Vous êtes sur la page 1sur 256

Pile Buck, Inc.

NOTICE:
The compilation of handbooks on this CD-ROM and the material contained within is
provided without warranty of any kind. Pile Buck, Inc. hereby disclaims any and all
express or implied warranties of merchantibility, fitness for any general or particular
purpose or freedom from infringement of any patent, trademark, or copyright in
regard to information or products contained or referred to herein. Nothing herein
contained shall be construed as granting a license, express or implied, under any
patents. Anyone making use of this material does so at his or her own risk and
assumes any and all liability resulting from such use. The entire risk as to quality
or usability of the material contained within is with the user. In no event will Pile
Buck, Inc. be held liable for any damages including lost profits, lost savings or other
incidental or consequential damages arising from the use or inability to use the
information contained within. Pile Buck, Inc. does not insure anyone utilizing this
material against liability arising from the use of this information and hereby will not
be held liable for "consequential damages," resulting from such use.

To obtain passwords to unlock


secured manuals, contact:

Pile Buck, Inc.


P.O.Box 1736
Palm City, FL U.S.A. 34991

Phone: (561) 223-1919 Fax: (561) 223-1995


e-mail: pilebuck@pilebuck.com www: http://www.pilebuck.com


Copyright 1999 Pile Buck, Inc. All Rights Reserved.

NEWSPAPER
Steel Sheet Piling Timber Piles Foundations Geotechnical Engineering Marine Docks Marinas Bridge Building Dredging
Pile Hammers Pipelines Jetties Breakwaters Barges Cranes Cofferdams Harbors Piers Wharves Soils Moorings
BOOKS
Bulkheads Earth Retaining Structures Ports Seawalls Steel Sheet Piling Timber Piles Foundations Geotechnical Engineering
Marine Docks Marinas Bridge Building Dredging CD Pile ROMS
Hammers Pipelines Jetties Breakwaters Barges Cranes
SOFTWARE
Cofferdams Harbors Piers Wharves Soils Moorings Bulkheads Earth Retaining Structures Ports Seawalls Steel Sheet
Handbook for
Marine Geotechnical Engineering

Technical Editor
Karl Rocker, Jr.

March 1985

DEEP OCEAN TECHNOLOGY

NAVAL CIVIL ENGINEERING LABORATORY


PORT HUENEME, CA 93043

Approved for public release; distribution unlimited.


ACKNOWLEDGMENTS

Much of the background material for the Handbook for Marine Geotechnical Engineering was
developed by the Naval Civil Engineering Laboratory (NCEL). These materials were updated and
supplemented with experience from the private sector by a number of contractors. Initial editing
and consolidation of the chapters was carried out by Brian Watt Associates, Inc. Technical review
of the Handbook was also made by Dr. Robert H. Mayer, Jr. (U.S. Naval Academy) and Mr. Homa
J. Lee. The primary contributor for each of the chapters is as follows:

Chapter Primary Contributor


1 Brian Watt Associates, Inc., Houston, TX
2 The Earth Technology Corporation, Long Beach, CA
3 Prof. I. Noorany, San Diego State Univ., San Diego, CA
4 The Earth Technology Corporation, Long Beach, CA
5 Woodward-Clyde Consultants, Santa Ana, CA
6 Brian Watt Associates, Inc., Houston, TX
7 Brian Watt Associates, Inc., Houston, TX
8 Brian Watt Associates, Inc., Houston, TX
9 Mr. H.J. Lee, U.S. Geological Survey, Menlo Park, CA
10 Mr. M.J. Atturio, Naval Civil Engineering Laboratory,
Port Hueneme, CA
11 Dr. D.A. Sangrey, Carnegie-Mellon Univ., Pitts, PA

iii
CONTENTS
Page
Chapter 1 - INTRODUCTION 1-1
1.1 OBJECTIVE 1-1
1.2 HANDBOOK ORGANIZATION 1-1
1.3 SELECTION OF FOUNDATION/ANCHOR TYPE 1-3
1.4 REFERENCES 1-6
Chapter 2 - SITE SURVEY AND IN-SITU TESTING 2-1
2.1 INTRODUCTION 2-1
2.1.1 Purpose 2-1
2.1.2 Factors Influencing the Site Survey 2-1
2.2 PRELIMINARY STUDY 2-3
2.2.1 Information Sources 2-3
2.2.2 Typical Ocean Soils 2-4
2.3 REGIONAL SURVEY 2-12
2.3.1 General 2-12
2.3.2 Seismic Profiling 2-12
2.3.3 Limited Sampling 2-14
2.3.4 Side-Scan Sonar 2-14
2.3.5 Visual Observation 2-14
2.3.6 Survey Line Spacing 2-14
2.4 SITE-SPECIFIC SURVEY 2-15
2.4.1 General 2-15
2.4.2 Shallow Sampling 2-16
2.4.3 Deep Sampling 2-19
2.4.4 Location, Number, and Depth of Sampling 2-20
2.4.5 Sample Handling 2-21
2.5 IN-SITU TESTING 2-21
2.5 1 General 2-21
2.5.2 Vane Shear Tests 2-22
2.5.3 Cone Penetration Tests (CPT) 2-23
2.5.4 Pressuremeter Tests 2-25
2.5.5 Dynamic Penetrometer 2-25
2.5.6 Borehole Logging Techniques 2-25
2.6 SEISMICITY SURVEY 2-25
2.7 REFERENCES 2-26
2.8 SYMBOLS 2-27
Chapter 3 - LABORATORY DETERMINATION OF SOIL PROPERTIES 3-1
3.1 INTRODUCTION 3-1
3.1.1 Scope 3-1
3.1.2 Special Considerations 3-1
3.2 SOIL CLASSIFICATION 3-1
3.2.1 Classification by Origin 3-1
3.2.2 Clarification by Grain Size 3-1
3.2.3 Classification by Grain Size and Behavior 3-2
3.3 INDEX PROPERTY TESTS 3-4
3.3.1 General 3-4
3.3.2 Sample Preparation 3-4
3.3.3 Water Content 3-4
3 3.4 Unit Weight 3-7

v
Page
3.3.5 Specific Gravity 3-7
3.3.6 Liquid Limit, Plastic Limit, and Plasticity Index 3-7
3.3.7 Grain Size Analysis 3-9
3.3.8 Carbonate and Organic Carbon Content 3-9
3.4 ENGINEERING PROPERTY TESTS 3-9
3.4.1 General 3-9
3.4.2 Vane Shear Test 3-10
3.4.3 Unconfined Compression Test 3-12
3.4.4 Unconsolidated, Undrained Triaxial Compression Test 3-13
3.4.5 Consolidated-Undrained and Consolidated-Drained Triaxial Compression Tests 3-13
3.4.6 Consolidated-Drained Direct Shear Test 3-13
3.4.7 Considerations for Triaxial Testing of Marine Soils 3-13
3.4.8 One-Dimensional Consolidation Test 3-13
3.5 PROPERTY CORRELATIONS 3-14
3.5.1 General 3-14
3.5.2 Nearshore Sediments 3-14
3.5.3 Deep Sea Sediments 3-16
3.6 REFERENCES 3-17
3.7 SYMBOLS 3-18
Chapter 4 - SHALLOW FOUNDATIONS AND DEADWEIGHT ANCHORS 4-1
4.1 INTRODUCTION 4-1
4.1.1 General 4-1
4.1.2 Definitions/Descriptions 4-1
4.2 DESIGN CONSIDERATIONS 4-3
4.2.1 General 4-3
4.2.2 Site 4-4
4.2.3 Structure 4-4
4.2.4 Loading 4-4
4.2.5 Geotechnical 4-4
4.2.6 Factor of Safety 4-5
4.3 DESIGN METHODOLOGY AND PROCEDURE 4-5
4.3.1 General 4-5
4.3.2 Bearing Capacity 4-7
4.3.3 Lateral Load Capacity 4-11
4.3.4 Overturning Resistance 4-13
4.3.5 Shear Key Design 4-14
4.3.6 Foundation Settlement 4-15
4.3.7 Installation and Removal 4-17
4.4 EXAMPLE PROBLEMS 4-17
4.4.1 Problem 1 - Simple Foundation on Cohesive Soil 4-17
4.4.2 Problem 2 - Simple Foundation on Cohesionless Soil 4-24
4.5 REFERENCES 4-28
4.6 SYMBOLS 4-28
Chapter 5 - PILE FOUNDATIONS AND ANCHORS 5-1
5.1 INTRODUCTION 5-1
5.2 PILE DESCRIPTIONS 5-1
5.2.1 Pile Types 5-1
5.2.2 Mooring Line Connections 5-1
5.2.3 Modifications for Increasing Lateral Load Capacity 5-1
5.3 DESIGN PROCEDURES FOR SIMPLE PILES IN SOIL SEAFLOORS 5-1
5.3.1 General 5-1
5.3.2 Soil Properties 5-5
vi
Page
5.3.3 Pile Design Loads 5-6
5.3.4 Lateral Load Analysis 5-7
5.3.5 Axial Load Analysis 5-9
5.3.6 Steel Stress Analysis 5-11
5.3.7 Special Cases 5-12
5.4 DESIGN OF PILE ANCHORS IN ROCK SEAFLOORS 5-13
5.4.1 Lateral Capacity 5-14
5.4.2 Uplift Capacity 5-15
5.5 PILE INSTALLATION 5-15
5.5.1 Driven Piles 5-15
5.5.2 Drilled and Grouted Piles 5-17
5.5.3 Jack-in Piles 5-18
5.5.4 Jetted Piles 5-18
5.6 EXAMPLE PROBLEMS 5-18
5.6.1 Problem 1--Pile Design in a Cohesive Soil 5-18
5.6.2 Problem 2--Pile Design in a Cohesionless Soil 5-25
5.7 REFERENCES 5-29
5.8 SYMBOLS 5-30
Chapter 6 - DIRECT-EMBEDMENT ANCHORS 6-1
6.1 INTRODUCTION 6-1
6.1.1 Purpose 6-1
6.1.2 Function 6-1
6.1.3 Features 6-1
6.2 DIRECT-EMBEDMENT ANCHOR TYPES AND SIZES 6-1
6.2.1 Propellant-Driven Anchor 6-1
6.2.2 Vibratory-Driven Anchors 6-6
6.2.3 Impact-Driven Anchors 6-6
6.2.4 Jetted-In Anchors 6-7
6.2.5 Auger Anchors 6-7
6.3 SITE DATA NEEDED 6-8
6.3.1 General 6-8
6.3.2 Preliminary Penetration Estimate 6-8
6.3.3 Topography, Strata Thicknesses, Type 6-8
6.3.4 Engineering Properties 6-8
6.3.5 Complicating or Hazardous Conditions 6-10
6.3.6 Specialized Survey Tools 6-11
6.4 FLUKE PENETRATION AND KEYING 6-11
6.4.1 Penetration Prediction 6-11
6.4.2 Keying Prediction 6-11
6.5 STATIC HOLDING CAPACITY 6-12
6.5.1 Loading Conditions 6-12
6.5.2 Deep and Shallow Anchor Failure 6-12
6.5.3 Short-Term Capacity in Cohesive Soils 6-13
6.5.4 Long-Term Capacity in Cohesive Soils 6-14
6.5.5 Short- and Long-Term Capacity in Cohesionless Soils 6-15
6.5.6 Disturbance Corrections 6-15
6.5.7 Factors of Safety 6-16
6.6 DYNAMIC HOLDING CAPACITY 6-16
6.6.1 Loading Conditions 6-16
6.6.2 Cyclic Loading 6-16

vii
Page
6.6.3 Earthquake Loading 6-19
6.6.4 Impulse Loading 6-19
6.7 OTHER INFLUENCES ON HOLDING CAPACITY 6-23
6.7.1 Holding Capacity on Slopes 6-23
6.7.2 Creep Under Static Loading 6-24
6.8 HOLDING CAPACITY IN CORAL AND ROCK 6-24
6.8.1 Coral 6-24
6.8.2 Rock 6-24
6.9 EXAMPLE PROBLEMS 6-24
6.9.1 Problem 1 - An Embedment Anchor Used in Cohesive Soil 6-24
6.9.2 Problem 2 - An Embedment Anchor Used in Cohesionless Soil 6-28
6.10 REFERENCES 6-31
6.11 SYMBOLS 6-32
Chapter 7 - DRAG-EMBEDMENT ANCHORS 7-1
7.1 INTRODUCTION 7-1
7.1.1 Purpose and Scope 7-1
7.1.2 Drag Anchor Description 7-1
7.1.3 Types of Drag Anchors 7-2
7.1.4 Application of Drag Anchors 7-3
7.2 FUNCTIONING OF A DRAG ANCHOR 7-3
7.2.1 General 7-3
7.2.2 Tripping 7-4
7.2.3 Embedment 7-5
7.2.4 Stability 7-6
7.2.5 Soaking 7-8
7.3 SITE INVESTIGATION 7-8
7.3.1 Site Data Needed 7-8
7.3.2 Topography and Layer Thickness 7-8
7.3.3 Sediment Type and Strength 7-9
7.3.4 Site Investigation Summary 7-10
7.4 SELECTING A DRAG ANCHOR 7-11
7.4.1 General 7-11
7.4.2 Tripping and Penetration Performance 7-12
7.4.3 Stability Performance 7-12
7.4.4 Holding Capacity Performance 7-12
7.4.5 Selection of Anchor Type 7-12
7.5 SIZING A DRAG ANCHOR 7-12
7.5.1 Efficiency Ratio Method 7-12
7.5.2 Power Law Method 7-13
7.5.3 Analysis Based on Geotechnical Considerations 7-17
7.5.4 Factor of Safety 7-18
7.6 TROUBLESHOOTING 7-19
7.6.1 Soft Sediments 7-19
7.6.2 Hard Sediments 7-20
7.7 PIGGYBACKING 7-20
7.7.1 Field Practice 7-20
7.7.2 Results and Field Problems 7-22
7.7.3 Recommended Practice 7-23

viii
Page
7.8 REFERENCES 7-23
7.9 SYMBOLS 7-24
Chapter 8 - PENETRATION OF OBJECTS INTO THE SEAFLOOR 8-1
8.1 INTRODUCTION 8-1
8.1.1 Purpose 8-1
8.1.2 Scope 8-1
8.2 STATIC PENETRATION 8-1
8.2.1 Application 8-1
8.2.2 Approach 8-1
8.2.3 Method for Predicting Shallow Static Penetration 8-2
8.2.4 Methods for Predicting Deep Static Penetration 8-5
8.3 DYNAMIC PENETRATION 8-7
8.3.1 Application 8-7
8.3.2 Approach 8-7
8.3.3 Method for Predicting Dynamic Penetration 8-7
8.4 EXAMPLE PROBLEMS 8-11
8.4.1 Problem 1--Slow Penetration of a Long Cylinder 8-11
8.4.2 Problem 2--Rapid Penetration of a Long Cylinder 8-14
8.5 REFERENCES 8-19
8.6 SYMBOLS 8-20
Chapter 9 - BREAKOUT OF OBJECTS FROM THE SEAFLOOR 9-1
9.1 INTRODUCTION 9-1
9.1.1 Applications 9-1
9.1.2 General Concepts 9-2
9.1.3 Definitions 9-2
9.2 REQUIRED INFORMATION 9-4
9.2.1 Object Embedment Characteristics 9-4
9.2.2 Sediment Characteristics 9-4
9.2.3 Bearing Capacity (Cohesive Sediments) 9-5
9.3 SHORT-TERM (IMMEDIATE) BREAKOUT 9-5
9.4 LONG-TERM BREAKOUT TIME PREDICTION 9-6
9.5 BREAKOUT AIDS 9-7
9.5.1 Jetting and Drainage Tubes 9-8
9.5.2 Eccentric Loading 9-9
9.5.3 Cyclic Loading 9-9
9.5.4 Rocking or Rolling 9-9
9.5.5 Breakaway Parts 9-9
9.5.6 Altering Buoyant Weight 9-9
9.6 OTHER FACTORS 9-10
9.6.1 Irregular Shape or Nonuniform Embedment Depth 9-10
9.6.2 Waiting Time 9-10
9.6.3 Foundation Skirts 9-10
9.7 EXAMPLE PROBLEMS 9-10
9.7.1 Problem 1 - Recovery of a Large Long Cylinder 9-10
9.7.2 Problem 2 - Recovery of a Skirted Foundation 9-13

ix
Page
9.8 REFERENCES 9-16
9.9 SYMBOLS 9-17
Chapter 10 - SCOUR 10-1
10.1 INTRODUCTION 10-1
10.1.1 Objectives 10-1
10.1.2 Theory 10-1
10.1.3 Modeling 10-1
10.2 SCOUR TYPES 10-2
10.2.1 Seasonal and Local Scour 10-2
10.2.2 Deep Water Wave-Induced Scour 10-2
10.3 ESTIMATING SEASONAL SCOUR 10-2
10.4 ESTIMATING LOCAL SCOUR AT SEAFLOOR STRUCTURES 10-3
10.4.1 Effect of Seafloor Factors on Scour 10-3
10.4.2 Structures Piercing the Water Surface 10-5
10.4.3 Structures Resting on the Seafloor 10-9
10.4.4 Time for Scour Development 10-11
10.5 MINIMIZING SCOUR 10-12
10.5.1 Scour- Resistant Structures 10-12
10.5.2 Scour Protection Measures 10-12
10.6 REFERENCES 10-16
10.7 SYMBOLS 10-18
Chapter 11 - SLOPE STABILITY ASSESSMENT 11-1
11.1 INTRODUCTION 11-1
11.2 FORMS OF INSTABILITY 11-2
11.2.1 Translational Slides 11-3
11.2.2 Rotational Slides 11-3
11.2.3 Flow Slides 11-3
11.2.4 Turbidity Currents 11-5
11.3 LOADING 11-5
11.3.1 Loading Mechanisms 11-5
11.3.2 Probabilistic Prediction of Load 11-5
11.4 IMPORTANT SOIL PROPERTIES 11-6
11.4.1 General 11-6
11.4.2 Special Conditions: Underconsolidated Sediments 11-6
11.4.3 Repetitive and Dynamic Loading Response of Sediments 11-7
11.5 LEVEL OF ANALYSIS 11-7
11.6 SITE INVESTIGATION 11-7
11.6.1 General 11-7
11.6.2 Preliminary Studies 11-8
11.6.3 Acoustic Surveys 11-9
11.6.4 Sampling of Sediments 11-11
11.7 EVALUATION PROCEDURES 11-12
11.8 REFERENCES 11-15
11.9 SYMBOLS 11-16

x
LIST OF TABLES
Table Page
1.3-1 Features of Shallow Foundations and Deadweight Anchors 1-3
1.3-2 Features of Pile Foundation and Anchor Systems 1-4
1.3-3 Features of Direct-Embedment Anchors 1-5
1.3-4 Features of Drag-Embedment Anchor Systems 1-5
1.3-5 Performance of Foundation and Anchor Types as a Function of Seafloor and Loading Conditions 1-6
2.1-1 Site Data Requirements for Categories of Geotechnical Engineering Applications 2-2
2.1-2 Soil Engineering Parameters Normally Required for Categories of Geotechnical Engineering Applications 2-2
2.1-3 Historical Environmental Information Needed to Assess Geotechnical Hazards 2-3
2.3-1 Steps in a Typical Regional Survey 2-12
2.4-1 Steps in a Typical Site-Specific Survey 2-15
2.4-2 Shallow Soil Sampler Types and Applications 2-16
2.4-3 Characteristics of Some Free-Fall and Lowered Corers 2-19
2.5-1 In-Situ Tests, Applications, and Some Equipment Characteristics 2-22
3.2-1 Size Range Limits for Two Soil Classification Systems 3-2
3.3-1 Requirements for Index Property Tests 3-5
3.3-2 Some Index and Engineering Properties of Ocean Sediments (Most Data Limited to Upper 2 Meters of Seafloor) 3-6
3.4-1 Requirements for Engineering Property Tests 3-11
4.2-1 Soil Properties Required for Analysis and Recommended Factors of Safety 4-5
4.3-1 Summary of Steps in the Design of Shallow Foundations and Deadweight Anchors 4-7
4.3-2 Coefficient of Friction Between Cohesionless Soils and Some Marine Construction Materials 4-12
5.2-1 Pile Types 5-2
5.2-2 Mooring Line Connections 5-3
5.2-3 Techniques to Improve Pile Lateral Load Capacity 5-3
5.3-1 Properties of Cohesionless Soil Useful in Pile Design 5-5
5.3-2 Properties of Cohesive Soils Useful in Pile Design 5-5
5.3-3 Properties of Calcareous Soil Useful in Pile Design 5-6
5.3-4 Bearing Capacity Factors for Chain Lateral Force in Sand 5-7
5.3-5 Recommended Limiting Values for Unit Skin Friction and End Bearing for Cohesionless Soils 5-9
5.4-1 Rock Properties 5-14
6.2-1 Propellant-Driven Embedment Anchors for Ocean Use 6-3
6.2-2 Parameters for Navy Propellant-Embedded Anchors 6-4
6.3-1 Conditions Complicating or Hazardous to Direct-Embedment Anchor Use 6-10
6.5-1 Values for Strength Reduction Factor for Use in Equation 6-3 6-16
6.6-1 Average Values of Soil Permeability 6-18
6.7-1 Factors Associated With Direct-Embedment Anchors Which Can Influence Submarine Slope Stability 6-24
7.3-1 Estimated Maximum Fluke Tip Penetration of Some Drag Anchor Types in Hard and Soft Soils 7-9
7.4-1 Rating of Drag Anchor Types Based on Tripping and Dig-In, Roll Stability, and Holding Capacity Efficiency 7-11
7.5-1 Parameters HR and b Used in Equation 7-2 7-14
7.5-2 Parameters Nc and f Used for Clays and Cohesive Silts in Equation 7-3 7-18
7.6-1 Troubleshooting Procedures for Correcting Drag Anchor Performance Problems 7-19
8.2-1 Correlation Coefficients Between Dutch Cone Penetration Resistance and Thin-walled Skirt and Dowel
Penetration Resistance 8-6
8.3-1 Values of Constants Used in Equation 8-12 8-9
8.4-1 Summary of Calculations for Problem 1 8-14
8.4-2 Summary of Calculations for Problem 2 8-19
11.2-1 Movement Models for Submarine Slides in Soft and Loose Sediment 11-2
11.7-1 Level I Slope Stability Survey 11-13
11.7-2 Level II Slope Stability Survey 11-14

xi
LIST OF FIGURES
Figure Page
1.3-1 Simplified anchor types 1-3
2.2-1 Ocean sediment distribution 2-6
2.2-2 Topography of the carbonate compensation depth (CCD) 2-7
2.2-3 Marine geological provinces and probable soil types 2-8
2.2-4 Typical strength profile for hemipelagic terrigenous silty clay 2-9
2.2-5 Typical strength profiles for proximal and distal turbidites 2-9
2.2-6 Typical strength profiles for calcareous ooze 2-10
2.2-7 Typical strength profiles for pelagic clay 2-10
2.2-8 Typical strength profile for siliceous ooze 2-11
2.3-1 Acoustic profiling systems 2-13
2.4-1 Grab samplers and dredges 2-16
2.4-2 Box corer and its operation sequence 2-17
2.4-3 Long piston corer operation sequence with a short corer used as a trigger weight 2-18
2.4-4 Alpine vibracore sampler 2-20
2.5-1 Correction factor for vane determined shear strength 2-23
2.5-2 Electric friction-cone penetrometer tip 2-23
2.5-3 Correlation between cone tip resistance and sand relative density 2-24
2.5-4 Estimation of sand friction angle, , from material relative density 2-24
2.6-1 Seismic risk map of United States coastal waters 2-26
3.2-1 Trilineal soil classification plot - normally used with Wentworth grade limits 3-2
3.2-2 Unified soil classification chart 3-3
3.4-1 Miniature vane blade geometry 3-12
3.5-1 Relationship between su/P yo and PI for normally consolidated late glacial clay 3-14
3.5-2 Relationship between friction angle and PI for normally consolidated fine-grained soils 3-15
3.5-3 Correlation between coefficient of consolidation and liquid limit 3-15
3.5-4 Range of PI values for pelagic clay 3-16
3.5-5 Correlation between water content and Cc /(1 + e 0) for pelagic clay and calcareous ooze 3-17
4.1-1 Features of simple shallow foundation 4-1
4.1-2 Types of shallow foundations 4-2
4.1-3 Types and significant characteristics of deadweight anchors 4-3
4.3-1 Flow chart for the design of shallow foundations and deadweight anchors 4-6
4.3-2 Bearing capacity factors Nq, Nc, and Ny as a function of the soil friction angle 4-8
4.3-3 Area reduction factors for eccentrically loaded foundations 4-10
4.3-4 Possible failure modes when sliding resistance is exceeded 4-11
4.3-5 Forces considered in the overturning analysis 4-14
4.3-6 Soil stress increase beneath a rectangular foundation 4-16
4.4-1 Foundation sketch for example problems 1 and 2 4-18
4.4-2 Forces considered in the overturning analysis for example problem 1 4-23
4.4-3 Forces considered in the overturning analysis for example problem 2 4-27
5.3-1 Flow chart for the pile design procedure 5-4
5.3-2 Design values for nh for cohesionless soils 5-7
5.3-3 Design values for nh for cohesive soils 5-8
5.3-4 Deflection coefficients Ay and By at the ground surface 5-8
5.3-5 Influence values for a pile with applied lateral load or moment 5-11
5.4-1 Failure modes for pile anchors in a rock seafloor 5-13
5.5-1 Pile installation techniques 5 16
5.5-2 Drilled and grouted pile 5-17
5.6-1 Problem sketch for example problems 1 and 2 5-19
5.6-2 Soils data for example problem 1 5-19
6.2-1 Installation sequence for a propellant-embedment anchor 6-2
6.2-2 NCEL 10K propellant-embedment anchor showing sand and clay flukes 6-5
6.2-3 Coral and rock flukes for NCEL propellant-embedment anchor systems 6-5
6.2-4 Impact-driven anchors 6-6
6.2-5 Jetted-in anchors 6-7
6 3-3 Flow chart for predicting the holding capacity of a direct-embedment anchor 6-9
6.5-1 Failure modes for shallow and deep embedded plate anchors 6-12
6.5-2 Short-term holding capacity factors for cohesive soil where full suction develops beneath the plate 6-13
6.5-3 Long-term holding capacity factors and short-term no-suction factors for cohesive soils 6-14
6.5-4 Holding capacity factors for cohesionless soils 6-15

xii
Figure Page
6.6-1 Nomenclature for types of non-steady loading 6-17
6.6-2 Time required for dissipation of stress-induced excess pore pressure 6-18
6.6-3 Maximum cyclic load capacity without soil strength loss 6-19
6.6-4 Maximum (lifetime) cyclic load capacity without development of cyclic creep 6-19
6.6-5 Strain-rate factor, I, for cohesive soil 6-20
6.6-6 Inertial factor, If, for cohesive and cohesionless soils 6-21
6.6-7 Strain-rate factor, I, for cohesionless soil 6-22
6.9-1 Mooring sketch for example problem 1 6-25
6.9-2 Soil strength profile for example problem 1 6-25
6.9-3 Mooring sketch for example problem 2 6-28
7.1-1 Features of a drag anchor 7-2
7.1-2 Example of a movable fluke anchor: STEVIN cast 7-2
7.1-3 Example of a fixed fluke anchor: BRUCE cast 7-2
7.1-4 Examples of bilateral fluke anchors 7-3
7.1-5 Example of a soft soil anchor: STEVMUD 7-3
7.2-1 Development of a tripping problem in soft seafloors with an improperly set anchor 7-4
7.2-2 Proper anchor setting sequence using two floating platforms 7-5
7.2-3 Development of a tripping problem in hard seafloors 7-5
7.2-4 Penetration and orientation behavior of an anchor in hard and soft seafloors 7-7
7.2-5 Forces on unstabilized and stabilized anchors in sand 7-7
7.2-6 Anchor in soft soil, after balling-up and pulling-out 7-8
7.3-1 Site survey plan decision flow chart 7-10
7.5-1 Anchor chain system holding capacity at the mudline in soft soils 7-15
7.5-2 Anchor chain system holding capacity at the mudline in hard soils 7-16
7.6-1 Typical performance of drag anchors when operating properly and improperly 7-20
7.7-1 A pendant line and buoy arrangement for semisubmersibles 7-21
7.7-2 Chain chaser used to assist anchor deployment and recovery 7-21
7.7-3 Tandem/piggyback anchor arrangements 7-22
7.7-4 Parallel anchor arrangement 7-23
8.2-1 Shallow static penetration model 8-2
8.2-2 Relationships for calculating sinkage resistance in cohesionless soils, for = 30 o and 40 o 8-3
8.2-3 Flow chart of the calculation procedure for predicting static penetration 8-4
8.2-4 Location of the critical shear strength zone B for blunt and conical penetrators 8-5
8.3-1 Forces acting on a penetrator before and after contact with the seafloor 8-8
8.3-2 Flow chart of the calculation procedure for predicting dynamic penetration 8-10
8.4-1 Problem sketch and soils data for example problem 1 8-11
8.4-2 Plot of predicted soil resistance to EPS penetration 8-14
8.4-3 Sketch for example problem 2 8-15
9.1-1 Flow chart for procedures to determine immediate breakout force and time required for long-term breakout under a
lower force 9-3
9.3-1 Normalized immediate breakout force as a function of relative embedment depth 9-6
9.4-1 Normalized long-term breakout force as a function of breakout time parameter 9-8
9.7-1 Problem sketch and data for example problem 1 9-11
9.7-2 Problem sketch and data for example problem 2 9-14
10.4-1 Clear water scour and general sediment transport near a pile 10-3
10.4-2 Variation of maximum clear water scour depth with seafloor material diameter at a cylindrical pier 10-4
10.4-3 Idealized wave-induced flow and scour patterns around a vertical cylinder 10-6
10 4-4 Summary plot of field and model scour depth data at single piles and pile groups 10-7
10.4-5 Scour comparison for very large circular, square and hexagonal cylinders of equal cross-sectional area where a/D << 0.2 10-8
10.4-6 Scour development at a pipeline exposed to a uniform current of 1.2 to 1.5 ft/sec 10-10
10.4-7 Burial of a small object on the seafloor from a progressive scour sequence 10-11
10.5-1 Example of an inverted filter for scour protection at an offshore stricture 10-14
10.5-2 Example of a filter material specification (Diamond Shoals Light) 10-14
11.1-1 Loading and resistance forces on a potential slide mass 11-1
11.2-1 Enhanced seismic profile of a translational slide 11-3
11.2-2 Seismic profiles of a rotational slide 11-4
11.2-3 Seismic profiles of a submarine flow 11-4
11.3-1 Probabilistic description of loading and resistance and risk assessment in slope stability evaluation 11-6
11.5-1 Summation of forces and moment for stability analysis 11-8
11.6-1 Examples of high resolution seismic records of slides caused by slope instability 11-10
11.6-2 Examples of a shallow coring plan used in stability analysis 11-12
11.7-1 Features on seismic profiles indicative of bottom instability problems 11-15

xiii
Chapter 1

INTRODUCTION

1.1 OBJECTIVE aspects of engineering problems associated with the facilities


are difficult for Navy engineers to address because of the highly
Marine geotechnical engineering is the application of
specialized nature of most geotechnical topics. Also, due to a
scientific knowledge and engineering techniques to the
general lack of historical precedence for seafloor construction, a
investigation of seafloor materials and the definition of their
low level of understanding of seafloor soil behavior exists.
physical properties. The responses of these seafloor materials to
Much of what does exist is published in documents not widely
foundation and mooring elements are the engineering aspects of
distributed. The Handbook brings this information together. It
marine geotechnology which are addressed in this document.
is intended for use by Navy engineers who do not have an
This Handbook for Marine Geotechnical Engineering brings
extensive background in geotechnical engineering. The
together the more important aspects of seafloor behavior for
Handbook is not an all-inclusive design manual. Rather, the
application to Navy deep ocean engineering problems.
objective of the Handbook is to familiarize engineers with
The Navy installs, or may require installation of, a
geotechnical aspects of problems, serve as a design guide for
variety of facilities fixed to the continental shelves and slopes,
relatively uncomplicated problems, and be a technical directory
to the submarine slopes of seamounts and islands, and to the
to more complete discussions and to more sophisticated
deep ocean floor. Some of these facilities rest on shallow
analysis and design procedures. Although it is intended for use
foundations resembling a spread footing or on pile-like
with deep ocean problems (nominally beyond the continental
foundations. Others may be surface- or subsurface-moored types
shelf or below about 600 feet), the information contained in the
where a buoyant element is tethered to the seafloor by uplift-
Handbook is applicable to problems in shallow water as well.
resisting foundations or piles or by propellant-embedment or
drag-embedment anchors. Behavior of the mooring elements
1.2 HANDBOOK ORGANIZATION
lying on or embedded in the seafloor is dependent on the
This Handbook has 11 chapters; an Introduction, and 10
physical properties of the materials making up the seafloor in
technical chapters grouped into three major sections:
the immediate area. In addition, scour and slope stability
PROPERTIES DETERMINATION, DESIGN OF
problems may exist or may be created by the placement of these
FOUNDATIONS AND ANCHORS, and OTHER SEAFLOOR
elements.
PROBLEMS.
Navy military and civilian engineers will be required to
The Introduction serves as a guide to the remaining
plan for, design, supervise construction of, or have technical
chapters. It lists generalized features for each type of foundation
responsibility for these facilities. Geotechnical
and anchor,

1-1
and can assist the reader in selection of an appropriate In the Other Seafloor Problems section four other aspects of
foundation or anchor type based on environmental conditions marine geotechnical engineering are discussed. Chapter 8
and structural requirements. describes techniques for predicting the depth of penetration of
The Properties Determination section, consisting of objects into the seafloor. The techniques can be used for
Chapters 2 and 3, discusses on-site and laboratory determination penetration predictions with large and small objects of various
of soil properties and presents physical property models for shapes (such as lost hardware, instrument packages, or founda-
major seafloor soil types. Chapter 2 describes the various tion elements) impacting the seafloor at high or low initial
aspects of surveying a site, including preliminary reference velocities. The procedures of Chapter 8 can also be used to
information gathering and survey planning through brief predict the force required to embed a given object to a specified
descriptions of remote survey equipment, shallow and deep subbottom depth (shear keys below a bottom-resting
sampling equipment, and in-situ soil properties testing foundation, for example). Chapter 9 presents techniques for
equipment. Additional information on site surveying for the predicting the force or time required for breakout of objects
assessment of slope stability is given in Chapter 11. Chapter 2 embedded in the seafloor and discusses conditions that can have
also contains a section on estimating soil properties for use in a a significant effect on breakout. Analytical techniques are given
preliminary design when no field data are available. Chapter 3 for two significantly different cases -- full-suction and zero-
describes the laboratory tests performed on recovered soil suction--along with a discussion of mechanical techniques that
samples to generate index and engineering properties data can reduce the breakout forces and time requirements Chapter 10
required for analysis and design of seafloor structures. Use of describes scour prediction techniques. It is directed primarily
index properties to classify the soil and to correlate with toward scour problems around objects on the seafloor (local
engineering properties is also described. scour), but includes a discussion of nearshore seasonal seafloor
The Design of Foundations and Anchors section, profile changes. Most information on scour is drawn from
consisting of Chapters 4 through 7, describes the use of historical observations and model studies of nearshore and river
topographic, stratigraphic, and soil properties information conditions. Insight from these studies is extrapolated to
necessary to predict capacities of foundation and anchor conditions more likely to exist in deeper marine environments.
systems. Chapter 4 covers the design of foundations and Chapter 11 discusses slope stability in qualitative terms.
deadweight anchors bearing on the seafloor surface. Design of Techniques are described for surveying a site to determine its
piles for use as foundations or anchors is discussed in Chapter potential for slope instability. The methods of stability analysis
5. Plate-shaped anchors embedded in the seafloor are treated in are described but are not presented in detail. Considerable
Chapter 6, with emphasis given to the propellant-embedment technical interpretive skills are required for evaluation of site
anchor systems recently introduced into the Navy's Ocean Con- information and application of the analytical procedures.
struction Equipment Inventory. Chapter 7 covers the selection Each chapter has a list of references and symbols used in
and sizing of drag-embedment anchors; only the resistance that chapter. Example problems, which outline design or
developed from anchor and chain interaction with seafloor calculation procedures, are presented at the end of each chapter
materials is discussed and not the design of a complete mooring that includes design procedures.
system. References 1-1, 1-2, and 1-3 can be consulted for
information regarding complete mooring systems.

1-2
1.3 SELECTION OF FOUNDATION/ANCHOR TYPE

Chapters 4, 5, 6, and 7 each describe a different type of


foundation or anchor--deadweight, drag-embedment, pile, and
direct-embedment [Figure 1.3-1]. Each of these foundation and
anchor types has strong points or features. This section
summarizes these features [Tables 1.3-1 through 1.3-4] to
provide guidance on selecting the optimum foundation or
anchor type for a given set of problem conditions.
Shallow foundations and deadweight anchors are widely
used in the deep ocean environment because they are simple and
readily sized for most seafloor and loading conditions. However,
they do not perform well on steep sloping sea-floors. In
addition, deadweight anchors are not very efficient (that is, the
ratio of lateral load resistance to anchor weight is very low
compared to other lateral-load-resisting anchor types). Table
1.3-1 lists these and other features of the shallow foundations
and dead-weight anchors.

Figure 1.3-1. Simplified anchor types.

Table 1.3-1. Features of Shallow Foundations


and Deadweight Anchors

Features of Shallow Foundations and Deadweight Anchors


1. Simple, on-site construction feasible, can be tailored to task.
2. Size limited only by load-handling equipment.
3. Reliable on thin sediment cover over rock.
4. Lateral load resistance decreases rapidly with increase in seafloor slope.
Additional Features of Deadweight Anchors
1. Vertical mooring component can be large, permitting shorter mooring line
scope.
2. No setting distance required.
3. Reliable resisting force, because most resisting force is directly due to
anchor mass.
4. Material for construction readily available and economical.
5. Mooring line connection easy to inspect and service.
6. A good energy absorber when used as a sinker in conjunction with
"nonyielding" anchors (pile and plate anchors)
7. Works well as a sinker in combination with drag-embedment anchors to
permit shorter mooring line scopes.
8. Lateral load resistance is low compared to other anchor types.
9. In shallow water, the large mass can be an undesirable obstruction.

1-3
Pile foundations and anchors are used where less Direct-embedment anchors can be driven into seafloor
expensive types of shallow foundations and anchors cannot soils by propellant, vibratory, or hammer-driving systems. The
mobilize sufficient resistance. A principal drawback of piles for propellant-embedment systems are particularly amenable to use
the deep ocean is the highly specialized equipment needed for in the ocean because of the relatively short time period required
installation and the associated very high mobilization and for installation and the few limitations on operating depth.
installation costs. Table 1.3-2 lists features of piles used for Other features of direct-embedment anchors are listed in Table
foundations and anchors on the seafloor. 1.3-3.

Table 1.3-2. Features of Pile Foundation and Anchor Systems

Features of Pile Foundations and Pile Anchors


1. Requires highly specialized installation equipment.
2. Transmits high axial loads through soft surficial soils down to competent
bearing soil or rock.
3. Can be designed to accommodate scour and resist shallow mudflows.
4. Can be installed and performs well on substantial slopes.
5. Can be installed in hard seafloors (rock and coral) by drill-and-grout
technique.
6. Drilled-and-grouted piles require more specialized skills and installation
equipment and incur high installation costs.
7. Wide range of sizes and shapes are possible (pipe, structural shapes).
8. Field modifications permit piles to be tailored to suit requirements of
particular applications.
9. Costs are high and increase rapidly in deeper water or exposed locations
where more specialized installation vessels and driving equipment are
required.
10. Accurate soil properties are required for design.
Additional Features of Pile Anchors
1. High lateral capacity (greater than 100,000 lb) achievable.
2. Resists high uplift as well as lateral loads, permitting use with short
mooring line scopes.
3. Anchor setting not required.
4. Anchor dragging eliminated.
5. Short mooring line scopes permit use in areas of limited sea room or where
vessel excursions must be minimized.
6. Pile anchor need not protrude above seafloor.
7. Driven piles are cost competitive with other high-capacity anchors when
driving equipment is available.
8. Special equipment (pile extractor) may be required to retrieve or refurbish the
mooring, or new pile and pendant must be installed.
9. More extensive and better quality site data are required than the data required
for other anchor types.
10. Pile capacity goes to zero when its capacity as an anchor is exceeded and
pullout occurs (is a 'nonyielding' anchor).

1-4
Table 1.3-3. Features of Direct-Embedment Anchors

Features of All Direct-Embedment Anchors


1. High capacity (greater than 100,000 lb) achievable.
2. Resists uplift as well as lateral loads, permitting moorings of short scope.
3. Anchor dragging eliminated.
4. Higher holding-capacity-to-weight ratio than other anchor types.
5. Handling is simplified due to relatively light weight.
6. Accurate anchor placement is possible; no horizontal setting distance
necessary.
7. Does not protrude above the seafloor.
8. Possibly susceptible to strength reduction accompanying cyclic loading
when used in taut moorings in loose sand and coarse silt seafloors.
9. For critical moorings, soil engineering properties required.
10. Anchor typically not recoverable.
11. Anchor cable may be susceptible to abrasion or fatigue.
Features Unique to Propellant-Driven Plate Embedment Anchors
1. Can be placed on moderate slopes, and in rock and coral seafloors.
2. Installation is simplified as compared to other types because they can be
embedded immediately on seafloor contact.
3. Special consideration needed for ordnance.
4. Gun system not generally retrieved in water deeper than approximately
1,000 ft.
Features Unique to Screw-In, Vibrated-In,
and Hammer-Driven Plate Anchors
1. Can better accommodate layered seafloors (seafloors with variable resistance)
because of continuous power expenditure during penetration.
2. Penetration is controlled and can be monitored.
3. Surface vessel must maintain position during installation.
4. Operation with surfaced-powered equipment limited to shallow depths by
power and strength umbilicals as well as the mooring line.
5. Operation limited to sediment seafloors.

Table 1.3-4. Features of Drag-Embedment Anchor Systems

1. Wide range of anchor types and sizes available.


2. High capacity (greater than 100,000 lb) achievable.
3. Most anchors are standard off-the-shelf equipment.
4. Broad experience with use.
5. Can provide continuous resistance even though maximum capacity has been
exceeded.
6. Is recoverable.
7. Does not function well in rock seafloors.
8. Behavior is erratic in layered seafloors.
9. Low resistance to uplift loads; therefore, large line scope required to cause
near horizontal loading at seafloor.
10. If dragging is not acceptable, anchor must be pulled horizontally at high
loads to properly penetrate and set.
11. Dragging of anchor to achieve penetration can damage pipelines, cables, etc.
12. Loading must be limited to one direction for most anchor types and
applications.
13. Exact anchor placement limited by ability to estimate setting distance.

1-5
Table 1.3-5. Performance of Foundation and Anchor Types as a
Function of Seafloor and Loading Conditions

Table 1.3-4 lists features of drag-embedment anchors. Although 1.4 REFERENCES


these anchors can develop high capacities, the load on a drag
1-1. Design manual, harbor and coastal facilities, Naval
anchor is usually limited to one direction, and the mooring line
Facilities Engineering Command, NAVFAC DM-26.
angle at the seafloor must be virtually horizontal. The holding
Washington, D.C., Jul 1968.
capacity of drag anchors decreases very quickly as mooring line
angles exceed approximately 6 degrees. To assist in 1-2. API recommended practice for the analysis of spread
understanding the advantages and disadvantages of the various mooring systems for floating drilling units, American
anchor types, Table 1.3-5 compares how well they function Petroleum Institute, API RP 2P Dallas, Tex. , May 1981.
under different conditions. Judgments of expected performance
1-3. Rules for building and classing mobile offshore drilling
have been made primarily on the basis of holding capacity and
units, American Bureau of Shipping. New York, N.Y., 1980.
relative cost. It should be noted that Table 1.3-5 is an
expeditious guide for general use, and special circumstances can
shift the performance ratings.

1-6
Chapter 2

SITE SURVEY AND IN-SITU TESTING

2.1 INTRODUCTION topography on drag anchor performance or (2) a technical


inability to use this data element in design because analysis
2.1.1 Purpose
techniques are not developed, as in the inability to use dynamic
This chapter summarizes considerations and methods for soil properties in drag anchor design due to an absence of a
selecting and characterizing a site for bottom-resting or moored performance-related model
platforms in the deep ocean. Table 2.1-2 lists geotechnical parameters required for each
of the applications.*
2.1.2 Factors Influencing the Site Survey

2.1.2.1 Constraints . The type and detail of site data 2.1.2.3 Regional Versus Site-Specific Surveys.

sought will be a function of. Some projects or project phases require general information
from a large region, whereas others require more accurate data
Value and replacement cost of platform from a small geographic area. For example, a manned habitat
Impact of platform failure (primarily) on human life
Purpose of the platform installation may require low-precision data resulting from a
Topography and seafloor material type regional survey over a large area to determine an adequate or a
Any presurvey requirements for an exact geographical
location best location for its placement, while design for the habitat's
Types of man-induced and environmental loadings foundation requires high-precision data from the selected site.
Type and size of the foundations or anchors
Availability of personnel, equipment, and survey Since regional surveys compare sites or cover large
support platforms distances, detailed information is not possible or is not needed
as much as for site-specific studies, which provide design
2.1.2.2 Minimum Required Data. Table 2.1-1 is a
information. Regional surveys typically include geophysical
summary of site data requirements for various geotechnical
data collection with limited soil sampling, such as gravity
engineering applications.
coring. Deep soil borings and in-situ tests are not typically used
The level of importance or need for each site data element
for site selection. Once a Site has been selected however, site-
is also identified. "Low in Table 2.1-1 indicates a low
specific data are collected, usually from soil samples and from
requirement level, which may result from either: (1) a low
in-situ tests, when the project warrants such tests. Geophysical
impact of this data element on the system design and
and geological information
performance, as in the low impact of micro-
______________________
* Symbols used in Table 2.1-2 and other tables are defined
where used in the text and summarized in Section 2.8.

2-1
Table 2.1-1 Site Data Requirements for Categories of Geotechnical Engineering Applications

a High, mandatory; low, can design without; 0 = not needed

Table 2.1-2. Soil Engineering Parameters Normally Required for Categories of


Geotechnical Engineering Applications

a Symbols explained in Section 2.8.

2-2
Table 2.1-3. Historical Environmental Information Needed
to Assess Geotechnical Hazards

Hazard Information Needed


Earthquakes Frequency
Magnitude
Peak accelerations
Response spectra
Wind Velocity distribution
Direction distribution
Maximum wind velocity
Waves Wave height distribution
Maximum wave height
Direction distribution
Currents Vertical velocity profile
Distribution of current velocity

beyond that collected during the regional survey may also be respect to the surface vessel , and location of the vessel with
needed, depending on the project complexity. respect to geographical coordinates. Due to great distances and
lack of bench marks, positioning offshore often requires more
2.1.2.4 Hazardous Conditions, Scope of the site
sophisticated techniques than those used on land. Recent
investigation will be influenced by the sensitivity of the developments in electronic communication systems have
platform to geotechnical hazards, including earthquake loading, enabled the surveying of positions at most points on the
faulting, liquefaction, submarine landslides, erosion, and worlds oceans to a maximum accuracy of 15 meters. However
presence of underconsolidated sediments, These conditions must equipment availability may be the primary limitation on
be assessed for sites being evaluated. positioning accuracy. Relative seafloor-to-surface positions are
Table 2.1-3 outlines historical environmental information measured by sonar pingers and echo-sounding devices to
needed for assessment of geotechnical hazards from earthquakes, accuracies on the order of 0.1% of the distance being measured.
winds, waves, and currents.
Investigation of environmental factors and hazardous 2.2 PRELIMINARY STUDY
features can be pursued initially by an examination of existing 2.2.1 Information Sources
maps, charts, and bottom environmental data (Section 2.2.1).
In a preliminary survey of a site, one important step,
2.1.2.5 Positioning Capability . The ability to generally referred to as a "desk top study," is the search for

reference a site survey and position a platform on the seafloor available information from previous investigations near the

may dictate the scope of the site investigation. A precise site. Findings from the desk top survey can provide areal

determination of horizontal and vertical position is a critical information, as well as site-specific data, and aid in planning for

aspect of both geophysical and geotechnical investigations for a a more detailed survey.

specific installation. Information can be obtained from a variety of

Positioning an object on the seafloor usually requires governmental, industrial, and educational

location of the object with

2-3
institutions. Sources of information on geological and Sources of Data on Earthquakes and Earthquake Effects
geotechnical properties of ocean sediments and on earthquake American Meteorological Society, Bibliography on marine
seismics, by O. Leenhardt, Washington, D. C., 1967. (Special
and earthquake effects are given in the following lists.
bibliographies on oceanography. Contribution No. 3)
Universities and Government Organizations
Bulletin of the Seismological Society America, P. O. Box 826,
Chief, Atlantic Branch of Marine Geology, U. S. Geological Berkeley, CA 94701
Survey, Bldg. 13, Quissett Campus, Woods Hole, MA 02453
Journal of Geophysical Research, American Geophysical
Chief, Pacific Arctic Branch of Marine Geology, U.S. Union, Suite 435, 2100 Pennsylvania Avenue, N.W.,
Geological Survey, 345 Middlefield Road, Menlo Park, CA Washington, DC 20037
94025
Kresge Seismological Laboratory, California Institute of
Chief of Operations Division, National Ocean Survey, National Technology, 220 North San Rafael Avenue, Pasadena, CA
Oceanic and Atmospheric Administration, 1439 W. York 91105. A prime source of earthquake information.
Street, Norfolk, VA 23510
Navy Mine Defense Laboratory. Report 181. Bibliography on
Chief of Operations Division, National Ocean Survey, National sieches, by F C.W. Olson, Panama City, Fla. , Aug 1962.
Oceanic and Atmospheric Administration, 1801 Fairview (AD 283510)
Avenue, East Seattle, WA 98102
University of California, Seismological Department, Berkeley,
Lamont-Doherty Geological Observatory of Columbia CA 94720.
University, Palisades, NY 10964
University of Hawaii, Hawaii Institute of Geophysics,
National Geophysical and Solar-Terrestrial Data Center, Honolulu, Hawaii 96822.
Environmental Data Service, National Oceanic and Atmospheric
Administration, Boulder, CO 80302 University of Hawaii, Hawaii Institute of Geophysics, Report
HIG6725. Bibliography to the preliminary catalog of tsunamis
Naval Oceanographic Office, Code 3100, National Space occurring in the Pacific Ocean, by K. Iida, D. C. Cox, and G.
Technology Laboratories, NSTL Station, MI 39522 Pararas-Carayannis, Honolulu, Hawaii, Dec 1967.

