cimentación en roca.

M.Sc. Civil Engineering 8.7. Recomendaciones para la mejora de las normativas


Chapter 9

Further Research

For the sake of practicability of the investigations and in lieu of the elevated computa-
tional cost of the case studies presented here, it was necessary to renounce a rigorous
modelling of the physical reality of the offshore foundation in several aspects, so many
simplifications had to be introduced inevitably. Despite the fact that the relevance of the
simplification is believed to be high for the quantities given here the orders of magnitude
should be appropriate.

• Constitutive models. Despite the fact that in-depth research was carried out in
this particular topic, it was found that no constitutive model met the requirements.
Furthermore, even if such law existed, provided that ABAQUS has it implemented
on its library, it would have been useless to try and create an UMAT for it. Not
to mention the fact that the majority of the numerical problems came from the
use a plastic model when computing dynamic simulations. This is a topic in which
further research and numerical implementation has to be done to properly model
and interpret the behaviour of the soil. There is a data base called soilmodels.info in
which different authors has published for free UMATs of their constitutive models
for the public to use. This approach was also tried as an option, but the models
did simply not solve due to problems with the material laws, so no more time was
allocated for this.

• Pore water pressure. Here the importance of considering the coupling effects
between soil stress and pore water pressure in design, even for those cases where
a soil liquefaction could in principle be disregarded on the grounds of the high
relative density of the soil. For long term simulations, it has been shown that

149
Chapter 9. Further Research 150

the increase in the pore water pressure leads to a decrement in the soil resisting
properties [Cuéllar (2011)], so this has to be addressed in more detail. Another
problem of using coupling formulation of stresses and pore water pressure is the
need to use more complex elements in the FEM model, and the requirement of
using pore-elasticity in the elastic range. This will add computational cost, so
better machines would be needed.

• Densification of the soil. The densification of sandy soils around the monopile
is another effect that has been discovered by some authors during experimentation
[Cuéllar (2011)]. Apparently, because of the cyclic movement of the pile and the
nature of the displacement history, the pores of the surroundings of the monopile
are likely to be filled with small particles already present in the soil or with crushed
matter from existing grains likely formed during the installation process. This
will create a filter effect in the soil, finding a stiffer soil around the pile. It is
very difficult to simulate numerically with FEM and apart from the experimental
evidence of this phenomena, no other data exist.

• Seabed crust. This is a rare phenomena found experimentally and in field invest-
igations. Apparently due to the movement of the monopile, a soil crust is likely to
form around the monopile. It is true that this has not been reported as a source
of structural or geotechnical problems, but is a rare effect that should be studied,
and it is believed to be related to the densification of the surrounding soil.

• Scour. Compared to the seabed crust, the scour may create structural inefficiency,
as it produces a reduction in the mass of the soil acting against the loads, reducing
the effective resistance in so far the scour is more accused. Scour is usually related
with sea currents and bottom drag, and can be stopped with scour protection, like
boulders around the base of the monopile in the mudline. The scour is accounted for
in the codes, but further modelling should be recommended, specially for particular
case studies or particular locations in which strong currents are reportedly present.
The models usually does not take into account scour, and it may cause problems
if it is very pronounced (order of meters).

• Long term knowledge. Numerical long term simulations are almost non-existent.
During the life of the monopiles more than 109 load cycles are expected, and the
numerical results barely made it to the 103 range, which means that we are still
far from assessing properly the long term behaviour. It is true that experimental

M.Sc. Civil Engineering


151 Chapter 9. Further Research

tests have been carried out with this respect, but usually those are scale test and
helped to achieve qualitative findings, and not quantitative.

• Numerical problems. The computational time and cost have been the most
important constraints of the analyses, as results had to be produced in time. Re-
garding the constitutive model, the fact that the state parameters have not been
included yet means that the single calibration set for the model parameters is
only “right” for a certain depth and not for the whole soil. Without the state
parameter, a strict modelling would require that the soil elements at each depth
incorporate the material parameters calibrated for their corresponding confining
pressure, which is obviously not practicable. A perhaps more sensitive issue is the
size and degree of refinement of the finite element mesh. In this respect, the adop-
tion of a bigger model incorporating smaller elements perhaps might have rendered
a smoother solution and more accurate results.