Scripps Institution of Oceanography, La Jolla, CA 92093 University of Tokyo, Hongo, Bunkyo-Ku, Tokyo, Japan.

Woods Hole Oceanographic Institution, Woods Hole, MA U. S. Geological Survey, 345 Middlefield Road, Menlo Park,
02543 CA 94025
Journals and Conference Proceedings
2.2.2 Typical Ocean Soils
Canadian Geotechnical Journal, National Research Council of
Canada, Ottawa, Canada 2.2.2.1 Sediment Types. Seafloor soils are referred to

Civil Engineering in the Oceans (I through IV), American by origin as either terrigenous (land-derived) or pelagic (ocean-
Society of Civil Engineers, 345 East 47th Street, New York, derived).
NY 10017-2398
Terrigenous soils, described below, include gravels,
Geotechnique, The Institution of Civil Engineers, London, sands, silts, and clays. These soils are formed on or adjacent to
England
land; are transported by currents, wind, or iceberg rafting to the
Journal of Geotechnical Engineering, American Society of deep sea; and contain >30% silt and sand-sized particles of land
Civil Engineers, 345 East 47th Street, New York, NY 10017-
2398 origin.

Marine Geotechnology, Crane, Russals & Co., Inc., 3 East Terrigenous Silty Clays, or Muds - bordering
44th Street, New York, NY 10017 continents
Turbidites - sand, silt, and clay deposits transported
Ocean Engineering, Pergamon Press, Fairview Park, Elmsford,
great distances into deep water areas by turbidity
NY 10523
currents; characterized by graded beds -- sands at
bottom grading to clays at top
Offshore Technology Conference, Houston, Tex. (yearly)
2-4
Slide Deposits and Volcanic Ash - derived from eous ooze (a sediment composed of at least 30% by weight of
slumps on marine slopes or from volcanoes remains of organisms whose hard parts are calcium carbonate) is
Glacial Marine Soils - coarse-grained sediments usually found at depths shallower than the CCD. Siliceous
produced by glacial scouring of land features oozes are found in ocean areas of high surface productivity,

Pelagic soils, described below, include pelagic clays, usually where the seafloor depth is below the CCD.

siliceous oozes, and calcareous oozes. These soils are formed in


2.2.2.2 Sediment Consistency. Three terms are used to
the sea; are composed of clays or their alteration products or
describe the existing state of a soil: overconsolidated, normally
skeletal material from plants or animals; and cover 75% of the
consolidated, and underconsolidated. Overconsolidated sediments
seafloor.
have been subjected to a greater load (overburden) in the past
Pelagic Clays - contain <30% biogenous material, than exists at present. They have been compressed and become
silty clays of very high plasticity stronger. The overconsolidation (OC) phenomenon can also
Authigenic Deposits - minerals precipitated and result from many chemical or physical processes. Normally
crystallized in seawater, predominantly manganese consolidated (NC) materials have never been loaded by
nodules and phillipsite
overlying material more than they are now. Underconsolidated
Biogenous Sediments - derived from marine organisms sediments are "young." That is, they have not come to
and plants
equilibrium with the weight of overlying materials and are
- Calcareous Oozes - contain >30% biogenous weaker than they will be when this equilibrium is reached.
calcium carbonate material, includes coralline
deposits, calcareous sands (oolithes) and shells, and Permeability characteristics generally limit underconsolidation
fine-grained remains of microscopic animals to the fine-grained cohesive soils.
(oozes)
The degree of consolidation is important to a site
- Siliceous Oozes - contain >30% siliceous fine- investigation because it dictates the existing state and, therefore,
grained remains of plants and animals
the strength of the material. In a normally consolidated soil,
Figure 2.2-1 Shows the distribution of surface sediments strength generally increases with depth -- in proportion to the
over the world's oceans. weight of soils which lie above. For a particular soil, strength
The continental shelves and slopes are characterized by at equal depths below the seafloor will be greatest for
terrigenous deposits. Terrigenous sediment may also be found overconsolidated soils and least for underconsolidated soils. Soil
beyond the continental slope, on the continental rise and compressibility will vary inversely with the degree of
abyssal plain, as a result of transport by slope failures and consolidation, being least for overconsolidated and most for
turbidity currents. The composition of terrigenous sediment underconsolidated.
will vary according to its source, the distance from its source, For pelagic sediments, it is usually correct and
and the regional geology. conservative to assume that they are normally consolidated. In
The pelagic sediments are composed primarily of wind- contrast, nearshore sediments are often overconsolidated,
blown dust (pelagic clays) or calcareous or siliceous biogenous particularly those sediments that were exposed when the sea
materials. The pelagic clays are usually found at water depths level was significantly lower than at present. No consistent rule
deeper than the carbonate compensation depth (CCD) [Figure exists for locating overconsolidated sediments except that
2.2-2]. The CCD is defined as the depth above which the exposed locations (such as tops of rises or passages)
calcium carbonate dissolution is less than the carbonate supply.
Calcar-

2-5
Figure 2.2-1. Ocean sediment distribution.

2-6
Figure 2.2-2. Topography of the carbonate compensation depth (CCD) (after Ref 2-1).

2-7
are more likely to be overconsolidated than are protected areas Additional discussion of sediment types and properties can be
(such as basins). Underconsolidated sediments are almost always found in other chapters where they apply to specific chapter
found where fine-grained soils are being deposited at a very high subjects. *
rate. In active river deltas, such as near the Mississippi River When the site is on the continental shelf or slope, the
Delta, there may be little-to-no increase in soil strength with sediment is assumed terrigenous. Available National Ocean
increasing depth below the sea bottom. Survey charts should be consulted to determine whether the
sediment is primarily cohesionless (sandy) or cohesive (mud or
2.2.2.3 Estimating Soil Properties. For planning a
clay). If the sediment is cohesive, Figure 2.2-4 is used for the
geotechnical survey or design of a nonsensitive, bottom-resting strength distribution of a normally consolidated sediment. A
device, estimates of the soil engineering properties can be literature search is made for strong indications of
developed when the marine geological province is known. First, overconsolidation (e.g., recorded outcrops of older sediments or
the probable soil type for that province is identified from Figure an exposed location such as rise top, high recorded bottom
2.2-3; then the soil shear strength and buoyant unit weight currents). If sufficient evidence exists that overconsolidated soils
parameters are estimated from data extrapolations presented in are suspected, it would be prudent to drop some penetrometers
Figures 2.2-4 through 2.2-8. or short gravity corers.

Figure 2.2-3. Marine geological provinces and probable soil types.

____________
* Examples: In Chapter 3, Section 3.5,
engineering property correlations with easily
measured soil index properties are given. In
Chapter 5 Section 5.3, typical soil properties
of cohesionless and cohesive materials are
listed for use in pile problem solutions. In
Chapter 6, Section 6.6, typical values for soil
permeability are listed for estimating cyclic
load capacity of direct-embedment anchors.

2-8
Figure 2.2-4. Typical strength profile for hemipelagic terrigenous silty clay.

Figure 2.2-5. Typical strength profile for proximal and distal turbidites.
2-9
Figure 2.2-6. Typical strength profiles for calcareous ooze.

Figure 2.2-7. Typical strength profiles for pelagic clay.


2-10
Figure 2.2-8. Typical strength profile for siliceous ooze.

Nonpenetration or slight penetration with attainment of abyssal plain, it must be determined whether its water depth lies
minimal sample length suggests that overconsolidated sediment above or below the CCD [Figure 2.2-2].
does indeed exist. Typical sand properties are given in Figure
2.2-5. If the location is near a large active river delta, the site 1. If above the CCD, the sediment is probably calcareous
ooze. Figure 2.2-6 gives the typical properties; it
must be surveyed directly.
should be noted that a further subdivision between
When the site location is beyond the continental margins, coarse and fine ooze is made at the 10,000-foot level.
the probable sediment type can be identified from Figure 2.2-1.
2. If the site is below the CCD, the sediment is probably
If the sediment is classed as a turbidite, Figure 2.2-5 gives pelagic clay or a siliceous ooze. Figure 2.2-7 presents
estimates of properties for a pelagic clay.
typical parameters for proximal and distal turbidites. The
distinction is made based on the distance from a source of sand
Whenever possible, specialists in seafloor soils behavior
(e.g., the edge of the continental shelf) as follows: if the
should be consulted as they can provide information difficult to
distance is greater than about 30 miles, the sediment is
glean from the open literature. Many parts of the seafloor have
probably a distal turbidite.
been mapped for sediment distributions, and much more detailed
If the sediment is classed as a siliceous ooze (diatom or
information than is given in this discussion may be available.
radiolarian ooze, Figure 2.2-1), the typical properties can be
In addition, many core sample descriptions are available.
found in Figure 2.2-8.
Sources for experts, maps, and core descriptions are listed in
If the site is in the deep ocean and not an
Section 2.2.1.

2-11
2.3 REGIONAL SURVEY layers) with different acoustic properties, a portion of the energy
is reflected [Figure 2.3-1(a)]. The times of arrival of the
2.3.1 General
reflected waves are recorded, producing a profile of acoustic
Regional survey techniques include seismic profiling, interfaces in the sediments.
limited soil sampling, side-scan sonar surveys, and direct High-resolution, continuous, seismic reflection profiling
(visual) observation. Data obtained from the regional survey are is commonly used for studying the upper 300 feet of soil.
often qualitative only from the standpoint of physical soil These devices are often referred to as sediment sounders,
behavior. For example, seismic profiling can be used to boomers, or sparkers. They are characterized by their
delineate major soil layers but gives little if any information transmission frequency and, consequently, the penetration of the
about soil consistency. Table 2.3-1 summarizes the steps in a signal and its resolving power. In general , lower frequencies
typical regional survey. produce greater penetration with lower resolution, while high-
frequency systems yield less penetration but have a greater
2.3.2 Seismic Profiling
resolution. Typically, geotechnical engineering needs are best
Seismic, or acoustic, profiling techniques use reflected served using a 12-kHz sounder system to develop profiles of the
sound signals to develop a profile of the seafloor topography seafloor surface and a 3.5-kHz sounder system to delineate the
and subbottom layering. Compressive waves formed by a sediment strata, near-surface rock contact, and surface faulting.
controlled sound source propagate outward through the seawater Additional discussion of acoustic profiling, and the
at a certain velocity. When the compressive wave encounters important role it plays in evaluating slope stability problems,
another medium (i.e., surficial sediments or deeper, denser is presented in Chapter 11, Section 11.6.3.1.
sediment
Table 2.3-1. Steps in a Typical Regional Survey
Steps Procedure
1 Perform literature and data bank search
2 Identify facility to be installed
3 Identify parameters that impact on siting and design of facility
4 Plan acoustic reconnaissance program for identifying relevant
geological hazards
(a) Acoustic subbottom profiler
(b) Survey line spacing, 500 ft
5 Plan shallow sampling program consistent with geotechnical hazard
identification and soil parameter evaluation
(a) Gravity cores
(b) One core per 1/4 mi2
6 Perform acoustic and shallow sampling survey
7 Assign soil tests, test soil, and select design soil parameters; for
each 6-1/2 ft of core
(a) Two bulk unit weight tests
(b) Two water content tests
(c) Two grain size analyses (for sand and silt)
(d) One Atterberg Limit test (cohesive soils only)
(a) One shear strength test, (cohesive soils only)
8 In post survey, evaluate cost versus risk tradeoff for proposed sites
and select and rank the sites

2-12
Figure 2.3-1. Acoustic profiling systems.

2-13
2.3.3 Limited Sampling slope stability problems, is presented in Chapter 11, Section
11.6.3.2.
Regional surveys should include some bottom sampling
in order to provide examples of surficial sediment types and 2.3.5 Visual Observation
consistencies. In areas of outcropping sediment layers, acoustic
Visual observations of the seafloor are made by three
profiling and surficial sampling can be used together to provide
general methods, (1) direct observation, (2) use of a remotely
information on the sediment type and projected properties of
controlled camera, and (3) underwater television. Divers can
subbottom layers of an area. Samplers used for such surveys
make direct observations in shallow water. In deep water, direct
would generally be of the less sophisticated variety -- the grab
observations are made from a deep-diving vehicle, such as the
samplers, short gravity corers, and rock dredges. A discussion of
Deep Quest or Turtle. Visual observations become necessary
sampling equipment is included in Section 2.4.
when other survey techniques cannot provide necessary site data.
2.3.4 Side-Scan Sonar For example, remote survey techniques are not able to delineate
the extent of a talus deposit at the base of a steep rock slope. A
Side-scan sonar systems provide graphic records that
survey by a manned submersible, possibly with some
show two-dimensional plan views of the seafloor topography.
waterjetting to remove sediment infilling, could provide such
Seafloor objects as well as gas bubbles are detected and
information.
displayed as in an aerial photograph.
Figures 2.3-1(a) and (b) show the operation of a side-scan
2.3.6 Survey Line Spacing
sonar system. The side-scan sonar operates by emitting high-
For small and relatively low risk (unmanned, low cost)
frequency sound waves in narrow beam pulses from a transducer
installations, a regional survey may not be required if general
"fish" that is towed off the stern of a ship. The fish is towed
information on bathymetry and stratigraphy at the site is
above the seafloor at a height dictated by the chosen range. The
available.
returning acoustic signals are received by the same fish and
Major structures (such as manned gravity platforms or
transmitted by electrical cable to the ship. On deck, data
pile-supported platforms), however, may require a full-scale
acquisition systems transform these reflected signals into an
regional survey. The following guidelines are suggested for
acoustic image on a dual-channel recorder. Depending on the
mesh spacing of regional surveys for such structures where the
scale selected, this image can record a continuous path of the
overall survey area is to cover a 1-by 1-mile area.
seafloor from 300 to 3,000 feet wide. Figure 2.3-1(b) shows an
example of a side-scan record.
For bathymetric surveys, a mesh spacing of about 100
The side-scan sonar technique can be used to detect
feet is recommended at locations where significant
seafloor obstructions, such as sunken ships, pipelines, sediment bottom change is occurring, and a spacing of 500 feet
is recommended at other locations.
flows, and rock outcrops. By studying the results of a side-scan
sonar survey, an undesirable site can sometimes be avoided in The mesh spacing for subbottom profiling is
essentially the same as for bathymetric survey. If the
preliminary site selection.
soil is generally homogeneous, the profile spacing can
Additional discussion of side-scan sonar techniques, and be increased.
the role they play in evaluating

2-14
2.4 SITE-SPECIFIC SURVEY cated equipment and techniques than sampling on land or near
shore. In deeper waters, sampling must be done from a floating
2.4.1 General
vessel. Gravity corers and vibracorers are usually used to obtain
Soil properties at a site can be established by sampling samples in the upper 10 to 20 feet. Below this soil depth,
the soil and returning it to a laboratory for testing. Generally, drilling rigs and wire-line sampling techniques are normally
in shallow waters the techniques of drilling and sampling used used. The performance of these sampling techniques in the deep
on land can be adapted, utilizing a jackup barge or a fixed ocean is limited by the handling capability of the supporting
platform. In the deep ocean, because of the great water depth, vessel and the weather condition. Table 2.4-1 summarizes the
the sampling of soil sediments often involves more compli- steps of a typical site-specific survey.

Table 2.4-1. Steps in a Typical Site-Specific Survey

Steps Procedure
1 Check data bank and literature search for completeness

2 Identify facility to be installed

3 Identify soil parameters and geotechnical hazards impacting on design of facility

4 Identify types of information needed to complement existing data (from preliminary


studies and regional survey) [see Tables 2.1-1 and 2.1-2]

5 Perform acoustic data collection as necessary to complete data collection


(a) Close survey line spacing to 100 ft
(b) Make additional or more accurate seafloor profiling (deep tow) and subbottom
profiling

6 Complete shallow sampling; may be sufficient for platform type and size
(a) Gravity corer generally acceptable
(b) Spacing of coring locations, 300 ft
(c) Core to a depth of 1.5 times expected width of foundation or to maximum expected
penetration of anchor, if possible

7 Perform deep sampling where necessary; use in-situ tests for high-risk platforms
(a) Core to a depth of 1.5 to 2.0 times expected width of foundation or to maximum
expected penetration of anchor
(b) Sampling frequency (within single boring):
(1) One sample/5 ft, up to 30 ft
(2) One sample/10 ft, between 30- and 200-ft interval
(3) One sample/20 ft, over 200 ft
(c) In-situ tests:
(1) Use vane for clay undrained properties
(2) Use cone penetrometer for sand or clay and to define strata boundaries
(3) Use pressuremeter for soil deformation properties
(4) Define strata boundaries by borehole logging

8 Assign soil tests, test soil samples, and select design parameters. For every core sample:
(a) One water content test
(b) One bulk unit weight test
(c) One grain size analyses
(d) One Atterberg Limit test every other sample (cohesive soils only)
(e) Shear strength tests (cohesive soils only)

2-15
Table 2.4-2. Shallow Soil Sampler Types and Applications

Maximum
Sampler Type Sample Sample Application Comments
Quality Length
(ft)
Grab Sampler low 2 Soil identification Inexpensive; no water depth limitation
Index property tests
Box Corer very high 2 Soil identification, No water depth limitation; pretripping
Sample for: causes delays in deep water; best for
strength test seas below 7 ft
index test
Gravity Corer high 4 Soil identification, Free-fall is limited to 20,000 ft and
Free Fall 10 Sample for: may be difficult to find upon
Short Corer 30 strength test surfacing; others can use piston for
Long Corer index test higher sample quality; no water depth
limitation; specialty piston corers can
sample deeper soils
Vibracorer moderate 20-40 Soil identification, Used primarily in sands; water depth
Sample for: limited by power umbilicals
index test

2.4.2 Shallow Sampling 2.4.2.2 Box Corers. Box corers obtain large volume

of relatively undisturbed surficial soil. Sample sizes range from


Shallow soil sampling is usually conducted as part of the
1 to 2 feet in height and 0.25 to 3 ft 2 in area. A box corer
regional survey. For small seafloor installations, shallow-
typically consists of a weight column, sample box, spade lever
penetration samples may provide the soil parameters required for
arm, and a tripod support frame [Figure 2.4-2].
foundation design. The equipment used for shallow sampling in
the deep ocean includes grab or dredge samplers, box corers,
gravity corers, and submersible platform corers. A summary of
shallow sampling tools and their application is given in Table
2.4-2.

2.4.2.1 Grab or Dredge Samplers. Grab or dredge

samplers [Figure 2.4-1] offer the simplest method for obtaining


seafloor soil samples. Because samples obtained by this method
suffer from significant disturbance, grab or dredge samples have
little value in evaluating soil strength characteristics. Large
grab samplers are, however, often the only means for taking
samples of gravels and pebbles for surficial sediment
identification. The washing out of fines during sample recovery
can be a problem with grab samplers. This can be minimized
by use of samplers designed to minimize such sample loss [Ref Figure 2.4-1 Grab samplers and dredges (from Ref 2-13).

2-3].

2-16
Figure 2.4-2. Box corer and its operation sequence.

2-17
2.4.2.3 Gravity Corers. Gravity corers are tube samplers short corers and their use is limited to approximately the upper

that are driven into the soil by the kinetic energy of their falling 10 feet of the seafloor. The piston corer incorporates a piston

mass. Most gravity corers are lowered by winch through the fixed at the mudline during penetration to improve soil recovery

water column and free-fall through only the last 10 to 20 feet of and recovered soil quality. Most gravity corers incorporate

the water column after being released by bottom contact of a plastic barrel liners to promote sample quality and post-recovery

trigger weight [Figure 2.4-3] or, when a trigger weight is not soil handling, and use core catchers to limit core loss during

used, by free-wheeling of the winch. One, the Benthos retrieval. A comprehensive discussion of corer performance and

boomerang corer, is not attached to the surface by cable and coring techniques can be found in Reference 2-4. Table 2.4-3

free-falls through the entire water column. lists many gravity corer types, their characteristics, and

Gravity corers may be of open-barrel or piston type comments on their performance.

Open-barrel corers are relatively

Figure 2.4-3. Long piston corer operation sequence with a short corer used as a trigger weight.

2-18
Table 2.4-3 Characteristics of Some Free-Fall and Lowered Corers

2.4.2.4 Vibracorers. Vibracorers are used in cohesionless less than 600-foot water depths because of the necessity to

sediments, where gravity corers often fail to retain core or are supply power from a vessel down to the corer. Ability to

capable of only limited penetration. The typical vibracorer penetrate varies with the strength of material being cored.

consists of a core barrel and driver-vibrator unit supported by a Unless very soft sediment is encountered, maximum coring

stand, as shown in Figure 2.4-4. The driver-vibrator may be of length is limited to 40 to 50 feet.

a rotating eccentric weight or reciprocating piston variety, and it


2.4.3 Deep Sampling
may be powered by a pneumatic, hydraulic, or electrohydraulic
source. Most government and commercial organizations Offshore site investigations for major engineering
interested in marine geotechnical investigations maintain structures, especially those placed on piles, require expensive
vibratory corers -- some with only slight differences to deep soil sampling and recovery of representative samples. Deep
accommodate company preferences. Most are limited to boring techniques require a higher level of sophistication in
operation in equipment and in the drill

2-19
samplers. The wire-line hammer sampler uses successive
vertical blows to advance the sampler and provides somewhat
disturbed samples. However, it is economical and provides
adequate information for pile design. Wire-line push samplers
advance the sampler with a continuous motion, which produces
less soil sample disturbance. However, the operation of a push
sampler requires a fairly calm sea since the motion of the ship
is transferred to the sampler through the drill pipe. A limited
number of drill ships do compensate for motion with specialty
equipment, but this is an expensive complication to the basic
drilling operation.
The depth to which soil sampling is necessary depends on
the type and size of the structure. For most pile foundations,
the borehole should be at least as deep as the anticipated pile
penetration plus three pile diameters. For gravity structures the
borehole depths are usually between one and two times the
diameter of the foundation, depending on soil conditions. In
Figure 2.4-4. Alpine vibracore sampler (from Ref 2-5). general the softer the subsurface soils, the deeper the borehole
required. The sampling frequency for a borehole normally varies
vessel or platform than is needed to perform shallow sampling. with depth as approximated in Table 2.4-1.
Drilling and sampling operations are generally carried out from A discussion of borehole logging techniques for
a fixed platform, a jackup platform, an anchored barge or ship, additional information about oil stratigraphy is given in Section
or a dynamically positioned ship. 2.5.6.
The most versatile and economical approach today to
2.4.4 Location, Number, and Depth of Sampling
obtain deep samples at sea is wire-line sampling, which
involves use of an anchored ship with a centerwell. Regular oil The recommended location and number of borings are
platform supply boats, which are about 150 to 200 feet long, based on structure purpose and size, seafloor bathymetry, slope
are generally suitable for this type of operation. Drilling is angle of the soil strata, and uniformity of the acoustic profiles.
performed with a rotary rig mounted on the ship deck over the For most structures, if the acoustic profiles are practically
centerwell. The drill bit is advanced into the soil by rotating a uniform over an extensive area, one deep boring at the center of
drill pipe while pumping drilling fluid down the pipe. The fluid the structure location along with other, more shallow, data is
carries the soil cuttings to the seafloor surface and is not often sufficient. If the seismic records display a number of
returned to the ship (open circulation). The drill bit and casing irregularities, more deep borings should be drilled. For small,
are advanced to the elevation where a sample is desired and then low-risk installations, a few gravity cores taken at the site may
the sample is taken. provide sufficient information for foundation design.
One of two sampler types is used: wire-line hammer The depth to which soil sampling is necessary depends on
samplers and wire-line push the type and size of the structure. For most pile foundations,
the borehole

2-20
should be at least as deep as the anticipated pile penetration plus liner is then sealed with plastic caps, tape, and wax.
three pile diameters. For gravity structures the borehole depths Cohesionless sediments are usually stored loose in bags or jars
are usually between one and two times the diameter of the because they are quite disturbed in the coring process
foundation, depending on soil conditions. In general, the softer
2.4.5.2 Storage. Proper sample storage prior to testing is
the subsurface soils, the deeper the borehole required. The
sampling frequency for a borehole normally varies with depth as somewhat controversial. It is usually best that samples are

approximated in Table 2.4-1. stored vertically, when possible, to maintain their natural

A discussion of borehole logging techniques for orientation and to limit mixing or changes in stress conditions.

additional information about soil stratigraphy is given in Cushioning should be used to protect the samples from

Section 2.5.5. vibration. Samples should be stored at 5 2oC in near 100%


relative humidity and away from direct sunlight to prevent
2.4.5 Sample Handling biological growth and other physical alterations that might
otherwise occur.
2.4.5.2 Preparing and Packaging. To minimize sample

disturbance, cores must be prepared and packaged for shipment 2.4.5.3 Transit. Samples should be shipped to the
as soon as possible after recovery.
laboratory as soon as possible after the vessel arrives in port.
For gravity cores, liners should be pulled out of the carrel
Undisturbed samples are best "hand-carried," either as carry-on
and cut into sections 3 to 6 feet long. The top and bottom of
luggage in an airplane or in a private vehicle rather than by
each section and the position of each section in the core should
commercial carrier. For larger shipments, where personal
be marked on the liner. The ends of the liner tube should be
control is not possible, air freight is recommended since it
sealed with plastic caps and electrical tape and preferably sealed
minimizes the time in transit and also reduces vibration and
again with an appropriate wax.
shocks that might further disturb the sample.
With box corers, the metal corer box is not watertight. It
Samples should be tested as soon as possible because
is very difficult to maintain box corer samples at their natural
even proper storage will only slow down and not stop sample
water content. For quality samples, shipboard subsampling of
property changes that occur with time.
the recovered sample should be done as soon as possible. Many
large-diameter liners with sharpened ends are pushed from top to 2.5 IN-SITU TESTING
bottom of the box sample with a rapid and continuous motion.
2.5.1 General
When all liners have been pushed into the box, the sediment
around the liners is removed and can be saved in jars or plastic In-situ testing procedures involve making measurements

bags for tests that do not require undisturbed samples. The within the soil with specialized equipment. The most common

subsamples in the liners are then sealed and marked. measurements are for the undrained strength of clays, the drained

Wire-line samples are usually extruded onboard the ship. friction angle for sands, and soil modulus of elasticity.

Clay sediments are normally cut into sections about three The attraction of the in-situ test approach is that soils are

diameters long. Each section is then wrapped with wax paper not removed from their native environment during property

and aluminum foil and sealed with wax in a plastic container. If evaluation. Marine soil samples are normally subjected to an

liners are used, the liners are extruded and separated using a wire appreciable disturbance and a decrease in hydrostatic

saw. Each

2-21
and confining soil pressure in bringing them from the seabed to tests, soil information derived from each, and data evaluation
the sea surface. This decrease results in further disturbance to will be discussed in subsequent sections. A more detailed
the internal structure of the soil. When combined with discussion of in-situ testing and equipment is given in
disturbance introduced by sampling and subsequent sample Reference 2-6.
handling, the total disturbance can mask the actual in-situ
2.5.2 Vane Shear Tests
properties of the soil. This is particularly true in soft cohesive
soils containing appreciable dissolved gases, such as those In the vane shear test, the torque required to rotate a four-
found in the Gulf of Mexico. For these soils, decreases in bladed vane embedded in the soil is measured and converted into
hydrostatic stress often result in gases coming out of solution, a measure of the shear strength of the soil. Selection of the
which completely disturbs and remolds the soil. vane size is dictated by the anticipated soil strength and by the
Relatively few pieces of equipment are available to do in- torque measuring limits of the equipment. Vane geometry and
situ testing, and these tests generally represent at least a tenfold rotation rate are standardized [Ref 2-7]. However, offshore
increase in cost over gravity or piston coring. Thus, in-situ operations typically vary somewhat for the standardized rotation
tests are employed only when data of satisfactory quality cannot rate of 6 deg/min due to the operator's experience and preference.
be obtained through shallow coring and acoustics -- as, for In-situ vane shear test devices are available for either
example, when measuring the engineering properties of sands wire-line operation through a drill string at the bottom of a
and gassy sediments. borehole or from a seafloor resting platform [Table 2.5-1]. The
Table 2.5-1 summarizes types of in-situ tests and maximum depth of test for the seafloor resting systems is
equipment capabilities. The types of limited by the available reaction force (the underwater weight of
the system) to

Table 2.5-1. In-Situ Tests, Applications, and Some Equipment Characteristics

aLimited to cored and cased boring holes.

2-22
about 20 feet in soft clays for existing devices. rate, and sensitivity can often affect vane results out only to the
The shear strength can be calculated from measured torque extent that they can be used as a rough measurement of strength
values using: variability.

(2-1) 2.5.3 Cone Penetration Tests (CPT )

where: suv = vane shear strength [F/L2] The type of cone penetrometer used for offshore
T = applied torque [LF]
geotechnical investigation is the quasi-static cone.* This cone
d = diameter of vane blade [L]
H = height of vane plane [L] is pushed into the soil at a constant rate while point resistance
and side friction are measured. These data are then related to soil
To obtain undrained shear strength, su, a correction factor,
strength and other soil properties. The cone penetrometer
accounting for effects of anisotropy and strain rate, should be obtains continuous in-situ information with depth in both
applied to the vane shear strength determined from Equation 2-1 cohesionless and cohesive soils.
such that: The quasi-static cone is standardized, and has a 60-degree
apex and a base diameter of 1.4 inches [Figure 2.5-2]. Reference
su = s uv (2-2)
2-9 contains additional information on this test. The

where is a correction factor from Figure 2.5-1 based on the


soil plasticity index (PI).

Figure 2.5-1. Correction factor for vane determined shear


strength (after Ref 2-8)

Many factors can affect the measurement of shear


strength. Values obtained in sensitive, overconsolidated, or
cemented clays may not be correct due to disturbance of the clay
during insertion of the vane. Erratic results are obtained in soils
containing shells, gravel, or wood fragments. Effects of
anisotropy, strain
____________
Figure 2.5-2. Electric friction-cone penetrometer tip (after Ref
* This differentiates between a dynamic seafloor penetration 2-9).
test, Section 2.5.5, and a conventional standard penetrometer
test (SPT) on land. The latter utilizes successive hammer
blows to advance the cone.

2-23
cone can be equipped with one or more friction sleeves of the
same outer diameter as the cone to measure the unit friction
along the soil-sleeve interface. Cone penetrometers can also be
equipped with electrical piezometers at the cone tip or along the
friction sleeve to monitor the magnitude and dissipation
characteristics of the induced soil pore water pressure during or
after cone penetration. End resistance and side friction during a
CPT sounding can be recorded mechanically or electrically.
Quasi-static cone penetrometers are used both from a wire-line
system down a drilled borehole and from special seafloor-resting
platforms [Table 2.5-1].
Strength parameters and consolidation and permeability
characteristics of the soils can be estimated through
interpretation of CPT data. Soil stratigraphy is obtained from
correlations between the friction ratio (sleeve friction/cone
Figure 2.5-3. Correlation between cone tip resistance and sand
bearing) and cone tip resistance [Ref 2-10 and 2-11]. These relative density (after Ref 2-10).
correlations vary between seafloors because of differences in
material types and stress history and can vary because of
equipment operational status. Existing correlations should be
calibrated with known or determined seafloor material changes
in a given marine locale before use.
The cone penetration resistance in granular soils is
influenced by the state of stress and by the relative density.
Figure 2.5-3 shows a correlation between cone bearing pressure
(tip resistance), qc , vertical effective stress, pvo, and relative

density, Dr , for normally consolidated medium to silty fine

uniform sands. This relationship is used to determine Dr from

pvo and qc , which are determined during testing. For

overconsolidated sands, a correction is made to qc by use of the

dashed lines on Figure 2.5-3. The estimated relative density


from these curves is used to estimate the maximum effective
friction angle from Figure 2.5-4.
The value of the constrained modulus of elasticity, Es, for

sand to be used in settlement computations may be estimated


from:
Figure 2.5-4. Estimation of sand friction angle, , from
material relative density (after Ref 2-10).
Es = (1.5 to 2.0) qc (2-3)

2-24
The higher value applies to silty sands, and the lower value velocity-dependent shift in measured frequency of a constant
applies to finer grained but noncohesive soils. frequency sound signal emitted by the penetrometer. A
In clays the undrained shear strength is estimated from: shipboard hydrophone is used to monitor the penetrometer
signal and provides an analog of penetrometer velocity as it
embeds into the seafloor. The expandable Doppler penetrometer
(2-4)
system has the following components:
where: t = total unit weight [F/L3]
1. A penetrometer with its constant frequency sound
Z = depth below seafloor [L] source

um = hydrostatic pressure at mudline m [F/L2] 2. A receiving hydrophone with a preamplifier

Nc = bearing capacity factor = 15 for nonfissured and


not highly sensitive clays with overconsolidation 3. A receiver for processing the incoming data
ratio (OCR) < 2 and Plasticity Index (PI) > 10
The Navy system is capable of operating in water depths of
2.5.4 Pressuremeter Tests
20,000 feet and has penetrated up to 45 feet in some soils
The pressuremeter test measures soil deformation as a
function of expansion pressure when a membrane is expanded 2.5.6 Borehole Logging Techniques

out into the soil from its position down a borehole. Soil Borehole logging techniques are used to enhance
parameters that can be estimated from the test are shear knowledge of soil changes in lieu of more expensive and time-
strength, deformation modulus (modulus of elasticity), and the consuming testing. They can be performed in the drill pipe after
horizontal earth pressure at rest. The pressuremeter may be the drilling and sampling operations. In one type of test, an
placed either in a predrilled hole or pushed into the soil [Table electric probe is lowered in the drill pipe while constantly
2.5-1]. Reference 2-6 presents details in the test and on data measuring natural gamma emissions of the soil. The higher
interpretation. radio-active mineral contents of clays yield higher gamma
emissions than sands. Thus, a correlation between natural
2.5.5 Dynamic Penetrometer
gamma logs and soil types (stratigraphy) can be established.
A dynamic penetrometer is a hydrodynamically shaped The importance of this type of testing will be greatly increased
cylinder that free falls through the water column and penetrates where coring problems are encountered and core recovery
into the seafloor. Penetrometer velocity is monitored as it is (percent) is very low.
slowed by the soil during penetration. Undrained shear strengths Information on thermal and magnetic properties of the
are calculated in a manner similar to the reverse of the dynamic soil mass around the borehole can be gathered by other logging
penetration prediction procedure of Chapter 8. In this method, techniques.
the change in velocity over a short penetration interval is used
to calculate the kinetic energy consumed, and then to calculate 2.6 SEISMICITY SURVEY

the soil strength required to consume that energy. The level of seismicity can be determined on the basis of
For the Navy's Doppler penetrometer, the change in previous records of earthquake activity both in magnitude and in
velocity is calculated from the frequency of occurrence (see Section 2.2.1 for sources).

2-25
Figure 2.6-1. Seismic risk map of United States
coastal waters (after Ref 2-12).

Site-specific seismic considerations can include acoustic 2.7 REFERENCES


seafloor and subbottom profiling and dynamic shear tests for
2-1. W.H. Berger and E.L. Winterer. "Plate stratigraphy and
instability due to liquefaction and for mud slides triggered by
the fluctuating carbonate line," in Pelagic Sediments on Land in
earthquake activity, location of nearby faults, and characteristics
the Oceans, K.J. Hsu and H. Jenkyns, editors. Blackwell,
of ground motion.
Oxford, England, International Association of Sedimentation,
In the absence of historical seismic data, Figure 2.6-1 and
Special Publication 1, 1974, pp 11-48.
the following relation may be used as a guide in determining
the seismicity level for design. 2-2. P. LeTirant. Seabed reconnaissance and offshore soil
mechanics for the installation of petroleum structures. Houston,
z= 0 1 2 3 4 5 Tex., Gulf Publishing Co., 1979. (English translation by John
g= 0 0.05 0.10 0.20 0.25 0.40 Chilton-Ward)

where: z = zone given in Figure 2.6-1


g = effective horizontal ground acceleration in terms
of gravitational acceleration [L/T2]

2-26
2-3. R. McQuillin and D.A. Arders. Exploring the geology of proceedings of a session sponsored by the Geotechnical
shelf seas. London, England, Graham and Trotman, Ltd., 1977. Engineering Div, American Society of Civil Engineers
National Convention, St. Louis, Mo. , 26-30 Oct 1981.
2-4. H.J. Lee and J. E. Clausner. Seafloor soil sampling and
geotechnical parameter determination handbook, Civil 2-12. Planning, designing, and constructing fixed offshore
Engineering Laboratory. Technical Report R-873. Port platforms, American Petroleum Institute, API RP 2A.
Hueneme, Calif., Aug 1979. Washington, D.C., Jan 1982.

2-5. G.B. Tirey. "Recent trends in underwater soil sampling 2-13. I. Noorany. "Underwater soil sampling and testing - A
methods," in Underwater Soil Sampling, Testing, and state-of-the-art review," in Underwater Soil Sampling, Testing,
Construction Control, American Society for Testing and and Construction Control, American Society for Testing and
Materials, ASTM STP 501. Philadelphia, Pa , 1972, pp 42-54. Materials, ASTM STP 501. Philadelphia, Pa., 1972, pp 3-41.

2-6. J. Briaud. In Situ tests to measure soil strength and soil 2.8 SYMBOLS
deformability for offshore engineering, Texas A&M Research
Cc Compression index
Foundation. College Station, Tex., Oct 1980.
CCD Carbonate compensation depth, below which carbonate
materials will dissolve
2-7. "Standard method for field vane shear test in cohesive
c Drained cohesion intercept (effective soil cohesion)
soil," in Part 19, 1982 Annual Book of ASTM Standards, [F/L2]
American Society for Testing and Materials, ASTM D 2573- cv Coefficient of consolidation [L/T]
72. Philadelphia, Pa. , 1982. d Diameter of vane bland [L]
Dr relative density
2-8. L. Bjerrum. "Embankments on soft ground," in
E modulus of elasticity [F/L 2]
Performance of Earth and Earth-Supported Structures, vol II,
Es Constrained modulus of elasticity (for sand) [F/L2]
American Society of Civil Engineers. New York, N.Y., 1972,
pp 1-54. g Effective horizontal ground acceleration [L/T2]
H Height of vane blade [L]
2-9. "Standard method for deep, quasi-static, cone and friction- k Permeability [L/T]
cone penetration tests of soil," in Part 19, 1982 Annual Book Nc Bearing capacity factor
of ASTM Standards, American Society for Testing and PI Cohesive soil plasticity index
Materials, ASTM D-3441. Philadelphia, Pa., 1982. pvo Vertical effective stress (soil overburden pressure)
[F/L2]
2-10. J.H. Schertmann. Guidelines for cone test performance
qc Cone tip resistance [F/L 2]
and design, U. S. of Transportation, FHWA-75-78-209.
SP Normally consolidated, saturated, recent uncemented
Washington, D.C., 1978. fine sands
St Sensitivity
2-11. B.J. Douglas and R.S. Olson. "Soil classification using
Su Undrained shear strength [F/L2]
electric cone penetrometer," in Cone Penetration Testing and
Experience, G.M. Norris and R.D. Holtz, editors;

2-27
suv Vane shear strength [F/L2] t Total unit weight [F/L3]

T Torque applied to vane [F] Correction factor for vane shear strengths
um Hydrostatic pressure at mudline [F/L3] Drained, or effective, friction angle [deg]

Z Depth below seafloor surface [L] u undrained friction angle for sands rapidly sheared [deg]

z API seismic risk zone (see Figure 2.6-1)


b Buoyant unit weight of soil

2-28
Chapter 3

LABORATORY DETERMINATION OF SOIL PROPERTIES

3.1 INTRODUCTION 3.2 SOIL CLASSIFICATION

3.1.1 Scope 3.2.1 Classification by Origin

Table 2.1-2 of Chapter 2 identified engineering properties Marine geologists classify seafloor soil types by origin.
of soils required for analysis and design for several applications This classification system was introduced in Section 2.2.2,
in the deep ocean environment. Chapter 2 outlined elements of where characteristic soil strength profiles were discussed for
preliminary "desk-top" and field surveys, including the each major type, based on origin. A soil sample in this system
acquisition of some engineering properties from in-situ tests. can often be classified with visual examination of core material
This chapter describes the laboratory phase of soils by experienced personnel. Classification by inexperienced
classification and engineering properties determination. personnel or where soil does not cleanly fit into one of the
major types requires properties testing. This is done with tests
3.1.2 Special Considerations
to establish grain size distribution (Section 3.3.7), to determine
Most considerations and concepts developed in carbonate and organic carbon content (Section 3.3.8), and to
conventional onshore geotechnical engineering apply also in the identify and sort constituents by visual microscopic
marine environment. Differences in handling, testing, and data examination. With soils found in the deep ocean, classification
evaluation techniques arise because of the new soil materials by origin is often a necessary element in predicting the
encountered (primarily the biogenous remains and the engineering behavior.
authigenic precipitates); the very low effective stresses in
3.2.2 Classification by Grain Size
surficial materials (and resulting very soft or loose physical
state); and, to a lesser extent, the salt content of the pore fluid. The marine geologist also classifies sediments strictly by
This chapter often cites conventional soils testing references grain size, according to the Wentworth scale [Table 3.2-1(a)] or
[Ref 3-1 and 3-2], with most of the material devoted to according to an American Society for Testing and Materials
presenting necessary deviations from standard procedures. (ASTM) scale that has slightly different grade limits and
subdivides material types [Table 3.2-1(b)]. The portion of the
sample

3-1
Table 3.2-1. Size Range Limits for Two Soil below gravel size from each of the size groups (sand, silt, and
Classification Systems
clay), in percent of sample dry weight, is often reported on a
trilineal plot [Figure 3.2-1]. This Trilineal Classification
System by itself normally does not provide an adequate
description of a soil for engineering purposes, but it is a rapid,
size classification tool.

3.2.3 Classification by Grain Size and Behavior

The Unified Soil Classification System is based on the


soil's grain size distribution and its index properties. A sieve
analysis for grain size distribution (Section 3.3.7) and simple
index tests (Section 3.3.6) are necessary for classification. Data
from these tests are input for the Unified Soil Classification
Chart [Figure 3.2-2] in developing a soil's classification.

Figure 3.2-1. Trilineal soil classification plot -


normally used with Wentworth
grade limits.

3-2
Figure 3-2.2 Unified soil classification chart (after Ref 3-3)
3-3
The System first divides soil into three groups: coarse- data reduction procedures necessary to properly evaluate marine
grained (gravels and sands), fine-grained (silts and clays), and soils. Some index and engineering properties determined from a
highly organic materials. The classifications indicate that more range of marine soil types are shown in Table 3.3-2.
than 50% of the sample grains based on dry weight, are larger
3.3.2 Sample Preparation
(coarse-grained) or smaller (fine-grained) than 0.074 mm in
diameter (no. 200 sieve). Highly organic soils are identified by ASTM D 421 describes the standard method for dry
their black or dark gray color and by their hydrogen sulfide odor. preparation of soil samples for grain size and Atterberg limit
The coarse-grained soils are further subdivided by their tests. This dry preparation technique can be used for
predominant grain size and by the index properties of their fine cohesionless terrigenous soils having no more than a trace of
fraction. The fine-grained soils are subdivided entirely based on biogenous material. However, for those soils with measurable
their index properties [see Figure 3.2-2]. The Unified System is proportions of biogenous material, dry preparation should not
described in detail in Reference 3-1. be used. Air drying removes water from intra-particle voids of
biogenous material and these voids are not resaturated later
3.3 INDEX PROPERTY TESTS
during the Atterberg limit tests. Further, the mechanical dis-
3.3.1 General aggregation technique used (grinding with mortar and rubber-
covered pestle) is far too abrasive for use with fragile biogenous
Index tests provide information on the present condition
materials, including coralline sands.
(water content) and on the physical and chemical composition
Instead, marine soil samples are normally prepared in a
(grain size distribution, Atterberg limits, carbonate content) of a
wet state and are only dried when the test is completed to obtain
soil sample. The index tests can be run quickly and
dry sample weight. Disaggregation, if required, is best
inexpensively, compared to most tests for engineering
accomplished using an ultrasonic bath, with the sample
properties. Empirical relationships have been developed between
immersed in a deflocculant solution.
several index properties and engineering properties [Table 2.1-2]
of soils found on land. Most of these empirical relationships
3.3.3 Water Content
can be expected to apply to terrigenous marine soils because
ASTM D 2216 describes the standard method for the
these soils are essentially similar soils moved offshore or
laboratory determination of water content for terrestrial soils. In
submerged due to sea level or land elevation changes. However,
marine soils, salt comprises a small portion of the fluid phase
when dealing with pelagic soils, the previously established
in the natural state. For highly accurate computations, a
empirical correlations between index and engineering properties
correction should be applied to the equation for calculating the
may not be applicable.
water content. This correction, however, is most often not
Table 3.3-1 lists pertinent information on the most
made.
widely used index property tests and the standard ASTM
The water content of the soil sample, corrected for
references that can be found in Reference 3-1. This section will
dissolved salts in the pore fluid, is obtained from:
discuss particularly important aspects of these test procedures as
applied to marine soils and will describe modifications to the
(3-1)
standard test and

3-4
Table 3.3-1. Requirements for Index Property Tests (after Ref 3-4)

aDisturbed or undisturbed indicates that the source sample may be of either type.
bWeights of samples for tests on air-dried basis. (Dry weight estimated before test and determined after index test is run. )
cMaterial for these tests should not be dried before the test is run.

3-5
Table 3.3-2. Some Index and Engineering Properties of Ocean Sediments
(Most Data Limited to Upper 2 Meters of Seafloor) (after
Ref 3-5)

3-6
where: w = water content corrected for salt content [%] The soil sample is first leached of soluble salts by
r = salinity of pore fluid defined as the ratio
placing the sample on filter paper in a Bchner funnel and
between the weight of dissolved salt and the
weight of seawater; for engineering, it is washing the sample with distilled water. This sample is then
sufficient to assume a value of 0.035
washed into the pycnometer, and the test is run as described in
W1 = weight of container and moist soil [F]
W2 = weight of container and oven-dried soil [F] the ASTM standard.
Wc = weight of container [F] The specific gravity of soil grains is calculated from:

3.3.4 Unit Weight


(3-4)
The wet and dry unit weights are determined from
relatively undisturbed soil samples obtained directly from core where: Gw = specific gravity of distilled water w at the
temperature t of the pycnometer and contents
tube or liner sections of known length and diameter (ASTM D (see ASTM D 854-58)
2937) or from a carved sample of known volume, such as a
Wd = weight of oven-dry soil [F]
consolidation sample or a triaxial cylindrical sample. The wet,
Wa = weight of pycnometer filled with a distilled
or bulk, unit weight of the soil sample is: water at temperature t [F]

(3-2) Wb = weight of pycnometer filled with distilled


water and soil sample at temperature t [F]

where: W = wet weight of the soil sample [F]


For most seafloor soils (except for siliceous oozes and
V = volume of the soil sample [L3]
pelagic clays of high iron oxide content), the specific gravity
can be estimated as 2.7 without incurring significant error.
The adjusted dry density, corrected for salt content, is:
Table 3.3-2 lists some measured values for specific gravity of

(3-3) ocean sediments [Ref 3-5].

3.3.6 Liquid Limit, Plastic Limit, and Plasticity Index


3.3.5 Specific Gravity

The liquid limit and plastic limit are water contents of


Two methods are used for determining the specific
soils at borderlines used to describe significant changes in
gravity, or grain density, of materials that make up soil
physical properties. They are known as Atterberg limits, after
samples. For that portion of a sample finer than the no. 4 sieve
the man who designed the test. Although the test specifics were
[4.76 mm], the pycnometer method is used (ASTM D 854-58).
somewhat arbitrary, they are now a primary standard for
For the coarser portion of a sample, a technique better suited to
indexing behavior of fine-grained soils.
the larger grain sizes is used (ASTM C 127-80, Ref 3-6). Most
marine soils are finer than 4.76 mm, and a discussion of the
3.3.6.1 Liquid Limit . The liquid limit wL, is the
pycnometer method will suffice here. ASTM D 854-58 provides
water content at the transition between the liquid and plastic
guidance on computing the weighted average specific gravity for
states of a soil. This is determined from testing with a cupped
those samples containing both coarse and fine materials.
device into which remolded soil is placed. A groove is scoured
into the soil, separating it

3-7
into two halves. The liquid limit is arbitrarily defined as the is agitated up and down. Remains of the original sample are
water content at which the two halves will flow together when worked until all the fine material has passed through the sieve
the finely calibrated cup containing the soil halves is dropped a into the pan. Most water from the pan sample is removed by
specific distance 25 times at a specified rate (ASTM D 423-66- passing through filter paper in a funnel. The moist soil fines
72). retained on the filter paper are then warmed by heat lamp or air
The liquid limit test is intended to be performed only on blower until the soil reaches a puttylike consistency suitable for
that portion of a soil sample passing a no. 40 sieve (less than the liquid limit test. However, because this method of sample
0.42 mm in diameter). It is very important that preparation of preparation removes most of the pore water salt from the
the test sample does not change the sample characteristics and sample, it may influence test results. If at all possible, direct
cause this boundary to shift. Shifting of the boundary is likely use of marine soils from the stored sample tubes and physical
with marked change of: (1) the soil's pore water salt content and separation of coarse materials is preferred for Atterberg limit
(2) the sediment grain characteristics. determinations.
To minimize the impact of salt concentration change in Except for the recommended changes in sample
the pore water, the liquid limit test on marine soils should be preparation and mixing, the liquid limit test is conducted as
run on material taken directly from the stored sample tubes. specified in ASTM D 423-66-72. It is most desirable to run the
Distilled or deionized water should be added to the sample to test at water contents just above and below that which would
raise the water content, or the sample should be allowed to lose require 25 drops to cause sample closure. Two tests should be
moisture by air drying with air blower or heat-lamp assistance run at each of these levels, and interpolation should be used to
until the liquid limit has been defined by running the test at determine the 25-drop water content. The one-point method
several water contents. Although this procedure will result in described in the standard should not be used for marine soils.
some change in water-salt concentration, the effect of this small
3.3.6.2 Plastic Limit . The plastic limit, wp, of a
change on the liquid-plastic boundary is minimal.
To minimize degradation of sediment grains, especially soil is the water content at the transition between the plastic

for the pelagic oozes, the mechanical agitation and remolding of and semisolid states. The plastic limit is arbitrarily defined as

the soil must be minimized. Dry preparation should not be used the lowest water content at which the soil can be rolled into

for specimens containing measurable percentages of biogenous threads 1/8 inch in diameter without the threads breaking into

or organic materials (Section 3.3.2). pieces (ASTM D 425-79).

For those cohesive soil samples containing significant All comments pertaining to the preparation and handling

amounts of coarser materials, separation of coarse and fine of the sample for the liquid limit test also apply to the plastic

material may be possible but is a laborious process. The limit test.

sample can be soaked in distilled water and gently pushed


3.3.6.3 Plasticity Index. The plasticity index (PI) is
through a no. 40 sieve. The empty sieve is first placed in a pan;
calculated as the difference between the liquid and plastic limits;
and then the soaked sample is poured into the sieve. Distilled
i.e.,
water is added to the pan, bringing the water level to 1 cm over
the screen. The soil is then gently stirred with the fingers while
PI = wL - w p (3-5)
the sieve
Values of the liquid limit and plastic limit for samples of
seafloor soil types are reported in Table 3.3-2.

3-8
3.3.7 Grain Size Analysis The grain size curves from the sieve analysis for larger
particles and from the hydrometer analysis for smaller particles
The determination of the grain size distribution of marine
may not agree exactly where the test data overlap. Part of this
soils is performed in the same way as for terrestrial soils
deviation arises because the theory on which the hydrometer
(ASTM D 422-63-72), with some variation in sample
grain size analysis is based (Stoke's theory) assumes a
preparation to limit grain particle degradation.* The distribution
spherical-shaped particle. Clay particles and much of the foram
of particle sizes larger than 0.075 mm (no. 200 sieve) is
fragments are plate shaped and do not conform to the theory.
determined by sieving. The distribution of sizes finer than
Further, the whole biogenous shells, especially the
0.075 mm is determined by hydrometer test. Grain sizes up to
foraminifera, are hollow spheres. Thus, their effective specific
2.0 mm (no. 10 sieve) can be included in the hydrometer test
gravity is lower than that measured (Section 3.3.5). The
sample to provide an overlapping of grain size distribution
specific gravity error causes the percentage of the sediment
curves from the two methods. Since most marine soils,
classified as fine-grained to be larger than it really is.
including the brown clays and oozes, are finer than 2.0 mm,
separation of the sample on the no. 10 sieve is not necessary; 3.3.8 Carbonate and Organic Carbon Content
the hydrometer test can be performed directly on the sample.
The organic and carbonate carbon contents of the marine
Samples of marine soils for grain size sieving are
sample should be measured for those soils with suspected high
prepared by the wet method. The soil is not oven-dried prior to
carbonate content (>30%) and for those giving off hydrogen
the test because this would remove the water from within the
sulfide gas. These tests have not yet been standardized by
biogenous structures and could alter the structure of clay-sized
ASTM. Carbonate and organic carbon content may be
particles or could cause particle bonding into larger particles.
determined with an induction furnace using procedures described
These effects are not reversed upon rewetting. When washing
in the instruction manual. The weight of soluble salts should
these sediments on the sieves, the agitation must be kept to a
be subtracted from the total uncorrected dry sample weight.
minimum to limit particle degradation.
Samples of marine soils for the hydrometer test are
3.4 ENGINEERING PROPERTY TESTS
prepared by rough mixing the sample with water to promote
3.4.1 General
separation of the very small particles. Some pelagic clay
samples, notably those having a high iron oxide content, are Engineering property tests define properties of soil or soil
very difficult to separate into individual particles. One technique samples at significant states of stress. The most important is
places the sample in a solution of dispersing agent, sodium the undisturbed state -- or as the soil exists in its natural
hexametaphosphate. The soil and dispersant are then mixed in a environment. Most testing attempts to establish properties for
blender, followed by centrifuging of the mixture to separate the this condition.
solid particles. This process is repeated perhaps three times to
The undrained shear strength, su , of samples of cohesive
reduce the natural flocculated structure to a dispersed structure. marine soils can be measured either by a vane shear test or by
Particle degradation for clay-sized material is not a significant an unconsolidated-undrained (UU) triaxial test.
problem, as the material is primarily minute plate-shaped
particles that do not break down.
____________
* Salt content of the water plays no part in this test.

3-9
The unconfined compression test, a special case of the UU undrained shear strengths, s u, greater than 14 psi because vane
triaxial test where confining pressure is zero, is often run failure conditions in these higher strength soils may deviate
because of its simplicity. Soil strength is measured differently from the assumed cylindrical failure surface and cause
in each of the tests, but the value will be approximately the significant error in the measured strength. This method includes
same as undrained shear strength. use of either conventional calibrated spring units or electrical
Laboratory vane shear testing is uniquely suited to very torque transducer units with a motorized miniature vane.
soft sediments that cannot stand under their own weight outside
2. Description of the Test. The miniature vane shear
the core liner (a prerequisite for sample preparation in most UU
test consists of inserting a four-bladed vane in an undisturbed
testing). In addition, the vane shear test is the only laboratory
tube sample or remolded sample and rotating it at a constant
test used to date for determining the sensitivity of soft marine
rate to determine the torsional force required to cause a
cohesive soils.
cylindrical surface to be sheared by the vane. Force is measured
Other, generally more complex, testing is done for soil
by a calibrated torque spring or torque transducer directly
parameters useful in predicting soil behavior under conditions
attached to the vane. This force is then converted to a shearing
different than those existing (for example, the different stress
resistance primarily on the cylindrical surface.
conditions created by placement of a structure on or in the soil).
The effective stress parameters, c and , define a generalized soil 3. Apparatus. The vane should consist of a rectangular
failure criterion (Mohr-Coulomb failure envelope) and are four-bladed vane as illustrated in Figure 3.4-1. It is
usually determined in a consolidated-drained (CD) or recommended that the height of the vane be twice the diameter,
consolidated-undrained (CU) triaxial test with pore pressure although vanes with other ratios can be used, including a height
measurements. equal to the diameter. Vane blade diameter typically varies from
The compression index, Cc , and coefficient of 0.5 to 1.0 inch. Variations from recommended values would be
consolidation, cv, are determined from the one-dimensional made where sample size presents constraints or where other
consolidation test. The permeability, k, can be calculated either special conditions exist. The vanes should be "thin" so as to
from consolidated triaxial tests or from one-dimensional displace no more than 15% of the soil when inserted into the
consolidation tests. soil.
A summary of test requirements is given in Table 3.4-1. Torque is applied to the vane by manual or motorized
Care should be taken that the testing conditions represent stress power. The shaft should be rotated at a constant rate of 60 to 90
states for the soil being investigated. deg/min. Another, slower standard rate (6 deg/min) is
sometimes used. Torque is measured through a spring or an
3.4.2 Vane Shear Test
electrical transducer rotating with the shaft.
ASTM has not, as yet, standardized the laboratory vane
4. Preparation of Samples. Soil samples into which the
shear test. The following is based on a draft of the standard
vane is inserted should be large enough to minimize influence
developed for Reference 3-1.
of container sides on the test results. The test should not be
1. Scope. The laboratory vane shear test is applicable centered closer to the edge of the container than 1.5 times the
to very soft to stiff saturated fine-grained soils. The laboratory vane diameter. Tests run in the same container should be at
vane shear test should not be used in soils with very high least 2.0 times the vane diameter apart from each other.

3-10
Table 3.4-1. Requirements for Engineering Property Tests (after Ref 3-4)

Reference Applicability and Size or Weight of


Test for Suggested Variations From Sample for Test
Test Suggested Test (Undisturbed or
Procedure Procedure Remolded)
Vane Shear Text, Section 3.4.2 Applicable to very soft Test usually run on exposed sample at
to stiff saturated fine-grained soils. end of core liner tube. Specimen height
3 x vane height. Specimen diameter
3 x vane diameter.

Unconfined ASTM D 2166 Pelagic clays from shallower than 5 m; Minimum cross-sectional area 10 cm2.
Compression (UC) biogenous oozes, and nearshore Length 2 to 3 x diameter.
terrigenous silts normally too soft to be
properly prepared for this test.

Triaxial ASTM D 2850 Information similar to that from UC test; Same as UC test.
Compression can be used with softer and more pervious
Unconsolidated sediments than UC.
Undrained (UU) Ref 3-2 Rate of shear limited to allow complete Same as UC test.
drainage.
Ref 3-2 More common and generally less time Same as UC test.
Consolidated- consuming than CD. Pressure lines
Drained (CD) leading to sample should be seawater
filled. Loading rod friction should be
Consolidated- minimized by using air bushing or
Undrained (CU) equivalent.

One-Dimensional ASTM D 2435 Must provide for very low initial load Sample diameter 50 mm or 2 to 5
Consolidation increments. Sample x height. Sample height 13mm.
should be submerged in seawater

Direct Shear ASTM D 3080 Limited to consolidated shear tests on Sample diameter 50 mm, or 2 x
fine-grained soils. height. Sample height 12.5mm.

3-11
rotation should be continued (at a more rapid rate) until two
complete revolutions have been completed from the original
position of the vane when it was inserted. Determination of the
remolded strength should be started immediately after
completion of rapid rotation. The procedure, outlined above, is
followed with vane removal and soil sampling done at its
conclusion.

6. Calculations. A graph is prepared showing the


applied torque versus the rotation angle. The vane shear
strength, suv, is computed from the maximum torque value
using the same equation previously introduced for computing
in-situ vane shear strength [Equation 2-1]. The remolded shear
strength is computed the same way
Soil sensitivity, S t, is derived from the ratio of

undisturbed to remolded shear strength as follows:

Figure 3.4-1. Miniature vane blade geometry.


(3-6)

5. Test Procedure. The vane shear unit should be The shear strengths measured by the laboratory vane shear test
securely fastened to a table or frame to prevent movement are influenced by effects of anisotropy and strain rate, as
during a test. The vane is inserted and fixed at an elevation in described in Section 2.5.2 for in-situ vane tests. For
the sample so that the vane top is embedded by an amount at comparison with undrained shear strengths determined by
least equal to the vane height. The sample should be held firmly triaxial testing, a correction factor should be applied as done for
to prevent rotation. Torque readings should be recorded at a the in-situ vane data through use of Equation 2-2 and the
frequency that will allow good definition of the torque-rotation correlation with plasticity index from Figure 2.5-1.
curve (approximately every 5 degrees of rotation) or until a
3.4.3 Unconfined Compression Test
maximum of 180 degrees of rotation is obtained. The vane
blade is removed and cleaned, and a representative sample of the The unconfined compression test for marine soils is run
specimen is taken from the vicinity of the test to determine the as defined in ASTM D 2166-66-79.
water content. The soil is inspected for sand, gravel and other The unconfined compression test should only be
inclusions that may have influenced test results. Care should be performed on samples that are cohesive, are relatively
taken to make notes of a sample or test peculiarities observed. impervious, and have sufficient strength to stand under their
The remolded vane strength, if desired, is obtained own weight. Pelagic clays from subbottom depths beyond 15
following the test on the undisturbed sample and prior to feet would usually meet these criteria; biogenous oozes and
removal of the vane and soil sampling. Following the initial nearshore silty sediments generally would not.
test, the vane

3-12
3.4.4 Unconsolidated, Undrained Triaxial Compression Test First, since marine sediments have seawater as a pore fluid, the
pressure lines leading to the sample should be filled with
The unconsolidated, undrained (UU) triaxial test for
seawater rather than freshwater, if possible, when this water
marine soils is run as described in ASTM D 2850-70. Although
may enter the sample. Although data have not been published
the UU test determines the same type of information as the
to show that freshwater changes soil behavior, in theory,
unconfined compression test, it can be run on somewhat softer
changes in pore water salt content accompanying water entry to
and more pervious sediments.
marine samples could significantly alter behavior. Using
3.4.5 Consolidated-Undrained and Consolidated-Drained saltwater in the pressure lines may, however, introduce a
Triaxial Compression Tests corrosion problem requiring use of stainless steel fittings at

The consolidated-undrained (CU) and consolidated-drained critical points. Secondly, many marine sediments are

(CD) triaxial tests are used to measure the effective strength considerably softer than those usually found on land. Load and

parameters, c and , of cohesive marine soils. These parameters pressure transducers used to make measurements need to be

can also be measured from the drained direct shear test (Section sensitive or accurate at abnormally low readings: accuracy to

3.4.6). The CU and CD triaxial tests are not standardized by within 0.02 pound for load and 0.02 psi for pressure. Devices

ASTM; however, a comprehensive test description and such as air bushings must be used to reduce friction between

procedure is given on pp 122-137 of Reference 3-2 and in loading rod and cell considerably below what is normally accep-

Reference 3-7. table for soils testing.

3.4.6 Consolidated-Drained Direct Shear Test 3.4.8 One-Dimensional Consolidation Test

The consolidated-drained direct shear test is another Procedures for the one-dimensional consolidation test

alternative for determining the effective strength parameters, c have been standardized in ASTM D 2435-80. Engineering

and , for marine cohesive soils. This test is standardized as properties determined by this test include the Compression

ASTM D 3080-72-79. The direct shear test is well-suited to a Index, Cc , the Recompression Index, C s; the coefficient of

consolidated-drained condition because the drainage paths consolidation, cv; the coefficient of permeability, k, and the
through the test specimen are short, thus allowing coefficient of secondary compression, c. These data are used for
consolidation to take place fairly rapidly. However, the test is estimating the amount and time rate of settlement under applied
not suited to the development of exact stress-strain relationships loads.
within the test specimen because of the nonuniform distribution The degree of overconsolidation is also typically
of shearing stresses and displacements. The slow rate of determined from this test. It is an engineering property of high
displacement provides for dissipation of excess pore pressures, value because of the high impact that soil stress history has on
but it also permits plastic flow of soft cohesive soils. shear strength and other soil behavior. A soil that has
undergone consolidation under a higher effective vertical
3.4.7 Considerations for Triaxial Testing of Marine Soils
overburden pressure than presently exists is overconsolidated.
Some special considerations must be made for triaxial The ratio of this past effective pressure to the present effective
testing of marine sediments. pressure is called the overconsolidation ratio (OCR). OCR may
also be determined through triaxial testing, which measures
consolidation.

3-13
The only major procedural difference in adapting the
standardized consolidation test to marine soils is a need for
(3-7)
applying very low loads, as low as 8 psf. This is normally
achieved by placing small weights on the loading cap for the This relationship and the data in Figure 3.5-1 can be used to
first few load increments. In addition, the soil sample should be make a rough estimate of the in-situ strength of
submerged in seawater rather than freshwater. overconsolidated nearshore marine clays.

3.5 PROPERTY CORRELATIONS

3.5.1 General

This section presents correlations of soil engineering


properties with index properties more easily measured. The
correlations can be used as a rough guide for estimating
properties when only limited site survey information is
available. They should not be used for design without
evaluation of actual material properties through laboratory or
in-situ tests. Figure 3.5-1. Relationship between su/P vo and PI for normally
consolidated late glacial clay (from Ref 3-8).
3.5.2 Nearshore Sediments

The nearshore terrigenous sediments are highly variable


Figure 3.5-2 shows a correlation between friction angle
in composition. A considerable amount of information in
(angle of shearing resistance, ) and plasticity index for
geotechnical engineering literature can be applied to these
normally consolidated, fine-grained soils. Although this was
sediments. Figure 3.5-1 shows a correlation between su/pvo
not developed with data from marine soils, it can be used to
versus plasticity index for normally consolidated (NC) glacial
make a rough estimate of for nearshore, terrigenous, fine-
clays on land and in coastal regions [Ref 3-8]. In this figure, su
grained soils.
is the undrained shear strength, and pvo is the effective vertical
Figure 3.5-3 shows a correlation between coefficient of
overburden pressure. In Figure 3.5-1, "young" refers to
consolidation, cv, and liquid limit for fine-grained soils.
normally consolidated recent sediments, and "aged" refers to
Although this was not developed with data from marine soils, it
clays that have developed higher strength due to higher inter-
can be used to make a rough estimate of cv for nearshore,
particle bonding that has occurred with aging. The effects on
terrigenous, fine-grained soils.
strength are similar to a mild overconsolidation.
The soil deposits in many river delta front regions of the
The su/pvo ratio for overconsolidated (OC) soils will be
continental shelf are underconsolidated (have not fully
higher than the range shown in Figure 3.5-1. Compared with
consolidated under their present effective overburden pressure)
normally consolidated (NC) soils, the ratio su/pvo has been
and have strengths lower than would be expected for their
observed to increase as a function of the over-consolidation ratio
existing stress condition. References 3-9, 3-10, and 3-11
(OCR) as follows
contain information on the properties of these special types of
sediments.

3-14
Figure 3.5-2. Relationship between friction angle and
PI for normally consolidated fine-grained
soils (after Ref 2-4).

Figure 3.5-3. Correlation between coefficient of consolidation and liquid


limit (from Ref 3-4).

3-15
The properties of nearshore calcareous sediments, engineering properties at equivalent depths. Chapter 2 presented
particularly their interaction with pile foundations, have been estimated strength profiles for soils found in deep ocean regions
found to be different from those of terrigenous sediments. to assist in planning for site surveys. Most data were
Calcareous sands are highly variable in character and behavior extrapolated from shallow soil samples by consolidating them
due to mode of deposition and alterations that take place after to the state of stress found at deeper elevations. The following
deposition. For this reason, typical properties cannot be sections present additional data on deep sea sediments, most of
suggested for nearshore calcareous sands. Reference 3-12 gives a which were obtained in connection with shallow (upper 60 cm)
summary of available information. exploration for manganese nodule deposits.

3.5.3 Deep Sea Sediments 3.5.3.1 Pelagic Clays. The pelagic brown or red clays are

very fine-grained silty clays typically with more than 60%


Although the deep sea region is a relatively calm
particles finer than 0.002 mm. Liquid limits range from 75 to
depositions environment, other processes, such as turbidity
275, and plasticity indices range from 40 to 180. As shown in
currents, can cause deposition in a considerably different
Figure 3.5-4, the plasticity data for pelagic clays indicate they
manner. Still, vast regions of the seafloor are covered with
behave like highly compressible clayey silts and silty clays
sediments relatively uniform in profile (compared to adjacent
(this is indicated by data plotting close to the "A" line and to
sediments at the same depth) with corresponding relatively
the right of the "B" line).
uniform

Figure 3.5-4. Range of PI values for pelagic clay.

3-16
Figure 3.5.5. Correlation between water content and C c /(1 + e o)
for pelagic clay and calcareous ooze.

The water content for pelagic clays is usually higher than 300% for siliceous ooze) and shear strengths in the range of 0.5
the liquid limit. The average undrained sheer strength within the to 1.5 psi.
upper 60 cm of the soil profile is in the range of 0.5 to 1.5 psi. Very limited test data indicate effective cohesion values,
The laboratory-measured sensitivity is in the range of 2 to 10. c, in the range of 0 (not measurable) to 4 psi and a friction
Effective cohesion values, c, in the range of 0.14 to 0.45 psi angle, , in the range of 27 to 37 degrees.
and a friction angle, , in the range of 27 to 37 degrees are The range of compression index, Cc , for calcareous oozes
indicated (from very limited test data). is shown in Figure 3.5-5.

The range of compression index, Cc , for pelagic clays is


shown in Figure 3.5-5. In consolidation tests, pelagic clays 3.6 REFERENCES
from shallow embedment depths exhibit mild overconsolidation
3-1. "Part 19, Soil and rock; building stones," in Annual
behavior.
Book of ASTM Standards, American Society for Testing and
3.5.3.2 Pelagic Oozes. Pelagic oozes are calcareous or
Materials, Philadelphia, Pa., 1982.
siliceous remains of tiny marine organisms or plants and have
properties based or the type of sediment and the amount of clay 3-2. T.W. Lambe. Soil testing for engineers. New York,

in the sediment. Only a limited amount of data is available on N.Y., John Wiley and Sons, 1951.

the engineering properties of pelagic oozes in the upper few feet


3-3. Earth manual, Bureau of Reclamation, U.S. Department
of the soil profile. These indicate that oozes have water contents
of the Interior. Denver, Colo., 1968.
in the range of 50 to 100% (or up to

3-17
3-4. Soil mechanics, Naval Facilities Engineering Command, Cc Compression index, also grain size coefficient
NAVFAC DM-7. Washington, D.C. 1971. CD Consolidated-drained triaxial compression test
CH Inorganic clay, high plasticity
3-5. P.J. Valent. Engineering behavior of two deep ocean
CL Inorganic clay, low to medium plasticity
calcareous sediments, including influence on the performance of
Cs Recompression index
the propellant driven anchor, Ph.D. thesis, Purdue University.
CU Consolidated-undrained triaxial compression test (with
Lafayette, Ind. , Aug 1979, pp 19-20. pore pressure measurement)
Cu Coefficient of (grain size) uniformity
3-6. "Part 14, Concrete and mineral aggregates," in Annual
cv Coefficient of consolidation [L2/T]
Book of ASTM Standards, American Society for Testing and
Materials. Philadelphia, Pa., 1982. c Coefficient of secondary compression

d Diameter of vane blade [L]


3-7. A.W. Bishop and D.J. Henkel. The measurement of soil D10 Sample grain diameter, below which 10% of material
properties in the triaxial test, second edition. London, England, fails
Edward Arnold Ltd., 1962. D30 Sample grain diameter, below which 30% of material
fails
3-8. L. Bjerrum. "Embankments on soft ground," in D60 Sample grain diameter, below which 60% of material
Proceedings of ASCE Specialty Conference on Performance of fails

Earth and Earth-Supported Structures, vol 2, Purdue University, eo, e o Void ratio

1972. GC Clayey gravel


GM Silty gravel
3-9. I. Noorany and S.F. Gizienski. "Engineering properties
GP Poorly graded gravel
of submarine soils, state-of-the-art review, " Journal of Soil
Gs Specific gravity of soil grains
Mechanics and Foundations Division, ASCE, vol 96, no.
Gw Specific gravity of distilled water
SM5, Sep 1970, pp 1735-1762.
GW Well graded gravel

3-10. A.F. Richards. Marine geotechnique. University of H Height of vane [L]

Illinois Press, 1967. k Permeability [L/T]


LI Liquidity index
3-11. K. Terzaghi. "Varieties of submarine slope failures," in MH Inorganic silt
Proceedings of Eighth Texas Conference on Soil Mechanics and ML Inorganic silt, low plasticity
Foundation Engineering, University of Texas, 1956. NC Normally consolidated
OC Overconsolidated
3-12. K.R. Demars, editor. Geotechnical properties behavior,
OCR Overconsolidation ratio
and performance of calcareous soils, American Society for
OH Organic clays, medium to high plasticity
Testing and Materials, STP 777. Philadelphia, Pa., 1982.
OL Organic silts and clays, low plasticity
PI Plasticity index
3.7 SYMBOLS

c Effective soil cohesion (drained cohesion intercept)


[F/L2]

3-18
pt Peat and other highly organic soils Wa weight of pycnometer filled with distilled water [F]

pvo Vertical effective stress, or soil overburden pressure Wb Weight of pycnometer filled with distilled water and
[F/L2] soil sample
r Salt content Wc Weight of container in water content determination [F]
SC Clayey sand Wd Weight of oven-dry soil [F]
S.G. Specific gravity W1 Weight of container and moist soil [F]
SM Silty sand
W2 Weight of container and oven-dried soil [F]
SP Poorly graded sand
w Water content corrected for salt content
St Sensitivity
wL Liquid limit
su Undrained shear strength [F/L2]
wp Plastic limit
suv Vane shear strength [F/L2]
wn Natural water content
SW Well-graded sand
Wet, or bulk, unit weight [F/L3]
t Reference temperature for pycnometer test weights
d Dry unit weight [F/L3]
UC Unconfined compression test
Effective, or drained, friction angle [deg]
UU Unconsolidated-undrained condition for triaxial testing
V Volume [L3]
W wet weight of soil sample [F]

3-19
Chapter 4

SHALLOW FOUNDATIONS AND DEADWEIGHT ANCHORS

4.1 INTRODUCTION Figure 4.1-2. Some shallow foundations for use in soft ocean
soils are constructed with shear keys or skirts that extend below
4.1.1 General
the foundation base to improve the lateral load resistance of the
Shallow foundations and deadweight anchors are typically foundation.
similar structures. Shallow foundations primarily resist
downward-bearing forces, while deadweight anchors resist
upward forces.
The design methods in this chapter are applicable to
shallow foundations and deadweight anchors located in the deep
and shallow ocean areas and follow an iterative or trial-and-error
process. The process starts with an estimation of reasonable or
"convenient" foundation or anchor dimensions and an analysis
is made to predict performance. If the proposed foundation or
anchor is found to be inadequate or to be excessively
overdesigned, the dimensions are changed and the analysis
process is repeated. In some cases the selected shallow Figure 4.1-1. Features of a simple shallow foundation.

foundation or deadweight anchor for the given soil conditions


may be found impractical or too costly. Other foundation types The loading on a shallow foundation will be the
(such as piles) must then be considered. combination of externally applied man-made forces, structure
weight, and environmental loading from current and wave forces
4.1.2 Definitions/Descriptions
(and possibly from wind and earthquake forces). Loadings
4.1.2.1 Shallow Foundations. Generally, shallow applied to the structural system may result in overturning
foundations have a depth of embedment, Df , less than the moments, which create uplift (tensile) loadings as well as
minimum lateral dimension of the foundation, B. The downward (compressive) loadings to a shallow foundation.
horizontal base dimensions of a shallow foundation are Virtually all loadings (except gravitational loadings on a
generally large relative to the foundation thickness. Figure 4.1- horizontal seafloor) develop some load component parallel to
1 is a sketch of a simple foundation. Other types of shallow the seafloor (lateral loads).
foundations are shown in

4-1
Figure 4.1-2. Types of shallow foundations.

The type of loading will determine the methodology used extent as the anchor is dragged) , the behavior of a deadweight
in design. If the foundation loading is compressive, resistance is anchor is practically the same as the behavior of a shallow
derived from the bearing capacity of the soil. If the foundation foundation subjected to an uplift load. The uplift resistance is
is loaded in tension, the uplift resistance will depend on the provided primarily by the net submerged weight of the anchor.
submerged weight of the foundation, the soil friction on the Often, these specially shaped anchors provide little additional
embedded surfaces, and "suction" beneath the foundation. uplift resistance above that from its submerged weight.
Ten types of deadweight anchors are shown in Figure 4.1-
4.1.2.2 Deadweight Anchors. A deadweight anchor
3. The anchors range from relatively sophisticated anchors with
can be any heavy object that is placed on the seafloor. shear keys to simple concrete clumps or clumps of heavy scrap
Deadweight anchors can rest on the seafloor or be partially or materials. Anchors with shear keys provide greater lateral load
even completely buried within it. The primary purpose of a resistance than do those without shear keys. However, the
deadweight anchor is to resist uplift and, often, lateral forces additional capacity of the more sophisticated deadweight anchors
from a mooring line connected to a buoyant object. With the may be offset by increased costs for fabrication and for
exception of a few specially shaped deadweight anchors accommodating special installation procedures.
(designed to dig into the soil to a limited

4-2
Figure 4.1-3. Types and significant characteristics of deadweight anchors.

4.2 DESIGN CONSIDERATIONS


Knowledge of site characteristics (water depth,
4.2.1 General topography, stratigraphy, environmental loading
conditions, potentially hazardous features, among
Information about potential sites and the loading others).
characteristics for the foundation or anchor must be determined. Structure or moored platform characteristics (and the
This information is used for the following purposes: nature and relative importance of the structure or
platform).

To define the appropriate types of geotechnical Soil characteristics (vertical and lateral extent of the
information needed which will allow foundation or soil investigation to determine existing conditions and
anchor design. parameters for design).

To select sites attractive from geotechnical Extent of knowledge and local experience or the
considerations and to avoid sites containing hazards. behavior of similar foundation or anchor types.

To obtain the needed specific soil parameters for Cost and level of risk of failure.
design.

Factors influencing a foundation or dead-weight anchor design The foundation design process is interactive and involves all
include the following items: these considerations in varying degrees. The listed
considerations should influence the choice of safety factors for
design calculations which contain uncertainties.

4-3
4.2.2 Site pore pressures generated by the loading do not have
time to dissipate). For example: mooring line loads
Site characteristics important to shallow foundations and created from storm waves, or earthquake loading of
the structure or of the soil mass itself.
deadweight anchor design include water depth, topographic
features, data on environmental conditions, stratigraphic 4. Slow cyclic loading (i.e., cyclic forces that occur
over a sufficiently long time period so that excess
profiles, sediment characteristics, and potentially hazardous pore pressures generated by the loading have
seafloor features. At the earliest opportunity, attention should sufficient time to dissipate between loads). For
example: mooring line load variations created by
be given to identifying seafloor features that suggest steep tide-related changes in current.
slopes, erosion, existing slumps, or under-consolidated
4.2.5 Geotechnical
sediment. These indicate that excessive settlement,
overstressing, or large foundation movement can occur. If a The design of shallow foundations and deadweight
more suitable site cannot be located, minimization of the effects anchors requires that the following items affecting geotechnical
of these problems should be considered during design. aspects of design be considered:
Foundation instability: bearing capacity failure and other
4.2.3 Structure failures due to uplifting, overturning, horizontal sliding, or
combinations of these.
Characteristics of the structure or moored platform drive
Slow foundation displacements -- primarily excessive
the foundation or anchor design weight, configuration, stiffness, consolidation settlement.
purpose, design life, and cost, as well as other information, are
Installation problems associated with the use of shear keys.
relevant. These dictate the loading conditions and other design
considerations and relevant factors of safety which determine the Recovery problems associated with high resistance to
breakout (Chapter 9).
type and size of the foundation or anchor.
Seafloor slope instability and potential massive soil
4.2.4 Loading movements (Chapter 11).

The following loading conditions should be determined: The soil data required include soil type, index properties,
density, strength under the conditions of the applied loads, and
1. Static long-term loading (i.e., relatively constant deformation characteristics under static and cyclic loading
loads applied for a long time period). For example:
the deadweight loading from the structure and conditions. Table 4.2-1 lists the property values necessary to
foundation. evaluate the loading conditions discussed in Section 4.2.4 for

2. Static short-term loading (i.e., relatively constant cohesive and cohesionless soils.
loads applied for a short time period). For example: At the site where the foundation or deadweight anchor is
a downward force applied during installation to
insure penetration of shear keys, or a load increase to be placed, the depth to which soils data are required is
on the foundation or anchor during infrequent approximately the foundation width or diameter below the
replacement of a subsurface buoy which had been
applying a significant buoyant force. foundation base. The soil characteristics and design parameters
should be obtained through on-site and laboratory testing. For
3. Rapid cyclic loading (i.e., significant repetitive
forces occurring over a relatively short period, so unmanned or other noncritical installations, and for small
that excess structures and low loads, where overdesign is

4-4
Table 4.2-1. Soil Properties Required for Analysis and
Recommended Factors of Safety

not costly, soil information can often be estimated from and for a critical installation, the safety factors should be
available literature (Chapters 2 and 3 include properties for increased by multiplying the table value by 1.5.
typical soil types and engineering property correlations with
4.3 DESIGN METHODOLOGY AND PROCEDURE
more easily obtained index properties). However, this lack of
high quality soils data must be reflected by use of a high factor 4.3.1 General
of safety.
The design procedure for a shallow foundation or
deadweight anchor is an iterative process. A foundation or
4.2.6 Factor of Safety
anchor trial size is selected, and then checked for adequacy.
A safety factor must be applied during the design of
When that size is found to be inadequate , it is modified and
foundations and deadweight anchors to account for uncertainties
checked again until a satisfactory design results. A flow chart
in loading, soils data accuracy, and analytical procedure
for design of shallow foundations and deadweight anchors is
accuracy. Table 4 2-1 lists recommended factors of safety to be
given in Figure 4.3-1. The individual steps in the design
applied to the loading conditions discussed in Section 4.2.4
process are summarized in Table 4.3-1. The design must
when soils properties are accurately known (by field and
consider those factors discussed in Section 4.2 which are
laboratory testing). Where these data are not well known,
applicable.

4-5
Figure 4.3-1. Flow chart for the design of shallow foundations and deadweight anchors.

4-6
Table 4.3-1. Summary of Steps in the Design of Shallow Foundations and Deadweight Anchors

4.3.2 Bearing Capacity 4.3.2.1 Static Short-Term Loading and Cyclic Loading in
COHESIVE SOILS. Under static short-term loadings and all
The bearing capacity of a seafloor soil is dependent upon
the following factors. cyclic loadings, failure on cohesive soils will occur before
excess pore pressure can dissipate. These are, therefore,

Engineering properties of the soil profile undrained failures and the soils properties used in the design are
Type and size of foundation undrained properties [Table 4.2-1].
Depth of embedment
Load direction The maximum downward vertical load, Qu, that a
Inclination of the ground surface foundation or cohesive soils can support under undrained
conditions is calculated by:
The maximum bearing capacity for the trial size
foundation is calculated as explained in Sections 4.3.2.1 Qu = A (su Nc Kc + b Df K1) (4-1)
through 4.3.2.6. This capacity is compared to the sum of all
forces acting normal to the seafloor surface, with an appropriate where: A' = effective base area of foundation depending on
the load eccentricity [L2] (Section 4.3.2.5)
safety factor applied to these normal forces.

4-7
su = undrained shear strength of cohesive soil- K = correction factor dependent on load
averaged over the distance B below the inclination, foundation shape,
foundation base [F/L2] inclination of embedment depth, inclination of ground
foundation
(Section 4.3.2.6). For foundations
Nc = bearing capacity factor; for undrained failure Nc
= 5.14 subject to essentially vertical loading,
b = buoyant unit weight of soil - above the K can be assumed to be 0.6 for a square
foundation base [F/L3] or circular foundation
Df = depth of embedment of foundation[L]
Kc , K q = correction factors that account for load
inclination, foundation shape, embedment
depth, inclination of foundation base, and
inclination of ground (Section 4.3.2.6). For a
nearly level surface, a nearly square or round
shape, and a vertical load, Kc = 1.2 and K q =
1.0.

For the simple case of a vertical load applied


concentrically on a square or circular foundation resting on the
seafloor surface, where both the foundation base and seafloor are
horizontal, Equation 4-1 reduces to:

Qu = (6.17) A su (4-2)

where A is the foundation base area.

4.3.2.2 Static Long-Term Loading in COHESIVE

SOILS. The static long-term loading condition on cohesive

soils exists after the excess pore water pressures have dissipated.
This is a drained failure condition, and the soils properties used
in the design are drained properties [Table 4.2-1].
The maximum downward vertical load of a shallow
Figure 4.3-2. Bearing capacity factors Nq, Nc , and N as a
foundation or deadweight anchor for the drained case is function of the soil friction angle (from Ref 4-1).
calculated by:

Qu = A' [c N c Kc + b Df Nq Kq 4.3.2.3 Static Short- and Long-Term Loading in

+ (1/2) b B N K ] (4-3) COHESIONLESS SOILS. Bearing capacity failure on

cohesionless soils occurs under drained conditions.


where c = effective soil cohesion as determined Cohesionless soils have a sufficiently high permeability to
from a Mohr-Coulomb envelope [F/L 2]
B = minimum foundation base dimension allow water drainage and rapid dissipation of excess pore
[L] pressures. Therefore, both the short- and long-term designs use
Nc , N q, N = bearing capacity factors that are
drained soil properties [Table 4.2-1].
functions of , obtained from Figure
4.3-2
= effective soil friction angle [deg]

4-8
The maximum downward vertical load of a shallow at the center of the reduced area. As shown in Figure 4.3-3, for
foundation or deadweight anchor on cohesionless soils is a rectangular base area, eccentricity can occur with respect to
calculated by: either, or both, axis. The altered effective dimensions of the
footing are:
Qu = A' [ b Df Nq Kq + (1/2) b B' N K] (4-4)

L' = L - 2 e1
where B' is the effective foundation minimum base dimension
(4-6)
(Section 4.3.2.5) [L].
B' = B - 2 e2
For the simple case of a vertical load applied
concentrically on a square or circular foundation resting at the The effective area is:
seafloor surface, where both the foundation base and seafloor are A' = B' L' (4-7)
horizontal, Equation 4-4 reduces to:
4.3.2.6 Correction Factors. The correction factors
Qu = 0.3 A b B N (4-5)
Kc , K q, and K , used in Equations 4-1, 4-3, and 4-4, each

where A is the foundation base area. represent a subgroup of factors, which account for the
following:

4.3.2.4 Rapid Cyclic Loading in COHESIONLESS


Item Designation
SOILS. Cyclic loading applied to a foundation on a
Inclination of resultant load vector from i
cohesionless soil may occur without sufficient time for
vertical
dissipation of generated excess pore pressures. The bearing Shape of foundation s
Depth of embedment d
capacity of a foundation or deadweight anchor under such rapid
Inclination of foundation base b
cyclic loading conditions may be lower or higher than the static Inclination of ground surface g
bearing capacity. The bearing capacity under rapid cyclic loading
The correction factors Kc , K q, and K are evaluated from:
can be determined by using Equation 4-4 with bearing capacity
factors Nq and N selected using an undrained friction angle u Kc = i c s c dc b c g c

determined from special cyclic undrained laboratory tests (which Kq = i q s q dq b q g q (4-8)

must be run if this condition is anticipated). K = i s d b g


For slower cyclic loading, where the excess pore
Seafloor foundations and deadweight anchors are often
pressures have time to dissipate, the drained analysis of Section
subjected to a large lateral load component arising from wave
4.3.2.3 is used with Nq and N obtained using the drained
and current loadings and, occasionally, from wind loading on a
friction angle [Table 4.2-1].
surface float which is connected by a mooring line. This large
lateral load component, combined with the gravity load
4.3.2.5 Effective Base Dimensions. If the loading is
component of a structure or deadweight, forms a resultant load
applied eccentrically or includes a moment, a reduced foundation
of substantial inclination to the vertical. This inclination of the
base-to-soil contact area is used to determine bearing capacity.
resultant load causes a change in the form of the bearing
Eccentric vertical loading and applied moments are handled by
capacity failure surface, permitting failure to take place at a
defining an eccentricity, e, which is the distance from the center
lower load. Subgroup correction factors, which account for
of the foundation to the center of the reduced area. The
inclination of the resultant load, are:
equivalent or resultant vertical load acts

4-9
Figure 4.3-3. Area reduction factors for eccentrically loaded foundations (after Ref 4-1).

and

where = angle between the line of action of F h and the


long axis of the foundation

For the case of an undrained hearing capacity failure in


(4-9)
cohesive soil ( = 0), the correction factor ic is obtained from:

where: F h = horizontal component of design load [F] (4-10)


F v = vertical downward component of all loads [F]

The bearing capacity factors of Figure 4.3-2 are derived


for the two-dimensional

4-10
failure case or for an infinitely long strip foundation. Shear keys are often incorporated in the foundation base
Corrections to the calculated two-dimensional bearing capacity design [Figure 4.1-1]. This is done to increase lateral load
prediction for the more likely rectangular and circular bearing capacity by forcing the failure surface, the surface on which the
areas are calculated from: foundation will slide, deeper into the seafloor where stronger
soils can resist higher lateral loads.
sc = 1 + (B/L) (Nq/Nc )
Three possible failure modes that can occur for shallow
sq = 1 + (B'/L') tan (4-11) foundations fitted with shear keys are shown in Figure 4.3-4.
s = 1 - 0.4 (B'/L') Generally, the shear keys should be designed sufficiently close
together to force the sliding failure to occur at the base of the
For circular foundations loaned without eccentricity, the
keys as shown in Figure 4.3-4(a). Procedures for evaluating this
shape correction factors are calculated by setting B'/L = 1.
base sliding resistance are described in the following two
Normally, conditions are such that the remaining factors
sections. Methods for analyzing the passive wedge failure mode
can be assumed equal to 1.0.
shown in Figure 4.3-4(c) are addressed in the shear key design
The correction factor for depth of embedment, subgroup
section (Section 4 3.6). The deep passive failure mode shown in
d, is very sensitive to soil disturbance along the sides of the
Figure 4.3-4(b) does not often occur and is not detailed in this
embedded base. Because installation procedures normally disturb
chapter.
the soil zone, it is wise to discount entirely the beneficial effect
Maximum lateral load capacity is calculated as explained
of overburden shear strength. This is done by setting dq = d =1.
in Sections 4.3.3.1 through 4.3.3.3. This capacity is then
The correction factors for inclination of the foundation base compared to the sum of all forces driving the foundation in the
subgroup b, and for the inclination of the seafloor, subgroup g, downslope direction with an appropriate safety factor applied to
can be set equal to 1 where the foundation is placed nearly level these driving forces
on a near-horizontal seafloor. Because nonsloping sites are
usually sought for foundations, and the foundation is usually
placed in a near-horizontal orientation, then: bc = b q = b = 1,

and gc = g q = g = 1.

For those instances where foundations will be installed


deeply in the seafloor at a severe inclination or on a slope,
relationships for the correction factor subgroups d, b, and g can
be found in References 4-1 and 4-2

4.3.3 Lateral Load Capacity

The lateral load that acts on the foundation or deadweight


anchor can result from: downslope gravity force caused by a
sloping seafloor, current drag on a foundation and structure,
nonvertical mooring line loading, or storm-wave and earthquake
loadings.
Figure 4.3-4. Possible failure modes when sliding
resistance is exceeded (after Ref 4.3).