• Installation effects and stress field alteration. Another issue is that the
initial stress state assumed for the soil is also a rough simplification of the real
stress conditions of the seabed, thus stress heterogeneities and pile installation
effects are not considered. But also the preloading caused by the water distributed
load on the soil surface has a certain impact on the results. Richard Jardine has
sufficiently demonstrated the huge impact of the installation in the behaviour of the
monopiles, an it is worth considering in more accurate analyses. The introduction
of a pre-set stress field according to the installation effects was tried, but had to be
abandoned due to the lack of promising results and the complexity of the modelling
of the installation process with finite elements. Perhaps this enters in the realms
of the numerical modelling with particles and DEM, rather than with FEM.

• Centre of rotation. There is a big debate about this. It is clear that the more
slender the monopile gets, the more likely is to behave controlled by its bending
stiffness and the higher will be the centre of rotation of the monopile. The wider
and shorter the monopile gets, the more it behaves as a rigid body, and the lower
the centre of rotation gets. This is true as extremes, but in the middle of this
considerations lie the majority of the designed monopiles. So the behaviour is
something in between a flexible and a rigid body. This has an important influence
on the p-y method, in which the failure criterion associated with the ultimate
capacity relies on the critical depth which determines the flexural behaviour of the
monopile. This should be further analysed.

M.Sc. Civil Engineering


Chapter 9. Further Research 152

• Loading conditions. The loading conditions have also been severely simplified,
particularly for the study case and parametric investigations, where the few cycles
of a purely symmetric two-way sinusoidal load with constant period and acting in
a single direction are only a crude approximation of the foundation’s resultant of
the actual loading produced by wind and waves on the turbine. Furthermore, the
changing water-level conditions at the seabed surface have also been neglected in
this study, which is a factor that can also lead to liquefaction phenomena in the
upper metres of soil, as shown by several authors in the past.

• Validation and calibration. This thesis is also constrained by the fact that the
results of the simulation have not been validated with a real structure (neither
field nor centrifuge tests, nor on-site measurements), as well as by the scarcity
of results, which are focused on a few pile configurations embedded in a certain
sand with a given relative density. The absence of data with this respect is an
important limitations of this study. This is the reason why so much effort was
put in understanding the model from the basis and build up from a single shallow
foundation problem.

• 1D-2D-3D. The p-y curve methodology is a simple a approximation to a 1D


model of a beam in non-Winkler foundations. Most of the research is focused on
developing full 3D models. Perhaps it would be interesting to work in 2D models
as well if the monopiles are sufficiently wider to use this approximation. Another
approach that would be welcomed might be a corrected 1D model based on 3D
simulation. This will constitute an in-depth review of the p-y method with modern
simulation techniques but without missing the simplicity of the p-y method.

• Non-linearities. The stiffness of the soil exhibits non-linear behaviours with re-
spect to depth, strain magnitude, stress history, and the frequency of loading. This
non-linear stress-strain relationship makes it very difficult to establish a rigorous
solution for soil stiffness using continuum theory and led to the original devel-
opment of the Winkler hypothesis and the p-y curve method. However, because
the p-y method was developed using a pseudo-static approach, only the stiffness
non-linearities arising from depth and strain magnitude are properly dealt with,
while the effects from stress history and loading frequency are not considered.
These ignored effects can have significant consequences on the accuracy of the
model, as large changes in soil stiffness can occur due to hardening, softening and
pore-pressure accumulations, all of which are stress history and loading frequency

M.Sc. Civil Engineering


153 Chapter 9. Further Research

dependent processes.

• Damping and dynamic effects. Another important behaviour of the soil-pile


system is soil damping, which can be divided into two main categories. The first
of these is material (hysteretic) damping, which occurs due to the aforementioned
softening and hardening of the soil during cyclic loading. The hysteretic nature
of these softening and hardening processes is apparent, as the main mechanism
for both is “damage” to the structure of the soil by a rearranging of the soil
particle layout and changes in the soil particle contact interaction. The second form
of soil damping is radiation (geometric) damping, which takes the form of wave
propagation through the soil. This process is highly dependent on the frequency of
loading and to a lesser extent the stress history. The pseudo-static approach used
in the development of p-y curves means that the damping from the soil is largely
ignored by p-y curves. While soil damping is relatively minimal, it can nevertheless
provide an important contribution in a wind turbine structure, particularly when
the turbine is parked and provides very little aerodynamic damping.