4-11
Table 4.3-2. Coefficient of Friction Between Cohesionless Soils
and Some Marine Construction Materials

4.3.3.1 Static Short-Term Loading and All Cyclic Wbst = buoyant weight of bottom-supported structure
[F]
Loading in COHESIVE SOILS. Static short-term loading and

all cyclic lateral loading of foundations on cohesive soils are F ve = design environmentalloading and mooring line
loading in the vertical direction (upward is
treated as undrained failure problems [Table 4.2-1].
positive) [F]
The maximum lateral load capacity (parallel to the
F h = design environmental loading and mooring line
seafloor), Qul, for a foundation on cohesive soil under undrained
loading in the horizontal direction (downslope is
conditions is calculated by: assumed positive) [F]

b = buoyant unit weight of soil [F/L3]


Qul = s uz A + 2 sua Df B (4-12)

= seafloor slope angle [deg]


where: suz = undrained shear strength of the soil at depth zs
[F/L2] R p = passive soil resistance on leading edge of base
sua = average undrained shear strength between the [F]
seafloor and depth Df [F/L 2]
The coefficient of friction depends on soil type and on
4.3.3.2 Static Short- and Long-Term Loading in base material type and material roughness. Table 4.3-2 lists
COHESIONLESS SOILS. For cohesionless soils, lateral load coefficient of friction values for typical construction materials

failure is a drained soil failure, and the maximum lateral load and marine cohesionless soils. Where there is no other

capacity in sliding is calculated by: guidance, the value of can be estimated as follows:

Qul = [(Wbf Wbst + b A Df - Fve ) cos = tan ( - 5 deg) for a rough steel or concrete base without
shear keys
- F h sin ] + Rp (4-13)
= tan for a base with shear keys
where: = coefficient of friction between foundation base
where: = soil friction angle ( = u for the undrained case
and soil or between soil and soil when shear
keys cause this type of sliding failure and = for the drained case)

Wbf = buoyant weight of foundation [F] Where shear keys are present or the foundation is
embedded deeply, a wedge of soil in passive failure develops in
front of the leading
4-12
foundation edge and provides resistance to sliding. For some soil is based on a drained soil failure analysis similar to that
foundations this passive wedge can contribute around 10% of described for cohesionless soil with an additional contribution
the total lateral resistance. However, because sediment from c, the effective soil cohesion. The maximum lateral load
comprising this passive wedge may be removed by current capacity (parallel to the seafloor in the downslope direction) for
scour or by animal burrowing activity (see Chapter 10), the a foundation on a cohesive soil in a drained condition is
contribution of the passive wedge to sliding resistance is often calculated by:
omitted.
In the design of a shallow foundation on cohesionless
soils, the weight of the foundation is often increased to raise the (4-17)
maximum lateral load capacity. To maintain stability against
The minimum foundation buoyant weight for this case is
sliding, a factor of safety is applied to Qul to account for
determined by:
uncertainties in soil data or failure mechanism. The factor of
safety in this case is determined by:

(4-18)

. . . . . . . . (4-14) This long-term drained cohesive soil condition may


control the design in a very few cases where the soils are
The minimum foundation buoyant weight for this case is heavily overconsolidated and very stiff. Normally, the short-
derived from Equations 4-13 and 4-14 (assuming Rp = 0) as: term (undrained) case will yield a lower capacity and, therefore,
will control lateral load aspects of foundation and deadweight
anchor design.
. . . . . . . . (4-15)
4.3.4 Overturning Resistance
Note: Where skirts but no shear keys are used, the sliding
A shallow foundation, if subjected to excessive lateral and
will more likely occur along the foundation base and not at the
uplift loading, may rotate about a point near the leading edge of
depth of the skirt. Therefore, the last term in the buoyant
its base. Figure 4.3-5 illustrates the forces considered in
weight of soil in the skirts/keys ( b A D f ) should not be used.
analyzing the capacity of a shallow foundation or deadweight
For the special case of = 0 where the seafloor is level,
anchor to resist overturning on a horizontal surface. Where there
Equation 4-15 becomes:
is a seafloor slope, the forces are resolved into components
(4-16)
parallel and normal to the seafloor surface.
Stability against overturning is achieved in the design by
Table 4.2-1 lists recommended Fs values to be used in insuring the resisting or stabilizing moment, Ms, is greater
Equations 4-15 and 4-16. than the overturning moment, Mo. To insure full contact

between the base and the supporting soil, the foundation or


4.3.3.3 Static Long-Term Loading in COHESIVE
anchor should be designed so that
SOILS. The long-term lateral load capacity of a foundation or

anchor on cohesive

4-13
Figure 4.3-5. Forces considered in the overturning analysis.

the resultant normal soil reaction, Rs, acts within the middle It is important, when this equation is used to check a

one-third of the base. Where Rs crosses the shear key line then foundation's stability, to use appropriate values for the forces

becomes the assumed point of foundation rotation. The involved. The highest values possible for Wbf and Wbst would

maximum stabilizing moment is then calculated as: not result in an accurate calculation for the maximum value of
Ms. If lower values for Wbf and Wbst are possible at the same
(4-19)
time the maximum values for Fve and Fh occur, they should be
The soil within the shear keys is assumed to separate used.
from the base during overturning and, therefore, does not For preliminary sizing, the minimum width of the
contribute to the stabilizing moment. foundation can be calculated from:
The overturning moment is calculated as:
(4-22)
Mo = F h (H' + zs) (4-20)
To minimize the potential for overturning, the distance H

where: H = vertical distance from Fh to the base of the shear + z s (the moment arm of the lateral load component should be
key kept as small as possible. This is most easily done by
zs = depth of the shear key tip below the foundation
minimizing the foundation base height. It is recommended that
base
H be limited to 0.25B where possible.
To maintain stability, Ms M o. Therefore:
4.3.5 Shear Key Design
(4-21)
The depth of shear keys or perimeter skirts is usually
limited by the net downward force

4-14
available to drive the keys. A penetration resistance calculation In cohesionless soils, a zs = 0.5B is appropriate for
should show that full skirt penetration is assured under only the internal shear keys. The shear key around the edge of the
submerged weight of the foundation. foundation, called a perimeter skirt, serves the added function of
When shear keys or perimeter skirts are used, venting preventing undermining of the foundation by scour and animal
holes are required in the base to allow the water and soft burrowing. For this reason, the perimeter skirt will normally be
surficial soils trapped by them to escape. Sharpening the deeper and a depth of 0.1B is recommended.
leading edge of keys will also aid penetration. The actual The number of shear keys required in each direction on

foundation placement should be smooth and continuous to cohesionless soils is again computed using Equation 4-23. (A

minimize disturbance to the seafloor soil and the possible minimum recommended shear key spacing is 2 zs.) The passive
resistance developed per shear key is computed for cohesionless
resulting creation of an eccentric foundation orientation.
soils as:
4.3.5.1 Depth and Spacing of Shear Keys. In
(4-25)
cohesive soils, the shear key depth, zs, and spacing of shear
Kp is the coefficient of passive lateral earth pressure and is
keys are dictated by the need to force a failure to occur along the
base of the shear keys as shown in Figure 4.3-4(a). The computed by:

recommended maximum depth of shear keys on cohesive soils Kp = tan2 [45 deg + (/2)] (4-26)

is 0.1B [Ref 4-4). This depth may have to be reduced if full


4.3.5.2 Penetration of Shear Keys. The force required
penetration cannot be achieved (Section 4.3.5.2). Other steps to
to ensure full penetration of shear keys and perimeter skirts, Qe ,
reach a satisfactory design for adequate lateral load capacity
can be calculated using methods presented in Chapter 8. In
include making the base larger and making it heavier -- if
making the analysis the highest expected values of soil strength
necessary to insure shear key penetration.
should be used to be on the conservative side. (In using
The number of shear keys, n, required in each direction is
Equations 8-4 and 8-5, undisturbed soil shear strengths should
computed by comparing the design load parallel to the seafloor
be used for cohesive soils; the adhesion factor should be set
to the passive resistance developed per key. n is computed by:
equal to 1.0, and the friction angle between shear key and sand
(4-23) should be set equal to the soil friction angle.)
After Qe has been calculated, it is compared to the sum of
where F hp is the resultant of applied loads in the downslope Wbst + W f , the forces driving penetration, to check if it is
direction [F]. (The minimum recommended shear key spacing is smaller than the driving forces. If it is larger, then the skirts
zs). must be made smaller or fewer, or the foundation weight will
For cohesive soils, the resistance developed by the shear need to be increased. (If the foundation weight is increased, a
key is calculated as: check must be made to see that this increase does not create a
bearing capacity problem.)
(4-24)

4.3.6 Foundation Settlement


where sua is the average undrained soil shear strength between
The bearing capacity of the surficial seafloor soil is
the foundation base and the tip of the key measured in F/L2.
assumed to be sufficient to

4-15
support the foundation. If the bearing capacity is not sufficient, will gradually settle as the excess pore pressure which developed
then the foundation will immediately penetrate into the seafloor in response to the foundation loading dissipates and the soil
until its weight and the supported structure weight are balanced consolidates. In cohesionless soils there is no significant
by the soil resistance. A method for predicting this penetration amount of consolidation settlement as pore pressure dissipates
is described in Chapter 8. almost immediately. In evaluating this time-dependent
Foundation settlements due to elastic deformations and consolidation settlement for cohesive soils, the soil to a depth
soil consolidation may still pose a significant problem, even in of 2B below the foundation base should be considered. This is
the absence of a bearing capacity failure, because such the depth where the effective stress change after consolidation is
settlements are rarely uniform. The occurrence of differential about 10% of the change immediately below the foundation
settlement is greatly aggravated by eccentric loading. The base (for a square foundation) [Figure 4.3-6].
resulting tilting could impair structure function.
Settlement of a deadweight anchor is, in contrast, not
normally considered a problem because the holding capacity is
unaffected, or is sometimes increased by such embedment. In
some cases, however, even excessive embedment of a
deadweight anchor is not desirable because it limits the ability
to inspect and maintain the mooring line connections to the
anchor.
Settlement is the summation of initial and consolidation
settlements, which are discussed in the following sections.

4.3.6.1 Initial Settlement. Initial settlement is the

instantaneous response of the soil to the foundation loading and


results primarily from elastic soil deformations. The initial
vertical settlement of a shallow foundation, i, can be

calculated, after Ref 4-5, by:

(4-27) Figure 4.3-6. Soil stress increase beneath a rectangular


foundation.
where: = Poisson's ratio of' the soil (assume = 0.35 for
cohesionless soils and 0.45 for cohesive soils) Only consolidation of the cohesive soil layers within the
R = radius of the base; for square footing, assume R =
B/2 [L] depth 2B needs to be calculated. The calculations are made by
F v = total downward vertical load on the foundation breaking the distance 2B into several layers. The total
base [F]
E = Young's modulus of the soil, a property which consolidation settlement is the sum of the settlement of the
must be determined by soils testing [F/L2] individual layers and is obtained from:
G = (E/2) (1 + ) and is called the elastic shear
modulus of the soil [F/L2]

4.3.6.2 Consolidation Settlement. In cohesive soils,

after installation, a foundation

4-16
where: vmax = maximum lowering velocity [fps]
(4-28) Wbv = submerged weight of installation [lb]
Av = vertical projected area of the package [ft2]

where: c = consolidation settlement [L] As the installation approaches the seafloor, the maximum
n = number of layers within distance 2B
Hi = thickness of cohesive soil layer i [L] lowering velocity must be reduced by at least a factor of four to
C ci = compression index of cohesive soil layer i prevent too hard an impact with the bottom. A hard impact may
pi = added effective vertical stress at midpoint of
cohesive soil layer i [F/L2] result in bearing failure, excessive tilting, or an instability
poi = initial effective overburden stress at midpoint of failure. It is usually desirable to approach the seafloor as slowly
soil layer i [F/L2]
eoi = initial void ratio of layer i as possible and not to lift the foundation off the seafloor once
the initial touchdown has occurred.
An accurate computation of consolidation settlement The recovery of a shallow foundation also requires careful
requires considerable knowledge of soils properties. For consideration to insure that adequate lifting force for breakout is
instance, high quality soil samples and at least one laboratory available. Prediction of breakout forces and a discussion of
consolidation test per layer are required to determine C ci for techniques used to minimize the breakout force are presented in
these computations. Other soil properties must be measured for Chapter 9.
each layer to perform settlement analysis. A rough estimate for
these settlements car be made by using some estimated
4.4 EXAMPLE PROBLEMS
properties. Values for Cc may be determined empirically as
4.4.1 Problem 1 - Simple Foundation on Cohesive Soil
outlined in Chapter 3. The initial void ratio, eo, can be

determined from soil specific gravity and water content values 4.4.1.1 Problem Statement: Determine the dimensions of

by the relationship eo = Gs w. The value of pi can be a concrete and steel square foundation, essentially a deadweight

estimated from Figure 4.3-6. p oi is determined by using the soil anchor, with shear keys to resist given environmental loadings

buoyant unit weight. It is suggested that a maximum soil layer at a deep ocean site. The seafloor is mildly sloping and is

thickness of 0.25B be used. composed of a cohesive soil which has well-established


properties. Use a factor of safety of 1.5 and minimize the base
4.3.7 Installation and Removal horizontal size.
Data: The foundation is to be placed where the seafloor
The installation of shallow foundations should be planned
slopes at 5 degrees and must resist loads from a mooring line
so that the foundation can be properly set down at the intended
that may reach 20,000 pounds in uplift and in any horizontal
site without excessive disturbance to the supporting soil.
direction. Figure 4.4-1 shows a sketch for this problem and for
The maximum lowering rate of the installation line
Problem 2. The seafloor is a cohesive silty clay material whose
should not exceed the free-fall velocity of the package to avoid
properties have been determined by laboratory tests on cored
unstable lowering and possible entangling of the lowering line
samples. The data for undrained shear
A rough estimate for maximum lowering rate is:

(4-29)

4-17
strength (su), buoyant unit weight ( b), sensitivity (St), and the solution are shown below. They follow the method presented

drained parameters of cohesion (c) and friction angle () are for the design of a square foundation to resist the loads under the

shown to the right of Figure 4.4-1. existing conditions. The forces acting on the foundation are
shown in Figure 4.4-2.
4.4.1.2 Problem Solution: The analytical and

computational procedures for the problems

Figure 4.4-1. Foundation sketch for example problems 1 and 2.

Problem 4.4-1

4-18
Problem 4.4-1

4-19
Problem 4.4-1

4-20
Problem 4.4-1

4-21
Problem 4.4-1

4-22
Problem 4.4-1

Figure 4.4-2. Forces considered in the overturning


analysis for example problem 1.

4-23
4.4.2 Problem 2 - Simple Foundation on Cohesionless Soil from surface waves. The seafloor is composed of a medium-
dense well-graded sand (cohesionless) with the following
4.4.2.1 Problem Statement: Determine the dimension of
properties having been determined from in-situ or laboratory
a simple concrete square foundation, essentially a deadweight
testing: buoyant unit weight of 60 pcf and friction angle of 35
anchor, with a perimeter skirt to resist given environmental
degrees. (This is the same problem as 4.2.1 except that the soil
loading, at a shallow ocean site. The seafloor is mildly sloping
is different and skirts without interior shear keys are to be used.)
and is composed of a cohesionless soil which has well-
established properties. Use a factor of safety of 1.5. 4.4.2.2 Problem Solution: The analytical and
Data: The foundation is to be placed where the seafloor computational procedures for the problem's solution are shown
slopes at 5 degrees and must resist loads from a mooring line below. They follow the method presented for the design of a
that may reach 20,000 pounds in uplift and in any horizontal square foundation to resist the loads under the existing
direction [see Figure 4.4-1]. The area is sheltered and no cyclic conditions. The forces acting on the foundation are shown in
effects are expected Figure 4.4-3.

Problem 4.4-2

4-24
Problem 4.4-2

4-25
Problem 4.4-2

4-26
Problem 4.4-2

Figure 4.4-3. Forces considered in the overturning


analysis for example problem 2.

4-27
4.5 REFERENCES E Young's modulus of soil [F/L2]

4-1. Planning, designing, and constructing fixed offshore e, e1, e 2 Eccentricity [L]

platforms, American Petroleum Institute, API RP 2A. eo ,e oi Initial void ratio


Washington, D.C., Jan 1982. Fh Applied horizontal load

4-2. H.F. Winterkorn and H-Y. Fang. Foundation engineering F hp Resultant of applied loads in the downslope
direction [F]
handbook. New York, N.Y., Van Nostrand Reinhold, 1975.
Fs Factor of safety
4-3. A.G. Young et al. "Geotechnical considerations in Fv Vertical load component (downward is positive)
foundation design of offshore gravity structures," in Proceedings [F]

of the 1975 Offshore Technology Conference, Houston, Tex., F ve Design environmental loading in vertical direction
(upward is positive) [F]
1975. (OTC 2371)
G Elastic shear modulus of soil [F/L2]
4-4. P.J. Valent et al. OTEC single anchor holding capacities
Gs Specific gravity of the soil grains
in typical deep sea sediments, Civil Engineering Laboratory,
g Correction factor for inclination of ground surface
Technical Note N-1463. Port Hueneme, Calif., Dec 1976.
H Height or thickness of the foundation base [L]
4-5. T.W. Lambe and R.W. Whitman. Soil mechanics. New
H' Vertical distance from point of application of Fh
York, N.Y., John Wiley, 1969. to the point or rotation (assumed to be at the
foundation base) [L]

4.6 SYMBOLS Hi Thickness of individual cohesive soil layer [L]

i Correction factor for inclination of resultant load


A Foundation or anchor base area [L2]
Kc , K q, K Bearing capacity correction factors
A Effective foundation or anchor base area [L2]
Kp Coefficient of passive earth pressure
Av Vertical projection area of installation [L2]
L Length of foundation or anchor base [L]
B Minimum foundation or anchor base dimension
(usually called the foundation width) [L] L Effective length of foundation or anchor base [L]

B Effective minimum foundation or anchor base Mo Overturning moment [FL]


dimension [L]
Ms (Stabilizing) resisting moment [FL]
b Correction factor for inclination of foundation or
anchor base m Exponential term used in Equations 4-8 and 4-10

C c , C ci Compression index Nc , N q, N Bearing capacity factors

c Effective cohesion intercept as determined by n Number of shear keys oriented in one direction
Mohr-Coulomb envelope [F/L 2]
o Assumed point of rotation
Df Embedment depth of foundation or anchor (depth
of shear key tip below the seafloor) [L] poi Initial effective overburden stress at midpoint of
ith soil layer [F/L2]
d Correction factor for depth of base embedment
pv Effective vertical stress [F/L2]

p Applied uniform vertical stress from foundation at


base level

4-28
p Increase in vertical effective stress at depth z Wbst Effective or buoyant weight of structure supported
[Figure 4.3-6] by a foundation [F]

pi Added effective vertical stress at midpoint of ith Wbv Submerged weight of installation being lowered
soil layer [F/L 2]
z Depth to the bottom of the foundation base in soil
Qe Embedment force necessary to fully penetrate [L]
shear keys [F]
zs Depth of shear key tip below the foundation base
Qu Ultimate bearing capacity resistance [F] [L]

Qul Ultimate lateral resistance of the foundation or Slope of seafloor [deg]


deadweight [F]
b Buoyant unit weight of soil [F/L3]
R Radius of circular foundation base [L]
bc Buoyant unit weight of concrete [F/L3]
Rp Passive soil resistance [F]
bf Buoyant unit weight of foundation [F/L 3]
Rs Normal soil reaction
c Total consolidation settlement [L]
St Sensitivity of cohesive soil
i Initial settlement [L]
s Correction factor for shape of base
Angle between the ins of action of Fh and the
su Undrained shear strength of cohesive soil [F/L2] long axis of the foundation

sua Average undrained shear strength between Coefficient of friction between soil and foundation
foundation base and tip of shear key [F/L2] or between soil and soil when shear keys cause
this type of sliding failure
suz Undrained shear strength of the soil at shear key
tip [F/L2] Poisson's ratio of soil

vmax Maximum allowable lowering velocity during Soil friction angle [deg]
installation [L/T]
Effective friction angle for drained analysis [deg]
w Water content
u Undrained friction angle of cohesionless soil [deg]
Wb Effective or weight of the soil trapped within the
shear keys [F]

Wbf Effective or buoyant weight of the foundation or


deadweight anchor [F]

4-29
Chapter 5

PILE FOUNDATIONS AND ANCHORS

5.1 INTRODUCTION 5.2.2 Mooring Line Connections

Piles are deep foundation elements installed by driving or Table 5.2-2 shows four common types of mooring line
by drilling-and-grouting. High installation costs dictate that connections. The selection of a type of mooring line connection
piles be used only where surface foundation or anchor elements, should be based on the relative importance of the factors listed
such as shallow foundations, deadweight anchors, or drag or and the cost. Connection points, especially, must be sturdy
plate anchors, cannot supply the support required. Use of piles enough to withstand installation stresses as well as service
as foundations or anchors becomes less frequent as water depth loads.
increases due to associated rapidly increasing installation costs
5.2.3 Modifications for Increasing Lateral Load Capacity
This chapter describes pile types, designs, installation
considerations, and presents a simplified design procedure The lateral load capacity of a pile anchor can be increased
in several ways (as shown in Table 5.2-3): by lowering the
5.2 PILE DESCRIPTIONS
attachment point along the pile length, by lowering the pile
5.2.1 Pile Types head beneath the soil surface into stronger soils, or by attaching
fins or shear collars near the pile head to increase the lateral
Steel is the most common pile material used offshore.
bearing area. The expected increase in pile capacity must be
Prestressed concrete and wood are used in nearshore and harbor
weighed against increases in cost for fabrication and for the
applications but are rarely used in deeper water because of
higher complexity of the installation procedure.
construction, handling, and splicing difficulties with long piles.
For use as foundations and anchors in deep water, pipe sections
5.3 DESIGN PROCEDURE FOR SIMPLE PILES IN SOIL
and H-piles are the most commonly used pile types. A number SEAFLOORS
of specially designed piles are also in use as anchors [Ref 5-1].
5.3.1 General
These special piles are designed to increase lateral or uplift
capacity of the anchorage. Table 5.2-1 lists features of the more Pile foundations or anchors may be subjected to one or

common pipe and H-piles and of several types of specialty more of the following loads:

piles.
Axial downward loads (compression)
Axial uplift loads (tension)
Lateral loads
Bending moments

5-1
Table 5.2-1. Pile Types

a Special anchor pile.

5-2
Table 5.2-2. Mooring Line Connections

Table 5.2-3. Techniques to Improve Pile Lateral Load Capacity

5-3
For simple pipe and H-piles, axial forces are resisted by
soil friction developed along the pile shaft and by bearing on
the pile tip (for downward loads). Lateral forces and moments
are resisted primarily by the pile shaft bearing on the near-
surface soils. For a foundation pile, design is normally
controlled by downward axial and lateral loads. However,
significant moment and uplift loads may also be present. For an
anchor pile, design is normally controlled by uplift and lateral
loads. A significant moment may be present, depending on the
point of application of the lateral load.
A simplified procedure for the design of uniform cross-
section piles embedded in a non-layered seafloor consisting of
sand, clay, or calcareous soils is presented in Section 5.3.2. The
objective of the procedure is to determine if the pile length,
width, and stiffness are capable of resisting applied moments
and lateral and axial forces at the seafloor without excessive
movement and without exceeding the allowable stresses for the
pile material. Design is a trial-and-error procedure, where a pile
is selected and then is evaluated for its ability to resist those
forces and movements under the existing soil conditions. A
check is then made to determine if the allowable pile material
stresses are exceeded. In the procedure, the pile is assumed to be
a beam on an elastic foundation with an elastic modulus that
increases linearly with depth. Further description of this type of
analysis is found in Reference 5-2.
The procedure applies to piles loaded at the pile head.
(Modifications to this procedure to account for other conditions
are discussed in Section 5.3.7.)

In the following text sections, steps for pile design are


denoted in parentheses to the left of where the procedure step is
presented. Figure 5.3-1 summarizes the design procedure. Note
that in this design a factor of safety is applied by increasing
expected loads to the design loads before the pile is designed.

Figure 5.3-1 Flow chart for the pile design procedure.

5-4
5.3.2 Soil Properties If site-specific soil data are not available, it may be
possible to extrapolate from other property data using geologic
(STEP 1) Determine required soil engineering properties. These
and geophysical information from similar nearby areas. In the
properties should be site-specific (i.e., based on in-situ or
absence of any test-determined data, estimates of soil properties
laboratory tests, or both). The soil properties required for design
may be made based on the geologic and depositional
are:
environment and geophysical data. Chapter 2 provides

For cohesionless soil (sands and silts), guidelines for estimating engineering properties and identifies

= drained (effective) friction angle [deg] some engineering properties of major sediment types.
b = buoyant unit weight [F/L3] Additional soil properties, which are important to pile design,
Dr = relative density [%]
are given in Tables 5.3-1 through 5.3-3 for cohesionless soils
Cohesive soils (clays), and cohesive soils cased on soil density or degree of
su = buoyant unit weight [F/L2] consolidation.
b = relative density [F/L3]
Where soil properties vary significantly with depth,

Calcareous soils, average properties in the uppermost four pile diameters are used

Carbonate content, degree of cementation, and degree of for lateral load analysis (STEPS 4 through 11) and average
crushing (in addition to the properties required for properties over the pile length for axial load analysis (STEPS
cohesionless soils)
12 through 18).

Table 5.3-1. Properties of Cohesionless Soil


Useful in Pile Design

Standard
Type Penetration Dr b
Blow Count, N (deg) () (pcf)
Very loose to loose <10 28-30 0-35 45-55

Medium dense 10-30 30-36 35-65 55-65

Dense 30-50 35-42 65-85 60-70

Very dense 50+ 40-45 85-100 60-70

Table 5.3-2. Properties of Cohesive Soil


Useful in Pile Design

Type Su 50 b
(psi) () (pcf)
Underconsolidated clays 0.35-0.10 2 20-25

Normally consolidated soils 1.0 + 0.0033z 2-1 25-50


at depth z, in.

Overconsolidated soils based


on consistency:
medium stiff 3.5-7.0 1.0
stiff 7.0-14 0.7 50-65
very stiff 14-28 0.5
hard over 28 0.4

5-5
Table 5.3-3. Properties of Calcareous Soil
Useful in Pile Design

Characteristic Property
Carbonate content low, 0 to 30%
medium, 30 to 45%
high, 45+%

Degree of cementation uncemented or lightly cemented


well cemented

Degree of crushing crushes easily


resistant to crushing

5.3.3 Pile Design Loads where: P h = design horizontal load at the foundation pile [F]
P t = design vertical uplift load at the foundation pile
(STEP 2) Determine the maximum load or load combinations [F]
P c = design vertical compression load at the
at the seafloor surface. Actual loads are determined from an
foundation pile [F]
analysis of what is attached to the pile. A safety factor, Fs, is
In unusual cases, a moment may be applied to the foundation
applied to the actual loads to determine those loads used in the
pile. This actual moment should similarly be multiplied by the
design. If the necessary soils data are accurately known (from
factor of safety to yield a design moment.
in-situ testing or laboratory testing of core samples), a safety
factor of 1.5 to 2.0 is recommended. If soil properties are not Ma = F s x moment applied to the pile head
accurately known, a higher safety factor of 2.0 to 3.0 should be
where: Ma = design applied moment [LF]
used. Within these recommended ranges, higher values should
be used where the installation is more critical and where the soil (STEP 3) Calculate load at a submerged anchor pile head. For
properties are more questionable. anchor piles to be driven below the seafloor surface, the
For anchor piles, the design loads are: mooring line angle at the pile is not the same as that angle at
the seafloor, due to soil bearing resistance against the mooring
Th = Fs x horizontal component of mooring line
tension at the seafloor line. The actual angle at the pile becomes higher and the force
exerted on the pile becomes more of an axial load. To account
Tt = Fs x vertical component of mooring line tension
at the seafloor for this, horizontal and vertical design loads determined in
STEP 2 must be corrected. The general effect of this correction
where: Th = design horizontal load at the anchor pile [F]
Tt = design vertical load at the anchor pile [F] is to decrease the horizontal load a pile must resist and increase
the vertical or axial load. The force corrections include several
For foundation piles, the design loads are: simplifying conservative assumptions and are made as follows:
P h = Fs x horizontal component of load at the pile The correction to the horizontal force is the soil force that
P t = Fs x vertical uplift load at the pile is exerted on the mooring line in a horizontal direction:
P c = Fs x vertical compression load at the pile

For cohesionless soils,

F cb = zc 2 db b Nq (5-1a)

5-6
For cohesive soil, (STEP 5) Select deflection criteria. For anchor piles, a
F cb = (11.0) s u db zc (5-1b) suggested maximum lateral deflection criterion is ymax/D =
where: F cb = horizontal force exerted on the mooring line by 10%, where ymax is the maximum allowable lateral deflection.
the soil [F]
zc = depth of pile connection below seafloor [L] For foundation piles, a more rigid deflection criterion (a lower
db = characteristic mooring line size [L] (for chain, value of ymax/D) is used justified by structural requirements.
use 3 x chain size; for wire, use wire diameter)
Nq = bearing capacity factor, from Table 5.3-4
(STEP 6) Determine coefficient of subgrade reaction. The

The corrected anchor pile design loads are: coefficient of subgrade soil reaction, nh, is determined for the
Th = Th - Fcb F s (5-2) selected value of ymax/D. For cohesionless soils, this is

Tt = ( Tt2 + 2 Th F cb F s - F cb 2 F s2 )1/2 obtained from Figure 5.3-2 and Dr , the soil relative density. For
. . . . . . . . . (5-3) cohesive soils, first Figure 5.3-3 and su are used to determine a

where Th and Tt are the values for Th and Tt, corrected for the value for the coefficient K1. Then n h is determined from the

effects of a pile head being a depth zc below the seafloor. equation on that figure. For calcareous soils, Figure 5.3-2 and a
value of Dr = 35% are used.

Table 5.3-4. Bearing Capacity


Factors for Chain
Lateral Force in Sand
(after Ref 5-3)

Soil Friction _
Angle, Nq
(deg)
20 3

25 5

30 8

35 12

40 22

45 36

5.3.4 Lateral Load Analysis

(STEP 4) Select a trial pile size. The trial and-error pile

selection process begins with a pile selection--usually made on


the basis of availability. This determines pile diameter, D; wall Figure 5.3-2. Design values for nh for cohesionless soils (after
thickness, tw, and stiffness, EI. Ref 5-4).

5-7
where a is the distance of the pile load attachment point above
the seafloor surface for foundation piles [L] (a 0, as
foundation piles will not be driven below the surface).

(Step 11) Compare pile capacity to required design lateral load

capacity.

If Ph(calc) = P h (or P h' or T h) or is slightly higher, then the


trial pile is adequate for resisting lateral loads.

If Ph(calc) >> P h (or P h' or T h) then the pile is overdesigned.


One or more of the following changes are made
and computations are repeated on the new trial
pile size.

(a) Reduce pile stiffness by reducing diameter or


thickness (repeat calculations from STEP 4)
(b) Reduce pile length, unless already very short (repeat
Figure 5.3-3. Design values for nh for cohesive soils. calculations from STEP 9)

If Ph(calc) < P h (or P h' or T h) then the pile is underdesigned.


One or more of the following changes are made
and computations are repeated on the new trial pile
size. Note: This is done until P h (calc) is only
(STEP 7) Determine pile-soil stiffness. The pile-soil relative slightly greater than P h.
stiffness, T, is computed by:

(5-4)

(STEP 8) Select pile length. A pile length, L p, is assumed.

(A length of L p = 3T is suggested as a minimum.)

(STEP 9) Determine deflection coefficients. The maximum

value of the depth coefficient zmax = L p/T is computed. Then

deflection coefficients Ay and By at the ground surface are

obtained from Figure 5.3-4.

(STEP 10) Calculate lateral load capacity. The lateral load

capacity of the trial pile, Ph(calc), for the value of y max selected

in STEP 5 is computed.

(5-5)
Figure 5.3-4. Deflection coefficients Ay and B y at the ground
surface.

5-8
(a) Increase pile stiffness by increasing diameter and/or soils , the average unit skin frictional resistance, f s (in uplift and
thickness (repeat calculations from STEP 4)
(b) Increase pile length, unless already very long (repeat compression), is calculated from:
calculations from STEP 9)
(c) Increase design depth of pile head (repeat calculations fs = k p vo tan ( - 5 deg) (5-7)
from STEP 3)

where k is 0.7 for compression and 0.5 for uplift.


5.3.5 Axial Load Analysis
Table 5.3-5 presents limiting or maximum values for fs.
(STEP 12) Calculate average effective overburden pressure. The
It must be checked to ensure that calculated values of fs do not
average effective overburden soil pressure (pvo) at the pile
exceed those limiting values. For piles driven into calcareous
midpoint is computed for:
soils, the tables limiting values shown should be used unless

(5-6) higher values are justified by on-site testing.


For cohesive soils , first determine if the soil is normally
If the pile is not fully buried, find pvo at the midpoint of the
consolidated or overconsolidated. To do this, an average value
buried length. If the density changes, average the densities along for su/pvo over the pile length is computed. If su/pvo 0.4, the
the pile.
soil is considered normally consolidated (NC); if su/pvo > 0.4,

the soil is considered overconsolidated (OC).


(STEP 13) Calculate skin frictional resistance per unit length of

pile. For cohesionless

Table 5.3-5. Recommended Limiting Values for Unit Skin Friction


and End Bearing for Cohesionless Soils

5-9
For NC soils, the average unit skin friction resistance, fs, where: Ap = gross end area of the closed pile [L2]
qp = unit soil bearing capacity of that pile tip [F/L2]
is equal to:

For cohesionless soils,


fs = p vo [ 0.468 - 0.052 ln (Lp/2.0) ] (5-8a)

qp = p vo (tip) N q (5-11)
If Equation 5-8a exceeds su then

For cohesive soils,


fs = s u
qp = 9 s u (tip) (5-12)
where Lp is in feet.
where: pvo (tip) = effective vertical stress at pile tip [F/L2]
For OC soils: Nq = bearing capacity factor from Table 5.3-5
su (tip) = soil undrained shear strength at pile tip
fs = [ 0.458 - 0.155 ln (su/pvo) ] s u (5-8b) [F/L2]

If su/pvo exceeds 2.0, then Piles that are not closed-ended will develop a soil plug
when installed. (When this soil plug is removed and replaced
fs = 0.351 su
with concrete, the pile is considered closed-ended.) The soil plug
will limit the value of Qp that can develop to the force required
(STEP 14) Compute uplift capacity and compare to design load.
to push a soil plug up into a thin-walled pipe. This limiting
For both anchor and foundation piles, the frictional resistance of
value is approximately equal to the frictional capacity of the
the pile, Qs, must exceed the design load for uplift, P t. Pile
pile, previously computed as Qs. Therefore, for open-ended
frictional resistance is:
piles, Q p is equal to Q s.
Qs = As fs (5-9)
(STEP 16) Compare pile capacity in compression with required
where As is the surface area of the pile below the seafloor [L2].
capacity. The pile capacity compression, Qc , is computed, and

If Qs P t, then the design is adequate for resisting uplift this is compared with the design load under vertical
forces. compression, P c .
If Qs < P t, then the trial pile length is increased. Although
increasing the pile diameter is another approach which Qc = Qs + Qp (5-13)
will increase Qs, increasing pile length is usually
preferred. If Qc P c , the pile is adequate in compression.

(STEP 15) Compute soil bearing capacity for foundation pile If Qc < P c , then the pile is not adequate in compression. The
tip . For foundation piles, resistance to compressive loading pile length must be increased (repeat the calculations
from STEP 14).
comes from frictional resistance along the pile and from
resistance to tip or end penetration. For closed-ended piles, the If using an open-ended pile and Qp was significantly limited by

soil bearing capacity for the pile tip, Qp, is computed as the value of Qs, it may be of significant benefit to plug the pile

follows: end. Recompute Qc for a closed-ended trial pile and again check

for adequacy in compression.


Qp = Ap qp (5-10)

5-10
5.3.6 Steel Stress Analysis Figure 5.3-5; then computing the maximum moment, Mmax, at

a point along the pile by;


(STEP 17) Calculate maximum moment. The maximum

moment (Mmax) in the pile is determined by combining any Mmax = Am P h T + M a Bm (5-14)


applied (design) bending moments, Ma , and moments created by

the design horizontal load. The latter is calculated by It may be necessary to determine Mmax at several locations (z)

recomputing T (STEP 7) and Lp/T (STEP 9) for the current along the pile in order to find the maximum value of Mmax.

trial pile; selecting the influence coefficients Am and B m from This is done by entering Figure 5.3-5 with the several values of

the appropriate curves on z/T.

Figure 5.3-5. Influence values for a pile with applied lateral load or moment (after Ref 5-2).

5-11
(STEP 18) Calculate maximum steel stress. Maximum stress side of an anchor pile at a distance of more than five pile

in the pile under tension (f maxt) and compression (fmaxc ) is diameters from the head, the simplified approach presented in

calculated by: Section 5.3.4 for lateral load is not adequate. For the anchor
pile with such a buried side connection, a lateral load analysis
fmaxt = - P t /A ps - M max /S (5-15) must be made by other means, such as with the assistance of a
fmaxc = P c /A ps + M max /S (5-16) computer program [Ref 5-5]. For the same lateral load, pile
deflection and bending moments are reduced significantly as the
where: Aps = cross-sectional area of the pile [L2] mooring line connection point is lowered from the pile top to a
S = section modulus of the pile [L 3]
point midway down the pile. The mooring line will also
Note: The terms Aps and S and allowable maximum approach the pile at a decreasing angle as the connection point
stress in tension and compression from below are available is lowered, thus greatly lowering the lateral force and raising the
from steel design manuals or manufacturers literature containing uplift force.
data on these pile shapes. Where the mooring line chain is almost parallel to the
pile, the presence of chain alongside the pile and the influence
(STEP 19) Compare calculated and allowable steel stress. The
of repeated mooring loads may also reduce the soil strength and,
values for fmaxt and fmaxc are compared with the allowable steel
therefore, reduce soil resistance above the connection point.
stress in tension and compression for the pile being used. For
most common structural shapes, fa , the allowable maximum 5.3.7.2 Piles With Enlarged Cross Section. Pile size

stress in tension and in bending (tension and compression), is at may be increased near the seafloor surface to increase resistance

least 0.6 Fy, where F y is the minimum yield point of the steel to lateral loads and bending moments. Analysis of the response

being used. of piles with variable cross section to lateral loads is much
more complex and is usually done with the aid of computer
If fmaxt and fmaxc allowable, then the pile is adequate. programs [Ref 5-6, 5-7]. in general, a large increase in pile
diameter over a small depth is more efficient in reducing
If fmaxt or f maxc > allowable, then the pile is inadequate. A new
trial pile size with a larger section modulus (larger deflections at the seafloor than is a small increase in diameter
tw or D) is selected (repeat the calculations from
STEP 18). over a larger depth. When enlarged pile sections at the pile head
are used, the length of the enlarged section should be limited to
It is also possible to reinforce the pile over that length where
three times the larger diameter.
high moments exist. This will result in a larger section
modulus and may be done rather than using a larger pile. While 5.3.7.3 Special Seafloor Conditions. A number of
this is often done because it is cost-effective, it is beyond the
seafloor conditions can be considered hazardous to the placement
scope of this handbook.
or functioning of piles in the offshore environment. Some of
these are listed below.
5.3.7 Special Cases
Steeply sloping seafloors and accompanying downslope
5.3.7.1 Load Applied Below the Anchor Pile Head.
soil movements can cause additional lateral pile forces in the
When the mooring line is connected to the downslope direction. In addition, steep sloping makes the area
more

5-12
subject to instability problems such as might be initiated by the moving soils. For important structures in areas where such
seismic activity or wave forces. problems can be anticipated, the design should be based on a
The presence of rocks, cobbles, or cemented cones can thorough evaluation of such hazards and their influence on the

make installation of driven piles difficult. Drilled and grouted structure.

piles may be the best method of pile installation in areas where


5.4 DESIGN OF PILE ANCHORS IN ROCK SEAFLOORS
these occur.
Scour of near-surface soils can occur and can be An approach to pile anchor design in rock is presented in

accentuated in the vicinity of piles. Scour extent depends on the this section. A detailed procedure for pile anchor design in rock

velocity of seafloor currents, the type of soil, and the size and is not presented because this cannot be done simply - primarily

configuration of pile groups. Generally, granular soils are more because of the difficulty in characterizing the material failure

susceptible to scour than cohesive soils. For granular soils, the mode.

upper 5 to 10 feet may be subject to scour. Removal of this Rock failure modes are illustrated in Figure 5.4-1. Failure

surface material can significantly affect pile behavior, and the of a pile anchor in rock may occur in lateral bearing due to rock

possibility of this occurring should be taken into account crushing under lateral loads. It may also fail in uplift because of

during pile design. Scour is discussed in more detail in Chapter grout-to-rock bonding or because of a rock-mass failure in

10. fractured material. Uplift failure can also occur due to the pile's

Earthquake-related hazards should be assessed in loosening and the loss of resistance as a result of repeated lateral
loads.
seismically active areas. Earthquake motions may cause partial
The mode of failure is difficult to establish or predict for
loss of strength or complete liquefaction in loose granular soil
a specific location. The strength of a core sample may be
zones, essentially removing soil support developed in these
misleading
zones. Additionally, a weakened zone may slide downslope and
overload the pile due to high lateral force applied by

Figure 5.4-1. Failure modes for pile anchors in a rock seafloor.

5-13
when applied to the prediction of pile anchor holding capacity develop all support from the soil layer. The pile design should
in jointed, bedded, faulted, or weathered rock masses. Gross rock be made with the soil analysis procedures presented in Section
characteristics will likely govern pile behavior, and these gross 5.3.
characteristics must be thoroughly evaluated at each pile If zs/T < 3.0, then the pile can be conservatively designed
location. as if the soil were not present. That is, the pile is considered
cantilevered out of the rock surface and resists all forces without
5.4.1 Lateral Capacity
assistance from the soil layer. If this is done, a check must be
In a rock or hard cemented soil seafloor, a soil cover may made to determine if the compressive strength of the rock, sc , is
be present above the rock, or the rock/cemented zone may be exceeded by stresses from the applied lateral load. That is, a
underlain by soil. For layered soil-rock sites, available check should be made to ensure that:
computer programs should be used [Ref 5-6, 5-7] to account for
the complexities introduced by these nonuniform conditions. (5-17)

where: P h = lateral force applied to the pile head [F]


5.4.1.1 Soil Overlying Rock. For soil overlying
D = width or diameter of the pile [L]
rock, pile capacity can be roughly estimated by the following Le = effective length of the pile bearing on the rock
strata equal to the smaller of: (1) the rock layer
procedure. The relative depth of the rock, zs/T, is determined, thickness, z r , or (2) a depth interval equal to the
where zs is the thickness of the soil over the rock and T is the diameter of the pile, D
sc = rock compressive strength (see Table 5.4-1 for
pile relative stiffness. If zs/T 3.0, the pile can likely be typical values of s c ) [F/L 2]

designed to

Table 5.4-1.Rock Properties (after Ref 5-8)

5-14
5.4.1.2 Rock Layer Overlying Soil . For the case of should be limited to 27 psi. The grout-to-rock bond strength

rock or other hard layer overlying soil, if the rock layer may vary from 0.3 to 1.0 times the rock shear strength,

thickness, z r , is less than 0.2T, where T = (EI/nh)0.2 and nh is depending on cleanliness of the drilled hole, type of rock, and

the coefficient of soil reaction of the underlying soil, the grouting procedure.

influence of the rock layer may be ignored and the pile designed
5.4.2.2 Fractured Rock. For fractured rock, anchor
by using the procedure of Section 5.3.2. For rock layer
uplift capacity is determined by the mass of the blocks of rock
thickness, z r , greater than 0.5T, the influence of underlying soil
which move with the anchor and by the frictional force
may be ignored and the pile designed for rock compressive
developed between the attached blocks and adjacent blocks.
strength as under Section 5.4.1.1 For intermediate values of zr ,
Because of the difficulty in estimating the normal forces acting
the designer must judge whether to use the soil or rock
on vertical joints and cracks, this frictional force is normally
procedure; the decision depends upon pile diameter, knowledge
ignored.
of rock layer strength and fracturing, and the layer thickness.
5.5 PILE INSTALLATION
5.4.2 Uplift Capacity
Piles are installed on the foreshore environment by one or
Failure due to uplift load may occur: (1) at the grout-pile a combination of the following methods:
interface, (2) at the grout-rock interface, or (3) along a rock
fracture zone outside the grouted area. In (3), a block of rock Driving
Drilling and grouting
containing the pile is assumed to be lifted free of the Jacking
surrounding rock as shown in Figure 5.4-1(c). The following Jetting
guidelines for design are suggested. 5.5.1 Driven Piles

Piles may be driven by impact hammers operated above


5.4.2.1 Massive, Competent R ock. For massive,
the water surface, by underwater impact hammers, or by
competent rock, uplift capacity will be governed by frictional
vibratory hammers.
forces developed by the grout bonding strength along the pile,
by the grout shear strength itself, or, less likely, by the grout 5.5.1.1 Conventional Hammers. Piles for piers,
bonding strength along the drilled shaft wall. For failures of
harbor structures, bridges, and many offshore structures in
these types, the uplift capacity, Ra , of the rock anchor is given
shallow water are driven from above the water surface with
by: conventional hammers used in pile driving on land. The pile is
Ra = s b Lr C p (5-18) made long enough to extend above the water surface when
driven to its design penetration depth, or alternatively, a pile
where: sb = the lesser of (1) the grout-to-pile bonding
strength, or (2) the shear strength of the grout follower is used [Figure 5.5-1]. The piles are commonly guided
[F/L2] by a template that rests on the seafloor, although floating
Lr = Length of pile embedded in rock [L]
C p = minimum perimeter transmitting the uplift load templates may be used for small, shallow-water installations.
[L] The pile-driving operation is conducted from a carefully moored

Unless higher bond strengths are verified by testing, the work barge that supports the necessary cranes and auxiliary

grout-to-steel bonding strength equipment.

5-15
Figure 5.5.1 Pile installation techniques

The pile hammers used for this construction are scaled-up A number of large hammers have been developed over
versions of land hammers. The types include single-acting that last 10 years that are capable of both above and underwater
steam, compressed air, diesel, and hydraulic hammers. The rated operation. Some of these hammers have rated striking energies
energy of these hammers varies from less than 100,000 ft-lb per of up to 1,700,000 ft-lb.
blow to over 1,500,000 ft-lb per blow. Surface-operated pile
5.5.1.3 Underwater Vibrators. Vibratory pile drivers are
drivers have been used in water depths in excess of 1,000 feet.
The success of the surface-driven method of pile becoming more common in American practice as experience is

emplacement in deep water is dependent upon the presence of gained with their use and as more powerful machines are

the template to act as a guide for the piles. Without the restraint developed. The machines usually use counter-rotating eccentric

offered by the template, most of the driving energy would be weights powered by electric or hydraulic motors to produce the

dissipated by lateral deflection of the pile. For anchor piles vibratory forces. The major depth-limiting factors on present

driven from the surface without lateral restraint, a reasonable systems are the difficulty in handling long lengths of large-

maximum water depth appears to be about 250 feet. diameter, high-pressure hydraulic lines and the large friction
losses in the line. These factors limit the maximum practical
5.5.1.2 Underwater Hammers. Several standard depth of a surface-powered, hydraulic vibratory drive to about
terrestrial pile hammers may be modified for operation 1,000 feet.
underwater. One manufacturer makes a total of 12
5.5.1.4 Selection of Hammers, When selecting a
team/compressed air hammers, with rated energies in air of
8,750 ft-lb to 60,000 ft-lb. These may be operated while hammer for "lighter" offshore tasks, where required hammer

submerged with little loss of efficiency. The modifications energies are less than 150,000 ft-lb, a general role of thumb for

consist primarily of providing exhaust hoses that extend to the hammer selection is:

water surface. Because steam cools too much when the hoses go
For steam/air hammers, the weight of the pile should be no
underwater, compressed air is usually used to operate the more than two times the weight of the hammer ram.
hammers.
For diesel hammers, the weight of the pile should be no
more than four times the weight of the ram.

5-16
For vibratory drivers, the driving amplitude of the driver/pile the drill string, and the pile is placed over the drill string and
system should be between 0.25 and 0.5 inch].
lowered into the hole. Portland cement grout is pumped down
For "heavier" tasks, where hammer energies in excess of the drill string and forced up outside of the pile to fill the
150,000 ft-lb are required, the maximum rated energy of a annular void and bond the pile to the soil. Then, the interior of
steam/iar hammer required to drive a steel pile should be the pile is filled with grout as the drill string is withdrawn
estimated by: [Figure 5.5-2].

Eh = 2,000 Aps (5-19)

where: Eh = maximum rated energy of the hammer [ft-lb]


Aps = area of steel cross section [in.2]

This recommendation is based on the allowable stress in steel


of 12,000 psi under working loads.
For preliminary selection of air/steam hammers, it is
recommended that Equation 5-19 be used to estimate maximum
hammer energy. (For diesel hammers, the maximum energy of
the hammer may be 20 to 35% higher than the value given by
Equation 5-19.) These recommendations assume that the pile
will be driven to the maximum axial capacity. In cases where
lateral load governs pile design and full axial capacity of pile
cross section is not mobilized, hammers with significantly less
rated energy than given by Equation 5-19 may be adequate to
drive the pile to its design penetration depth. For all important
installations, pile driveability must be investigated beyond the
level of Equation 5-19 by the use of wave equation analyses
using available computer programs [Ref 5-9, 5-10] to evaluate
the proper combination of pile-hammer-cushion system.

5.5.2 Drilled and Grouted Piles

By use of drilling and grouting procedures, piles up to 8


feet in diameter have been placed in water depths in excess of
600 feet. The method is essentially identical to that used to set
a casing for an oil well. A hole of somewhat larger diameter
than the pile is drilled to proper depth using rotary drilling
tools. The hole is cleaned out by pumping seawater through

Figure 5.5-2. Drilled and grouted pile.

5-17
5.5.3 Jack-in Piles 5.6 EXAMPLE PROBLEMS

Piles may be pushed or jacked into the seafloor if an 5.6.1 Problem 1--Pile Design in a Cohesive Soil
adequate reaction force can be supplied. To develop a
5.6.1.1 Problem Statement. Design an anchor pile for
satisfactory degree of safety against bearing capacity failure
specified lateral and uplift loads. Also, determine its axial
under design loads, jacking loads equal to two to three times the
capacity in compression and the maximum steel stresses. Can a
design load must be applied. With mobile offshore jack-up
pile which is on-hand be used for this mooring?
platforms, water ballast is used to develop this surcharge.
Data: A closed-ended anchor pile is to be designed for a
However, when the entire installation is submerged, water
floating drydock which will be placed in an area with a soft clay
ballasting of the addition and removal of deadweight ballast is
seafloor. The drydock will be used for repair of nuclear
generally not practical.
submarines. The pile must resist lateral loads up to 50,000
The actual jacking of the piles can be accomplished by a
pounds and axial uplift loads of 25,000 pounds applied by a
number of systems. A rack-and-pinion system may be used,
chain mooring system at the seafloor. At some later date, the
with the rack being an integral part of the pile and running its
pile may be used as a foundation pile and see pure compression
entire length. A chain acted on by a chain jack or cable acted on
loads. A sketch of this pile is shown in Figure 5.6-1. The soils
by a hydraulic cable puller may be used, with the chain or cable
data at the site are fairly well known and are shown in Figure
applying load to the top of the pile. A short-stroke hydraulic
5.6-2. Available pile sizes are very limited. A supply of 24-
jack equipped with a means of gripping the wall of a pile may
inch-diam and 48-inch-diam steel piles (Fy = 36,000 psi) of
also be used.
several wall thicknesses is on-hand.

5.5.4 Jetted Piles


5.6.1.2 Problem Solution: The trial-and-error
Jetting is used to place piles in cohesionless soil. The computational procedures for the problems solution are
piles are pushed or lowered into the soil area, which has been presented below. These follow the method presented in this
greatly weakened by jetting. The jetting action is generally chapter and outlined by the flow chart in Figure 5.3-1.
confined to the inside of a pile or to portions of the outside of
the pile several diameters above its tip. Jetting can also be used
in a form of reverse circulation in which both air and water are
forced down a pipe inside or outside the pile. The air/water
mixture is used to lift the displaced soil materials to the surface
or seafloor.

5-18
Figure 5.6-1. Problem sketch for example problems Figure 5.6-2. Soils data for example problem 1.
1 and 2.

Problem 5.6-1

5-19
Problem 5.6-1

5-20
Problem 5.6-1

5-21
Problem 5.6-1

5-22
Problem 5.6-1

5-23
Problem 5.6-1

5-24
5.6.2 Problem 2--Pile Design in a Cohesionless Soil and axial uplift loads up to 25,000 bounds at the seafloor. (The
pile sketch in Figure 5.6-1 is the same for this problem.) The
5.6.2.1 Problem Statement: Design an anchor pile for
soils data at this site are not well known. The soils have not
specified lateral and uplift loads. Also, determine its axial
been tested but are believed to be a cohesionless silty sand of
capacity in compression, and the maximum steel stresses. Can
medium to dense relative density. The 24- and 48-inch-diam
piles which are on-hand be used for this mooring?
steel piles (F y = 36,000 psi) of several wall thicknesses are also
Data: The anchor pile is to be used for the floating
available for this mooring.
drydock mentioned in Problem 1, but the use of the drydock is
different and the seafloor conditions are different. The drydock is 5.6.2.2 Problem Solution: The trial-and-error
to be moored while not in use, but it again must resist lateral
computational procedures for the problem's solution are
loads up to 50,000 pounds
summarized below. These follow the method presented in this
chapter and outlined by the flow chart in Figure 5.3-1.

Problem 5.6-2

5-25
Problem 5.6-2

5-26
Problem 5.6-2

5-27
Problem 5.6-2

5-28
Problem 5.6-2

5.7 REFERENCES 5-6. L. C. Reese and W. R. Sullivan. Documentation of


computer program COM624, parts I and II, Analysis of stresses
5-1. R.J. Taylor, D. Jones, and R.M. Beard. Handbook for
and deflections for laterally-loaded piles including generation of
uplift-resisting anchors, Civil Engineering Laboratory. Port
p-y curves, Geotechnical Engineering Center, University of
Hueneme, Calif., Sep 1975.
Texas at Austin. Austin, Tex., Aug 1980.
5-2. H. Matlock and L. C. Reese. "Non-dimensional solutions
5-7. H.L. Gill and K.R. Demars. Displacement of laterally
for laterally loaded piles, with soil modulus assumed
loaded structures in nonlinearly responsive soil, Naval Civil
proportionate to depth," in Proceedings of the Eighth Texas
Engineering Laboratory, Technical Report R-670. Port
Conference on Soil Mechanics and Foundation Engineering,
Hueneme, Calif., Apr 1970.
Bureau of Engineering Research, University of Texas at Austin,
Austin, Tex., 1956. (Special Publication No. 29)
5-8. I.W. Farmer. Engineering properties of rocks. London,
England, E. and F.N. Spon Ltd., 1968, p. 57.
5-3. G.G. Meyerhoff and J. I. Adams. "The ultimate uplift
capacity of foundations," Canadian Geotechnical Journal vol 5,
5-9. G.G. Goble and F. Rausch. Wave equation analysis of
no. 4, Nov 1968, pp 225-244.
pile driving - WEAP program, vols 1-4, Federal Highway
Administration, U.S. Department of Transportation, FHWA-IP-
5-4. K. Terzaghi. "Evaluation of coefficients of subgrade
14, Washington, D.C. , 1976.
reaction, " Geotechnique, vol 5, no. 4, 1955, pp 297-326.

5-10. T.J. Hirsch, L. Carr, and L. L. Lowery, Jr. Pile driving


5-5. H. Matlock and T.A. Haliburton. A program for finite-
analysis - Wave equation users manual TTI program, vols 1-4,
element solution of beam-columns on nonlinear supports,
Federal Highway Administration, U.S. Department of
University of Texas at Austin, BMCOL68. Austin, Tex. ,
Transportation, FHWA-IP-13. Washington, D.C., 1976.
1964.

5-29
5.8 SYMBOLS Lp Length of pile [L]

a Distance of pile load attachment point above the Lr Length of pile embedded in rock [L]
seafloor for foundation piles [L]
Ma Design applied moment [LF]
Am Nondimensional moment coefficient
Mmax Maximum bending moment in pile [LF]
Ap Gross end area of pile [L2]
NC Normally consolidated
Aps Cross-sectional area of pile [L2] _
Nq, N q Sand and a special case sand/chain dimensionless
As Surface area of pile below seafloor [L2] bearing capacity factors

Ay Nondimensional deflection coefficient nh Coefficient of subgrade soil reaction [F/L3]

Bm Nondimensional moment coefficient OC Overconsolidated

By Nondimensional deflection coefficient Pc Design vertical compressive load at the foundation


pile [F]
Cp Minimum perimeter transmitting uplift load [L]
Pc' P c [F]
D Pile width or diameter [L]
Ph Design horizontal load at the foundation pile [F]
db Characteristic mooring line size [L]
P h' P h corrected for effects on loading of a pile being
Dr Relative density [%] driven below the seafloor [F]

E Modulus of elasticity of pile [F/L2] P h(calc) Lateral load capacity of the trial pile [F]

Eh Maximum rated energy of hammer [LF] Pt Design vertical uplift load at the foundation pile [F]

EI Stiffness P t' P t corrected for effects on loading of a pile being


driven below the seafloor [F]
F cb Horizontal component of mooring line nearing _
resistance (lateral force exerted by mooring line on pvo Effective vertical stress in soil [F/L 2]
the soil) [F]
Qc Pile capacity in compression [F]
Fs factor of safety
Qp Soil bearing capacity for the pile tip [F]
Fy Minimum yield point of the pile material [F/L2]
Qs Frictional resistance of pile [F]
fa Allowable stress in pile material [F/L2]
qp Unit soil bearing capacity of the pile tip [F/L2]
fmaxc Maximum applied compressive stress in pile [F/L2]
qp(max) Maximum allowable qp from Table 5.3-5 [F/L2]
fmaxt Maximum applied tensile stress in pile [F/L2]
Ra Uplift capacity of rock anchor [F]
fs Unit skin resistance along pile shaft [F/L2]
S Section modulus of pile [L 3]
fs(max) Maximum allowable unit skin friction [F/L2 ]
sb Bond strength of grout to pile, or grout shear
I Moment of inertia of pile [L 4] strength, whichever is less [F/L2]

K1 nhD/su = nondimensional coefficient for clay sc Compressive strength of rock [F/L 2]

k Coefficient of lateral earth pressure su Soil undrained shear strength [F/L2]

Le Effective length of pile bearing on rock [L] su(z) Soil undrained shear strength at depth z [F/L2]

5-30
T Relative stiffness of pile-soil system [L] ymax Maximum allowable lateral pile deflection [L]

Th Design horizontal load at anchor pile [F] z Depth below seafloor [L]

Th' Th corrected for the effects on loading of a pile being zc Depth of pile connection below seafloor [L]
driven below the seafloor
zmax Maximum value of the depth coefficient
Tt Design vertical load at anchor pile [F]
zo Depth of pile head below seafloor [L]
Tt' Tt corrected for the effects on loading of a pile being
driven below the seafloor zr Thickness of rock layer [L]

tw Wall thickness of pile [L] zs Thickness of soil over rock [L]

y Lateral pile deflection [L] b Buoyant unit weight of soil [F/L3]

Drained (effective) friction angle

Friction angle for sand against pile

5-31
Chapter 6

DIRECT-EMBEDMENT ANCHORS

6.1 INTRODUCTION (2) their resistance to nonhorizontal loading, which permits


short mooring line scopes and tighter moorings.
6.1.1 Purpose

This chapter details procedures and considerations for use 6.2 DIRECT-EMBEDMENT ANCHOR TYPES AND
SIZES
of direct-embedment anchors. It deals primarily with plate-type
anchors, specifically the propellant-driven type used by the 6.2.1 Propellant-Driven Anchor
Navy. Flexibility in use of these anchors is restricted by the
Propellant-driven embedment anchors have been
very few sizes available. The "design" is really selection of an
developed by the Navy for both deep and shallow water uses.
adequately sized gun and plate anchor, and the calculation of the
This anchor type uses a fluke which is fired into the seafloor by
expected capacity for a set of expected soil and loading
a gun barrel. It has the significant advantage of instantaneous
conditions.
embedment on seafloor contact. In deep water uses, these
propellant-emdedded anchors do not need a support stand during
6.1.2 Function
installation, can be installed without difficulty on moderate
Most direct-embedment anchors are installed by inserting slopes, and do not require the deployment vessel to maintain
the anchor member vertically into the seafloor and then station for long periods as with vibratory or impact-driven
expanding or re-orienting this anchor member to increase its systems. The process of propellant-emoedded anchor installation
pullout resistance. Direct-embedment anchor types are described is illustrated in Figure 6.2-1. After firing on touchdown (step
in Section 6.2. 1), the anchor penetrates and comes to rest deep within the soil
mass (step 2). When a load is applied to the mooring line (step
6.1.3 Features
3), the anchor keys or rotates into a position of maximum
The five major types of direct-embedment anchors are: resistance to further displacement (step 4).
propellant-driven, vibratory driven, impact-driven, jetted-in, and Several propellant-embedded anchor systems are described
augered-in anchors. in Table 6.2-1. More detailed information, including
Features of direct-embedment anchors are summarized in photographs and drawings describing the anchor systems listed
Chapter 1. The more significant advantages of the direct- in Table 6.2-1, can be found in References 6-2
embedment anchors are: (1) their very high holding-capacity-to-
weight ratio (100:1 for propellant-embedded types) and

6-1
Figure 6.2-1. Installation sequence for a propellant embedment anchor.

through 6-5. The Navy systems are further described in Table Figure 6.2-3 illustrates typical coral and rock flukes.
6.2-2 for a preliminary design of mooring systems. The fluke Both A/S Raufoss Ammunisjonsfabrikker of Norway and
dimensions listed represent an optimized balance of gun- Ametek Offshore of Santa Barbara, Calif., have been licensed to
deliverable energy and anchor fluke mass, end area, and bearing market the Navy systems commercially. Raufoss has installed
area. Figure 6.2-2 illustrates components of a propellant- 100K anchors as elements of a permanent mooring system in
embedded anchor system with sand and clay flukes. the North Sea.

6-2
Table 6.2-1. Propellant-Driven Embedment Anchors
for Ocean Use [after Ref 6-1]

6-3
Table 6.2-2. Parameters for Navy Propellant-Embedded Anchors
(Data After Ref 6-1 through 6-6)

6-4
Figure 6.2-2. NCEL 10K propellant-embedment anchor showing
sand and clay flukes.

Figure 6.2-3. Coral and rock flukes for NCEL propellant-embedment anchor systems (from Ref 6-23).

6-5
6.2.2 Vibratory-Driven Anchors required for proper fluke embedment, typically 15 to 30
minutes. The handling platform must maintain its position
Vibratory-driven embedment anchors have been
directly over the anchor during installation, usually requiring
developed, but have not seen wide use. The Naval Civil
use of a multi-point mooring system.
Engineering Laboratory (NCEL) developed and tested an
experimental , battery-powered variation of this anchor for water 6.2.3 Impact-Driven Anchors
depths to 6,000 feet [Ref 6-2, 6-7]. Further use of this device
Various types of impact- or hammer-driven anchors,
was dropped as propellant-embedment anchors were developed.
which expand or rotate to achieve high capacity, have been
Another system was developed by Ocean Science and
developed as land anchors [Ref 6-8]. Two hammer-driven types
Engineering, Rockville, Md. It is powered from the surface and
have been used for seafloor anchorage: the Navy umbrella pile
has a maximum operating water depth of 500 feet [Ref 6-2].
anchor [Ref 6-2, 6-9] and the Menard rotating plate anchor [Ref
This depth limitation is dictated by power losses in the lines to
6-2] [Figure 6.2-4]. Both are limited in water depth by the
the hydraulic motors that drive a counter-rotating mass vibrator.
available pile-driving and pile-follower equipment and both
Holding capacities for both systems are approximately 40 kips
require expensive pile-driving equipment for installation.
in sand and 25 kips in soft, normally consolidated clays to
Holding capacities range from 100 kips in clay to 300 kips in
clayey silts. A disadvantage of the vibratory-driven systems is
sand.
the prolonged time period

Figure 6.2-4. Impact-driven anchors.

6-6
6.2.4 Jetted-In Anchors and cobble soils, penetration by jetting will likely be very slow
and uneconomical.
Jetted-in anchors are buried through water jet disturbance
of the soil. This allows the anchor to be placed or to be more 6.2.5 Auger Anchors
easily pushed into the seafloor. These systems range in holding
Auger anchors are screw-shaped shafts installed under
capacity from small diver-installed anchors of 1 to 10 kips [Ref
high torque and some vertical load. They have been used for
6-12] to larger systems of 150 kips and greater [Ref 6-11, 6-13]
anchoring pipe-lines on the seafloor [Ref 6-2]. Normally,
[Figure 6.2-5]. Jetted anchors function primarily in sands that
anchors are installed in pairs, one on each side of the pipeline at
are easily liquefied by the water jets. After the jetting action is
the same time to provide a torque reaction for each other.
stopped, the liquefied sands return to a more dense condition
Operational water depth is limited primarily by difficulties in
over the anchor plate. Penetration in clays is not as easily
supplying power, usually by hydraulic hoses, at the seafloor.
accomplished, and the resulting backfill is much weaker than
The experience limit is approximately 500 feet.
the undisturbed material. In hard clays and shell

Figure 6.2-5. Jetted-in anchors (after Ret 6-11).

6-7
6.3 SITE DATA NEEDED over the wide area by a 3.5-kHz acoustic profiling system. In
areas of large relief, such as areas of outcropping rock, erosion,
6.3.1 General
or slumping features, a deep tow profiling system may be
Figure 6.3-1 presents an outline of steps in the selection necessary to obtain an accurate picture of seafloor topography
of an adequate direct-embedment anchor size for a given set of and distribution of sediment infill between the relief features.
loading and site conditions. The first steps are the collection of The results of the acoustic surveys should be used during
adequate site data. the exploratory program to position coring locations so that:
Use of a direct-embedment anchor requires knowledge of (1) the core samples will provide data representative of the
sediment properties over the possible depth of penetration of the probable anchor locations and (2) the potential for lost time and
anchor fluke. Maximum penetration depth, keying distance, and damage to coring equipment is minimized. Core samples and
holding capacity under the specific conditions of applied in-situ tests (see Chapters 2 and 3) are used primarily to identify
loadings must be calculated from these properties. With soil type accurately and to determine soil engineering property
adequate site data, a direct-embedment anchor sufficient for data. They also provide a necessary control and calibration for
existing soil and loading conditions can be selected. interpretation of the acoustic data.
Some of the specific site survey information required for
6.3.2 Preliminary Penetration Estimate
selecting propellant-embedded anchors whether critical or
In many cases some preliminary data are available on the noncritical mooring uses are the same: seafloor material type,
general seafloor environment to permit identification of the layer thickness, and depth to rock. Where consequences of a
probable sediment type (see Chapter 2 for sediment data single anchor failure are not severe (noncritical applications), a
sources). Knowledge of a probable sediment type allows lower level of data on sediment type and thicknesses may be
estimation of probable anchor penetration depth (see Chapter 8 sufficient for anchor selection. Geotechnical properties may
and estimates for penetration of specific propellant-embedded then be estimated from soil property profiles (Chapter 2) to
anchors in Table 6.2-2). These estimates of soil type and make a rough estimate of capacity in lieu of accurate site-
penetration depth guide selection of site survey and sampling specific data.
equipment. For propellant anchors in critical moorings, where the
consequences of a single mooring failure are severe, data from
6.3.3 Topography, Strata Thicknesses, Type
in-situ tests and good quality soil cores are required.
In nonhomogeneous soils, strata material type and
thickness must be determined. The seafloor material type and 6.3.4 Engineering Properties

approximate consistency must be known in order to select the To make accurate predictions of fluke penetration and
appropriate anchor fluke type (i.e., clay, sand, coral, or rock holding capacity, several soil engineering properties must be
fluke). The thicknesses of the sediment strata must be known to known. These are sensitivity, natural water content, bulk
ensure that the anchor fluke has sufficient sediment thickness to density, grain size, carbonate content, origin and history,
develop the design capacity. These data are best obtained over a permeability, and shear strength (drained and undrained). Where
wide area through acoustic subbottom profiling and coring (see dynamic loans are significant, other, more specialized tests on
Chapter 2). Seafloor topography, surficial sediment layer core samples may be necessary. At least one
thicknesses, and depth to rock are best measured

6-8
Figure 6-3.1 Flow chart for predicting the holding capacity of
a direct-embedment anchor.

6-9
good quality sediment core is required to determine these below this sampled depth may be assumed to be similar, and
properties. In many cases, only one is required. For example, the soil property profile would be extended to the necessary
for a deep water mooring on a large abyssal plain, sediment depth. Expendable penetrometers can provide additional data
variability over a 4- to 5-mile diameter of the mooring area may where longer coring is not possible and can thus extend the
not be great. One core should suffice. However, where the depth of a survey below sampling depth.
sediment consistency or type may vary across the mooring site,
6.3.5 Complicating or Hazardous Conditions
cores should be obtained at each anchor location. Soil cores
should be obtained over the full estimated penetration depth of Direct-embedment anchor systems function well in a wide
the fluke. Table 5.2-2 suggests what this required depth is, range of seafloor conditions. They can be adapted to function
based on experience with different Navy anchors used under well where drag anchors and pile anchors are inefficient and even
various soil conditions. The range is considerable. nonfunctional. Extreme soil conditions such as very hard or
Long, heavy corers and specialized handling systems are very soft seafloors complicate the use of direct-embedment
required to penetrate and recover cores from the deeper soil anchors, making special efforts necessary during site survey,
depths. However, smaller corers, which achieve penetrations of positioning, design, installation and proof-loading. Table 6.3-1
10 feet in sands and 30 feet in clays, are often used to obtain the lists some complicating conditions and describes their impact
sediment and define the upper portion of the geotechnical on direct-embedment anchor performance. The approach to most
property profile. With the guidance of geophysical data, of these is to avoid the area
sediments

Table 6.3-1. Conditions Complicating or Hazardous to Direct-Embedment Anchor Use

6-10
by relocating the plate anchor so the problem is not acute or to Prediction of penetration for anchor flukes installed by other
than propellant embedment is not presented in this
select a different anchor system -- one less sensitive to the
handbook.
problem (e.g., using gravity anchors when on rock).
A technique for predicting penetration of vibratory-driven
flukes is given in Reference 6-2
6.3.6 Specialized Survey Tools
Methods for calculating hammer energies required to
Two specialized site survey tools have been developed penetrate plate anchors with followers (considered a type of
that can support the siting, design, and installation of the pile) are given in Reference 6-16 and Chapter 5.
Navy's propellant-embedded anchors. One, the expendable The penetration depth of jetted-in anchors is limited
Doppler penetrometer [Ref 6-14], indirectly measures the primarily by the occurrence of hard layers in the soil profile,
which stop or impede the jet erosion process.
undrained shear strength of the soil (see Chapter 2, Section
2.2.5). Although strength is not determined to the level of The power required for penetration of screw anchors requires
an evaluation of the fluke characteristics and soil frictional
accuracy as that from in-situ measurement or coring, the resistance in order to estimate required installation power.
method requires very little on-site ship time and becomes more
6.4.2 Keying Prediction
cost-effective as water depth increases.
The second device, the Pinger probe, is a 3.5-kHz battery- As is illustrated in Figure 6 2-1, the anchor fluke moves
powered sound source. It can be used with a propellant- upward a certain distance as it rotates into a horizontal
embedded anchor system to assist positioning the anchor in orientation -- the position of maximum holding capacity. This
complex seafloor conditions [Ref 6-15]. The Pinger probe is keying distance, zk, is a function of fluke geometry, soil type,
attached to the anchor system's lowering line about 100 feet soil sensitivity, and duration of time between penetration and
above the anchor. The probe provides a high quality, real-time keying. However, experience has shown that the Navys
image of subbottom seabed stratification The installation vessel propellant-embedded anchor flukes keys in about 2.0 fluke
then maneuvers the anchor system over a seabed profile until a lengths when embedded in cohesive soil and in about 1.5 fluke
location is found that maximizes chances for successful installa- lengths in cohesionless soils. Thus, given the initial fluke
tion. penetration, the keyed depth for a Navy anchor fluke is
estimated as follows:
6.4 FLUKE PENETRATION AND KEYING
In cohesive soil:
6.4.1 Penetration Prediction
z = zp - 2 H (6-1)
Table 6.2-2 includes estimates of propellant-embedded
In cohesionless soil,
fluke penetrations for generalized classifications of surficial
z = zp - 1.5 H (6-2)
soils. These estimates can be refined for established profiles of
undrained soil strength and other anchor fluke dimensions, where: z = depth of the fluke after keying [L]
zp = maximum penetration depth of fluke [L]
mass, and initial velocity by applying the predictive techniques
H = length of the fluke [L]
of Chapter 8. The following are pertinent to a discussion of
penetration prediction: Although no recommendation is made for altering the
above estimate for z, it is believed that keying distance may be
longer in highly sensitive soils but may be shorter as more
time elapses between penetration and keying.

6-11
6.5 STATIC HOLDING CAPACITY 6.5.2 Deep and Shallow Anchor Failure

6.5.1 Loading Conditions Holding capacity depends on the soil failure mode, which
is dependent on the anchor plate embedment depth and on the
Static loading is a relatively constant load maintained for
soil type and strength. Anchor failure is characterized as being
a long period of time. This is in contrast to dynamic loading,
either shallow or deep (illustrated in Figure 6.5-1). A shallow
where the short length of time the load is applied significantly
failure occurs when the seabed surface is displaced by the
affects anchor holding capacity. In reality, loadings on seafloor
upward motion of the anchor plate and the soil failure surface
anchors are rarely completely static but often have impulse or
continues up to the seabed surface. A deep failure is present
repetitive components. However, at dynamic load levels below
when the anchor plate is sufficiently deep within the seabed so
certain limits, the anchor-soil response will be as if the system
that the soil failure surface accompanying movement of the
were statically loaded. For cyclic repetitive loading such as that
anchor does not reach the seabed surface. The transition from
caused by wave effects, the loading can be approximated as
shallow to deep behavior has been found to be a function of
static if the change in load is less than 5% of the static load
relative embedment depth (the ratio of embedment depth to
component. For impulse loadings such as that caused by a short
anchor minimum dimension, z/B) and soil strength. It occurs
tug on a mooring line, the loading can be treated as static when
over a range of z/B values in cohesive soil from 2 to 5 and in
the load development occurs over more than 10 minutes
cohesionless soil from 2 to 10 [Ref 6-18].
duration in clays and more than 10 seconds in sands and coarse
silts [Ref 6-11, 6-17].

Figure 6.5-1. Failure modes for shallow and deep embedded plate anchors.

6-12
6.5.3 Short-Term Capacity in Cohesive Soils B = plate minimum dimension, usually width [L]
L = plate maximum dimension, usually length [L]
The short-term condition exists when the anchor-caused
_
soil failure is governed by a soils undrained shear strength. The value of the holding capacity factor, N cs , is obtained
Failure in the short-term condition occurs before significant
from Figure 6.5-2. It is a function of the soil's undrained shear
drainage of pore water can take place. It occurs immediately, or
strength and the relative embedment depth. For the deep failure
within a few minutes, of full load application. Plate anchor
mode, Ncs = 15. The disturbance correction is explained in
short-term holding capacity under static loading conditions in
Section 6.5.6.
cohesive soils, Fst, is:
In some instances, Ncs from Figure 6.5-2 may be too
_
F st = A s u h N cs [0.84 + 0.16(B/L) ] (6-3) high for the existing conditions. If drainage vents or tubes
allow water to flow rapidly to the underside of the plate anchor,
where: A = projected maximum fluke area perpendicular to
direction of pullout [L2] then the suction formed on the underside of the plate will be
su = soil undrained shear strength [F/L2]
h = correction factor for soil disturbance due to relieved. If this happens, the value of Ncs should be reduced to
_ penetration and keying that for the long-term holding capacity factor, Nc , from Figure
Ncs = short-term holding capacity factor in cohesive
soil 6.5-3 [Ref 6-18].

Figure 6.5-2. Short-term holding capacity factors for cohesive soil where full suction develops beneath the plate.

6-13
Figure 6.5-3. Long-term holding capacity factors and short-term no-suction factors for cohesive soils.

6.5.4 Long-Term Capacity in Cohesive Soils and maximum F lt = F st (the applied long-term load cannot

exceed the short-term load without initiating a failure).


The long-term condition exists when a static load is
The factor Nc is obtained from Figure 6.5-3 and the factor
applied to the anchor over a time sufficiently long to allow
near-complete dissipation of excess pore water pressures. This Nq from Figure 6.5-4.

time duration may be a day for silts, a week for silty clays, and For very soft underconsolidated sediments, such as delta
considerably longer for clays. In cohesive soils, the long-term muds, the shear failure mode may be different than with
holding capacity is governed by the effective soil drained normally consolidated sediments. Section 3.5.2 of Chapter 3
strength parameters: the drained cohesion intercept, c; and the gives a brief discussion of underconsolidated soils. This
drained friction angle, . The long-term static holding capacity, condition can result in a lower holding capacity than would be
F lt, is: calculated by Equation 6-4. The reduced holding capacities in

_ _ _ these very soft soils can be conservatively predicted by reducing


F lt = A (c Nc + b z N q) [0.84 + 0.16 (B/L) ] the c and values before entering Figures 6.5-3 and 6.5-4 to
. . . . . . . . . (6-4) obtain the holding capacity factors. The new c and values are
calculated as follows:
where: c = drained soil cohesion [F/L2]
(6-5)
Nc = long-term holding capacity factor in cohesive
soils
b = buoyant unit weight of the soil [F/L3]
(6-6)
Nq = holding capacity factor for a drained soil
condition

6-14
Figure 6.5-4. Holding capacity factors for cohesionless soils.

6.5.5 Short- and Long-Term Capacity in Cohesionless Soils reduced in the same manner as for the drained cohesive soil case
(i.e., by Equation 6-6) before Figure 6.5-4 is entered.
The short-term loading condition in cohesionless soils is
also the drained or long-term failure condition. In sands and 6.5.6 Disturbance Corrections
gravels, virtually all of the excess pore water pressures resulting
Equation 6-3 contains a correction factor for soil
from a static loading dissipate as the load is applied. Thus, for
disturbance to correct for soil remolding during fluke
cohesionless soils, the soil failure is assumed to be drained for
penetration and keying. Values for the factor, h, were
both static short- and long-term conditions.
determined for the four soil types listed in Table 6.5-1 by
The static holding capacity in cohesionless soils (sands
anchor tests. These values are recommended for application to
and gravels) for both short- and long-term conditions, F, is:
similar soil types in calculating propellant-embedded anchor
_
F = A b z N q [0.84 + 0.16(B/L) ] (6-7) holding capacity [Ref 6-19].
_ Sensitivity, St, (the ratio of undisturbed to remolded shear
The holding capacity factor Nq is obtained from Figure 6.5-4
strength), is an important indicator of the amount of disturbance
using the relative embedment depth and the soil friction angle.
likely. For soils with considerably different S t values than those
When dealing with very loose sands (i.e., relative density less
reported in Table 6.5-1, an estimate must be made for the h-
than 40%), the soil friction angle, , should be
value. More sensitive soils should display greater reductions in
strength and be assigned a lower value of h.

6-15
Table 6.5-1. Values for Strength Reduction the subsequent report sections. The calculation procedures
Factor for Use in Equation 6-3
presented in this section provide rough but conservative
Strength estimates of the effect of dynamic loadings on anchor capacity.
Soil Type Reduction A more complete discussion on these loadings and what may
Factor, h
Very soft, moderately sensitive, clayey silt 0.8-0.9 cause them is given in Reference 6-20.
su 1 psi, S t 3
Soft, normally consolidated, silty clay 0.8 6.6.2 Cyclic Loading
su 2 psi, S t 3
Pelagic clay 0.7 6.6.2.1 Definitions. Cyclic loading can be considered
su 1.2 psi, S t 3
an impulse loading that occurs in a repetitive manner rather
Foraminiferal sand-silt, 77-86% carbonate 0.25
su 2.2 psi, S t 10 than as a single event. For design purposes cyclic loadings are
separated into three categories: (1) cyclic line loading of the
anchor that may lead to a soil strength loss in the vicinity of
6.5.7 Factors of Safety the anchor and subsequent anchor failure; (2) cyclic line loading
that may cause anchor upward movement (creep), which could
The factor of safety to be applied to anchor holding
accumulate to move the anchor into more shallow soil and
capacity determined from Equations 6-3, 6-4, and 6-7 varies
thereby lower its short-term static holding capacity; and (3)
with the type and purpose of the mooring and with the level of
earthquake-caused cyclic loading of the soil mass with resulting
environmental data on the site. For those applications where
near-complete loss of strength in the entire soil mass and a
little is known about the soil conditions at the site, or for
sudden anchor failure. Cyclic loads are characterized by a pure
critical installations, a safety factor of 3 is recommended. When
cyclic "double-amplitude" loading component, Pc,
adequate site data permit a high level of design confidence, or
when that mooring element is noncritical, the factor of safety superimposed on a basically static loading component, P s.

may be reduced to 2. Cyclic and static load magnitudes are expressed as a percentage
of the static short-term anchor holding capacity (as determined
6.6 DYNAMIC HOLDING CAPACITY by Equations 6-3 or 6-7). Figure 6.6-1 illustrates this

6.6.1 Loading Conditions nomenclature with an example where Ps is approximately 18%

of the short-term holding capacity, P c is approximately 33% of


Dynamic loads are defined as those rapidly applied but of
that capacity, and four load cycles occur within approximately
short duration (less than 1 minute). They are divided into two
0.55 minute. In design, a cyclic load must have a double
categories: (1) cyclic or repetitive loadings and (2) impulse
amplitude greater than 5% of the static short-term holding
loading (basically a single event). Both types can alter plate
capacity. Smaller cyclic loads are difficult to measure or predict
anchor holding capacity by changing the existing conditions in
and can be ignored in the design.
the soil surrounding the anchor. These loading types are
Two additional parameters are required to describe a cyclic
illustrated by the anchor line load history shown in Figure 6.6-
loading condition. The first is the total number of load cycles
1 and discussed in
expected in the anchor's lifetime, nT. This parameter is needed

to evaluate the potential for anchor creep. The second is the


number of cycles, n c , that occurs in a limited time period

required

6-16
Figure 6.6-1. Nomenclature for types of non-steady loading (after Ref 6-20).

for dissipation of excess pore pressure, tcd . The parameter nc is For other soils, the susceptibility of a given plate anchor

used to evaluate soil strength loss and potential for liquefaction. to cyclic-load-caused strength reductions can be evaluated by
estimating the maximum cyclic load that can be sustained by

6.6.2.2 Strength Loss. Virtually all soils are subject to the anchor without pore pressure dissipation. This is done using

some strength loss from extended cyclic loading. The amount of the soil permeability, k, whose value must be determined from

strength loss, however, varies considerably depending on soil testing of undisturbed soil from the vicinity of where the anchor

type, state, and the nature of the cyclic loading. In general, the will lie in the soil mass. Table 6.6-1 shows typical values of k,

following lower the soil's susceptibility to strength loss: a which are used with Figure 6.6-2 to estimate tcd , the time

denser soil, a more plastic soil, a lower cyclic loading required for dissipation of excess pore pressures. The maximum

magnitude, a smaller number of load cycles, and a longer time number of double-amplitude cyclic loadings that can occur

period over which the cycles occur. within the time period t cd is estimated from the known or

Some low relative density cohesionless soils are expected loading conditions. Figure 5.5-3 is then used to
susceptible to complete liquefaction. Sediments of this type determine the maximum value of cyclic load, Pc , that can be
(such as uniform fine sands, coarse silts, and some clean deep- sustained without a significant soil strength loss.
sea oozes) can experience a near-total strength loss under cyclic Figure 6.6-3 can also be used to determine the maximum
loading. Use of plate anchors in these soils is not recommended number of cycles that can be sustained without soil strength
if significant cyclic loading is expected. loss for a given cyclic load level. The prediction curves of

6-17
Table 6.6-1. Average values of Soil Figure 6.6-3 apply where the average static load is less than
Permeability (after Ref
33% of the anchor's static short-term holding capacity. When
6-20)
the average static load is greater than 33% of the static holding
Soil Type Permeability, k capacity, an adjustment must be made to the cyclic and static
(fps) loads before using Figure 6.6-3. The amount of the static load
Uniform Coarse Sand 1 x 10 -2
above 33% (Ps - 33) is added to the value of Pc . This new value
Uniform Medium Sand 3 x 10 -3
of P c is then used to determine the maximum number of cycles
Well-Graded Clean Sand 3 x 10 -4 from Figure 6.6-3.

Uniform Fine Sand 1 x 10 -4


6.6.2.3 Cyclic Creep. Cyclic creep of an embedment
Well-Graded Silty (Dirty) Sand 1 x 10 -5 anchor can occur under loading conditions that appear quite safe

Uniform Silt 2 x 10 -6 relative to the above criteria for cyclic strength loss. To
evaluate cyclic creep potential, the number and magnitude of
Silty Clay 3 x 10 -8
significant loading cycles occurring during the lifetime of an
Low Plasticity Clay (Kaolinite), 3 x 10 -8 anchor are the controlling load factors. This loading usually
PI < 20
must be summarized by statistical techniques in a spectral or
Medium Plasticity Clay (Illite), 3 x 10 -19 quasi-spectral format, to estimate total number of uniform
PI = 20-60
loading cycles that can occur over the anchor's lifetime [Ref 6-
High Plasticity Clay, PI = 60-200 3 x 10 -10 20].
Very High Plasticity Clay, PI > 200 3 x 10 -11 The maximum number of cyclic loads that can occur over
the lifetime of an anchor without significant upward creep is
shown by the curves presented in Figure 6.6-4. These criteria
are applicable to cases where the average static load, Ps, is less
than 20% of static short-term anchor capacity. For cases where
the 20% static load criterion is exceeded, the double-amplitude
cyclic load, P c , should be adjusted by an amount equal to that

portion of Ps above 20% (i.e., by Ps - 20). For example, if the

static load component of a mooring is 28% of the static short-


term holding capacity, and the appropriate curve on Figure 6.6-
4 indicates an allowable maximum double-amplitude cyclic load
of 30% over the lifetime of the structure, then that allowable
double-amplitude cyclic load should be reduced by 8% (28%
minus 20%). Therefore, the allowable cyclic load becomes 22%
(30% minus 8%). This procedure is very conservative for
Figure 6.6-2. Time required for dissipation of stress-induced
excess pore pressure (after Ref 6-20). anchor systems expected to have a long service life and be
subjected to many cycles of significant cyclic loading. It was
designed to be conservative because cyclic creep of anchors is
not well-understood.

6-18
6.6.2.4 Factors of Safety. Because the above approaches

to design for cyclic loadings are quite conservative, a lower


safety factor is recommended for use with the cyclic loading
aspects of anchor design. It is recommended that the following
factors of safety be used: 1.75 for critical installations or where
the soils data are not well known and 1.25 for noncritical
installations or where the soils data are very well known.

6.6.3 Earthquake Loading

In contrast to line-applied cyclic loading of an anchor,


Figure 6.6-3. Maximum cyclic load capacity without soil
earthquakes load the entire soil mass. The loading, however,
strength loss (after Ref 6-20).
occurs for only some 10 to 30 significant and rapidly applied
loading cycles. Cohesive soils and most medium to dense sands
(materials that are less susceptible to liquefaction) subjected to
such loading are treated by the techniques outlined in Section
6.6.2. Relatively clean cohesionless soils, of medium to low
density, are considerably more susceptible to liquefaction under
earthquake loadings. Anchors embedded in these soils, even
when under very low static loads, will completely fail and pull
out if the soil liquefies. The prediction of such events is not
treated here, as it is not well-understood and is very complex;
however techniques for evaluating the liquefaction potential of
soils are presented in References 6-17 and 6-20.

6.6.4 Impulse Loading

6.6.4.1 Definitions. Procedures for predicting the holding

capacity under impulse loading are presented below. They are


appropriate for use with circular, square, or rectangular (L/B
2) anchor flukes only. The procedures consist of applying a
number of influence factors to Equations 6-3 and 6-7, the basic
equations for calculating static short-term anchor holding
capacity. The influence factors yield conservative results in all
cases.
An impulse load, PI , is defined as basically a single load
Figure 6.6-4. Maximum (lifetime) cyclic load capacity without that is applied quickly but does not remain for a long time (less
development of cyclic creep (after Ref 6-20).
than

6-19
10 minutes for clays and less than 10 seconds for sands). An F I = I R c R I If (F st) (6-8)
example is shown in Figure 6.6-1. These loads are considered
where: F st = static short-term anchor holding capacity
single events only where enough time elapses between similar
(Equation 6-3 for cohesive soil) [F]
events to allow the soil to return to its "normal" state without I = influence factor for adjusting the soil strength
for strain rate
any residual effect from other impulse loads. In the absence of R c = reduction factor for cyclic loading
other dynamic loads, impulse loading will result in a higher R I = reduction factor for repeated impulse loading
If = inertial factor for capacity increase under very
anchor holding capacity during the loading event than that rapid and short-duration loading (i.e., for
impulse duration less than 0.01 second)
computed for the static short-term holding capacity. That is, the
anchor will be able to resist an impulse loading higher than its The factor I is a strain rate used to adjust for an increase
static short-term capacity. in the soil undrained shear strength during impulse loading. The
value of I is obtained from Figure 6.6-5 on the basis of impulse
6.6.4.2 Cohesive Soil . The holding capacity under load duration and a general description of soil type.
impulse loading, FI , is the maximum load that can be applied

to an anchor under impulse conditions. It is determined for


cohesive soils by:

Figure 6.6-5. Strain-rate factor, I, for cohesive soil (after Ref 6-21).

6-20
The factor Rc depends on the loading history prior to the period, then RI = 1; or (2) if there is more than one impulse

impulse loading. It is used to adjust F I for the influence of load in a 4-hour period, then:

other nonstatic loads that are occurring at about the same time
(6-10)
and is determined as follows: (1) if the impulse load is the first _
event, then R c = 1.0; or (2) if cyclic loads immediately precede where fc is the average frequency, in impulses per hour, over a

the impulse event, then: 4-hour period.


The factor If is used to adjust (increase) FI for the inertia
R c = P c /F st (6-9)
of the soil mass at very short duration loadings (i.e., where the
loading is known to be applied for less than 0.1 second). It is
The factor RI adjusts FI for repeated impulse loadings. It is
determined from Figure 6.6-6.
determined as follows: (1) if there is only one impulse load in a
4-hour

Figure 6.6-6. Inertial factor, I f , for cohesive and cohesionless soils (after Ref 6-21).

6-21
6.6.4.3 Cohesionless Soil . The holding capacity The adjusted holding capacity factor, NqI , accounts for the effect

under impulse loading in cohesionless soils is also derived by of the impulse loading on the soil friction angle, . In order to
applying a series of influence factors to the calculated short- obtain NqI , first a new friction angle adjusted for the impulse
term static holding capacity. The maximum anchor capacity loading effect, I , is calculated as shown below [Ref 6-21].
impulse loading is given by: Then, NqI , is determined from the curves in Figure 6.5-4.

(6-11)
(6-12)

where: F = static short-term anchor holding capacity where I is the influence factor for adjusting the soil strength
(Equation 6-7 for cohesionless soil) [F] (from Figure 6.6-7 for cohesionless soil).
NqI = cohesionless soil holding capacity factor
adjusted for impulse loading The factor Rc is determined in the same manner as for
Nq = cohesionless soil holding capacity factor (from
cohesive soil. This is described in Section 6.6.4.2.
Figure 6.5-4)
R c = reduction factor for cyclic loading The factor for repeated impulse loadings, RI , is dependent
R I = reduction factor for repeated impulse loading
If = inertia factor for capacity increase under very on the frequency of those impulse loads, fs: (1) if f s one
rapid and short-duration loading (i.e., for impulse/10 min, then RI = 1.0; or (2) if f s > one impulse/10
impulse duration less than 0.01 second)
min, then RI is obtained from:

Figure 6.6-7. Strain-rate factor, I, for cohesionless soil (after Ref 6-21).

6-22
Table 6.7-1. Factors Associated With Direct-
(6-13) Embedment Anchors Which Can
Influence Submarine Slope
where fs is the average number of impulses per 10 minutes. Stability (after Ref 6-22)

The inertial factor If is determined in the same manner as Factor Reasons for Consideration
for cohesive soil (described in Section 6.6.4.2). If is determined Impact Loading Effect similar to earthquake loading but
During Embedment with greater local influence; more
from Figure 6.6-6. critical problem in loose soils.
Remolding or Needs consideration in all anchor
6.6.4.4 Factors of Safety. To calculate allowable Disturbance of Soils installations; effect varies significantly
During Installation from one soil to another.
design loadings from the maximum impulse loadings calculated Cyclic Loading by Important consideration in all anchor
from this section, the factors of safety recommended in Section Anchor installations.
Local Instability Potentially can progress to major slope
6.5.7 should be applied. After Anchor Pullout failure.
Direct Application of Probably not more significant than a
6.7 OTHER INFLUENCES ON HOLDING CAPACITY Anchor Load to local instability problem but can
Slope progress into large slide.
6.7.1 Holding Capacity on Slopes
ure and, therefore, lowering the holding capacity that can be
Two major items should be considered in predicting the
developed. A significant portion of the soil shear strength is
holding capacity of embedded plate anchors on slopes: first,
mobilized to support the soil slope against gravitational forces.
stability of the slope itself under the additional influence of the
For downslope loading, most of the soil shear stresses
embedment anchor; and second, the influence of the inclined
developed to resist anchor pullout will be in addition to those
seafloor on the soil resistance mobilized (i.e., holding capacity)
resisting slope failure. The result is that a smaller amount of
by the loaded anchor.
the soils shear strength is available to resist anchor pullout
The effect of an anchor on slope stability is extremely
than is available with a horizontal seafloor. This becomes less
complex, involving the effects of anchor installation and anchor
important as the anchor is loaded in a more vertical or more
loading on the slope. Although slope stability is treated
upslope direction.
separately in Chapter 11, no direct solution to the associated
Methods for calculating the reduced holding capacity of
problems is presented. Table 6.7-1 presents a list of factors
anchors on slopes have not been developed. However, as a
associated with direct-embedment anchor installation and
conservative approach, the holding capacity of a direct-
loading which influence submarine slope stability [Ref 6-22].
embedment anchor on a specific slope can be estimated by
All factors lead to a lower resistance to soil mass downslope
multiplying the holding capacity calculated to a horizontal
sliding and, therefore, greater slope instability. It is important
seafloor by a reduction factor, Rs, determined by:
to note that the influence of plate anchors on slope stability
depends to a high degree on the type and sensitivity of the
(6-14)
sediment. Slope angle itself is not a clear indicator of potential
problems. where Fs is the factor of safety against a slope failure without

The inclined seafloor also influences the anchor by the anchor. R s represents the

lowering the resistance to soil fail-

6-23
amount of soil strength remaining, or the amount not Note: This equation is completely empirical and is not
dimensionally stable. Values used for m and v must be in the
mobilized in maintaining the slope stability. Note that in
units described.
computing the anchor holding capacity, both anchor depth and
the holding capacity factors N c and Nq are based on a depth of where: F c = holding capacity [kips]
m = anchor fluke and piston mass [slugs]
embedment measured perpendicular to the seafloor surface.
V = initial fluke velocity [fps]

6.7.2 Creep Under Static Loading


This equation can be used for estimating the holding capacity of
Creep failure of direct-embedment anchors under static all the NCEL propellant-embedded coral-fluke anchors in coral
loading conditions is possible. In some onshore soils, data have seafloors similar to those where the anchor data were obtained.
been taken showing soil strength reductions of 60% for some The use of this equation is limited to the range of input
soft cohesive soils under only static loading [Ref 6-23]. parameters covered by the data from which it was developed.
However, tests on two pelagic clays [Ref 6-24] and a calcareous Thus, Equation 6-15 is valid only for anchor fluke shapes
ooze [Ref 6-25] indicate that, for the undrained condition, shear similar to the NCEL platelike coral fluke [see Figure 6.2-3(a)]
strength reductions may not be nearly so great for seafloor and for coral with unconfined compressive strengths ranging
sediments. Further, for soil overconsolidation ratios (OCR) from 1,500 to 4,200 psi.
reasonable in seafloor soils, negative pore pressures will not be
6.8.2 Rock
generated above the anchor plate, and the shear strength above
the anchor plate will always increase with time [Ref 6-24]. To date, efforts to develop a reliable holding capacity
Therefore, the creep failure potential for plate anchors under predictive equation for the propellant-embedded anchors in hard
only static loading is minimal. A safety factor of 2 against rock have not been successful. Local rock strength variations
creep failure is recommended on the maximum long-term within the rock types tested are believed to be largely
holding capacity [Ref 6-11]. responsible for the lack of consistent holding capacity
performance. A conical-shaped fluke design [Figure 6.2-3(b)] is
6.8 HOLDING CAPACITY IN CORAL AND ROCK
recommended for use in hard rock seafloors, based on test
6.8.1 Coral results with seven fluke designs in weathered granite and in
vesicular basalt [Ref 6-1]. Work is continuing in this area to
Prediction of direct-embedment anchor holding capacity in
better understand embedment anchor behavior and to allow
coral is difficult due to the variable nature of coral seafloors.
development of a predictive method. Until then, proof-loading
Tests of anchor flukes embedded by the NCEL 10K and 20K
must be used as the only method available for reliably
systems at Dahu and Kauai, Hawaii, were evaluated by a simple
estimating capacity.
regression analysis. An equation for predicting holding capacity
as a function of kinetic energy was developed from these anchor 6.9 EXAMPLE PROBLEMS
tests [Ref 6-1]. It estimates the short- and long-term holding
6.9.1 Problem 1--An Embedment Anchor Used in Cohesive
capacity as follows:
Soil

6.9.1.1 Problem Statement: Determine (1) the maximum


(6-15)
allowable steady uplift force, (2) the maximum allowable cyclic
load component,

6-24
and (3) the holding capacity under impulse loading of a specific conditions. The subsurface buoy is expected to remain in place
propellant-embedment anchor in a cohesive seafloor so that a for 10 years, with the surface buoy and its mooring line being
buoy and mooring system may be properly designed. replaced more frequently. The subsurface buoy installation
Data: An NCEL-style 20K propellant embedment anchor procedure will result in a single impulse load at the anchor of
with a 3- by 3-foot clay fluke, with 8.5 ft2 bearing area, is to be between 0.01 and 0.1 second duration. In addition, some cyclic
installed in 1,200 feet of water off the Southern California loading will result from wave loading on the surface buoy.
Coast to moor a subsurface buoy about 60 feet below the water (Historical records show that the maximum storm in a 10-year
surface and a surface data transmission buoy [see Figure 6.9-1]. period produces 13-second waves for about a 10-hour period. )
The installation is considered critical in nature. The site has
6.9.1.2 Problem Solution: The procedures and
been well-surveyed for prior work and the material is "very soft"
silty-clay (cohesive). A soil strength profile has been computations used to solve this problem are shown below.

established [see Figure 6.9-2]. The site does not contain any They follow procedures outlined in Figure 6.3-1 and Equations

hazardous or otherwise unusual 6.3 through 6.6 of this chapter.

Figure 6.9-1. Mooring sketch for example


problem 1. Figure 6.9-2. Soil strength profile for example
problem 1.

Problem 6.9.1

6-25
Problem 6.9.1

6-26
Problem 6.9.1

6-27
Problem 6.9.1

6.9.2 Problem 2--An Embedment Anchor Used in of 10 hours duration with wave periods of 10 seconds. In
Cohesionless Soil addition, 10 smaller storms are expected each year with

6.9.2.1 Problem Statement: Determine (1) the maximum approximately the same wave conditions but of only 6 hours

allowable static uplift force and (2) the maximum allowable duration each. The site is not in an active earthquake zone.

cyclic uplift force of a specific propellant-embedment anchor in


a sand seafloor for design of a subsurface buoys mooring
system.
Data: A Navy 100K propellant-embedment anchor is to
be installed in a sand seafloor to tautly moor a large subsurface
buoy 20 feet below the water surface [see Figure 6.9-3]. The
100K anchor sand fluke is 5 by 2.5 feet and has a surface area of
11 ft2. Corers have consistently recovered a well-graded dirty
silty-sand, which has been estimated to be in a medium-dense
state. A 3.5-kHz subbottom survey detects only one reflector,
believed to be a rock contact, at about 40 feet. A review of
available areal geologic data indicates nothing that would
contradict a simple sand-over- rock soil profile model. The
Figure 6.9-3. Mooring sketch for example problem 2.
anchor is scheduled to be used for 2 years. Impulse loading is
not expected. The worst storm expected is once in 2 years and is

6-28
6.9.2.2 Problem Solution: The analytical procedures dures outlined in Figure 6.3-1 and in Sections 6.3 through 6.6

and computations used to solve this problem are shown below. of this chapter.