M.Sc. Civil Engineering


Bibliography

Achmus, M., Albiker, J., Peralta, P., and tom Wörden, F. (2008). Scale effects in the
design of large diameter monopiles.

Achmus, M., Kuo, Y.-S., and Abdel-Rahman, K. (2009). Behavior of monopile founda-
tions under cyclic lateral load. Computers and Geotechnics, 36(1):725–735.

API (1993). Recommended practice for planning, designing, and constructing fixed off-
shore platforms - Working stress design, API RP2A-WSD. American Petroleum In-
stitute, Washington D.C., 21 edition.

Arias-Trujillo, J., Blázquez, R., and López-Querol, S. (2012). A methodology based on


a transfer function criterion to evaluate time integration algorithms. Soil Dynamics
and Earthquake Engineering, 37:1–23.

Ashour, M. and Norris, G. (2000). Modeling lateral soil-pile response based on soil-pile
interaction. Journal of Geotechnical and Geoenvironmental Engineering, 126(5):pp.
420–428.

Bourgeois, E., Rakotonindriana, M., Le Kouby, A., Mestat, P., and Serratrice, J. (2010).
Three-dimensional numerical modelling of the behaviour of a pile subjected to cyclic
lateral loading. Computers and Geotechnics, 37:999–1007.

BrØdbæk, K. T., MØller, M., SØrensen, S. P., Augustensen, H., and Anders, H. (2009).
Review of p-y relationships in cohesionless soil. Technical Report DCE Technical
Reports; No. 57, Aalborg University.

Byrne, B. and Houlsby, G. (2003). Foundations for offshore wind turbines. Philosophical
Transactions of the Royal Society of London. Series A: Mathematical, Physical and
Engineering Sciences, 361(1813):2909–2930.

155
Bibliography 156

Cuéllar, P. (2011). Pile Foundations for Offshore Wind Turbines: Numerical and Ex-
perimental Investigations on the Behaviour under Short-Term and Long-Term Cyclic
Loading. PhD thesis, Technischen Universität Berlin.

Damgaard, M., Bayat, M., Andersen, L., and Ibsen, L. (2014). Assessment of the
dynamic behaviour of saturated soil subjected to cyclic loading from offshore monopile
wind turbine foundations. Computers and Geotechnics, 61:116–126.

de Blaeij, T. (2013). On the modelling of installation effects on laterally cyclic loaded


monopiles. PhD thesis, Delft University of Technology.

Dijkstra, J. and Broere, W. (2009). Experimental investigation into plugging of open


ended piles. ASME International Conference on Ocean, pages 1–8.

DNV (1992). Foundations - Classiffcation Notes No 30.4. Det Norske Veritas, Det
Norske Veritas Classiffication A/S.

DOWEC (2003). Dowec 6 mw pre-design. Technical Report DOWEC-F1W2-HJK-01-


046/9, DOWEC Project. Kooijman, H.J.T. and Lindenburg,C. and Winkelaar, D. and
van der Hooft, E.L.

EC (2011). Com(2011) 112 final, a roadmap for moving to a competitive low carbon
economy in 2050. Technical report, European Commission. Brussels.

EUP (2009). Directive 2009/28/ec on the promotion of the use of energy from renewable
sources. Technical report, European Parliament and Council. Brussels.

Eurostat (2012). Renewable energy. analysis of the latest data on energy from renewable
sources. Technical report, Sturc, M.

EWEA (2010). Maritime spatial planning: supporting offshore wind and grid develop-
ment. Technical report, European Wind Energy Association.

EWEA (2011). Wind in our sails: The coming of europe’s offshore wind energy industry.
Technical report, Azau, S. and Bianchin, R.: European Wind Energy Association.

EWEA (2015). The european offshore wind industry: Key trends and statistics 2014.
Technical report, Corbetta, G. and Mbistrova, A.: European Wind Energy Associ-
ation.

GL (2005). Rules and Guidelines for the Certification of Offshore Wind Turbines. Ger-
manischer Lloyd, Germany.