They follow the proce-

Problem 6.9-2

6-29
Problem 6.9-2

6-30
6.10 REFERENCES 6-10. J.E. Smith. Umbrella pile-anchors, Naval Civil
Engineering Laboratory, Technical Report R-247. Port
6-1. J.F. Wadsworth and R.M. Beard. Propellant-embedded
Hueneme, Calif., May 1963, p. 4.
anchors: Prediction of holding capacity in coral and rock
seafloors, Civil Engineering Laboratory, Technical Note N-
6-11. R.J. Taylor. "Interaction of anchors with soil and anchor
1595. Port Hueneme, Calif. , Nov 1980.
design," presented at short course on Recent Developments in
Ocean Engineering, University of California, Berkeley, Calif.,
6-2. R.J. Taylor, D. Jones, and R.M. Beard. "Uplift-resisting
Jan 1981.
anchors," in Ocean Engineering, vol 6. Pergamon Press, 1979,
pp 3-137.
6-12. H.S. Stevenson and W.A. Venezia. Jetted-in marine
anchors, Naval Civil Engineering Laboratory, Technical Note
6-3. J.F. Wadsworth and R.J. Taylor. CEL 10K propellant-
N-1082. Port Hueneme, Calif., Feb 1970.
actuated anchor, Civil Engineering Laboratory, Technical Note
N-1441. Port Hueneme, Calif., Jun 1976.
6-13. N. Kerr. "The hydropin: A new concept in mooring,"
Transactions, North East Coast Institution of Engineers and
6-4. D.G. True and R.J. Taylor. The CEL 100K propellant
Shipbuilders, vol 92, no. 2, Nov 1975, pp 39-44.
anchor - Utilization for tanker moorings in soft coral at Diego
Garcia, Civil Engineering Laboratory, Technical Note N-1446.
6-14. R.M. Beard. Expendable Doppler penetrometer: A
Port Hueneme, Calif., Jul 1976.
performance evaluation, Civil Engineering Laboratory,
Technical Report R-855. Port Hueneme, Calif., Jul 1977.
6-5. D.G. True and R.J. Taylor. "New expedient anchors in
fleet applications," Naval Engineers Journal , Feb 1981, pp 53-
6-15. R.J. Malloy and P.J. Valent. Acoustic siting and
61.
verification of the holding capacity of embedment anchors,
Civil Engineering Laboratory, Technical Note N-1523. Port
6-6. J.C. Miller. 100K propellant anchor technology transfer -
Hueneme, Calif. , Jul 1978.
Diego Garcia fleet mooring installation, Naval Civil
Engineering Laboratory, Technical Note N-1638. Port
6-16. Design manual: Soil mechanics, foundations, and earth
Hueneme, Calif., Jul 1982.
structures, Naval Facilities Engineering Command, NAVFAC
DM-7, Washington, D.C., 1962.
6-7. R.M. Beard. Direct embedment vibratory anchor, Naval
Civil Engineering Laboratory, Technical Report R-791. Port
6-17. R.M. Beard. Holding capacity of plate anchors, Civil
Hueneme, Calif. Jun 1973.
Engineering Laboratory, Report R-882. Port Hueneme, Calif.,
Oct 1980.
6-8. A. Kovacs, S. Blouin, B. McKelvy, and H. Colligan. On
the theory of ground anchors, Cold Regions Research and
6-18. A.S. Vesic. "Breakout resistance of objects embedded in
Engineering Laboratory, U.S. Army Corps of Engineers,
ocean bottom," in Civil Engineering in the Oceans II. New
Technical Report 258. Hanover, N.H., Jan 1975.
York, N.Y., American Society of Civil Engineers, 1970, pp
137-165.
6-9. Design manual, harbor and coastal facilities, Naval
Facilities Engineering Command, NAVFAC DM-26.
Washington, D.C., Jul 1968.

6-31
6-19. P.J. Valent. Results of some uplift capacity tests on e Exponential base for natural logarithms
direct embedment anchors, Civil Engineering Laboratory, F Holding capacity in cohesionless soils [F]
Technical Note N-1522. Port Hueneme, Calif. , Jun 1978. Fc Anchor holding capacity in coral seafloors

6-20 H.G. Herrmann. Design procedures for embedment FI Holding capacity under impulse loading [F]

anchors subjected to dynamic loading conditions, Naval Civil F lt Long-term static holding capacity [F]
Engineering Laboratory, Technical Report R-888. Port Fs Factor of safety
Hueneme, Calif., Nov 1981.
F st Short-term static holding capacity [F]

6-21. B.J. Douglas. Effects of rapid loading rates on the fc Average number of impulse loadings per hour, in clay,
over a 4-hour period [l/T]
holding capacity of direct embedment anchors, Civil
fs Average number of impulse loadings per 10 minutes, in
Engineering Laboratory, P.O. No. M-R420. Port Hueneme, sand [l/T]
Calif., Oct 1978.
h Correction factor for soil disturbance due to penetration
and keying
6-22. F.H. Kulhawy, D.A. Sangrey, and S.P. Clemence. Direct
I Influence factor for adjusting soil strength for impulse
embedment anchors on sloping seafloors, state-of-the-art, Civil loading
Engineering Laboratory, P.O. No. M-R510. Port Hueneme, If Inertial factor for holding capacity increase under impulse
Calif., Oct 1978. loading

k Soil permeability [L/T]


6-23. A. Singh and J.K. Mitchell. "General stress-strain-time
L Fluke or plate maximum dimension, usually length [L]
function for soils," Journal of the Soil Mechanics and
m Fluke and piston mass in rock holding capacity equation
Foundations Division, American Society of Civil Engineers,
[FT2/L]
vol 94, no. SM1, Jan 1968, pp 21-46.
Nc Long-term holding capacity factor in cohesive soil

6-24. R.M. Beard. Long-term holding capacity of statically Ncs Short-term holding capacity factor in cohesive soil
loaded anchors in cohesive soils, Civil Engineering Laboratory, Nq Holding capacity factor for a drained soil condition
Technical Note N-1545. Port Hueneme, Calif. , Jan 1979.
NqI Cohesionless soil holding capacity factor adjusted for
impulse loading
6-25. P.J. Valent. Long-term stress-strain behavior of a
nc Actual number of cycles occurring during time period,
seafloor soil, Civil Engineering Laboratory, Technical Note N- tcd
1515. Port Hueneme, Calif. , Feb 1978. nT Total number of cycles occurring during anchor lifetime

Pc Double-amplitude cyclic load component [F]


6.11 SYMBOLS
PI Magnitude of the impulse line load [F]
A Projected fluke area perpendicular to direction of pullout
[L2] Ps Static or nearly static line load component [F]

Rc Reduction factor for holding capacity under cyclic loading


B Fluke or plate minimum dimension, usually width [L]
RI Reduction factor for holding capacity under repeated
c Drained soil cohesion [F/L2] impulse loading
c Reduced drained soil cohesion for very soft Rs Reduction factor applied to anchor holding capacity to
underconsolidated soils [F/L2] account for slope instability

St Soil sensitivity

6-32
su Soil undrained shear strength [F/L2] b Soil buoyant unit weight [F/L3]

tcd Time duration required for dissipation of excess pore Soil friction angle [deg]
pressure [T]
Drained cohesive soil friction angle [deg]
v Initial velocity of rock fluke [L/T]
Reduced drained cohesive soil friction angle for very soft
z Embedded depth of fluke after keying [L] underconsolidated soils [deg]

zk Distance required for fluke keying [L] I Soil friction angle adjusted for impulse loading [deg]

zp Maximum penetration depth of keying-type fluke [L]

6-33
Chapter 7

DRAG-EMBEDMENT ANCHORS

7.1 INTRODUCTION 7.1.2 Drag Anchor Description

7.1.1 Purpose and Scope Most drag anchors are referred to by manufacturer
tradenames as the anchor type. While specialty anchors may
This chapter consolidates performance, selection, and
have a very different shape, many widely used anchors share
design information for drag-embedment anchors. The
common features. All have a shank, through which the
functioning of the drag anchor is described, and a brief
mooring line load is applied, and a fluke or flukes, which are
description of geotechnical information necessary for design
the digging parts of the anchor and provide the bearing area to
follows. Criteria for selection of an anchor are listed; techniques
mobilize sediment resistance. These and other components
for designing or sizing are then presented. A section on field
common to most drag anchors are shown in Figure 7.1-1.
solution of anchor performance problems follows.
Tripping palms* improve the capability of the anchor fluke to
This chapter deals only with geotechnical aspects of
open and dig into the seafloor. Stocks or stabilizers are used on
mooring system design. Prediction of mooring system loads
many anchors to improve their biting into the seafloor and their
selection of mooring line materials, and maintenance and
rotational stability. The biting or leading edge of the fluke is
inspection of mooring system components are not discussed.
the fluke tip. The area where the fluke connects to the shank is
For prediction of environmental loadings on moored vessels,
the crown or head of the anchor.
the reader is referred to the Navys Design Manual DM-26 [Ref
The chain or wire rope mooring line attached to the
7-1], the Det norske Veritas "Rules for the Design,
anchor shank is an integral part of a drag anchor system. The
Construction and Inspection of Offshore Structures" [Ref 7-2],
section of this mooring line is dragged below the seafloor
the rules of the American Bureau of Shipping [Ref 7-3], and the
surface by the anchor, while a variable length of the mooring
API "Recommended Practice for the Analysis of Mooring
line usually rests on and is dragged along the seafloor surface.
Systems for Floating Drilling Units" [Ref 7-4]. Information on
Both mooring line sections develop resistance to horizontal
mooring system component selection and system design and
movement and contribute to the holding capacity of the anchor
analysis are generally available [Ref 7-5, 7-6]. Mooring
system.
inspection and maintenance rules have been drawn up by Det
norske Veritas [Ref 7-4] and summarized by Fjeld [Ref 7-7].

____________
*Often called "mud palms" or simply "palms."

7-1
Figure 7.1-1. Features of a drag anchor (after Ref 7-5).

7.1.3 Types of Drag Anchors

Drag anchors are classified as:


Movable or fixed fluke
Bilateral or unilateral fluked
Hard or soft seafloor anchors
Standard or high holding power anchors

Movable fluke anchors are hinged at the crown so that the


anchor can lie flat, with the plane of the flukes parallel to the
axis of the shank when on the deck of a work boat [Figure 7.1-
2(a)]. The flukes can still open fully for digging in when the
anchor is placed on the seabed [Figure 7.1-2(b)]. Fixed fluke
Figure 7.1-2. Example of a movable fluke anchor:
anchors are those with either movable flukes which have been STEVIN cast.
welded or blocked in the open position, or those where the fluke
and shank have been cast or fabricated as a single piece [Figure
7.1-3].
Bilateral fluked anchors are those constructed so the
flukes can open freely to either side of the anchor shank [Figure
7.1-4(a)]. Unilateral fluked anchors are constructed so the fluke
must remain only on one side of the shank [Figure 7.1-4(b)].
Unilateral fluked anchors require greater care in handling and
placement on the seafloor to ensure that the flukes are down for
digging; improper placement could result in the anchor being
on its back and unable to dig into the seafloor.
Figure 7.1-3. Example of a fixed fluke anchor:
BRUCE cast (after Ref 7-9).

7-2
Figure 7.1-4. Examples of bilateral fluke anchors.

Anchors which work best in hard soil seafloors (i.e., hard permanent mooring of floating platforms. They are the leading
clays and most sands and gravels) are sharp-fluked with close-set contenders for temporary moorings because they are efficient,
fluke tips to initiate penetration. They have long stabilizer bars require a minimum of specialized support, and are reusable.
to counter roll instability tendencies [Figure 7.1-4(a)]. Anchors Often, catenary system motion (in reaction to an increase in
for soft seafloors, on the other hand, maximize fluke area and line tension) and load characteristics of the drag anchor system
emphasize streamlining to achieve deep seafloor penetration are required in a mooring system to cope with dynamic load
[Figure 7.1-5]. components.
On the disadvantage side, the drag anchor is often a poor
7.1.4 Application of Drag Anchors
performer on very hard seafloors. Further, the drag anchor
Drag anchors are standard equipment for the temporary system is not a constructed anchor system, like the pile or
mooring of all mobile craft of substantial size and are often gravity anchor system. After placement, the drag anchor must
selected for the trip, dig in, and remain stable with drag. All of these are
statistical performance functions and have a probability of
occurrence lower than 100%, particularly in hard soils.

7.2 FUNCTIONING OF A DRAG ANCHOR

7.2.1 General

The drag anchor is analogous to an inverted kite made to


"fly downward" into the soil. The shank acts as the kite bridle,
maintaining the angle of attack of the fluke to the soil to
maximize the developed mooring line tension in the horizontal
direction at the seafloor. The
Figure 7.1-5. Example of a soft soil anchor:
STEVMUD (from Ref 7-11).

7-3
anchor is pulled along the seafloor until it digs in or penetrates On soft seafloors (i.e., soft clays and muds) those anchors
to its place of maximum holding capacity. The tripping palms with very heavy crowns, small or nonexistent tripping palms,
assist by causing the movable fluke anchors, when placed on or with the shank-to-fluke hinge far back on the fluke tend to
the seafloor, to begin digging in when dragged. The stabilizers have tripping problems. This holds true especially when the
assist, especially on harder seafloors in orienting and anchor is lowered and placed crown first by the mooring line
maintaining the anchor in the horizontal position. When [Figure 7.2-1(a)]. Anchors of this type (e.g., the Stockless,
working properly, the drag anchor will embed itself into the LWT, STEVFIX), when dragged to set them, will often not dig
soil to some equilibrium depth dependent on existing mooring into the seafloor but will instead slide at mudline level with the
line, anchor, and soil conditions. When not working properly, movable flukes oriented parallel to the shank or pointing
the anchor will not embed as deeply nor develop as high a slightly upward, serving only as a deadweight anchor at the mud
holding capacity, or it may not embed at all. surface [Figure 7.2-1(b)]. Even the newer anchor types are not
free of this soft soil tripping problem, e.g., the STEVFIX
7.2.2 Tripping
anchor has demonstrated a tripping problem in soft, clayey silts
Most drag anchors are the movable, bilateral fluke [Ref 7-12].
variety, where the flukes are free to move to either side of the This problem in soft soil can be largely eliminated by
shank and where the anchor can lie flat on the deck of a work proper anchor setting procedures [depicted in Figure 7.2-2].
boat with flukes nearly parallel to the shank and deck. Generally, implementation of this procedure requires two
Generally, the movable fluke anchor is easier to handle, deploy, vessels to lay a mooring leg: one paying out the mooring line
and recover. Occasionally, a movable fluke anchor on the and the second handling, lowering, and positioning the anchor
seafloor may not trip. for digging in when the mooring line is pulled.

Figure 7.2-1. Development of a tripping problem in soft seafloors with an improperly set anchor.

7-4
7.2.3 Embedment

Embedment, or penetration of the fluke into the seafloor


once digging-in is initiated, is governed by the fluke angle, the
soil type, the degree of anchor streamlining, and the
smoothness of the fluke surfaces. In addition, the mooring line
angle to the seafloor must be zero to ensure that even a properly
selected drag anchor will embed properly.
There is a critical fluke-to-shank angle (termed the "fluke
angle," ) at which the anchor holding capacity is maximized
This critical angle decreases as:

Fluke length increases


Shank length decreases
Fluke surface becomes rougher

Figure 7.2-2. Proper anchor setting sequence using


two floating platforms.

Anchor tripping difficulties may also occur in dense and


hard soils. Even with proper anchor handling procedures, the
fluke tips may not be able to develop sufficient local stress to
initiate digging in. Under these conditions, the anchor may
simply slide without tripping [Figure 7.2-3(a)] or may dig in
slightly to effect standing up [Figure 7.2-3(b)], and then fall on
its side and drag [Figure 7.2-3(c)]. Those anchors having a
relatively heavy crown and a shank connection well back of the
center of fluke area (Moorfast, Offdrill, etc.) appear susceptible
to this penetration problem in hard soils. The hard soil tripping
problem has been corrected by sharpening the fluke tips to
improve digging capability, by welding barbs on the tripping
palms to increase the tripping moment, and by reducing fluke
angle several degrees below the sand setting. Most drag anchor
types incorporate stocks or stabilizers to prevent the standing
anchor from falling completely on its side [Figure 7.2-3(c)].
Thus, the stocks serve to hold the fluke tips in a digging Figure 7.2-3. Development of a tripping problem in
hard seafloors.
position.

7-5
For fluke angles greater than critical, the standing anchor as with the more streamlined new anchors such as the
will penetrate only slightly into the sediments and will slide in STEVFIX. In soft soils (soft clayey silts and clays) the anchors
a standing or tipped orientation [Figures 7.2-3(b) and (c)] at a will penetrate more deeply [Figure 7.2-4(b)], from 45 feet for
very small holding capacity. For fluke angles less than critical, large STATO-type anchors to 60 feet for a 20,000-pound
anchor embedment depth will be reduced, and the anchor will STEVIN [as reported in Ref 7-8].
develop less than its maximum potential holding capacity in
7.2.4 Stability
that particular environment.
Data gathered for STATO anchors in the 1950s [Ref 7- A drag anchor may exhibit instability after the initial
14] showed an optimum fluke angle of 50 degrees for soft mud biting and digging-in due to differences in the soil resistance
(clayey silt) and 34 degrees for sand. More recent tests [Ref 7- encountered by the two flukes, initial differences in the fluke
11] in a denser sand showed that, for the STATO, a fluke angle penetration depths, a slight change in the direction of mooring
of 30 degrees is more appropriate for use in sands. Most line pull, or some other source of asymmetry. Drag anchors in
anchors are manufactured with a fluke angle of about 50 degrees sand will often become unstable and roll after only a few feet of
for soft soils, with bolted or welded wedges or inserts providing drag. Once the anchor begins to roll, the soil pressure on the
an option for fluke angle reduction for hard soils. Changes to rising fluke becomes much reduced, and the force couple in the
standard fluke angle settings may be advised to improve direction of the roll is increased, thus speeding the rolling of the
performance, especially on hard, dense sands and gravels. For anchor on its side [Figure 7.2-5(a)]. Stocks or stabilizers are
instance, where a fluke angle change to 28 degrees may be designed to develop a countering force couple to resist the roll
necessary to initiate the biting and digging-in of large Moorfast- motion [Figure 7.2-5(b)]. In soft clays and silts (mud), the
and offdrill-type anchors. stabilizers probably provide a significant stabilizing influence
Anchor penetration is influenced by the anchor's degree of only for the first few feet of drag, up to the point where both
streamlining. The newer anchor designs with tapered and stabilizers pass beneath the mudline. Even when fully
sharpened flukes, narrowed and chamfered shanks, and open embedded, every drag anchor is potentially unstable. Once
tripping palms have good penetration capability and can reach beneath a soft seafloor surface, anchor stability is primarily a
stronger soils deep in the soil profile. function of the symmetry of the anchor, the variation in the
An increase in fluke roughness limits anchor penetration direction of mooring line pull and the homogeneity of the soil.
and therefore influences holding capacity. Flukes with smooth At present, insufficient data exist to qualify the probability of
surfaces mobilize less soil resistance to penetration in the plane an instability developing in either hard or soft soils.
of the fluke. Thus, smooth flukes penetrate deeper and reach the Instability after anchor dig-in also includes the
stronger soils which usually occur at greater depths. phenomenon of anchor "balling up" and pulling out [Ref 7-15].
At optimum fluke angles, penetration behavior of an This balling up phenomenon refers to formation of a large ball
anchor is vastly different for hard and for soft soils. In hard soils of soil on the entire fluke and crown assembly [Figure 7.2-6],
(stiff clays and the denser sands), the drag anchor does not which can occur after dragging 50 to 200 feet in soft soils. This
penetrate deeply. Rather, the crown of the anchor may remain ball of soil or "dead zone" travels with the anchor, distorting its
above the surrounding seafloor surface, as with a STATO shape and significantly limiting anchor penetration capability
[Figure 7.2-4(a)], or it may penetrate only a few feet, and its stability. A

7-6
Figure 7.2-4. Penetration and orientation behavior of an anchor in hard and soft seafloors.

Figure 7.2-5. Forces on unstabilized and stabilized anchors in sand.

7-7
achieve their maximum capacity in a much shorter drag distance
because the soil shear surfaces developed by these anchors are
substantially greater per unit fluke area than those developed by
the newer streamlined anchors. This will be apparent in Section
7.5.3.1.

7.2.5 Soaking

"Soaking" of an anchor is the practice of allowing a


Figure 7.2-6. Anchor in soft soil, after balling-up and newly embedded anchor to rest for a period of time, typically 24
pulling-out (from Ref 7-15).
hours, before applying the required proof load. The mechanism
balled-up anchor that rises to the surface due to rotational that makes soaking work is similar to that causing balling up
instability will not re-embed with further dragging. It must be (i.e., consolidation of the silt and clay soil around the fluke,
recovered and cleaned before it can be reset. causing strengthening of that soil and increasing the anchor's
The process of development of this balling up is holding capacity with time).
hypothesized as follows. The soil in the dead zone is subjected
7.3 SITE INVESTIGATION
to a high total stress and the soil develops large positive excess
pore water pressures. In silt to clayey silt soils (muds), some of 7.3.1 Site Data Needed
this excess pore water pressure dissipates to the surrounding
Although this chapter does present the results of initial
soils as the anchor drags. The soil immediately in front of the
work toward a rational methodology for determining the
flukes, therefore, becomes stronger than the undisturbed soil and
capacity of drag anchors, present technology cannot utilize
is able to adhere to and build up on the flukes to become a
detailed soil strength and behavior information for drag anchor
large, compact lump.
design.
The balling-up phenomenon has been reported [Ref 7-13]
Drag anchor design usually consists of selecting an
with LWT, STATO, and Moorfast anchors, but not with the
anchor type and size. This selection and sizing requires the
more streamlined BOSS anchor. Performance differences lie in
following knowledge of site conditions:
the orientation of the fluke with respect to the anchor trajectory
[Ref 7-12, 7-11]. The STATO fluke surface, for instance, is Topography and thickness of significant soil layers
Sediment type (cohesive versus cohesionless)
oriented much more obliquely to its trajectory in the soil than
Undrained shear strength for cohesive soils
the fluke surface of the more streamlined Hook anchor. This
7.3.2 Topography and Layer Thickness
more oblique orientation and the STATOs large, tripping
palms are believed responsible for a tendency to ball up in some Knowledge of topography and sediment thickness at the
soft soils. The fluke, tripping palm, and shank streamlining of proposed anchor locations is necessary for determining whether
the newer anchors (e.g., BOSS, Hook, STEVMUD, and Flipper or not drag anchors can work at a given site. First, topography
Delta) are responsible for reducing the potential for this is an indicator of seafloor material type. Irregular or rugged
happening. topography may indicate outcrop-
A positive feature associated with balling up is that
anchors with this characteristic

7-8
ping of rock or hard strata or the existence of boulders or talus. depth of about one fluke length in dense sands and to three to
Drag anchors should not be expected to function well, if at all, six fluke lengths in soft clays and silts [see Table 7.3-1]. For
at such sites. Second, topography indicates the bottom slope less dense sands penetration will be slightly more deep, and for
gradient at the anchor location. An anchor being pulled stronger clays and silts the penetration will be less deep than
downslope will have a lower holding capacity than if it were indicated on Table 7.3-1. This table can be used to gauge the
pulled on a horizontal surface in the same soil. In lieu of depth to which a site investigation is needed. The ratios for the
rational and definitive procedures, it is recommended that the larger drag anchors, for example, indicate sediment thickness
siting of drag anchors be limited to downslopes of no more than requirements of 15 to 20 feet in dense sands and 60 to 80 feet in
10 degrees and, where it is practical , to slopes of less than 5 the soft clays and silts.
degrees. Techniques and equipment for obtaining the necessary
Topography is also important in the design of the total soils data are discussed in Chapter 2.
mooring system. Overall topography may limit anchor
7.3.3 Sediment Type and Strength
placement locations, and affects the required lengths of the
mooring legs, the proportions of chain and wire rope in hose Data on sediment type (clay, silt, or sand) and
mooring legs, mooring leg loads, and allowable anchor consistency (soft or hard) are necessary for selection of anchor
displacements. type and for sizing. This information can sometimes be
Sediment thickness data are necessary to determine that interpreted directly from subbottom acoustic records or
sufficient depth exists to permit full anchor penetration to experienced personnel. But, more often, soil samples are
equilibrium depth, where the maximum holding capacity will required for visual classification and index property testing (see
be developed. Drag anchors will penetrate to a Chapter 3).

Table 7.3-1. Estimated Maximum Fluke Tip Penetration of Some


Drag Anchor Types in Hard and Soft Soils

Normalized Fluke Tip


Penetration, (dt/L) max
Anchor Type Cohesive Soils Cohesionless Soils
(soft clays (dense sands)
and silts)
Stockless (movable fluke) 2 1
Stockless (fixed fluke) 3 1
Moorfast 4 1
Offdrill
BOSS 4-1/2 1
BRUCE Cast
Danforth
LWT (Lightweight)
STATO/NAVMOOR
STEVFIX
BRUCE Twin Shank 5 1
Hook
STEVMUD

7-9
In most instances, samples from short gravity corers will
provide the necessary sediment for determining classification
and consistency data. Acoustic survey data will usually suffice
to characterize the sediments over the expected depth of
penetration of the anchor, provided the acoustic data show no
significant sediment layering or changes over the planned drag
path of the anchor. In homogeneous soils, sediment parameters
measured on the surficial samples can be used to develop usable
prediction of soil strength over the full anchor penetration
depth.
In those situations where the sediment stratigraphy is
complicated, soil data may be required from samples taken from
deeper depths as, for example, in the following instances.
Where a thin 5- to 10-foot-thick layer of soft soil overlies a
hard layer, the anchor may drag in the soft surficial material but
be unable to penetrate into the underlying hard material A drag
anchor will also have some problems where a thin 2- to 5-foot
layer of sand overlies a soft silt or clay strata. In this soil
condition, the anchor fluke angle may have to be set for a sand
bottom to initiate dig-in. However, the anchor will then not
develop a high holding capacity in the underlying soft material.
Soil samples of the entire soil profile to be penetrated
will identify potential problems such as these and help in
selection of the most suitable drag anchor type and size.

7.3.4 Site Investigation Summary

Figure 7.3-1 presents a flow chart for developing a


Figure 7.3-1. Site survey plan decision flow chart.
geotechnical site survey plan for drag anchor design. It may be
advantageous or necessary to accomplish some of the on-site
tasks according to a different schedule based on known site
conditions or on the availability and cost of the survey and
positioning equipment and the vessel support.

7-10
7.4 SELECTING A DRAG ANCHOR To aid the anchor selection process, Table 7.4-1 rates
important anchor types, based on reports of field experience
7.4.1 General
with prototype anchors and on test findings with small anchors.
Designing an anchor system is a two-step process in Separate ratings are given for (1) tripping reliability and dig-in
which: performance, (2) roll stability during setting and dragging, and
(3) holding capacity "efficiency" (the ratio of holding capacity
One or more drag anchor types are selected for use
based on expected overall performance, availability, to anchor air weight). The anchors are rated as high, medium, or
and cost low for each category. Where data on a particular anchor type
The selected anchor type is then sized to develop the are not complete, the ratings have been partly based on the
required holding capacity. performance of geometrically similar anchor types.

Table 7.4-1. Rating of Drag Anchor Types Based on Tripping and


Dig-In, Roll Stability, and Holding Capacity Efficiency

7-11
7.4.2 Tripping and Penetration Performance These ratings can also be expected to vary for those soil profiles
which are more complex.
The ratings in Table 7.4-1 on tripping and dig-in
In soft soils (normally consolidated soft clays and silts),
performance are for anchors with fluke angles set according to
the STEVMUD, a specialized soft soil anchor, has the highest
recommendations for cohesive soils (clays and silts) or
holding-capacity-to-weight ratio. Next highest in developed
cohesionless soils (sand) respectively. (Mud palms are used for
holding capacity efficiency are the streamlined, deep-digging
the STATO on soft clay. )
anchors of substantial fluke area, including the STATO,
In soft clays, the tripping/dig-in rating is also a function
HOOK, BRUCE Twin Shank, and the STEVFIX anchors. Cast
of the installation method. If the anchor is installed using two
large anchors of the Moorfast and Offdrill II type exhibit lower
work platforms to keep the flukes open before touchdown as
holding capacity efficiencies because they are less streamlined
shown in Figure 7.2-2, then most anchors will trip and dig in
and do not penetrate as deeply in the sediment profile, and
properly. However, some anchors with limited roll stability
because their fluke areas per pound of steel are smaller.
may still roll during setting and may pull out. Once back at the
In sand, the STEVFIX, STEVDIG, STEVPRIS, BRUCE
seafloor surface the probability of anchor re-embedment would
(cast and twin shank), BOSS, and STATO anchors all perform
be small.
quite well, with no significant differences apparent among
7.4.3 Stability Performance them. Ine cast, bilateral anchors of the Moorfast and Offdrill II
type are projected as exhibiting good holding capacities, but not
The stability performance ratings presented in Table 7.4-1
as high as those developed by the anchors of newer design.
are based on recent field tests where the anchors were
instrumented to measure roll [Ref 7-11, 7-12] or are based on 7.4.5 Selection of Anchor Type
model tests in sand. Since the number of anchor tests used to
Selection of the anchor type to be used for a given
rate stability is small and the stability of an anchor is statistical
mooring should be based in large part on the considerations
in nature, the ratings should be treated as a best estimate -- one
previously discussed. The designer may wish to apply weighted
subject to revision as more data become available.
numerical ratings to the anchor types for each of the
7.4.4 Holding Capacity Performance performance characteristics to provide a "score" for each anchor
type. The designer must also consider other very important
The Table 7.4-1 holding capacity ratings are estimates
factors of (1) availability, (2) hardware purchase price, (3)
based on the measured performance of some of the anchor types
transportation, (4) vessel space requirements, and (5) handling
in comparative field tests [Ref 7-12, 7-11] and on the predicted
ease.
performance for the other anchor types based on their geometric
similarity to those which were field tested. This presentation of 7.5 SIZING A DRAG ANCHOR
drag anchor rating based on holding capacity is highly
7.5.1 Efficiency Ratio Method
generalized also because it is based on performance in two
uncomplicated soil profiles. These holding capacity ratings can The prediction of drag anchor capacities has traditionally
be expected to change and become more well defined as been by empirical approaches. That is, holding capacity is based
technical understanding and the predictive capability regarding on field experience with anchors. The most widely used
drag anchor performance improve.

7-12
of these approaches is the simple efficiency ratio method. In 7.5.2 Power Law Method
this method, the anchor's efficiency, e, is defined as a ratio of
The power law method is recommended for use as the
the horizontal load resistance developed divided by the anchor's
best method for predicting the holding capacity of drag-
weight (also called the "holding-capacity-to-weight ratio").
embedment anchors. It accounts for the nonlinear increase in
holding capacity with increasing anchor air weight [Ref 7-1].
This method produces a straight-line relationship between
(7-1)
anchor holding capacity and anchor weight on a log-log plot.
where: HM = holding capacity (horizontal load resistance at The validity of the power law method has been demonstrated in
the seafloor) field tests [Ref 7-15, 7-16].
WA = weight of the anchor in air
Expressed as an equation, the holding capacity, HM, is

This method assumes that the drag anchor efficiency is a determined by:

constant for a given soil type over a wide range of anchor sizes
(7-2)
(weights). It also assumes that the mooring leg is properly
designed and the anchor is properly installed and the anchor is
where: HR = holding capacity of a 10,000-lb air weight
dragged the necessary distance to develop its maximum holding
version of the reference anchor type [F]
capacity. WA = anchor weight in pounds for which HM is to be
determined
The efficiency ratio method remains widely used because
b = an exponent constant depending on the anchor
of its simplicity and familiarity within the user community. and soil
However, comprehensive field tests have shown that the method
Anchor holding capacity field test data have been used to
may produce unsafe holding capacity predictions for large
develop values for HR and b used in Equation 7-2. These data
anchors. In these tests, the efficiency was shown to decrease
with increasing anchor weight [Ref 7-1]. Therefore if an were obtained from 1950 NCEL testing of Navy Stockless,

efficiency constant was used to project anchor holding capacity STATO, and LWT anchors [Ref 7-14, 7-17 through 7-20]; from

for a larger anchor, the projection would overpredict that Exxon Production Research testing of LWT and BOSS anchors

capacity. [Ref 7-15, 7-16], from 1979-80 NCEL testing of Navy

To develop the holding capacities predicted by Equation Stockless, STATO, Moorfast, BRUCE, and STEVIN family

7-1 or by the two anchor holding capacity predictive techniques anchors [Ref 7-12, 7-11]; and from some offshore industry

given in Sections 7.5.2 and 7.5.3, the mooring line tension tests. The test data and previous interpretations [Ref 7-1, 7-15,

must be applied to the anchor system IN A DIRECTION 7-16] have resulted in the values of HR and b shown in Table

PARALLEL TO THE SEAFLOOR. This is accomplished, in a 7.5-1 for a wide range of drag anchors used in cohesive and

proper design, by using sufficient chain weight (sometimes cohesionless soils. Table 7.5-1 and Equation 7-2 can be used to

accompanied by a concrete sinker weight) in the mooring line predict the holding capacity for any weight of the anchors listed.
leg to keep the chain angle zero even at the highest load The mooring leg, which includes the anchor, however, must be
condition. In addition to this, an extra shot (90 feet) of heavy properly designed and the anchors must be properly installed.
chain is often added at the anchor end to ensure that some chain Figures 7.5-1 and 7.5-2 are a graphical presentation of the data
remains on the seafloor. from Table 7.5-1 for cohesive soils (clays and silts) and
cohesionless soils (sands), respectively.

7-13
Table 7.5-1. Parameters HR and b used in Equation 7-2:
HM = HR (WA/10,000 lb)b

7-14
Figure 7.5-1. Anchor chain system holding capacity at the mudline in soft soils.

7-15
Figure 7.5-2. Anchor chain system holding capacity at the mudline in hard soils.

7-16
It is noted that the curves for the Flipper Delta anchors are not diameter wire rope should allow the anchor to penetrate deeper
the result of field tests but have been estimated. into stronger soils, and an increased anchor resistance at that
In Figure 7.5-1, the curves describe the holding capacity depth is expected to compensate for any reduction in resistance
of drag anchors in normally consolidated clays and cohesive from the smaller mooring line.
silts. These curves were developed directly from measured
7.5.3 Analysis Based on Geotechnical Considerations
holding capacities. Note the difference in performance for the
Stockless anchor with flukes in both fixed and movable A complete method for the prediction of drag anchor
positions. The Stockless, with flukes left movable, will holding capacity based on geotechnical considerations is not
develop only half the capacity of the anchor with flukes fixed at presently available. The prediction of drag anchor and mooring
50 degrees in soft soils [Ref 7-21]. line penetration is beyond present analytical capability.
Figure 7.5-2 presents an interpretation of data from field However, recent work has improved the ability to predict
tests in several sand types. Only one curve is used to describe capacity based on geotechnical considerations for anchors
the performance of an anchor type for all types and densities of embedded in soft soils. The technique for calculating anchor
sands. A more accurate predictive scheme for these variations holding capacity is presented here for use with an established
has not yet been developed. While the holding capacity anchor, where the depth of penetration is known.
developed with a given anchor is expected to vary somewhat
with sand type and density, the relative capacities should remain 7.5.3.1 Calculations for Soft Cohesive Seafloors.
about the same. The predictive curves of Figure 7.5-2 are also Most soft cohesive seafloors defined as mud or soft silt to clay
believed applicable in very stiff to hard clays (i.e., in all hard size sediments, are normally consolidated to slightly
seafloors). overconsolidated. Drag-embedment anchors will penetrate deeply
In Table 7.5-1, parameters were assigned to some anchor in these sediments. In this situation, the drag anchors can be
types where data were not available to develop curves. Based on expected to behave similar to deeply embedded plate anchors.
geometric similarities, the Danforth, G.S., Offdrill II, Flipper Drag anchor flukes differ considerably in their plan shape.
Delta, and STEVIN have each been assumed to behave like one In order to adequately describe the projected plan area of the
of the other anchors which were tested and to have similar different shapes, a correction factor is applied to the anchor
values of holding capacity. fluke dimensions. The holding capacity of a drag anchor is then
The power law method, as used here, includes both buried expressed as:
chain and anchor holding capacity, but makes no separate
accounting of anchor and chain contributions. The method HA = Nc (f B L) s u (7-3)

assumes that the chain size used is compatible with the holding
where: Nc = a constant, depending on the failure mode
capacity to be developed. Larger chain sizes may develop around the anchor and on the anchor geometry
somewhat higher holding capacities than those predicted by [see Table 7.5-2]
f = Correction factor converting the rectangular area
using Table 7.5-1 [Ref 7-21]. The use of wire rope for the B x L to the actual projected fluke area [see
Table 7.5-2]
embedded portion of the mooring line, despite the smaller B = width of fluke [see Figure 7.1-5] [L]
projected area of comparable chain, is not expected to L = length of fluke [see Figure 7.1-5] [L]
su = undrained shear strength of cohesive soil at
significantly reduce anchor system holding capacity. The center of area of anchor [F/L2]
smaller

7-17
Table 7.5-2. Parameters N c and f Used for Clays and
Cohesive Silts in Equation 7-3:
HA = Nc (f B L) s u

Anchor Type Nc f
Stockless, fixed flukes 13.0 0.54

Danforth 11a 0.60

LWT 11a 0.60

STATO/NAVMOORb 12 0.95

Moorfast 12 0.95

Offdrill II 12a 0.95

STEVFIX 6.4 0.72

STEVMUD 6.8 0.77

Hook 6.5 0.80

BRUCE Cast 4.0 0.36

BRUCE Twin Shank 6.5 0.52


a Estimated values.
bNAVMOOR has a configuration similar to that of the
STATO anchor.

This predictive equation does not provide a complete maximum holding capacity occurs even before the entire fluke
solution to the problem of designing a mooring system because is embedded. Presently, a predictive equation which uses
the depth of penetration of the drag anchor/chain system is geotechnical considerations is not available.
beyond present predictive capability (save for the estimates
7.5.4 Factor of Safety
shown in Table 7.3-1). For established anchors, where anchor
penetration can be measured from a pendant line, Equation 7-3 The holding capacities predicted in the previous sections
can be utilized for evaluating the holding capacity. can be considered minimum ultimate holding capacities for
Equation 7-3 predicts only the holding capacity developed design purposes. When selecting an anchor size, a factor of
at the shank-to-mooring-line connection point. The section of safety must be applied to the design loading to arrive at a
mooring line embedded in the seafloor by the penetrating drag required minimum ultimate holding capacity. It is recommended
anchor makes a significant contribution to the anchor system's that a safety factor of 1.5 be applied to the peak load from
holding capacity. Based on available field data [Ref 7-12, 7-11], survival conditions, such as the 100-year storm. A factor of
the mooring line contribution, Hc , in soft muds can be safety of 2.0 is recommended for all conditions during which
conservatively estimated to be 0.2 HA. operations might be on-going. These recommendations follow
those of the API [Ref 7-23] for pile design and are recommended
7.5.3.2 Considerations for Sands and Stiff Clays. to ensure the anchor drags before the mooring line breaks.
Field measurements [Ref 7-12, 7-11, 7-22] show that drag
anchors do not penetrate deeply in sand and stiff clays. Often the

7-18
7.6 TROUBLESHOOTING anchors and identified several penetration and stability problems
that could occur in the use of drag-embedment anchors. This
Section 7.5 offers techniques for predicting the potential
section summarizes these problems and potential solutions in a
holding capacity of drag-embedment anchors, assuming proper
troubleshooting procedure. These are outlined in Table 7.6-1.
penetration and stability of the drag anchor system. Section 7.2
discussed the functioning of drag

Table 7.6-1. Troubleshooting Procedures for Correcting Drag Anchor


Performance Problems

Troubleshooting Proceduresa for--


Problem Cohesive Soils Cohesionless Soils
Failure to 1. Proper setting 1. Sharpen flukes
Initiate procedure used? 2. Reduce fluke angle to a
Penetration 2. Fix fluke in open minimum 27 deg
position 3. Increase stabilizer length
3. Increase stabilizer 4. Place barbs on tripping
length palms
5. Fix fluke open
6. Crush hardened surface soil
with explosives
Failure to 1. Recover and clean; 1. Increase stabilizer length
Develop Expected check for balling up; 2. Reduce fluke angle -to a
Holding Capacity clean off stud ball minimum 27 deg
2. Reset and soak 24 hr 3. Piggyback or change to
3. Piggyback or change to larger anchor
larger anchor
a Troubleshooting procedures listed in order of recommended
implementation

7.6.1 Soft Sediments setting procedure [Figure 7.2-2]. When limited support prevents
use of this procedure, then correction of a tripping problem
Drag anchors in soft seafloors may encounter (1) tripping
requires fixing of the anchor fluke in its open position and,
problems, (2) instability problems, or (3) inability to develop
possibly, the lengthening of stabilizers to help right the anchor
sufficient holding capacity without allowing for sediment
[Figure 7.2-5(b)].
strength gain through anchor soaking. Figure 7.6-1 illustrates
Instability problems are suggested by a rise in the line
the effects of tripping and instability problems on the line
tension followed by a fall to the "failure-to-trip level" [curve b
tension developed while setting an anchor. Failure to trip and
in Figure 7.6-1]. When the anchor behaves in this manner and
initiate penetration is suggested when line tension remains
fails to develop adequate capacity, the anchor should be
nearly constant at one-half to two times the combined weight of
recovered and cleaned. In soft soils it is likely that a mud ball
the anchor and the amount of mooring line on the seabed [curve
has formed on the anchor flukes. In hard soils, the fluke angle
c in Figure 7.6-1]. Tripping problems are normally avoided by
should be reduced. The anchor should be reset and dragged until
using a proper anchor
the line

7-19
Figure 7.6-1. Typical performance of drag anchors when operating properly and improperly.

tension approaches its prior peak value. The line tension should stabilizer length. A decrease of the fluke angle to increase
then be slacked off and the anchor allowed to soak for 24 hours. penetration has also proven beneficial to stability. Should these
If this delay is not possible or if, after soaking, the required line steps fail in gaining the required line tension, larger anchors or
tension cannot be developed, then an anchor of higher capacity piggybacked anchors will be required. In extreme cases
will be required or the original anchor must be piggybacked sufficient anchorage will require shifting to use of a pile.
(Section 7.7).
7.7 PIGGYBACKING
7.6.2 Hard Sediments
7.7.1 Field Practice
Tripping, penetration, and instability problems can occur
When an anchor will not develop the required capacity on
on hard seafloors (including all sands, gravel, and hard clay).
being proof-loaded, it is common practice to install a second
Tripping is enhanced by sharpening the fluke tips. Penetration
anchor on the same mooring leg, in-line, and beyond the first or
or dig-in is enhanced by reducing the fluke-to-shank angle to as
primary anchor. This is called a piggyback anchor, and
small as 25 to 27 degrees and increasing the stabilizer length to
sometimes called a tandem or backup anchor. The procedure for
keep the anchor from sliding on its side [Figure 7.2-3]. In
installing a piggyback anchor varies.
extreme cases, barbs have been added to the tripping palms to
In the offshore oil industry, many contractors simply
increase the tripping moment, and flukes have been fixed in the
detach the existing pendant line surface buoy from the pendant
open position. On some hard surfaces (e.g., coral or weak rock)
line wire rope of the first anchor [see Figure 7.7-1] and reattach
these measures cited to initiate penetration will not suffice.
the line to a second anchor. The second anchor is then lowered
Shaped charges have been used on these surfaces to crush the
and set by a second pendant line, while the line to the first
material in front of the anchor and provide an area for fluke
anchor is kept tightly stretched. Sometimes it is necessary to
embedment.
retrieve the first anchor, attach a heavier wire
The primary means for increasing anchor stability in hard
seafloors is increasing the

7-20
Figure 7.7-1. A pendant line and buoy arrangement for semisubmersibles (from Ref 7-24).

rope to connect the two anchors, and set both anchors in


sequence. In some cases a chaser system [Figure 7.7-2] is used
to install and recover the primary anchors. When using a chaser
system the piggyback anchor is attached directly to the chaser
wire. In all of the above techniques, the primary anchor is
dragged some short distance to set it before the piggyback
anchor is lowered. The Navy, working from available anchor
inventory, has found piggybacking necessary for high capacity
permanent ship moorings. Techniques for laying the mooring
leg in water depths to 100 feet using a crane barge have been Figure 7.7-2. Chain chaser used to assist anchor
demonstrated [Ref 7-11]. deployment and recovery.

7-21
7.7.2 Results and Field Problems

The results from piggybacking are strongly dependent on


how the primary and secondary anchors are attached to the
mooring line. Pendant lines are usually attached to the anchor at
the back of the fluke where possible or at the crown end of the
shank. This produces mixed results, illustrated by Figure 7.7-3.
With many anchors used in this arrangement, the line tension
from the piggyback anchor will cause rotation and breakout of
the primary anchor [Ref 7-25]. With other anchors, this
arrangement works fine and results in a piggyback system
capacity equal to or greater than the sum of the holding
capacities of the two anchors loaded separately [Ref 7-26]. Once
dislodged, however, primary anchors with movable flukes and
with load applied directly to the anchor crown will not dig in
again, but will slide with the flukes held parallel to the shank
by the line tension applied at the pendant padeye [Figure 7.7-
3(a)]. Navy tests in sand and mud have shown that the procedure
shown by Figure 7.7-3(b) is suitable provided the inbound
anchor is very stable (e.g. BRUCE, Hook, STATO,
NAVMOOR) and anchor flukes are prevented from closing up
under load. Anchors whose center of fluke area is close to the
fluke-to-shank connection point (e.g., Danforth, LWT,
STEVIN types, Flipper Delta) are inherently less stable anchors
and are not appropriate for shank-to-crown piggyback
connections.
For those anchors where the crown attachment (of the
piggyback anchor) creates problems, model tests have shown
that primary anchor stability will not be significantly affected Figure 7.7-3. Tandem/piggyback anchor arrangements
when the line to the piggyback is attached to the shackle end of (from Ref 7-25).
the shank of the primary anchor [Figure 7.7-3(c)]. The holding
capacity of the anchors attached this way reached the sum of the show that the anchors tend to come together, and that load
holding capacities of the individual anchors. This attachment equalization between anchors is a problem [Ref 7-19]. Once the
technique should be used only with fixed-fluke anchors to anchors do come together they will foul and will not re-embed.
minimize potential for fouling of the second anchor wire or If used in this manner, the parallel anchors should be staggered
chain with the primary anchor [Ref 7-25]. by using different length chain legs to avoid anchor interference
Parallel tandem anchor arrangements have been suggested Navy
[Figure 7.7-4]. Full-scale tests

7-22
7.8 REFERENCES

7-1. Design manual. Harbor and coastal facilities, Naval


Facilities Engineering Command, NAVFAC DM-26.
Washington, D.C., Jul 1968.

7-2. "Rules for the design, construction, and inspection of


offshore structures," Appendix E of Hydrostatic Stability and
Anchoring, Det Norske Veritas. Oslo, Norway, 1977.

7-3. Rules for building and classing mobile offshore drilling


units. New York, N.Y., American Bureau of Shipping, 1980.

7-4. API recommended practice for the analysis of spread


mooring systems for floating drilling units, American
Petroleum Institute, API RP 2P. Dallas, Tex., May 1981.

7-5. M.A. Childers. "Spread mooring systems," in The


Technology of Offshore Drilling, Completion and Production,
compiled by ETA Offshore Seminars, Inc. Tulsa, Okla.,
Petroleum Publishing Company, 1976, pp 94-128.
Figure 7.7.4. Parallel anchor arrangement.

7-6. ____________. "Deep water mooring - Parts I, II, and


tests [Ref 7-26] with staggered anchors (separated by at least
III," Petroleum Engineer, Sep and Oct 1974 and May 1975.
four fluke lengths) have shown system capacities 15 to 20%
greater than the sum of the individual anchor capacities. 7-7. S.A. Fjeld. "Certification aspects," in Anchoring of
Offshore Structures, proceedings of symposium organized by
7.7.3 Recommended Practice
Q.M.C. Anchor Technology Ltd., Sep 1980, pp 6/1-6/26.
Careful placement of piggyback anchors is required so
7-8. Vryhof ankers. The Netherlands, Vryhof Ankers, 1980.
that orientation and stability are properly controlled.
Attachment of the piggyback anchor at the crown is proper
7-9. Bruce International Ltd. The Bruce anchor. Isle of Man,
when the primary anchor is of the stabilized Stockless, Hook,
England, 1978.
STATO, or BRUCE type. Attachment of the piggyback anchor
at the shackle end of the primary anchor is proper when the 7-10. Baldt Inc. Baldt. Chester, Pa., 1977
primary anchor is one of the other anchor types. The primary
7-11. R.J. Taylor. Conventional anchor test results at San
anchors should be set and be well-stabilized before tension from
Diego and Indian Island, Civil Engineering Laboratory,
the piggyback anchor is applied.
Technical Note N-1581. Port Hueneme, Calif., Jul 1980.

7-23
7-12. R.J. Taylor. Test data summary for commercially 7-22. I. Foss, T. Kvalstad, and T. Ridley. "Sea bed anchorages
available drag embedment anchors, Civil Engineering for floating offshore structures," FIP Commission on Sea
Laboratory. Port Hueneme, Calif., Jun 1980. Structures, Working Group on Foundations, Feb 1980.

7-13. K.J. Saurwalt. "Movements and equilibrium of anchors 7-23. API recommended practice for planning, designing, and
holding on an impervious sea bed, section I," Schip en Werf, constructing fixed offshore platforms, American Petroleum
Rotterdam, no. 9, 1971, pp 627-634. Institute, API RP 2A. Dallas, Tex., Jan 1982.

7-14. R.C. Towne and J.V. Stalcup. New and modified anchors 7-24. P.G.S. Dove. "Methods in anchor handling," Offshore,
for moorings, Naval Civil Engineering Laboratory, Technical Mar 1980, pp 114-128.
Report R-044. Port Hueneme, Calif. , Mar 1960.
7-25. P.J. Klaren. "Anchors in tandem or the use of back-up
7-15. R.W. Beck "Anchor performance tests," in Proceedings of anchors (piggybacks)," Holland Shipbuilding, Anker Advies
the Offshore Technology Conference, Houston, Tex. 1972, pp Bureau, pp 230-232.
268-276.
7-26. R.J. Taylor and G.R. Walker. Model and small-scale tests
7-16. M.W. Cole and R.W. Beck. "Small-anchor tests to to evaluate the performance of drag anchors in combination,
predict full scale holding power," Society of Petroleum Naval Civil Engineering Laboratory, Technical Note N-1707.
Engineers, SPE 2637, 1969. Port Hueneme, Calif., Oct 1984.

7-17. R.C. Towne and J.V. Stalcup. Tests of BUSHIPS


anchors in mud and sand bottoms, Naval Civil Engineering 7.9 SYMBOLS
Laboratory, Technical Note N-195. Port Hueneme, Calif., Aug
1954. B width of anchor fluke [L]

b An exponent constant in the holding capacity predictive


7-18. R.C. Towne. Test of anchors and ground tackle design, relationship; value varies with anchor/soil combination
Naval Civil Engineering Laboratory, Technical Memorandum [Table 7.5-1]

066. Port Hueneme, Calif., Jun 1953. dt Depth of penetration of the fluke tip [L]

e Efficiency of drag anchor system, based on anchor air


7-19. R.C. Towne and J.V. Stalcup. Tests of moorings and weight
ground tackle design in mud bottom, Naval Civil Engineering
f Correction factor converting the rectangular area B x L to
Laboratory, Technical Memorandum 097. Port Hueneme, the actual projected fluke area
Calif., Dec 1954. HA Drag anchor holding capacity (without chain
contribution) [F]
7-20. R.C. Towne. "Mooring anchors," paper presented at the
HC Contribution of embedded chain to holding capacity of
Annual Meeting, The Society of Naval Architects and Marine drag anchor system [F]
Engineers, New York, N.Y., 1959. HM holding capacity of drag anchor system, anchor plus
chain [F]
7-21. R.J. Taylor. "Performance of conventional anchors," in
HR Holding capacity of a 10,000-lb air weight version of the
Proceedings of the Offshore Technology Conference, Houston, reference anchor, WA [F]
Tex., 1981, pp 363-372.

7-24
k A constant in anchor holding capacity predictive equation
for sands; value varies with crown embedment, soil
friction angle and effective unit weight, and anchor shape
[F/L3]

L Length of anchor fluke [L]

NC A constant in anchor holding capacity predictive equation


for clays; varies with anchor geometry

su Undrained shear strength of soil [F/L2]

WA Air weight of anchor (does not include chain weight) [F]

Angle between fluke and shank of drag anchor [deg]

7-25
Chapter 8

PENETRATION OF OBJECTS INTO THE SEAFLOOR

8.1 INTRODUCTION 8.1.2.3 Penetrator Velocities. The techniques presented

8.1.1 Purpose are applicable for penetration velocities from near zero
(normally called "static" penetration) to 400 fps. This
This chapter presents techniques for predicting the depth
encompasses the full velocity range expected in Navy deep
of penetration of objects pushed into or impacting on sediments
ocean applications.
found in the deep ocean. The techniques presented can also be
used to predict the force required to push an object to a specified 8.2 STATIC PENETRATION
depth within the seafloor and to predict the sediment-related
8.2.1 Application
forces acting on a rapidly moving object after it impacts the
seafloor surface. The term "static penetration" identifies a category of soil
penetration events in which the objects' penetration velocity on
8.1.2 Scope impact with the seafloor is less than 3 fps. For velocities less
than this, fluid drag force in the water at the time of impact,
8.1.2.1 Seafloor Types. Techniques presented here are
soil strain rate, and inertia factors have negligible influence on
limited in application to seafloors of unlithified sediments (i.e.,
penetration. This static problem is adequately treated by
the terrigenous and pelagic clays and silts, sands, and deep sea
conventional pier and pile bearing capacity methods. Objects in
oozes). Special techniques not covered here, are required to
this category are footings, structure bearing pads, mud mats,
predict penetrations in lithified sediments, coral basalt, and
raft foundations, gravity anchors, spud cans, jacked piles for
other rock types. These special techniques are highly empirical
offshore platforms, and many bottom-resting devices used to
and are normally limited in application to a particular projectile
test for properties of seafloor sediments.
[Ref 8-1]. Some discussion on penetration of propellant-
embedded anchors in coral and rock is given in Chapter 6. 8.2.2 Approach

8.1.2.2 Penetrator Types. The prediction techniques The static penetration is subdivided into shallow and deep

used are applicable to objects of all shapes and sizes, ranging penetration cases, as defined by the ratio of an objects

from long streamlined objects such as instrumented penetration depth to its diameter (or minimum lateral dimension

penetrometers, free-fall corers, and propellant-embedded anchor if not circular), z/B. In shallow penetration, this ratio is less

flukes, to large blunt objects such as gravity anchors, structure than 2.5. The shallow condition is treated as a

bearing pads, and ship hulls.

8-1
conventional bearing capacity problem with the influence of b = buoyant unit weight of soil [F/L3]
friction on the sides of the penetrator set equal to zero [Ref 8-2, z = depth of penetration of the effective bearing
surface [L]
8-3] (see Section 8.2.3 and Figure 8.2-1). Nc ' = a dimensionless bearing capacity factor
determined by

Nc ' = 5 [1 + 0.2(B/L)] [1 + 0.2(z/B)] (8-2)

where: maximum N c ' = 9


B = penetrator diameter, or the minimum
value of the bearing area horizontal
dimensions
L = penetrator diameter, or the maximum
value of the bearing area horizontal
dimensions

For cohesionless soils (e.g., coarse silts, sands, and gravels),


the resistance to penetration is described by the relationship
[Ref 8-4]:

Qu = 0.5 At (B b Nq) (8-3)

Figure 8.2-1. Shallow static penetration model (after


Ref 8-2 and 8-3). where: Nq = bearing capacity factor
= penetrator bearing area shape factor

In deep penetration (i.e., where z/B 2.5), the influence Figure 8.2-2 presents curves for obtaining values for N q and .
of side friction on the penetrator becomes substantial and must These values depend on , the soil friction angle, and on KH,
be included in computations. This is done by the same methods
the coefficient of lateral earth pressure. Both values are difficult
used for predicting pile capacities. Techniques for predicting
to obtain. In-situ measurement is not likely to be possible, and
deep static penetration are presented in Section 8.2.4.
material sampling with subsequent laboratory testing is difficult
because of the easily disturbed nature of cohesionless soils.
8.2.3 Method for Predicting Shallow Static Penetration
Unless other information is available, the values shown for KH
8.2.3.1 Descriptive Equations. For cohesive soils
under average conditions in Figure 8.2-2 should be used. A best
(e.g., clays, muds, fine silts), the resistance to penetration is estimate for should be made. In Figure 8.2-2, graphical
described by the simplified relationship [Ref 8-3]: values of Nq can be obtained for = 30 degrees and = 40

degrees. For most fine sands and silts, a friction angle of 30


Qu = At (su Nc + b z) (8-1)
degrees can be considered typical of a low to medium relative
where: Qu = bearing force resisting penetration [F] density. An angle of 40 degrees can be considered typical of a
At = end area of penetrator (the effective bearing very high relative density.
surface) [L2]
The bearing capacity factors, Nc ' and Nq, used in
su = soil undrained shear strength - in this equation,
measured at a depth B/2 below z [F/L2] Equations 8-1 and 8-3, respectively, are both functions of the
depth of penetration, z. To determine the maximum depth of
penetration of

8-2
Figure 8.2-2. Relationships for calculating sinkage resistance in cohesionless soils; for 30 o and 40o (after Ref 8-4).

an object subject to a given force, the applicable equation must depth increment and then increasing this increment each time
be solved by trial-and-error. An assumed depth is entered into the resistance forces are calculated. In making these
the equation, and it is solved to find the corresponding resisting calculations, it is helpful to take the assumed penetration depth,
forces at that penetration. If the resistance forces are smaller z, and corresponding penetration resistance, Qu, of each
than the penetration force, then a deeper penetration is assumed, calculation set and develop a plot of z versus Qu. Then,
and the forces are recalculated. knowing the force available to cause object penetration, the
Although the solution is actually trial-and-error, it is expected depth of penetration can be obtained from the plotted
often useful to use an iterative approach. This is done by curve. Otherwise, the expected depth can be determined from
selecting a direct interpolation between the

8-3
values of Qu higher and lower than the driving force. Should the method for predicting deep static penetration in Section 8.2.4).

ratio z/B exceed 2.5, then the soil friction on the sides of the A flow chart of the calculation process for both shallow and

object must be considered in the calculation (see the deep penetration is given in Figure 8.2-3.

Figure 8.2-3. Flow chart of the calculation procedure for predicting static penetration.

8-4
8.2.3.2 Influence of Penetrator Shape. Tapered or 8.2.3.3 Strength Parameter Selection. Methods for

rounded penetrators that present an increasing width and area the development of a soil strength profile for undrained shear
during penetration should have this increasing area considered at strength, su, and effective stress strength parameters, c and ,
each new calculation. The increasing area affects the bearing are given in Chapters 2 and 3. These methods are applicable to
force, not only directly through the increased bearing area, but the penetration problem when dealing with relatively uniform
also by changing the ratio, z/B, determining the bearing deposits of clays or sands. However, deep uniform sand deposits
capacity factors. are not found often in the deep sea environment. Those sands
The penetrator nose shape also influences the shape of the that do occur are normally the lower portions of turbidite layers
soil failure zone mobilized to resist the penetration. For blunt and have interbedded layers of graded silt to clay sized soil. For
objects, the soil zone governing the penetration performance such complex sediment profiles, present technology is not
reaches a depth of one object width or diameter below the sufficiently developed to describe the penetration phenomena.
bearing surface. The average of the soil strength values, su, over Given the present state-of-the-art, such complex profiles should
a depth interval of B below z is used in the bearing capacity be treated as cohesive soil when excess penetration is of
equation [Figure 8.2-4(a)]. For tapered objects, such as the primary concern and as cohesionless soil when inadequate
conical-pointed spud cans on some offshore jack-up platforms, penetration is the primary concern.
su is determined over a depth, B, measured from the bottom of
8.2.3.4 Prediction Accuracy. Some field data are
the cylindrical or full section [Ref 8-3] [Figure 8.2-4(b)], rather
available to examine the accuracy of Equation 8-1 in predicting
than from below the point of tip penetration.
penetration. Equation 8-1 was used to predict the penetration of
spud cans for offshore jack-up platforms [Ref 8-3] in clay. Out
of 120 sets of data, 70% of the predictions were found to be
within 25% of the measured penetrations. No prediction was
less than 50% of the measured value, and only five predictions
exceeded the measured value by more than 50% [Ref 8-3].
Penetration predictions in sand, however, will be considerably
less accurate due to difficulties in obtaining accurate in-situ soil
strength parameter data in cohesionless soils.

8.2.4 Methods for Predicting Deep Static Penetration

8.2.4.1 Descriptive Equations. Where z/B 2.5, the

soil friction on the walls of the penetrator becomes a significant


resistance factor and must be considered. Such penetrations are
not normally the case with initial penetration velocities of less
than 3 fps except for slender, very heavy objects or for objects
that

Figure 8.2-4. Location of the critical shear strength


zone B for blunt and conical penetrators
(after Ref 8-3).

8-5
are being pushed by external forces (e.g., the spud cans of a compared to the known driving forces (object underwater weight
jack-up rig or soil test devices). and any external force), and new selections are made for the
In cohesive soils, when z/B exceeds 2.5, Equation 8-1 is penetration depth to improve the agreement between calculated
supplemented with a side friction term to become: resistance and known driving forces. A block diagram of this
calculation process is given in Figure 8.2-3.
Qu = At (su Nc + b z) + A s s u (8-4)
8.2.4.2 Skirt Penetration Prediction. The offshore oil
where: As = side area in contact with the soil [L2]
industry has developed an alternative technique for the prediction
= adhesion factor (soil against object); use = 0.5
or = 1.0 as explained in the following of forces required to embed the thin-walled steel skirts and
paragraph dowels used beneath the bases of offshore gravity platforms
su = average soil undrained shear strength over the
side area in contact with the soil [F/L 2] [Ref 8-7]. Specific in-situ soil tests are required. Correlations
are developed with a Dutch cone penetration resistance measured
The adhesion factor is difficult to determine and is a function of during the tests [Ref 8-8]. The skirt tip resistance is:
the soil plasticity, shear strength, and degree of over-
consolidation [Ref 8-5, 8-6]. Use = 0.5 when too much qu = k t qc (8-6)

penetration is the primary concern, and use = 1.0 when


where: kt = tip correlation factor
inadequate penetration is the primary concern. Note that for
qc = Dutch cone penetration resistance obtained from
square and round penetrators, N c reaches its maximum value in-situ tests [F/L2]
(Nc = 9.0) at z/B = 2.5 and remains constant at higher values
The unit wall friction-is;
of this ratio. For other shapes this occurs at a slightly higher
z/B ratio. f = kf qc (8-7)

In cohesionless soils, when z/B exceeds 2.5, Equation 8-3


where kf is the friction correlation factor.
is supplemented with a side friction term to become:
Value of kt and kf are listed in Table 8.2-1. In general,
Qu = 0.5 At (B b Nq) + As KH p o tan the higher values listed should be used in predicting penetration
. . . . . . . . (8-5)
resistance. The lower values were found in the upper 3 feet of

where: KH = coefficient of lateral earth pressure (from Figure the sediments and are believed to reflect the effect of piping
8.2-2) from water trapped beneath the platform base. These
po = soil overburden pressure [F/L2]
= friction angle between penetrator and sand (for correlations are applicable to very slow penetration rates of 0.3
steel = - 5 deg) [deg] to 3 ft/hr [Ref 8-7].
= soil friction angle
Table 8.2-1. Correlation Coefficients Between Dutch Cone Penetration
Resistance and Thin-Walled Skirt and Dowel Penetration
Equations 8-4 and 8-5 for deep static penetration are used Resistance (from Ref 8-7)

to determine penetration depths as described in Section 8.2.3.1 Soil Description kt kf


Very stiff, silty clay 0.4-0.6 0.03-0.045
for shallow penetrations. First, a penetration depth is assumed;
Intermixed dense sand/clay layers 0.5 0.006-0.014
parameters are calculated; and a corresponding resistance to
Dense to very dense silty fine sand
penetration is calculated. This calculated resistance is 0.3-0.6 0.001-0.003

8-6
8.3 DYNAMIC PENETRATION layer. The calculation is repeated for each successive layer until
the kinetic energy of the penetrator has been consumed and its
8.3.1 Application
velocity reaches zero. The depth at zero velocity is the predicted
The methods presented in this section are used to predict penetration depth.
the penetration of objects entering the seafloor at velocities of 3 The distance penetrated at each step (z) must be assigned
fps and greater. Examples of objects that undergo dynamic prior to beginning the calculations. It is suggested that z be
penetration are: objects being lowered rapidly or free falling to assigned by dividing an expected (guessed) total penetration into
the seafloor, such as gravity anchors and ship hulls; other free- approximately 10 equal increments.
falling objects, such as gravity corers and penetrometers; and
objects propelled at higher speeds than terminal free-fall 8.3.3 Method for Predicting Dynamic Penetration
velocities, such as propellant-embedded anchor plates.
8.3.3.1 Forces Acting on the Penetrator. The forces

8.3.2 Approach acting on an object penetrating at moderate velocity deep in a


soil mass are shown in Figure 8.3-1. The net downward force
The technique presented here predicts both total
after full object entry into the soil at distance z i is:
penetrations and decelerating forces versus depth. Early versions
of this technique [Ref 8-9, 8-10, 8-11] have been modified [Ref F i = F di + W bi - Q ni - Fsi - Fhi (8-8)
8-12, 8-13] to adapt the technique to velocities up to 400 fps.
The approach is basically the same as that used for static where: F i = net downward force exerted by the penetrator [F]
F di = external driving force, if any (e.g. , rocket
penetration prediction, but is extended to account for strain rate motor) [F]
effects on the soil shear strength, remolding of the soil on the Wbi = penetrator buoyant weight [F]
Qni = tip or nose bearing resistance [F] (see Section
sides of the penetrator, and transition effects in passing from the 8.3.3.2)
overlying fluid drag regime into the soil penetration regime. F si = side friction or adhesion [F] (see Section
8.3.3.3)
An iterative procedure is used for the solution of the F hi = fluid drag force [F] (see Section 8.3.3.5)
dynamic penetration problem because the resisting force terms subscript i = ith increment of soil depth
for nose bearing, side friction, and hydraulic drag, as well as the Two of these forces, F di and Wbi, are driving the
driving kinetic energy term, are velocity dependent. The
penetrator into the soil mass. The other three, Qni, F si, and F hi,
procedure is to step the penetrator into the soil in equal finite
are resisting that penetration.
depth increments or layers (z). Resisting soil and
hydrodynamic forces are calculated on the basis of the entry 8.3.3.2 Nose Resistance. The tip or nose bearing
velocity and the soil properties for each step. The energy lost resistance force for the ith layer is obtained from:
by the penetrator in overcoming the first layer resistance forces
is calculated and subtracted from the penetrator kinetic energy Qni = s ui(nose) N ti A t (8-9)
on entering the underlying layer. The kinetic energy remaining
where: sui(nose) = soil undrained shear strength at a depth
with the penetrator and its reduced velocity are used for
D/2 below z, averaged over ith increment
computations on penetrating the second of penetration
= strain rate factor (to be defined in Section
8.3.3.4)
At = end area of penetrator (the effective
bearing surface)[L2]

8-7
the undrained shear strength of a sand can be found in Reference
8-14. (It is a very difficult property to measure or estimate.)

8.3.3.3 Side Friction. The side friction or adhesion

force in the ith layer is obtained from [Ref 8-12]:

(8-11)

where: S ti = soil sensitivity or ratio of undisturbed


undrained shear strength to remolded
undrained shear strength (obtained normally
from soils testing)
As = side area of the penetrator [L2]
sui(side) = soil undrained shear strength averaged over
the length of the penetrator in contact with
the soil

This assumes no separation between the soil and the side of the
object being penetrated during penetration.

8.3.3.4 Strain Rate Factor. Cohesive soil undrained

shear strength increases with an increase in the rate of strain.


This increase may be as high as a factor of 5.0 when soil shears
Figure 8.3-1. Forces acting on a penetrator before in response to a rapidly penetrating object [Ref 8-11, 8-15].
and after contact with the seafloor.
This is more commonly referred to as the strain rate effect on
shear strength. A somewhat cumbersome formulation, based on
and Nti is a dimensionless nose resistance factor determined by: a best fit to penetration test data [Ref 8-12, 8-13], has been
developed for calculating the strain rate factors :
Nti = 5 [1 + 0.2(B/L)] [1 + 0.2(zi/B)] (8-10)

(8-12)
where maximum Nti = 10.0.

where minimum = 1 and


The term s ui represents the undrained shear strength of

sands as well as clays. Dynamic penetration in sands is rapid


= maximum strain rate factor, from Table 8.3-1
enough to be considered undergoing an undrained type of shear = empirical strain rate coefficient, from Table 8.3-1
failure. When sands are sheared during a dynamic penetration [FT/L2]
event, the pore water does not have time to flow, and failure vi-1 = velocity entering the ith layer [F/L]
Co = empirical strain rate constant, from Table 8.3-1
occurs when either the sand grains are crushed or cavitation of
sui = soil undrained shear strength equal to sui(nose) or
the pore water occurs. The forces required to do this are large sui(side), depending on which of these that is
and hence the undrained shear strength is high (e.g., on an order modifying

of magnitude of 50 psi). A method for obtaining

8-8
Table 8.3-1. Values of Constants Used in Equation 8-12

Parameter Value
Condition for Use in Co
Rapid Penetration Problems (lb-sec/ft2)
Problems with long, cylindrical 4 4 0.11
penetrators [Ref 8-13]

All other object shapes where 3 10 0.25


inadequate penetration is of concern
[Ref 8-16]

All other object shapes where excess 2 40 1.0


penetration is of primary concern
[Ref 8-16]

D = equivalent diameter of penetrator [L] as determined 8.3.3.6 Method of Solution. An iterative process is
by
used for the solution of penetration problems because the major
D = (4 At/)0.5 (8-13) resisting force terms (Qni, F si, and F hi) are velocity or depth

dependent, requiring the input of new values with penetrator


The appropriate values of , , and C o to use in travel. The net downward force is an inertial force related to the
Equation 8-12 are specified in Table 8.3-1, as are the conditions deceleration of the penetrator and can be obtained from the
of their use. In Table 8.3-1, long cylindrical penetrators are following modification of Newton's second law (modified to
categorized separately from "all other object shapes." This eliminate the parameter time):
simple geometric shape has a large side surface area in
comparison to its frontal area and lends itself better to this (8-15)

predictive technique (see Section 8.3.3.7).


where: M = penetrator mass [FT2/L]
8.3.3.5 Fluid Drag Force. The fluid drag force acting on = instantaneous change in velocity

a penetrator while moving through water is assumed to


continue to exist as it moves through the soil. The fluid drag For making incremental calculations, dv/dz is replaced with
force is calculated as: (2v)/(2z). The double increments are used to minimize
deviations in the prediction caused by minor errors in the
F hi = (0.5) CD At (vi)2 (8-14) assumed penetration velocity [Ref 8-12]. Then, after
reorganization of terms:
where: C D = dimensionless fluid drag coefficient (the same as
that in seawater)
(8-16)
= mass density of the soil, the "fluid" being
accelerated [FT2/L4]
vi = penetrator velocity after penetrating the ith layer The new velocity for the (i+1)th increment is:
[F/L]
vi+1 = vi-1 + 2 vi (8-17)
Values for CD values are best obtained, where possible,

by backfiguring from measured terminal velocities in water. In


To begin the incremental calculations, the velocity, v1, at the
absence of measurements, use CD values from a hydrodynamics
end of the first increment
reference (e.g., Ref 8-17).

8-9
of seabed penetration, z1, must be generated. An approximation to obtain. Prediction of penetration in oozes is particularly

for v1 can be obtained from: difficult. In foraminiferal oozes, limited data suggest this
method will underpredict penetration--possibly by as much as a
v1 = v o + (1/vo)[(z/M)(Fd1 + W b1 - Q n1 - Fs1 - Fh1 )] (8-18) factor of two [Ref 8-18].
Penetration of the Navys Doppler penetrometer, a long
where vo is the initial penetrator velocity on entering the
slender penetrometer shape, in various seafloors is compared
seafloor (F/L], Fd1, W b1, Q n1, F s1, and F h1 are the respective with in-situ conditions in Reference 8-18.
parameter values for the first layer of penetration [F].
This approximation for v1 is then used to recalculate the

fluid drag force (Equation 8-14). At this point, all the forces
necessary to calculate the net downward force have been
determined. The first iteration for values of Fi, 2vi, and v i+1

can then be completed (Equations 8-14, 8-16, and 8-17).


Subsequent iterations are made by recalculating Fdi, W di, Q ni,

F si, and F hi for the next increment of penetration. When the

computation for v i+1 produces a negative velocity, the iterative

is completed. The maximum penetration of the object is then


obtained by interpolating between the last two velocity values
as follows:

(8-19)

A flow chart for this procedure is shown in Figure 8.3-2.

8.3.3.7 Prediction Accuracy. The dynamic penetration


prediction technique provides reasonable penetration estimates
for long slender penetrators in cohesive sediments, such as
hemipelagic and pelagic clays and in fine silts. This is true
because the predictive technique uses empirical data ( , ,
C o) from field tests with penetrators of this shape in soils of
these types. For differently shaped penetrators and for
penetration in sands and other noncohesive materials, the
accuracy will not be as good. Little field data exist for object
penetration into these granular sediments and for their in-situ
properties. In addition, the undrained shear strength developed
during the rapid penetration is a very difficult parameter
Figure 8.3-2. Flow chart of the calculation procedure for
predicting dynamic penetration.

8-10
8.4 EXAMPLE PROBLEMS container is 4 feet in diameter by 12 feet high and weighs 20
kips in seawater. The sediment at the site is a pelagic clay. Soil
8.4.1 Problem 1--Slow Penetration of a Long Cylinder
properties at the site have been determined from a 40-foot
8.4.1.1 Problem Statement. Determine how deep into the piston core. A schematic diagram for this problem is shown in
seafloor a cylinder will penetrate after being slowly lowered to Figure 8.4-1a. The results of laboratory testing for soil shear
the bottom. strength and density are shown in Figure 8.4-1b.
Data: An electric power source (EPS) is to be lowered to
a specific seafloor location at a speed of 2 fps. An estimation of 8.4.1.2 Problem Solution: The analytical procedures and

the depth of static penetration z is required to verify that cooling computations used to solve this problem are shown below.
water intakes remain above the mudline and to provide input They follow the procedure outlined in Section 8.2 and
data for calculating lateral stability and breakout load. The EPS summarized by the flow chart in Figure 8.2-3.

Figure 8.4-1. Problem sketch and soils data for example problem 1.

NOTE: It is helpful to keep track of calculated values in a tabular form. This is shown for Problem 1 in Table 8.4-1.

8-11
Problem 8.4-1

8-12
Problem 8.4-1

8-13
Problem 8.4-1

Table 8.4-1. Summary of Calculations for Problem 1.

Figure 8.4-2. Plot of predicted soil resistance to EPS


penetration.

8.4.2 Problem 2--Rapid Penetration of a Long Cylinder or to abandon or destroy the device, the depth of its penetration
into the seafloor needs to be estimated. The EPS is suspected to
8.4.2.1 Problem Statement: Determine how deep into the
have hit the seafloor in the same orientation in which it was
seafloor a large long cylinder will penetrate if it falls to the
being lowered due to a concentrated mass at its lower end and
seafloor at terminal velocity.
drag from the severed lowering line. The terminal velocity and
Data: The electric power source (EPS) in Problem 1 was
drag coefficient for this orientation have been calculated as 40
being lowered to the seafloor when the lowering line was
fps and 1.0, respectively. (The soil properties are the same as
severed. The EPS could not be located after falling to the
those shown in Figure 8.4-1b.) A schematic diagram of this
seafloor. In order to decide whether to attempt a recovery
problem is shown in Figure 8.4-3.

8-14
8.4.2.2 Problem Solution. The analytical procedures and dures outlined in Section 8.3 and summarized by the flow chart

computations used to solve this problem are shown below. in Figure 8.3-2.

They follow the proce-

Figure 8.4-3. Sketch for example problem 2.

Problem 8.4-2

8-15
Problem 8.4-2

8-16
Problem 8.4-2

8-17
Problem 8.4-2

8-18
Problem 8.4-2

Table 8.4-2. Summary of Calculations for Problem 2

8.5 REFERENCES 8-4. Design manual: Soil mechanics, foundations, and earth
structures, Naval Facilities Engineering Command, NAVFAC
8-1. J.F. Wadsworth III and R.M. Beard. Propellant-embedded
DM-7. Washington, D.C., 1962, p. 7-12-5.
anchors: Prediction of holding capacity in coral and rock
seafloors, Civil Engineering Laboratory, Technical Note N-
8-5. API recommended practice for planning, designing, and
1595. Port Hueneme, Calif., Nov 1980.
constructing fixed offshore platforms, 12th edition, American
Petroleum Institute, API RP 2A. Washington, D.C., Jan 1981,
8-2. K. Terzaghi. Theoretical soil mechanics. New York,
p. 40.
N.Y., John Wiley and Sons, 1943, p. 121.

8-6. J. de Ruiter and F.L. Bergen. "Pile foundations for large


8-3. J.P. Gemenhardt and J.A. Focht Jr. "Theoretical and
North Sea structures," Marine Geotechnology, vol 3, no. 3,
observed performance of mobile rig footings on clay," in
1979, pp 267-314.
Proceedings of the Offshore Technology Conference, Houston,
Tex., 1970. (Paper 1201)

8-19
8-7. O. Eide, O. Kjekstad, and E. Brylawski. "Installation of 8-16. D. G. True. Personal communication, Dec 1982.
concrete gravity structures in the North Sea," Marine
8-17. S. F. Hoerner. Fluid dynamics drag. Brecktown, N.J.,
Geotechnology, vol 3, no. 4. 1979, pp 315-368.
Hoerner, 1965.
8-8. Standard method for deep, quasi-static, cone and friction-
8-18. R.M. Beard Expendable Doppler penetrometer for deep
cone penetration tests of soil, part 19, American Society for
ocean sediment strength measurements, Naval Civil
Testing and Materials, ASTM Standard D3441-79.
Engineering Laboratory, Technical Report R-905. Port
Philadelphia, Pa., 1981, pp 515-522.
Hueneme, Calif., Feb 1984.
8-9. W.B. Schmid. "The penetration of objects into the ocean
bottom," in Civil Engineering in the Oceans II. New York,
8.6 SYMBOLS
N.Y., American Society of Civil Engineers, 1970, pp 167-208.
As Side area of penetrator in contact with soil [L2]
8-10. R.J. Smith. Techniques for predicting sea floor
At End area of penetrator (the effective bearing surface) [L2]
penetration, U. S. Naval Post Graduate School. Monterey,
Calif., Jun 1969. B Minimum dimension of the penetrator, width or diameter
[L]

8-11. H.J. Migliore and H.J. Lee. Seafloor penetration tests: CD Fluid drag coefficient
Presentation and analysis of results, Naval Civil Engineering
Empirical strain rate coefficient [FT/L2]
Laboratory, Technical Note N-1178. Port Hueneme, Calif.,
Aug 1971, p 59. Co Empirical strain rate constant
_
8-12. D.G. True. Undrained vertical penetration into ocean c Effective soil cohesion [F/L2]
bottom soils, Ph.D. thesis, University of California, Berkeley.
D Equivalent diameter of penetrator [L]
Berkeley, Calif., 1976.
F Net downward force on penetrator [F]
8-13. D.G. True. Penetration of projectiles into seafloor soils,
Fd External driving force on penetrator [F]
Civil Engineering Laboratory, Technical Report R-822. Port
Hueneme, Calif. , May 1975, p. 45. Fh Fluid drag force on penetrator [F]

Fi Net downward force exerted by the penetrator [F]


8-14. H.B. Seed and K.L. Lee. "Undrained strength
characteristics of cohesionless soils," Journal of the Soil Fs Side friction or adhesion force on penetrator [F]
Mechanics and Foundations Division, American Society of
f Unit wall friction for skirt [F/L2]
Civil Engineers, vol 93, no. SM6, Nov 1967.
H Length of penetrator, in direction of travel [L]
8-15. U. Dayal, J.H. Allen, and J.M. Jones. "Use of the impact
i ith increment of soil depth
penetrometer for the evaluation of the in situ strength of marine
sediments," Marine Geotechnology, vol 1, no. 2, 1975, pp 73- KH Coefficient of lateral earth pressure

89. kf Friction correlation factor for standard cone

kt Tip correlation factor for standard cone

L Length of bearing surface of penetrator [L]


8-20
M Penetrator mass [FT2/L] v Penetrator velocity [L/T]

N'c Bearing capacity factor for static penetration in cohesive vo Penetrator velocity at soil contact [L/T]
soils
Wb Buoyant unit weight of penetretor [F]
Nt Nose resistance factor for dynamic penetration
z Depth of penetration of At, the effective bearing surface
Nq Bearing capacity factor for static penetration in [L]
cohesionleess soils
_ z Incremental change in penetration
po Soil overburden pressure [F/L2]
Adhesion factor (soil against object)
Qn Nose resistance during dynamic penetration [F]
b Soil buoyant unit weight [F/L2]
Qu Net resistance to (static) penetration [F]
t Soil bulk wet density [F/L2]
qc Standard cone penetration resistance [F/L2]
Friction angle between penetrator and sand [deg]
qu Tip resistance for skirt [F/L 2]
Penetrator bearing area shape factor
Strain rate factor
Mass density of the material being accelerated [FT2/L4]

Maximum strain rate factor Soil friction angle [deg]


_
St Soil shear strength sensitivity Effective or drained friction angle [deg]

su Soil undrained shear strength [F/L2]

8-21
Chapter 9

BREAKOUT OF OBJECTS FROM THE SEAFLOOR

9.1 INTRODUCTION 9.1.1 Applications

Removing objects embedded in or resting on the seafloor Determination of breakout force is important in two
requires determination of the force needed for an immediate general situations--where retrieval is desired, and where it is
breakout, or the amount of time that a specified lower force has desired that the object remain in place. The more common of
to be applied for object breakout to occur. Either force level or the two is the retrieval of objects embedded in the seafloor.
application time may be limited by ship or equipment These objects include sunken ships, submarines, airplanes
capabilities or by structural limitations of the object being weapons, previously placed instrumentation packages, or
lifted. The basis for most of what is presented in this chapter is foundations for seafloor structures. The objective of a retrieval
Navy research and development work [Ref 9-1 through 9-4], operation may be the recovery of the object for study or to deny
although the concepts are similar to those proposed elsewhere its use to others. The goal in planning the operation would be
[Ref 9-5]. The total uplift force required for breakout will to predict and provide sufficient lifting force for immediate
always include the underwater weight of the object plus the breakout, or to determine an adequate amount of time for
buoyant weight of any bottom sediment adhering to or lifted application of a lower force to cause breakout. A conservative
with the object. It is, however, the prediction of sediment- estimate for this operation would be the upper limit or required
generated breakout force that is particularly difficult and is of or time required for breakout to occur.
concern in this chapter. The second type of breakout situation involves objects
The force and time predictions by the methods presented that are expected to remain in place and not become dislodged
are not highly accurate due to a poor understanding of the from the seafloor. An example would be a foundation which
complex nature of what happens to the soil during the breakout could experience an uplift force in excess of its underwater
process, and the variable nature of how different soils react weight. For this application, a conservative estimate for
during that process. To improve predictive capabilities would breakout would he a lower force limit. That is, the object would
require much general study of the breakout process, and a be expected to remain in place if no uplift forces were exerted in
knowledge of soil properties in specific situations far above excess of this value. This chapter presents methods for
what could reasonably he expected to be available in most obtaining the best estimate and upper and lower limits of the
actual field conditions. breakout force, and discusses the accuracy of

9-1
that estimate. It also presents a method for estimating the time forces along the object's sides can resist its removal. However,
(and upper and lower limits) required for breakout to occur at a these forces will probably be considerably smaller than object
force below the best estimate. weight and can be treated, if necessary, as friction along a
The chapter considers only the breakout of objects that friction pile (Chapter 5).
are embedded to a depth less than their width. Most objects that With impervious or low permeability soil (such as
would fall outside this category could be treated adequately as cohesive materials like clays and many silts), water flow into
embedment anchors or as piles (Chapters 6 or 5), depending on the space below the object is impeded. Breakout may be
object size and depth of embedment. Some unusual situations achieved with a relatively small force, but the amount of time
in which an object is deeply embedded are not well-described by needed will be intolerably long for most applications. For a
these other treatments (a ship, for example); but for the present more rapid breakout, a greater force must be exerted: one high
these situations must also be approached as if the object were enough to cause sediment flow into the space left behind by the
an embedment anchor or a pile. object. This failure mechanism is somewhat analogous to
The procedures presented for evaluating the problem and movement accompanying foundation bearing capacity failures
determining values for breakout force or time for breakout are (Chapter 4). Object breakout, however, may be from a
outlined in Figure 9.1-1. combination of sediment and water flow. For example, initial
sediment flow may create sediment cracks, which can promote
9.1.2 General Concepts
more rapid water flow and lower the amount of time required for
An object cannot be removed from the seafloor unless breakout under the applied load.
either water or sediment moves into the space that the object
9.1.3 Definitions
occupies. Whether breakout is achieved by sediment or water
flow depends on the amount of time allowed and the overall 9.1.3.1 Breakout Force. The breakout force is the net
permeability of the sediment. An uplift force in excess of the uplift force (the force in excess of an objects submerged
buoyant weight of the object will induce change in the soil pore weight) applied to lift an object out of a shallow seafloor
water pressure around the object. As time passes, water will embedment.
flow within the soil mass to normalize the pore pressure
(dissipate negative pore pressures that were developed) and allow 9.1.3.2 Immediate Breakout Force. The immediate

the object to move upward. Breakout will eventually occur with breakout force is the breakout force in excess of the object's

even low net uplift forces if enough time is allowed. With submerged weight required to dislodge an object "immediately"

pervious or highly permeable soil such as cohesionless (i.e., within a few minutes). For cohesionless soils such as

materials like sand and gravel , the rate of water flow will be sand and gravel, the immediate breakout force is caused by water

rapid. Only a small nearly negligible, net breakout force will be flow and is negligible. For cohesive soil such as clay and many

needed to achieve breakout in a few minutes time. It is common silts, the immediate breakout force is the result of soil flow,

practice to plan for no breakout forces beyond the object's which is primarily dependent on soil shear strength, and can be

buoyant weight when lifting shallowly embedded objects from a very high.

cohesionless sand seafloor. For a large object partially embedded


9.1.3.3 Long-Term Breakout Force. The long-term
in cohesionless seafloors, significant frictional
breakout force is a force less than the immediate breakout force
that is applied to an object embedded in cohesive soil to
dislodge

9-2
Figure 9.1-1. Flow chart for procedures to determine immediate breakout force and
time required for long-term breakout under a lower force.

9-3
it in some finite amount of time ("breakout time" defined in aids. They include rocking, lifting from the end, water-jetting,
Section 9.1.3.6). The long-term breakout force must be applied and installing drainage tubes.
over that amount of time to allow pore water flow, sediment
9.1.3.9 Waiting Time. Waiting time is the time that
creep, and possible crack development. Long-term breakout is
the process that occurs during that time and can be hours or elapses between object embedment and application of a breakout

more. force.

9.1.3.4 Shallow or Partial Embedment. Shallow or 9.2 REQUIRED INFORMATION

partial embedment is embedment to a depth that is less than the 9.2.1 Object Embedment Characteristics
object width. This is the only type of breakout considered in
The existing object embedment characteristics must be
this chapter.
determined or estimated. An estimate is made of the object's
volume below the seafloor. The cross-sectional area, A, of the
9.1.3.5 Soil Suction . Soil suction is the sediment flow
object at the mudline is determined. It is then represented by a
into the space left by an object, as occurs with immediate
rectangle having the same area and an equivalent length, L, and
breakout from a cohesive sediment. Because pore water flow
width, B, as best fits the existing situation. (B must always be
into the space beneath the object does not occur quickly
less than or equal to L. ) For nonrectangular cross sections, the
enough, negative pore pressure or "suction" forces pull the
object's longest embedded dimension is used for L, and A/L is
sediment along with the object and into the space beneath the
calculated as the value of B. An equivalent embedment depth is
object as it is extracted.
calculated by dividing the embedded volume by the cross-
9.1.3.6 Breakout Time. Breakout time is the time sectional area.

required to achieve breakout from a cohesive sediment when a If possible, the downward velocity of the object as it

force less than the immediate breakout force is applied. became embedded in the seafloor should be estimated. If the
object was placed relatively gently, this velocity is assumed
9.1.3.7 Breakout Line Force. The breakout line force is equal to zero.
the total retrieval line force required to extract an object from
9.2.2 Sediment Characteristics
the seafloor. For cohesionless soils, it is approximately equal
to the object's submerged weight. For cohesive soils, it is equal Characteristics of the seafloor soil must be determined.
to the immediate or long-term breakout force plus the object's First, it should be determined whether the bottom is cohesive or
submerged weight minus the buoyant weight of sediment cohesionless. Likely sediment type can be estimated from the
displaced. Although the displaced sediment term is included, it worldwide ocean sediment distribution chart. [Figure 2.2-1,
is usually minor and becomes smaller as the object moves Chapter 2] or by more accurate local charts, if available. Actual
upward during the breakout process. sediment type should be confirmed by sampling (see Chapter 2
for methods), if possible. Sediment type may be able to be
9.1.3.8 Breakout Aids. The breakout forces discussed in
deduced from topography, photographs, and material adhering to
this chapter are considered to be applied approximately at the
remote vehicles if they are used to observe the object. IF THE
object's center of mass. Any actions that reduce the breakout
SEDIMENT IS COHESIONLESS AND EMBEDMENT IS
force below the value calculated for this simplified case are
considered breakout

9-4
NOT DEEP, IT IS LIKELY THAT THE REQUIRED depth was measured within a short time (days) after the object
BREAKOUT FORCE WILL BE NEGLIGIBLE. If the sediment penetrated the seafloor, as continued downward movement of the
is cohesive, it is necessary to estimate its undrained shear object and soil consolidation will likely occur in such soft
strength and the (downward) soil bearing capacity for the object materials. In this case, the bearing capacity is assumed to have
determined. These are discussed below. been exceeded by the object during placement, and:

9.2.3 Bearing Capacity (Cohesive Sediments) F q = Wb - Ws (9-2)

Bearing capacity for an object vertically loaded on


where: Wb = object's submerged weight [F]
cohesive sediments (Fq) is described [after Ref 9-6] by: Ws = buoyant weight of sediment displaced by the
embedded object [F]
(9-1) = b Vs
b = buoyant unit weight of sediment [F/L3]
Vs = volume of the sediment displaced by the
where: su = undrained shear strength of the sediment (for
embedded object [L3]
this equation it is su averaged a from the
seafloor surface to depth D+B) [F/L2]
A = horizontal area at the mudline [L2]
D = embedment depth [L] 9.3 SHORT-TERM (IMMEDIATE) BREAKOUT
B,L = object's lateral dimensions, with L > B [L]
The short-term or immediate breakout force, F Ib , is
Note: A, D, B, and L, for the purposes of this chapter, are the
determined [after Ref 9-1] from:
values determined in Section 9.2.1.

F Ib = F q[1.0 - 0.97 e-2.75 (D/B) ] (9-3)


It will be necessary to measure or estimate su in order to

determine F q. If no site strength data are available, an


The recovery line force for immediate breakout, FlIb, is:
estimation can be made by using the strength profiles for
seafloor sediments presented in Chapter 2. If the sediment is F lIb = F Ib + W b - Ws (9-4)
unusual or the operation is particularly critical, the strength
should be measured. This can be done by testing piston or Equation 9-3 is empirical and based on field and

gravity core samples or strength can be obtained from an in-situ laboratory breakout tests [Ref 9-1] with objects of various sizes

test (see Chapters 2 and 3 for these procedures). If the object and shapes. The objects varied from some as small as a 3-inch

penetrated more rapidly and the entry velocity is accurately cube to those as large as a 6.4-foot-diam sphere or a 17.5-foot-

known, the average undrained shear strength can be estimated by long by 4.5-foot-wide cylinder. Other shapes in intermediate

inverting the penetration prediction equations of Chapter 9. sizes were included in a total of 57 individual breakout tests. It

However, this is a complicated hand-procedure. can be seen from Equation 9-3 that FIb will approach F q as D/B

Another, and simpler, method for obtaining Fq can be approaches 1. As D/B approaches zero, indicative of very

used where an object is placed gently on a very soft cohesive shallow embedment, the breakout force becomes very small.

seafloor. It must be known that the object impacted the bottom The normalized immediate breakout force (FIb /F q)

slowly (at < 2 ft/sec), and the object settled to an embedment computed by Equation 9-3 is plotted on Figure 9.3-1 as a
depth greater than 0.25 B. This method should only be used if function of relative embedment
the embedment

9-5
Figure 9.3-1. Normalized immediate breakout force as a function
of relative embedment depth (after Ref 9-1).

depth. When data from Reference 9-1 for these field and estimates of the immediate breakout force when the object's
laboratory breakout tests are compared to the breakout force breakout is or is not desired (the upper and lower limits
predicted by Equation 9-3, 77% of the data all fall between respectively). If the operation is one of object retrieval, for
limits 50% greater and 33% less than the equation's curve (i.e., instance, the conservative value for immediate breakout force is
within the area bounded by 1.5 FIb and F ib/1.5). If these limits twice the value computed by Equation 9-3 and required retrieval
are extended to 100% greater and 50% less than the equation's line force at the object would become 2 FIb + W b - W s. If the
curve (2.0 F Ib and FIb /2), 93% (all but 4) fall within the limits. operation requires a line load that will not cause breakout, the
These data are for objects whose shape and embedment were conservative value to prevent immediate breakout is one-half
accurately known. It is probable that object dimensions, the value computed by Equation 9-3, and the maximum line
embedment depth, and average shear strength will not be known force at the object should be FIb /2 + Wb - Ws.
as well for problems in the field. Therefore, the agreement
between predicted and actual breakout may also not be as good. 9.4 LONG-TERM BREAKOUT TIME PREDICTION

The upper and lower limits, (2.0 FIb )/F q and F Ib /(2.0 F q) Long-term breakout considerations are most important
are also plotted on Figure 9.3-1. The predicted immediate when it is not desirable or not possible to apply a load as high
breakout force should fall between these limits when object as that required for immediate breakout. This might be because
dimensions, embedment depth and soil strength are reasonably of structural limitations of the object or limitations imposed by
well-known. The upper and lower limits are also used for available lines or lifting gear. The recovery line force applied at
conservative a

9-6
roughly constant level over an extended period of time in an require accurate determination of several other sediment
effort to cause object breakout is called the long-term breakout engineering properties and a vastly more complex problem
line force F lLb. The long-term breakout force, F Lb, is that definition than would be possible in most Navy field breakout

portion of the long-term breakout line force carried by the problems.

sediment. F Lb is determined by: Equation 9-6 is graphed on Figure 9.4-1 in a plot of


relative force level (FLb/F Ib ) versus the time parameter. Also
F Lb = F lLb - Wb + W s (9-5) shown are conservative limit curves recommended for
estimating breakout time if breakout is or is not desired (upper
Unlike the immediate breakout force, the long-term
and lower curves, respectively). These are plots of Equation 9-6
breakout force is not a specific value but can be any significant
where 2 FLb and FLb/2 are used, respectively, for the upper and
force lower than the immediate breakout force. The unknown
lower curves. They should be used to estimate a conservative
quantity in long-term breakout problems is the time that this
value of breakout time for field problems. Because of the
load will need to be applied for breakout to occur.
limited data base available for these empirical relationships,
The following relationships are used to determine this
some values of actual t b (which occur in the field) will fall
time, tb, for a specified long-term breakout force [after Ref 9-1].
outside the area bounded by the two limit curves.
log10 (F Lb/F Ib ) = -0.193(log10 T - 3.84) (9-6) The use of the curves in Figure 9.4-1 requires
determination of the immediate break-out force. Therefore, all
T = (p t b/D2)(B/D) 2 (9-7)* information required for calculating the immediate breakout
force is also required for the time calculation. After determining
p = F Lb/A (9-8)
the ratio F Lb/F Ib the appropriate value of T is determined from

where: T = time parameter Figure 9.4-1. The breakout time (in minutes) is calculated from
p = the average breakout pressure applied to the Equation 9-7 by rearranging the terms as follows. (As in
sediment [psf]
tb = the time required for breakout to occur when the Equation 9-7, distance must be in feet and force must be in
long-term breakout force is applied [min] pounds.)

The basis for the relationship defined by Equations 9-6


(9-9)
and 9-7 is entirely empirical. That is, it is based entirely on data
for breakout times from field and laboratory breakout tests. The
curve is a best fit of data from 76 individual tests, with more 9.5 BREAKOUT AIDS
weight given to the field tests.
Breakout aids are any operations that reduce the force
Presently, no theoretical solution for determining
required for immediate breakout or the time required for long-
breakout time is possible. If and when one becomes available,
term breakout. The following discussion provides suggestions
it will likely
for achieving breakout when the immediate breakout force is
____________
too high to be applied or the
* Because this relationship is empirically based and uses units
not dimensionally coherent, it is essential that the following
units are used: force in pounds, time in minutes, and length
in feet.