M.Sc. Civil Engineering Bibliography


157 Bibliography

Haiderali, A., Lau, B., Haigh, S., and Madabhushi, S. P. G. (2014). Lateral response
of monopiles using centrifuge testing and finite element analysis. Taylor and Francis
Group, page Conference Paper.

Helwany, S. (2007). Applied Soil Mechanics: with ABAQUS Applications. John Wiley
and Sons, Inc.

Hetenyi, M. (1946). Beams on elastic foundations. The University of Michigan press.

Hu, M., O’Sullivan, C., Jardine, R., and Jiang, M. (2010). Stress-induced anisotropy in
sand under cyclic loading. Granular Matter, 12:469–472.

Igoe, D., Gavin, K., and O’Kelly, B. (2011). Shaft capacity of open-ended piles in sand.
J. Geotech. Geoenviron. Eng., 137(10):903–913.

Jardine, R., Chow, F., Overy, R., and Standing, J. (2005). Icp design methods for driven
piles in sands and clays. Thomas Telford, London., 97.

Jardine, R., Zhu, B., Foray, P., and Dalton, C. (2009a). Experimental arrangements for
investigation of soil stresses developed around a displacement pile. Soils and Founda-
tions. JGS., 49(5):661–673.

Jardine, R., Zhu, B., Foray, P., and Yhang, Z. (2009b). Measurements of stresses around
close-ended displacements piles in sand. Geotechnique, 63(1):1–17.

Jardine, R., Zhu, B., Foray, P., and Yhang, Z. (2013). Interpretation of stress measure-
ments made around close-ended displacements piles in sand. Geotechnique, 63(8):613–
627.

Knappett, J. A. and Madabhushi, S. P. G. (2009). Influence of axial load on lateral pile


response in liquefiable soils. part i: physical modelling. Geotechnique, 59(7):571–581.

Labenski, J. and Moormann, C. (2015). A comparison of different installation methods


of monopiles for wind power plants and their influence on the bearing behaviour.
Symposium on Energy Geotechnics.

Lau, B., Lam, S., Haigh, S., and Madabhushi, S. (2014). Centrifuge testing of monopile
in clay under monotonic loads. Taylor and Francis Group, pages 689–695.

Lehane, B., Schneider, J., and Xu, X. (2008). The design of displacement piles in siliceous
sands using the cpt. Australian Geomechanics Journal, 43(2):21–40.

Bibliography M.Sc. Civil Engineering


Bibliography 158

Lesny, K. and Wiemann, J. (2005). Design aspects of monopile foundations in german


offshore wind farms. Frontiers in Offshore Geotechnics.

Liu, R., Zhou, L., Lian, J., and Ding, H. (2016). Behavior of monopile foundations
for offshore wind farms in sand. Journal of Waterway, Port, Coastal, and Ocean
Engineering, American Society of Civil Engineers, 142(1):1–11.

López-Querol, S., Peco, J., and Arias-Trujillo, J. (2014). Numerical modeling on vi-
broflotation soil improvement techniques using a densification constitutive law. Soil
Dynamics and Earthquake Engineering, 65:1–10.

Lüking, J. and Kempfert, H. (2013). Plugging effect of open-ended displacement piles.


Proceedings of the 18th International Conference on Soil Mechanics and Geotechnical
Engineering, Paris, 18:2363–2366.

Malhotra, S. (2011). Selection, design and construction of offshore wind turbine found-
ations. Technical report, Parsons Brinckerhoff, Inc. Intech.

Matlock, H. (1970). Correlations for design of laterally loaded piles in soft clay. Pro-
ceedings of the Sixth Annual Offshore Technology Conference, Houston, Texas(OTC
1204):pp. 577–594.

McClelland, B. and Focht, J. (1958). Soil modulus for laterally loaded piles. ASCE,
123(8):1026–11049.

M.E.L.T (1993). Technical rules of design and calculation of the foundations of the works
of civil engineering. Technical report, Ministere de l’equipement, du logement et des
Transports.

Murchison, J. and O’Neill, M. (1983). An evaluation of p-y relationships in sands (re-


port). Technical report, American Petroleum Institute.

Paikowsky, S. (1990). The mechanism of pile plugging in sand. Offshore Technology


Conference.

Pando, M. A. (2013). Analyses of lateral loaded piles with p-y curves - observations on
the effect of pile flexural stiffness and cyclic loading.

P.H.R.I (1980). Technical standards for port and harbour facilities in japan. Technical
report, Port an Harbour Research Institute.

M.Sc. Civil Engineering Bibliography


159 Bibliography

Poulos, H. G. and Davis, E. H. (1974). Elastic Solutions for Soil and Rock Mechanics.
NY, John Wiley and Sons, Inc.

Reese, L. and Matlock, H. (1956). Non-dimensional solutions for laterally loaded piles
with soil modulus assumed proportional to depth.

Reese, L. P., Cox, W. R., and Koop, F. D. (1972). Analysis of laterally loaded piles
in sand. Proceedings of the Sixth Annual Offshore Technology Conference, 2(OTC
2080):Houston, Texas.

Reese, L. P., Cox, W. R., and Koop, F. D. (1974). Analysis of laterally loaded piles
in sand. Proceedings of the Sixth Annual Offshore Technology Conference, 2(OTC
2080):Houston, Texas.

Reese, L. P. and Van Impe, W. F. (2001). Single piles and pile groups under lateral
loading. Taylor and Francis Group plc.

Rimoy, S., Jardine, R., and Standing, J. (2013). Displacement response to axial cycling
of piles driven in sand. Geotechnical Engineering. ICE, 166:GE2.

Sim, W., Aghakouchak, A., and Jardine, R. (2013). Cyclic triaxial tests to aid offshore
pile analysis and design. Geotechnical Engineering. ICE, 166:GE2.

Simulia (2010). Abaqus Theory Guide and Manual.

Terzhagi, K. (1955). Evaluation of coefficients of subgrade reaction. Harvard soil mech-


anics series, Harvard University.

Trochanis, A., Bielak, J., and Christiano, P. (1991). Three-dimensional nonlinear study
of piles. Journal of Geotechnical Engineering, ASCE, 117(3):429–447.

Van der Temple, J. (2006). Design of support structures for offshore wind turbines. PhD
thesis, Delft University of Technology.

White, D. and Lehane, B. (2004). Friction fatigue on displacement piles in sand. Geo-
technique, 54:645–658.

White, D. J., Schneider, J. A., and Lehane, B. M. (2005). The influence of effective area
ratio on shaft friction of displacement piles in sand. proceedings of the international
symposium on frontiers in offshore geotechnics. Rotterdam.

Winkler, E. (1897). Die lehre von elasticzitat und festigkeit (on elasticity and plasticity).
(Prague):182 p.

Bibliography M.Sc. Civil Engineering


Bibliography 160

Xu, X., Schneider, J., and Lehane, B. (2005). Evaluation of end-bearing capacity of
open-ended piles driven in sand from cpt data. Proceedings of the 1st International
Symposium on Frontiers in Offshore Geotechnics, Australia, pages 19–21.

Yu, F. and Yang, J. (2011). Base capacity of open-ended steel pipe piles in sand. J.
Geotech. Geoenviron. Eng., 138(9):1116–1128.

M.Sc. Civil Engineering Bibliography


Chapter 10

Appendices

10.1 Code for Infinite Elements

*Heading
** Job name: Infelem Model name: InfElem
** Generated by: Abaqus/CAE 6.14-1
*Preprint, echo=NO, model=NO, history=NO, contact=NO
**
** PARTS
**
*Part, name=Part-1
*Node
1, 1., -2.4000001, 1.
(...)
3000, 1280, 192, 11, 209, 3872, 1442, 228, 1523
*Element, type=CIN3D8
3001, 146, 957, 17, 1, 237, 1524, 272, 13
(...)
3100, 73, 5, 164, 1037, 254, 15, 255, 1604
*Nset, nset=Set-1, generate
1, 3872, 1
*Elset, elset=Set-1, generate
1, 3100, 1
** Section: Section-1
*Solid Section, elset=Set-1, material=Material-1

161
Chapter 10. Appendices 162

1.,
*End Part
**
**
** ASSEMBLY
**
*Assembly, name=Assembly
**
*Instance, name=Part-1-1, part=Part-1
*End Instance
**
*Nset, nset=Set-1, instance=Part-1-1
(...)
*Elset, elset=Set-1, instance=Part-1-1, generate
3001, 3100, 1
*Nset, nset=Set-2, instance=Part-1-1
4,
*Elset, elset=_Surf-1_S5, internal, instance=Part-1-1
(...)
*Surface, type=ELEMENT, name=Surf-1
_Surf-1_S5, S5
*End Assembly
**
** MATERIALS
**
*Material, name=Material-1
*Density
2100.,
*Elastic
9.8e+07,0.
**MC ----------------------------------------------------------------
**
** STEP: Step-1
**
*Step, name=Step-1, nlgeom=NO
*Dynamic,direct

M.Sc. Civil Engineering 10.1. Code for Infinite Elements


163 Chapter 10. Appendices

0.01,1.,
**
** BOUNDARY CONDITIONS
**
** Name: BC-1 Type: Displacement/Rotation
*Boundary
Set-1, 1, 1
Set-1, 2, 2
Set-1, 3, 3
**
** LOADS
**
** Name: Load-1 Type: Pressure
*Dsload
Surf-1, P, 3000.
**
** OUTPUT REQUESTS
**
*Restart, write, frequency=0
**
** FIELD OUTPUT: F-Output-1
**
*Output, field, variable=PRESELECT, frequency=1
**
** HISTORY OUTPUT: H-Output-1
**
*Output, history, frequency=1
*Node Output, nset=Set-2
U1, U2, U3, UR1, UR2, UR3
*End Step*Heading
** Job name: Infelem Model name: InfElem
** Generated by: Abaqus/CAE 6.14-1
*Preprint, echo=NO, model=NO, history=NO, contact=NO
**
** PARTS
**

10.1. Code for Infinite Elements M.Sc. Civil Engineering


Chapter 10. Appendices 164

*Part, name=Part-1
*Node
1, 1., -2.4000001, 1.
(LINES)
*Element, type=C3D8R
1, 146, 957, 1605, 615, 1, 17, 273, 72
(LINES)
*Nset, nset=Set-1, generate
1, 3872, 1
*Elset, elset=Set-1, generate
1, 3100, 1
** Section: Section-1
*Solid Section, elset=Set-1, material=Material-1
1.,
*End Part
**
**
** ASSEMBLY
**
*Assembly, name=Assembly
**
*Instance, name=Part-1-1, part=Part-1
*End Instance
**
*Nset, nset=Set-1, instance=Part-1-1
(...)
*Elset, elset=Set-1, instance=Part-1-1, generate
3001, 3100, 1
*Nset, nset=Set-2, instance=Part-1-1
4,
*Elset, elset=_Surf-1_S5, internal, instance=Part-1-1
(...)
*Surface, type=ELEMENT, name=Surf-1
_Surf-1_S5, S5
*End Assembly
**

M.Sc. Civil Engineering 10.1. Code for Infinite Elements


165 Chapter 10. Appendices

** MATERIALS
**
*Material, name=Material-1
*Density
2100.,
*Elastic
9.8e+07,0.
** ----------------------------------------------------------------
**
** STEP: Step-1
**
*Step, name=Step-1, nlgeom=NO
*Dynamic,direct
0.01,1.,
**
** BOUNDARY CONDITIONS
**
** Name: BC-1 Type: Displacement/Rotation
*Boundary
Set-1, 1, 1
Set-1, 2, 2
Set-1, 3, 3
**
** LOADS
**
** Name: Load-1 Type: Pressure
*Dsload
Surf-1, P, 3000.
**
** OUTPUT REQUESTS
**
*Restart, write, frequency=0
**
** FIELD OUTPUT: F-Output-1
**
*Output, field, variable=PRESELECT, frequency=1

10.1. Code for Infinite Elements M.Sc. Civil Engineering


Chapter 10. Appendices 166

**
** HISTORY OUTPUT: H-Output-1
**
*Output, history, frequency=1
*Node Output, nset=Set-2
U1, U2, U3, UR1, UR2, UR3
*End Step

M.Sc. Civil Engineering 10.1. Code for Infinite Elements

Vous aimerez peut-être aussi