9-7
Figure 9.4-1. Normalized long-term breakout force as a function
of breakout time parameter (after Ref 9-1).

long-term breakout time for the force that can be applied is too With water jetting, a pump is connected to the end of the
long. Although most discussion is on reducing breakout force, tube and water is forced through the tube into the soil. Jetting
most breakout aids also reduce breakout time at a particular is preferable because the positive pressure is more effective, and
force level. it should also be noted that reduction in breakout the turbulence of water flow may cause some soil erosion and
force by one-half, for example, is not the same as a lowering of reduce the object-to-soil contact area.
the retrieval line load by one-half. Line load is equal to breakout With drainage tubes no pumps are used. In this technique,
force plus object buoyant weight minus displaced soil weight, the act of pulling up on the object draws water through the
as defined in Equation 9-5. tubes into the spaces below the object. It is essential that the
free end of the tube does not clog.
9.5.1 Jetting and Drainage Tubes
Field tests [Ref 9-7] have demonstrated that a 50 to 77%
An effective way to reduce the force required for breakout reduction of the breakout force is possible with both water jets
is to improve the flow of water into the soil around the object. and drainage tubes. The reduction for rounded objects was
This can he done either passively with drainage tubes or greater than for square blocks. It should be expected that the
actively with water jets. Both consist of tubes that are forced effectiveness of these aids will vary significantly, depending on
beneath the object, or may be a part of the object if a breakout specific soil and embedment conditions.
problem is anticipated. Preferably, the tubes should have No guidelines exist for the design of either water jets or
openings along the side to improve water flow all along the drainage tubes. Where used, however, the spacing between
tube. openings (both along the tubes and between the tubes)

9-8
should be minimized and kept relatively constant over the entire develop along the objects side. These openings will facilitate
contact area of the object. water flow and reduce the object-soil contact area. A lateral force
greater than the passive resistance of the sediment (estimated as
9.5.2 Eccentric Loading
2DLsu) is likely necessary to achieve adequate movement. If
A reduction in the required breakout force will result if possible, the object should be loaded alternately in opposite
the object is lifted from one end rather than through its center. directions to increase the development of openings along the
In Chapter 4, the concept of effective area is introduced with sides. This process of rocking or rolling may also result in a
respect to bearing capacity prediction for eccentrically loaded reduction of object embedment depth and, therefore, an
foundations. This concept also applies to breakout prediction. additional reduction in necessary breakout force.
The uplift force is essentially applied to a reduced area, raising
9.5.5 Breakaway Parts
the applied stress in that area. After that section is broken free,
the force is transferred to the remaining area and breaks it free. Objects that are meant to be placed on the seafloor can be
A rough estimate of the effect of this procedure is that breakout designed with "breakaway parts." That is, the portion of the
force may be reduced by up to 50%. Long, narrow objects object in contact with the sediment is designed to separate from
should see the most effect. In many field cases, this may be an the remainder of the object upon application of uplift loading.
easy breakout aid to use; however structural limitations on the The immediate breakout force is effectively reduced to the level
object could prevent eccentric attachment of the applied load. required to break any connectors to the parts that remain behind.
This situation can be achieved with weak links or connectors
9.5.3 Cyclic Loading
that corrode in seawater and leave the object parts to be
The strength of some cohesive sediments can be reduced recovered unrestrained for vertical motion. This procedure can be
through cyclic application of uplift force and resulting useful for instrumentation packages that are on the seafloor for
development of positive sediment pore water pressures. For long periods of time before retrieval.
example, based on earthquake stability analysis [Ref 9-8],
9.5.6 Altering Buoyant Weight
breakout force may be reduced by 30 to 40% with as few as 10
load applications for clayey silt soils. This procedure will be The retrieval line force may be significantly reduced if it
less effective with soils which are more fine-grained and more is possible to decrease the object's underwater weight. This can
plastic. During the time between load applications, an uplift be done by pumping air into enclosed spaces, attachment of lift
force equal to the buoyant weight of the object should be bags, or removal of heavy parts of the object prior to
maintained, and the loads should be applied every 10 minutes or attempting breakout. These methods do not alter the breakout
less. Benefits from this effect may occur where wave actions force required but may be useful where the breakout force is not
affect the recovery vessel and cause unsteady recovery line loads. the largest component in the retrieval line load. Air- or lift-bag-
assisted recovery may introduce potentially dangerous situations
9.5.4 Rocking or Rolling
where retrieval line control can be lost. For example, trapped air
Prior to lifting, if a significant lateral force can be applied will expand in volume as the object rises from the seafloor or
to an object embedded in relatively stiff soil (su > 3 psi), can be lost if the
openings may

9-9
objects orientation changes. A positively buoyant object will the object. The procedures outlined in this chapter apply as if
become dangerous to anything above it as it moves toward the the foundation plus trapped soil is a solid object and embedment
water surface. is considered to be at the depth of skirt penetration. The skirted
foundation may also introduce problems during recovery
9.6. OTHER FACTORS
operations. Soil trapped within the skirt system may drop away
9.6.1 Irregular Shape or Nonuniform Embedment Depth at any time following breakout. (It usually ends up on the deck
of the recovery vessel.)
Laboratory and field tests used to define the empirical
equations in this chapter have not been conducted with objects 9.7 EXAMPLE PROBLEMS
that are irregular in shape or that have a nonuniform embedment
9.7.1 Problem 1 - Recovery of a Large Long Cylinder
depth. The equations presented, and the procedures of Section
9.2.1 for calculating effective D, B, and L, can still be used for 9.7.1.1 Problem Statement. Determine if it is possible to
these unusual cases. However, this introduces a higher level of quickly lift an embedded, cylindrically shaped object with the
uncertainty and results in the computational procedures limited capacity of an available recovery vessel, and, if this is
becoming less accurate. not possible, to estimate how long it will take for the object to
break free of the seafloor at a lower force.
9.6.2 Waiting Time
Data: A large, long cylinder--20 feet in diameter, 200 feet
An object at rest in the seafloor for an extended period of long, and an underwater weight of 1 million pounds fell to the
time before recovery is attempted may require a higher seafloor, impacting at an unknown velocity. It was located
immediate breakout force (calculated by Equation 9-3) than if immediately and found lying approximately horizontal in 200
recovery is attempted relatively soon after the object is feet of water and embedded to approximately half its diameter in
embedded. This occurs primarily because the weight of the the seafloor [see Figure 9.7-1]. A recovery vessel can be at the
object consolidates the sediment and increases sediment site within 2 weeks. It can apply a 4.5-million-pound uplift
strength. If the object is in place for an extended time (perhaps a force at 200-foot water depths for only 10 to 20 minutes
month or more), a significant strength increase can occur. It can without seriously damaging the recovery gear. But, it can apply
be compensated for by increasing the soil strength used in the up to 3 million pounds of uplift force for several days. The
calculations. No recommendation is made here because of the seafloor was determined to be a terrigenous silty clay when the
complexity of soil-related factors that influence the amount of cylinder was located; no other data on the sediment were
the change. obtained.
9.7.1.2 Problem Solution: The analytical and
9.6.3 Foundation Skirts
computational procedures used to solve this problem are shown
If a foundation has skirts extending below its base (useful below. They follow the procedures outlined by the flow chart in
to provide an increased resistance to lateral loads), the skirts Figure 9.1-1 and discussed in this chapter.
will raise the required immediate breakout force. The skirts force
the soil failure surfaces to a greater subbottom depth. The
sediment contained within the skirts must be considered as part
of

9-10
Figure 9.7-1. Problem sketch and data for example problem 1.

Problem 9.7.1

9-11
Problem 9.7-1

9-12
Problem 9.7-1

9.7.2 Problem 2 - Recovery of a Skirted Foundation 3-day experiment and then recovered. Skirts were added to the
perimeter of the footing to prevent horizontal movement from
9.7.2.1 Problem Statement: Determine if it is reasonable
forces applied during the experiment. Full embedment of the
to expect quick recovery of a small foundation from the seafloor
skirts is expected, but the footing base will rest approximately
with the limited capacity of a specific workboat.
on the seafloor surface. The footing is 4 feet in diameter, and
Data: The foundation is a heavy, small circular footing
has an underwater weight of 3,000 pounds [see Figure 9.7-2].
that will be installed for a
The skirts extend 0.5 foot below the footing

9-13
base. The location of the experiment is on an abyssal hill at seafloor and maintaining that force level indefinitely.
approximately 165oE longitude, 47oN latitude. The vessel that
9.7.2.2 Problem Solution:. The analytical and
will install and remove the foundation costs $10,000/day to
lease and is capable of lifting 7,000 pounds off the computational procedures used to evaluate this problem are
shown below.

Figure 9.7-2. Problem sketch and data for example problem 2.

Problem 9.7.2

9-14
Problem 9.7-2

9-15
Problem 9.7-2

9.8 REFERENCES 9-4. C.L. Liu. Ocean sediment holding strength against
breakout of embedded objects, Naval Civil Engineering
9-1. H.J. Lee. Breakout of partially embedded objects from
Laboratory, Technical Report R-635. Port Hueneme, Calif.,
cohesive seafloor soils, Naval Civil Engineering Laboratory,
Aug 1969.
Technical Report R-755. Port Hueneme, Calif. , Feb 1972.

9-5. P.M. Byrne and W.D.L. Finn "Breakout of submerged


9-2. ___________ . "Breakout of partially embedded objects
structures buried to a shallow depth," Canadian Geotechnical
from cohesive seafloor soils," in Proceedings of the Offshore
Journal, vol 15, 1978, pp 146-154.
Technology Conference, Houston, Tex., 1973. (OTC 1904)

9-6. A.W. Skempton. The bearing capacity of clays, Building


9-3. B.J. Muga. Ocean bottom breakout forces, including field
Research Congress, Collected Papers, Div 1, Pt 3. London,
test data and the development of an analytical method, Naval
England, 1951, pp 180-189.
Civil Engineering Laboratory, Technical Report R-591. Port
Hueneme, Calif. , Jun 1968.

9-16
9-7. K.D. Vaudrey. Evaluation of bottom breakout reduction F lIb Line force required to achieve immediate breakout [F]
methods, Naval Civil Engineering Laboratory, Technical Note
F lLb Line force applied for long-term breakout [F]
N-1227. Port Hueneme, Calif., Apr 1972.
Fq (Downward) bearing capacity [F]
9-8. H.J. Lee, B.D. Edwards, and M.E. Field. "Geotechnical
L Object length (maximum horizontal dimension) [L]
analysis of a submarine slump, Eureka, California," in
Proceedings of the Offshore Technology Conference, Houston, p Long-term breakout pressure applied to the sediment
[F/L2]
Tex., 1981. (OTC 4121)
su Undrained shear strength of the sediment [F/L2]

9.9 SYMBOLS T Time parameter used in evaluating long-term breakout


(not dimensionally coherent)
A Horizontal cross-sectional area of an embedded object at
the mudline [L2] tb Time required for breakout to occur when the long-term
breakout force is applied [T]
B Object width (minimum horizontal dimension) [L]
Vs Volume of object below the mudline (or sediment
D Equivalent embedment depth of the object [L] displaced) [L3]

F Ib Immediate breakout force carried by the sediment [F] Wb Submerged weight of object [F]

F Lb Long-term breakout force carried by the sediment [F] Ws Buoyant weight of sediment displaced [F]

b Buoyant unit weight of sediment [F/L 3]

9-17
Chapter 10

SCOUR

10.1 INTRODUCTION sent selected experimental and field data helpful in structure
design, (3) evaluate common scour prevention techniques, and
10.1.1 Objectives
(4) present methods for minimizing scour.
Structures located on the seafloor are susceptible to
failure caused by sediments eroding from beneath or around their 10.1.2 Theory
foundations. The term "scour" is used in a general sense
The shear stress approach is the analytical method most
without reference to the cause or mechanism of the process to
often used to describe sediment erosion. The basis of this
describe this erosion of sediments that results in a local
approach is that a mean shear stress can be described based on
depression or a general lowering of the seafloor. For example,
mean flow parameters, and erosion of sediment can be related to
both sediment erosion in the vicinity of a structure and the
this mean shear stress. In the shear stress approach the flow
general lowering of the seafloor resulting from a change in
field and boundary layer are defined, and then lift, drag, and
wave climate are referred to as scour.
inertia forces on individual sand grains are calculated. The major
Where, when and how much scour will occur on the
fault of the theory is that the role of turbulent fluctuations in
seafloor are not well-understood and cannot be calculated or
initiating sediment motion is not considered. Turbulence in the
estimated with a high degree of accuracy. In addition to this
flow can erode sediment even though the mean shear stress is
uncertainty, most of what is known is from historical
near zero. Therefore, two flows with the same mean shear stress
observation of scour effects along piles and piers in relatively
can have substantial differences in velocity and pressure
shallow ocean water and in rivers. In many of these cases, the
fluctuations.
significant or dominant forces causing the scour are not typical
The macroscopic (stochastic) approach to predicting
of more tranquil deep ocean conditions (e.g., oscillatory bottom
sediment motion has been suggested for sediment transport in
currents from short-period waves in shallow water and high-
streams (Ref 10-1 and 10-2) and may be applicable to oceans.
velocity currents in rivers cause significant scour, but are not
Further, a turbulence spectrum may be correlated with sediment
typically found in deep water). Even if parallel deep ocean
suspension, much as mean wind velocity is now used to
situations are rare, this chapter presents an overall picture of
calculate a spectrum of waves [Ref 10-3].
scour in order to give the reader a more complete understanding
of the processes. The objectives of this chapter are to; (1) 10.1.3 Modeling
describe scour mechanisms in cohesionless sediments, (2) pre-
The usual procedure when analytical approaches are
difficult or not satisfactory is

10-1
to rely on physical model testing; but, for evaluating scour, sonal seafloor level changes that would occur, even if the
proper scaling of all the important modeled elements (flow, structure were not present, and local scour caused by the
sediment, structure) simultaneously is extremely difficult. presence of the structure. The relative occurrence of these two
Laboratory and field data have not correlated well, scour types depends primarily on the water depth. A structure
particularly when modeling small structures and waves. Some located in deep water, well beyond the breaker zone, would
success in modeling sediment scour around structures in uni- likely experience little seasonal deposition or removal of
directional or slowly reversing flow without waves has been sediment. However, if significant currents are present, local
achieved. Laboratory test results of modeling scour at bridge scour at the structure could be substantial. Conversely, a
abutments and piers [Ref 10-4] compare favorably with field structure located in the highly active area of breaking waves
data. This supports the contention that, at least for uni- could experience substantial seasonal scouring and
directional flow, model test results can be extrapolated to field comparatively minor local scour. Reference 10-6 presents a
scale. collection and brief discussion of field scour observations.
Unfortunately, the same cannot be said for model studies Sections 10.3 and 10 4 present techniques for estimating
of scour from nonsteady flow, such as turbulence caused by seasonal and local scour. Each scour type may be estimated
wave action. Significantly greater modeling problems exist. independently and combined to provide a rough estimate of the
The height and period of waves that can be produced in a total scour likely at a structure.
laboratory flume are limited to values much smaller than
10.2.2 Deep Water Wave-Induced Scour
prototype waves. If similarity in model and prototype is to be
maintained, the bed material--sand grains--in the model must Another mechanism that contributes to scour is wave-
also be reduced in size. Attempts have been made to model sand induced excess pore pressure and subsequent water flow in the
(angularity, weight, etc.) by using other materials , such as seabed [Ref 10-7 through 10-9]. The effects of these phenomena
gilsonite (specific gravity = 1.06). This is very difficult and on the seafloor are not fully understood, but they can be
may introduce other uncertainties in applying the results to field dramatic in deep water, even causing sediment liquefaction. The
situations. In at least one instance [Ref 10-5], however, effect on scour is complicated by the structure itself and by
gilsonite was effectively used in a model study to reproduce a simultaneous action of currents and forces on the structure.
field scour situation at an offshore pile-supported platform. But, Although the effects can not be quantified, they increase with
replication of the scour pattern observed in the field was increasing wave length and decreasing water depth. In a field test
achieved only after a long series of modeling attempts. In of scour protection systems for a research platform in water 100
general , although turbulence is a significant factor in sediment feet deep, some hinged concrete scour protection slabs, 17-1/2
scour, its analytical description is not possible, and model study by 6-1/2 feet in size, were found raised from their horizontal
results have not produced reliable quantitative scour estimates. positions to vertical positions [Ref 10-10]. Seven slabs had
broken away from their hinges.
10.2 SCOUR TYPES
10.3 ESTIMATING SEASONAL SCOUR
10.2.1 Seasonal and Local Scour
The best method for estimating seasonal scour at a
Sediment erosion (scour) from around a structure may be
location is from historical data for
divided into two types: sea-

10-2
the site. If these are not available and data cannot be gathered, 10.4 ESTIMATING LOCAL SCOUR AT SEAFLOOR
STRUCTURES
maximum expected seasonal scour can be estimated from
empirical correlations [Ref 10-11]. The maximum seasonal An accurate estimate of the amount of scour to be
sediment level change is roughly estimated by: expected for a footing of given size and shape in a specified
environment is not yet possible. Most knowledge on the
ymax = 1.15 He - 4.1 (10-1)
initiation and extent of scour has been gained from field
observations or from model testing (where appropriate modeling
where H e is the annual extreme wave height (which is the
is possible).
significant wave height exceeded 12 hr/yr) [L, in feet].
The annual extreme wave height is [after Ref 10-12]: 10.4.1 Effect of Seafloor Factors on Scour
_ _
He = H + 5.6 4.5 H (10-2) Observations of laboratory modeling together with the
techniques of dimensional analysis have provided insight into
_
where: H = average monthly significant wave height [L in the effect of some important structural and environmental
ft]
factors on scour. The effects of the most significant factors are
= standard deviation of the monthly average
significant wave heights [L, in ft] summarized below.

The water depth at which ymax occurs is: 10.4.1.1 Flow Velocity and Equilibrium Scour Depth.

Model tests on scour at bridge piers show that scour ceases to


dm = He (10-3)
increase after current velocity reaches a critical value. Figure
10.4-1 demonstrates the concept of equilibrium scour depth for
The seaward limit of significant sediment level change,
a coarse material at a pier in uni-directional flow. Scour begins
dc , also termed the close-out depth, has been determined to be
at
[Ref 10-13] the water depth:

dc = 2 H e (10-4)

or equals the depth of the most shallow depth contour parallel


to shore if this condition is present and > than 2 He .

Equation 10-1 provides a rough estimate (20%) of


maximum seasonal sediment level change. A single
bathymetric survey will often show the presence of one or more
longshore bars. The position of the largest longshore bar can be
used to estimate the position of maximum scour. The
magnitude of maximum scour averages twice the maximum bar
height but may be as much as three times the maximum bar
height in some locations [Ref 10-11]. Site-specific data for the
beach of interest are required to obtain appreciably more accurate
predictions of scour than are given in Equations 10-1 to 10-4 Figure 10.4-1. Clear water scour and general sediment
transport near a pile.
[Ref 10-11].

10-3
an obstruction when the velocity in the undisturbed area outside
the influence of the structure is roughly 0.5 times the critical
velocity (velocity that will initiate sediment movement on an
unobstructed flat bed). As velocity increases, scour in the
vicinity of the obstruction increases rapidly until the critical
velocity is reached. Scouring that occurs before the critical
velocity is reached is termed clear water scour since sediment
upstream of the obstruction is not being eroded and carried
downstream. The maximum scour depth at the structure occurs
at the critical velocity. At velocities above the critical velocity,
sediment over the entire bed is in motion and scour depression Figure 10.4-2. Variation of maximum clear water scour
actually decreases to a depth about 10% less than its maximum depth with seafloor material diameter at a
cylindrical pier (after Ref 10-16).
value. Some recent model data [Ref 10-14] suggest that at very
high velocities scour may actually begin to increase again, but
bulence due to the presence of the structure, it is more likely to
at a much lower rate than the initial part of the curve in Figure
erode. If erosion does occur and it extends to the depth of the
10.4-1. This concept of equilibrium depth has also been
finer material, the scour depth will increase dramatically because
observed in the ocean [Ref 10-15].
of the finer materials lower resistance to scour.

10.4.1.2 Layered sediment is not the only way shielding may


Influence of Sediment Size. For sand-size
affect erosion. The peak in the curve in Figure 10-4.2 for the
sediment, the maximum scour depth at a pile may vary from
0.55-mm sand is thought to reflect shielding of the fine sand by
2.3 to as little as 0.3 pile diameters, depending on the
the coarser material within the sample.
composition of the sand--specifically, the ratio of the standard
deviation of the grain size distribution to the median diameter 10.4.1.4 Scour Due to Waves and Currents
[Ref 10-16]. The curves in Figure 10.4-2 show that relative
Rational scour predictive techniques for piles or small
scour depths as much as two pile diameters will occur for sand
objects in the ocean are not yet available. The depth and lateral
sizes common in the ocean. The curves also indicate that, in
extent of scour reported for structures in the ocean normally
general, well-graded materials with higher standard deviation
exceeds the values reported for scour in areas without waves
will experience less scour than poorly graded, uniform
[Ref 10-5, 10-15, 10-18 through 10-20]. However, some model
materials.
investigations [Ref 10-21] show that scour under the action of
10.4.1.3 Shielding. Severe scour can occur when the waves and currents is no greater than for currents alone. At the

surface sediment is composed of material more erosion-resistant Diamond Shoals Light, current was noted to be primarily

than material below it within the scour role. For example, responsible for the scour that occurs [Ref 10-15], and scour

when a structure is located on a layered sediment with a coarse depth sometimes decreased during storms. Presently, no

sand or a gravel material over fine sand, the coarser surface layer adequate explanation has been made for this apparent

ordinarily may not erode. However, when it experiences higher contradiction.

velocities and greater tur-

10-4
10.4.2 Structures Piercing the Water Surface large structures. As structure size is reduced or as wave length is
increased, vortices begin to form at the back of the cylinder, and
More data exist for piling and for large cylindrical
the lateral extent and depth of scour increase. For a/D = 1.0,
structures that pierce the water surface than for structures that do
scour depth is about 0.4 D. Most often, the situations that exist
not. Much of the available data and general scour observations
for piles and piers is where a/D >> 1.0. In this flow pattern, the
are derived from model data, which must be carefully evaluated
vortices are shed alternately from each side of the cylinder,
before application to a field problem. More field information,
producing a highly turbulent wake behind the cylinder. The
however, should be forthcoming as observations of scour
turbulent wake and continual vortex shedding for a/D >> 1.0 is
occurring at large, deep water, gravity platforms, such as those
similar to that for steady flow around a cylinder. For oscillatory
used for offshore oil recovery in the North Sea, are reported.
flow, the scour pit at individual piles develops symmetrically
Figure 10.4-3 shows flow and scour characteristics at a
and extends two to three diameters from the edge of the pile, as
circular cylinder. The figure illustrates the fluid flow field and
shown in the figure. In flow from only one direction, the scour
scour differences between very large structures (like deep water
pit extends several diameters downstream of the obstruction,
gravity platforms) and small cylindrical structures (like single
producing an asymmetrical scour pit.
piles) in waves. Types of flow and patterns of scour in this
The scour pit around pile groups can be much larger than
figure are characterized according to the ratio a/D, where a is the
scour that develops around individual piles. Figure 10.4-4
wave semi-orbital length at the seafloor, and D is the cylinder
combines field and model data collected for scour at piles and
diameter. Given the water depth, wave height, and wave length,
pile groups. It can be used to roughly estimate maximum scour
the semi-orbital length is obtained from:
at single piles, or at pile groups in currents only, or at pile
(10-5) groups in currents and waves. In the figure, the ratio of scour
depth to pile diameter is plotted against Ns, a dimensionless
where. H= wave height [L]
d = water depth [L] sediment number commonly used for describing sediment
L = wave length [L]
erosion [Ref 10-3]. Ns is determined by:
In deep water, the wave period is used to calculate the
deep water wave length as follows: (10-7)

where: u = mean undisturbed current (or maximum wave


(10-6) orbital velocity) [L/T]
S = specific gravity of the bed material
where: T = wave period [T] g = gravitational acceleration [LT-2 ]
g = gravitational acceleration [L/T2I D50 = median sediment diameter [L]
Several texts (e.g., Ref 10-3, 10-12) list values of d/L and sinh
(2d/L) for various values of d/Lo. For values of a/D When waves and currents are present, the mean current
considerably less than 0.2, the flow approximates the smooth velocity and the maximum wave orbital velocity may be added
frictionless flow depicted in the figure. This flow situation and to give a conservative value of u for Equation 10-7. Two design
the relatively mild scour pattern shown are expected to occur for limits are indicated in Figure 10.4-4, both of
very

10-5
Figure 10.4-3. Idealized wave-induced flow and scour patterns around a vertical cylinder (after Ref 10-22).

10-6
Figure 10.4-4. Summary plot of field and model scour depth data
at single piles and pile groups.

which are in terms of the maximum value for the relative scour wave-induced velocities and uni-directional current velocities at
depth (a ratio of scour depth to the pile diameter): (1) maximum the bottom exceed 5 and 3 fps, respectively. For severe ocean
relative scour depth for single piles in harsh environments and environments with both currents and waves, the higher
for pile groups in only current or in mild ocean environments: maximum limit curve in Figure 10.4-4 is recommended.
and (2) maximum relative scour depth for pile groups in current The effect of structure shape on scour for very large
and waves. structures (where a/D << 0.2) is shown in Figure 10.4-5. Scour
A somewhat complex method is available for computing depth and extent for square- and hexagon-shaped structures are
the maximum scour for single piles and pile groups in a steady compared to these data for a cylindrical shape under conditions
current that takes into account water depth, sediment number, of waves only and waves and currents. The hexagon with a
pier length, pier shape, and pier angle of attack [Ref 10-4]. point in the direction of the flow and the circular shape result in
However, a relative scour depth value of 2.0, the lower the least scour of the shapes in this study. For a square shape,
maximum limit in Figure 10.4-4, is recommended as a however, the orientation producing the least scour is a flat side
conservative design value [Ref 10-4] for all but very large perpendicular to the flow.
structures (where a/D << 0.1). The low magnitudes of relative scour shown indicate that
A simple rule for estimating scour at pile groups in erosion around small structures (like piles with high a/D values
severe ocean environments is not yet possible. Relative scour for the same a) and large structures differs considerably. The
depths from three to over four pile diameters have been observed relationship between scour depth and cylinder diameter indicated
[Ref 10-15, 10-17, 10-19, 10-20] at pile-founded structures in by Figure 10.4-4 are applicable to cylinders up to a few meters
the Gulf of Mexico and in the Atlantic Ocean near Cape diameter but do not reflect the smaller relative scour
Hatteras, N.C. It appears that the most severe scour has
occurred at structures located in ocean areas where

10-7
Figure 10.4-5. Scour comparison for very large circular, square and hexagonal cylinders of equal
cross-sectional area where a/D << 0.2 (after Ref 10-22).

10-8
depths near very large objects, like shipwrecks or rock outcrops was not; no explanation was given in the study. The other
[Ref 10-24]. For very large structures, model data can provide pipelines were not undermined.
reasonably good estimates of scour in waves and currents. A For an initial estimate, the scour depth for a pipe,
study made of 13 large gravity platforms placed in the North whether it will be undermined or not, can be estimated from the
Sea concluded that a square-shaped foundation is much more following equation [Ref 10-26]:
sensitive to scour than a cylindrical one [Ref 10-25].
(10-8)
10.4.3 Structures Resting on the Seafloor
where u is the mean flow velocity [L/T].
A great deal more information is available about scour at It is notable that Equation 10-8 indicates that scour depth
bottom-piercing structures, like tiles and piers, than is available is more sensitive to changes in the value of pipe diameter than
for scour at bottom-resting structures, like footings and it is to changes in flow velocity.
pipelines. A general lack of data and apparent lack of coherence
among the available data for scour at footings preclude 10.4.3.2 Footings . Scour at footings is also difficult to

construction of a plot similar to Figure 10.4-4 for guidance. In assess. In addition to the flow complexities, the footing moves,
particular, the mechanisms for scour around bottom-resting or buries, as significant scour occurs. Figure 10.4-7 shows a
structures due to waves are not clear. likely sequence of events for scour-initiated burial of a small
object placed on the seafloor. Flow and turbulence around the
10.4.3.1 Pipelines . Scour at a pipeline can remove
object cause local scour to develop. As the material beneath the
soil from beneath the pipe, causing free spans of pipe*. These structure is removed and the object becomes unstable, it drops
unsupported spans must withstand flow-induced forces as well progressively lower into the deepening scour pit. The object
as their own weight. Pipeline burial and pipeline movement can may also rock in response to ocean waves and pump sediment
also occur and have received considerable attention. Scour as a from beneath itself, or animal undermining can cause or
causative mechanism however has not been adequately increase settlements. Small hiding places under a footing
evaluated, and the problem of settling as a result of scour is become habitats for marine animals or are created by them for
difficult to predict. habitats. Larger burrowing animals and animals preying on
In model studies of pipeline scour, the pipe is usually others can increase scour further. Scour accelerates, and
fixed in the flow. Therefore, scour holes under the pipe do not eventually the footing subsides to a level below the original
result in settling. Figure 10.4-6 shows scour development in a height of the seafloor. At this elevation, infilling of the scour
model study of pipelines initially: completely buried, buried to depression occurs. Eventually, the footing can be completely
three-fourths of its diameter, buried to one-half of its diameter, buried.
and placed on the seafloor surface. The data were taken for A summary of observations on scour at small footings
steady uni-directional flow at velocities from 1.2 to 1.5 fps. In and objects on the seafloor [Ref 10-1] shows that objects placed
the study, all pipelines lying on the surface were undermined. in ocean water depths less than 30 feet are buried more quickly.
One of two pipes buried initially to half a diameter was At depths of 60 feet or more in relatively mild environments,
undermined, while the other scour caused settlement and tilting, but not burial of structures.
____________ Field data
* These unsupported spans of pipeline can occur as a result of
either seasonal or local scour.

10-9
Figure 10.4-6. Scour development at a pipeline exposed to a uniform current of 1.2 to 1.5 ft/sec
(after Ref 10-26).

10-10
Figure 10.4-7. Burial of a small object on the seafloor from a progressive scour sequence.

suggest that, in general, when steady or oscillatory currents Army Coastal Engineering Research Center (CERC) pier, 6 feet
exceed 0.7 fps, significant scour and eventual burial is likely to of sediment was removed in a 48-hour period in which the
occur. Higher velocity currents will undermine and bury the significant wave height was 6 feet [Ref 10-11]. These and other
footings more rapidly. Lower velocity currents are not expected data indicate that scour may occur very rapidly, usually in a
to result in significant scour under the footing. period of hours or days, in response to changing current
velocity or wave height. Also, it appears that the more severe
10.4.4 Time for Scour Development
the flow conditions, the faster the equilibrium scour depth will
Much evidence exists that the maximum scour depth for a be reached; however, at some future time, additional scour can
particular environmental condition is approached relatively occur in response to a higher level of waves or current. On the
quickly. In studies evaluating scour, 90% of the observed other hand, higher flow conditions may not increase the scour
scouring has been estimated to occur in a period of days to depth if the critical flow velocity has been exceeded.
weeks [Ref 10-27]. In another study, scour at a model pile For scour development near pipelines, it may also take
exposed to waves occurred rapidly at first, reaching 75% of its less time to reach maximum scour depth at high than at low
maximum depth after 1,000 waves and 100% after 2,000 to velocities. In one test series maximum scour depth was reached
6,000 waves [Ref 10-28]. At the Diamond Shoals Light a considerably more quickly at a higher velocity (100 minutes at
change in scour depth of 2 to 3 feet was observed in a single 1.7 fps) than at a lower velocity (1,600 minutes at 0.65 fps).
day [Ref 10-15]. At the

10-11
10.5 MINIMIZING SCOUR allowed to settle into the bottom as scour occurs. This will
result in a generally stable structure that undergoes small
Minimizing the effects of scour is usually accomplished
settlements. Properly designed, the tripod will not be harmed by
in one of three ways: (1) avoiding areas where the flow
the small reorientations. The primary disadvantage of the tripod
conditions are likely to produce scour, (2) using scour-resistant
is its open framework and height off the bottom, which make it
structures, and (3) employing scour protection measures.
more susceptible to snagging and damage by anchors and
As previously mentioned, field data suggest that, on a
commercial fishing equipment.
cohesionless seafloor, significant scour will occur when current
velocity or wave orbital velocity is greater than about 0.7 fps.
10.5.2 Scour Protection Measures
Below this value, reported scour has been negligible. If a
Scour protection measures can be successfully taken in
structure is to be placed in a nearshore area known to experience
many instances. To be most effective, scour protection should
large seasonal changes in sediment elevation, installation can be
be installed before, or immediately after, a structure is installed.
planned for the most advantageous season. For example,
Since scour most likely occurs soon after installation, failure to
pipeline to be lain and buried from onshore to some offshore
install scour protection early may limit its effectiveness.
location will benefit greatly by installation when sediment
In the case of offshore pile-supported structures, waiting
elevation is at a seasonal minimum. Frequently, however, the
until a large scour hole develops can also greatly increase the
best scour prevention technique is simply to anticipate the
cost of scour protection. The hole may have to be filled to
scour that will occur and design to this level.
provide adequate lateral restraint for the piling before the scour
10.5.1 Scour-Resistant Structures protection system is installed. Retrofit operations are generally
much more costly than operations included as part of an
Piles are very scour-tolerant structures. They are the
installation. For example, installing mats for scour protection
traditional structures used in locations where scour is expected
for 18 footings during a retrofit operation at the Port Everglades
to be severe. By anticipating the occurrence of scour around the
Shallow Water Range consumed approximately the same
piles and designing accordingly, the scour can be tolerated in
amount of time required for installing and aligning the footings
most cases. In applications where lateral load on the pile is
themselves. Had the scour protection been installed before or
high, the loss of a significant amount of material from the
during footing installation, the time required for cutting and
seafloor surface could significantly reduce pile lateral capacity.
fitting scour mats would have been substantially reduced.
This scour-caused loss in lateral resistance may not be
However, it is extremely difficult to know how much
economically recovered through use of larger diameter or longer
scour will occur and, therefore, how much protection is
piles. In these cases, reliance on scour protection measures may
necessary until after the structure is installed. The scour
be necessary.
protection system, for complete protection, should extend to the
Tripod structures (structures supported by frame with
limit of the largest scour hole expected. For large structures
three footings) are well-suited to areas where significant scour is
(a/D < 0.2), protection extending one radius from the edge of
expected. Since a tripod can have relatively small diameter
cylindrical structures and one diameter from angular structures
structural members above the seafloor, the resulting scour will
has been recommended [Ref 10-22]. For small isolated
be small. Footings or pads on each tripod leg can be articulated
structures (a/D > 1.0),
and

10-12
protection extending at least two diameters from the edge of the three filter requirements with the previous, lower, filter material
structure is recommended. For pile groups in the ocean, the treated as the base material. A properly graded, cubical or
protection should extend approximately one radius of the pile rounded material with a specific gravity of 2.5 or higher can be
group beyond the outermost piles. For large structures, model placer in layers as thin as practical but maintaining a minimum
tests can be used to provide a reasonably good estimate of the thickness based on grain size of the layer below [Ref 10-17].
extent of scour for designing protection measures. Several The material should be placed by tremie or similar method to
techniques used to minimize scour are described below. prevent size segregation of the particles. An example of a filter
It must be kept in mind that scour can overwhelm a design is shown in Figure 10.5-1 for the Diamond Shoals light
system that is insufficient for protection from existing forces. station. The filter design criteria [Ref 10-17] are satisfied by the
An unusual example is the reported scour history of the filter material used; bed material and filter material grain sizes
Christchurch Bay Tower [Ref 10-25], a low-profile, circular, are shown in Figure 10.5-2.
flat-based footing placed in 27-1/2 feet of water. Shortly after Filters that depart significantly from "proper design" are
placement it was undermined severely during storms, causing a common; however, even the considerable departure from
"rocking" erosion with only the footing center eventually standard filter requirements did not significantly affect the
resting on the sand. It was removed, fitted with skirts, and performance of filters used in the Gulf of Mexico [Ref 10-15,
reinstalled, but scour again took place and undermined the 10-17]. In at least one case, oyster shells have been successfully
skirts. The scour pit under the footing was then grouted and a used as backfill [Ref 10-15]. Steel mill slag has also been used
protective barrier of sand bags in a grouted matrix was placed successfully in an inverted filter in the North Sea [Ref 10-31].
outside the skirts, extending beyond the footing. The last Also, in the Gulf of Mexico [Ref 10-5] where the slag has been
reported information [Ref 10-25] was that scour was occurring user as the armor layer above an oyster shell layer, it has
under the edge of the sand bags, which were likely to be withstood the elements for more than 10 years.
undermined. The inverted filter is usually coverer with a layer of
protective stone sized to resist storm conditions. A formula for
10.5.2.1. The Inverted Filter . The inverted filter and
sizing this rip-rap, presented in Reference 10-12, is based on
protective rip-rap cover is a method used successfully [Ref 10- variables of water velocity, stone specific weight, and stone
29, 10-30] for preventing scour. The filter consists of one or diameter. Similar formulas and design curves are available in
more layers of cohesionless soil placed on top of the bed other references such as Reference 10-32). Other published
material to be protected, each layer being more coarse than the recommendations for sizing rip-rap (e.g., for bridge piers) call
previous layer. The principle is to produce a surface that will for design to resist a flow velocity at least twice that expected
not be eroded by scour. The filter layers may transition from in the undisturbed bed area [Ref 10-4]. This would result in use
material as large as rip-rap size down to the size of the bed of a rip-rap approximately four times as large as that sized on
material in sufficient layers to prevent the smaller grain size the basis of expected velocity. However, on the basis of
materials from being leached out of their layers. Grading criteria experiences at the Diamond Shoals Light [Ref 10-15] and at
have been established for this type of filter on the basis of platforms in the Gulf of Mexico [Ref 10-5], this
particle size distribution in each adjacent layer [Ref 10-17]. recommendation appears overly conservative for application to
Each successive layer must satisfy offshore

10-13
Figure 10.5-1. Example of an inverted filter for scour protection at an offshore structure (after Ref 10-15).

Figure 10.5-2. Example of a filter material specification (Diamond Shoals Light).

10-14
ocean structures. In these cases, rip-rap sized on the expected - Open area should not exceed 40%.
flow velocity without a multiplier, or sized smaller than that
Maximum open area: When adjacent to soils containing
value [Ref 10-5], has been successfully used. Until more field more than 50% silt by weight and having little or no
cohesion:
data are available, maximum expected flow velocity at the
undisturbed bed, without the multiplier of two, is recommended - EOS should be no larger than the opening in a no. 70
sieve.
for use in resign formulas for sizing rip-rap at structures in the
ocean. - Open area should not exceed 10%.

10.5.2.2 Filter Fabric. Filter cloth has been used Various methods have used to hold filter fabric in place.
successfully to combat scour [Ref 10-33], sometimes to The best and most commonly used method is to place a
overcome drawbacks that can be present with inverted filters. protective rock layer over the fabric.
For example, the latter are costly, and sometimes the materials
10.5.2.3 Artificial Seaweed (Polypropylene Fronds).
required to meet filter specifications are not available. The cloth
Artificial seaweed has been used to arrest erosion by lowering
replaces the graded soil filter to prevent migration of finer soils
current velocity in the vicinity. Results indicate that it also
up through the rock cover. Although a variety of filter cloth
encourages reposition of sediment [Ref 10-35, 10-10]. Fronds
materials are available, very little published information is
arc more likely to be used in conjunction with a filter fabric;
available regarding their performance. Filter fabric is made from
how to keep the lightweight fronds in place in harsh
various man-made materials, including polyester, poly-
environments is a primary consideration. Unlike filter fabric,
propylene, and polyvinylidene. Although two basic cloth
they cannot be weighted from above by a dense surface layer
categories are available--woven and nonwoven--only the woven
because they must be the top layer of protection. No successful
fabrics are recommended [Ref 10-34].
preferred method is presently apparent.
The two primary requirements for filter cloth used in
scour protection are: (1) it must not clog and (2) it much have
10.5.2.4 Flat Plate. A horizontal flat plate or collar with
and maintain good tensile strength in seawater and sunlight.
a minimum diameter of about three pile diameters may be used
Based on clogging tests, the following recommendations
to reduce scour at a pile. The plate should not be rigidly
are made for selecting filter cloth [after Ref 10-34]:
attached to the pile, but should be free to slide down the pile to

Only woven filter cloths with distinct openings are to be adjust for any undermining that occurs. Initial burial to one-
used. third of the pile diameter, if possible, is recommended. Plates at

Minimum open area: or below the undisturbed bed protect the seafloor soils adjacent
to the pile from pile-created turbulence and have reduced scour
- Equivalent opening size (EOS) of the fabric should not
be less than the openings in a no. 100 sieve by as much as 50% [Ref 10-36].

- Open area should not be less than 4% 10.5.2.5 Peripheral Skirt. A method of limiting scour

Maximum open area: When adjacent to granular materials around footings in nonsevere scour areas has been incorporation
containing 90% or less silt by weight and with little or no of a peripheral skirt to penetrate the seafloor and prevent minor
plasticity, passing the no. 200 sieve:
scouring from undermining the footing. A peripheral skirt
- around small footings also produces an anaerobic environment
below the footing, which helps prevent scour and settling

10-15
caused by burrowing sea creatures [Ref 10-37]. This method, 10-3. R. Silvester. Developments in geotechnical engineering
however, it is not recommended for severe scour environments. 4B, coastal engineering II, sedimentation, estuaries, tides,
If scour exceeds the depth of the skirt, erosion of material from effluents, and modeling. New York, N.Y., Elsevier Scientific
beneath the footing will still occur [Ref 10-23, 10-25]. Publishing Co., 1974.
Subsequent rocking, settling, and possible burial of the footing
10-4. H. Breusers, G. Nicollet, and H. Shen. "Local scour
may then follow.
around cylindrical piers," Journal of Hydraulics Research, vol
10.5.2.6 Protection of Pipelines . Pipelines supported 15, no. 3, 1977, pp 211-252.

by the seafloor surface can be undermined, creating excessive


10-5. C.J. Posey and J.H. Sybert. "Erosion protection of
stresses due to unsupported spans of pipe. Protection from
production structures," in Proceedings of the Ninth Convention,
scour around pipelines is difficult. If a protection system that
IAHR, Dobrovnik, Yugoslavia, 1961, pp 1157-1161.
uses a heavy cover over the pipeline is less than 100%
successful, an unsupported span problem may be greater than 10-6. J.M Atturio, Estimating and minimizing sediment scour
would exist without protection. at seafloor structures, Naval Civil Engineering Laboratory,
Protection from local scour and from potential pipeline Technical Memorandum M-42-81-3A. Port Hueneme, Calif.,
snagging is recommended for relatively permanent oil pipelines Oct 1981.
[Ref 10-10] by means of a complete covering of heavy concrete
mats extending many diameters out from the pipe. This is 10-7. T. Yamamoto. "Sea bed instability from waves," in

extremely costly. Proceedings of the Offshore Technology Conference, Houston,

Seasonal scour can remove an entire layer of material Tex., 1978. (OTC 3262)

over a large area and protection may not be possible. Much of


10-8. H. Moshagen and P.L. Monkmeyer. "Wave-induced
this problem can be avoided by placement of the pipeline after
seepage forces on embedded offshore structures," Civil
the seasonal scour has occurred when bottom elevations are at a
Engineering in the Oceans IV, vol II. New York, N.Y.,
minimum.
American Society of Civil Engineers, Sep 1979, pp 758-773.

10.6 REFERENCES
10.9. O.S. Madsen. "Wave-induced pore pressures and effective
10-1 C.F. Nordin and E.V. Richardson. "The use of stochastic stresses in a porous bed," Geotechnique, vol 28, no. 4, Dec
models in studies of alluvial channel processes," in Proceedings 1978, pp 337-393.
of the Twelfth IAHR Conference, Ft. Collins, Colo. 1967, vol
2, pp 96-102. 10-10. I.E. Maidl. "New experiences in scour protection for
offshore platforms and pipelines," in Proceedings of the Second
10-2. W.W. Sayre and W.J. Conover. "General two Canadian Geotechnical Conference on Marine Geotechnical
dimensional stochastic model for the transport and dispersion of Engineering, Halifax, Nova Scotia, Jun 1982.
bed material sediment particles," in Proceedings of the Twelfth
IAHR Conference, Ft. Collins, Colo., 1967, vol 2, pp 88-95. 10-11. A.E. Dewall and J.A. Christenson. Guidelines for
predicting maximum nearshore sea level changes on
unobstructed beaches, Army Coastal Engineering Research
Center, Unpublished Report. Fort Belvoir, Va., Dec 1979.

10-16
10-12. Shore protection manual, Army Corps of Engineers, 10-22. Wave Scour potential around large idealized shapes,
Coastal Engineering Research Center. Fort Belvoir, Va., 1977. Hydraulics Research Station, Report No. EX832. Wallingford,
England, Sep 1978. (Offshore Energy Technology Report OT-
10-13. R.J. Hallermeier. Calculating a yearly depth to the
R7826)
active beach profile, Army Coastal Engineering Center,
Technical Paper No. 77-9. Fort Belvoir, Va., Sep 1977. 10-23. A. E. DeWall. Experiments on nearshore scour around
small footings and pile-supported piers, Army Coastal
10-14. S.C. Jain and E.E. Fischer. "Scour around bridge piers
Engineering Research Center. Fort Belvoir, Va., Jul 1981.
at high flow velocities," Journal of the Hydraulics Division,
Proceedings of the American Society of Civil Engineers, vol 10-24. T. Carstens. "Wave-seabed structural interaction by
106, no. HY11, Nov 1980, pp 1827-1842 Bijker (Vol. 1, p. 830)," in Proceedings of BOSS '76, vol 2,
Trondheim, Norway, Aug 1976. Trondheim, Norway,
10-15. R.C.K. Au. Scour investigation Diamond Shoals light
Norwegian Institute of Technology, 1976, pp 558-562.
station, Cape Hatteras, North Carolina, Coast Guard, Contract
Report. Atlanta, Ga., Dames and Moore, Sep 1975. (Contract 10-25. R. Dhalberg. "Observation of scour around offshore
No. DOT-CG05-988). structure," Det norske Veritas, in Proceedings of the Second
Canadian Conference on Marine Geotechnical Engineering,
10-16. Raudkivi and Ettema. "Effect of sediment gradation on
Halifax, Nova Scotia, Jun 1982.
clear water scour," Journal of the Hydraulics Division,
American Society of Civil Engineers, Oct 1977, p. 1209. 10-26. S.P. Kjeldsen, O. Gjorsvik, and K.G. Bringaker.
(With discussion in May 1979) Experiments with local scour around submarine pipelines in a
uniform current, River and Harbour Laboratory. Trondheim,
10-17. C.J. Posey. "Protection against underscour," in
Norway, Mar 1974.
Proceedings of the Offshore Technology Conference, Houston,
Tex., 1970. (OTC 1304) 10-27. H.N.C. Breusers. "Local scour near offshore structures,"
in Proceedings of the Symposium on Offshore Hydrodynamics,
10-18. H.D. Palmer. Wave induced scour around natural and
Wageningen, The Netherlands, Aug 25-26, 1971.
artificial objects, Ph.D. thesis, University of Southern
California. Los Angeles, Calif., 1970 10-28. D.R. Wells and R.M. Sorenson. Scour around a circular
pile due to associated wave motion, Coastal and Ocean
10-19. J.C. Kreig. "Criteria for planning an offshore pipeline,"
Engineering Division, Texas A&M University, Report No.
Journal of the Pipeline Division, American Society of Civil
113-COE. College Station, Tex., Jan 1970. (Sea Grant
Engineers, vol 91, no. PLI, Jul 1965, pp 15-37.
Publication No. 208).

10-20. R. Blumberg. "Design environmental extremes,"


10-29. Laboratory investigation of filters for Enid and Grenada
Pipeline Industry, vol 25, no. 4, Oct 1966, pp 31-34.
Dams, Waterways Experiment Station, Technical Memorandum
No. 3-245. Vicksburg, Miss., Jan 1948.
10-21. E W Bijker. "Wave - seabed - structure interaction," in
Proceedings of BOSS '76, vol 1, Trondheim, Norway, Aug
10-30. C.J Posey. "Erosion prevention experiments," in
1976. Trondheim, Norway, Norwegian Institute of Technology,
Proceedings of the Thirteenth Congress, IAHR, Kyoto, Japan,
1976.
1969, vol. 2, pp 211-219.

10-17
10-31. R.N. Van Dijk. "Experience of scour in the southern D Diameter of pipe and pipe piles
North Sea," Journal of the Society of Underwater Technology,
D15 Filter material grain diameter below which 15% of the
Mar 1981, pp 18-21. material falls

10-32. C.D. Ponce-Campos. The stability of protective cover D50 Median grain diameter of a soil, or filter material [L]

layers for embedded underwater structures, Ph.D. thesis, d50 Standard mesh size through which 50% of material
University of Michigan. Ann Arbor, Mich., 1978. will pass (median diameter) [L]

d85 Bed material grain diameter below which 85% of the


10-33. N.D. Wilson and W. Abel. "Seafloor scour protection
material falls
for a semi-submersible drilling rig on the Nova Scotian Shelf,"
in Proceedings of the Offshore Technology Conference, dc Water depth of seaward limit of significant sediment
level change, also called the close-out depth [L]
Houston, Tex. , 1973. (OTC 1891)
dm Water depth at which maximum sediment level change
10-34. C.C. Calhoun, Jr. Development of design criteria and occurs [L]
acceptance specifications for plastic filter cloths, Army
EOS Equivalent opening size
Waterways Experiment Station. Technical Report S-72-7.
g Gravitational acceleration [L/T2]
Vicksburg, Miss., Jun 1972. (AD745085)
H Wave height [L]
10-35. P.A. Tufto and E. Tesaker. ICI scour prevention system _
H Average monthly significant wave height [L]
hydraulic flume study, River and Harbour Laboratory,
Norwegian Institute of Technology. Trondheim, Norway, Apr He Annual extreme wave height (significant wave height
1976. exceeded 12 hr/yr) [L]

L Wave length
10-36. S. Tanaka and M. Yano. "Local scour around a circular
cylinder," in Proceedings of the Twelfth IAHR Congress, Ft. Lo Deep water wave length [L]

Collins, Colo., pp 193-201. Ns Sediment number (defined by Equation 10-7)

10-37. J.S. Muraoka. Animal undermining of seafloor S Specific gravity


installations, Naval Civil Engineering Laboratory, Technical
T Wave period [T]
Note N-1124. Port Hueneme, Calif., Nov 1970.
u Mean undisturbed current velocity or maximum wave
orbital velocity [L/T]
10.7 SYMBOLS
xs Lateral extent of scour from the pile or pier [L]
a Semi-orbital length [L]
ys Maximum scour depth at an obstruction [L]
d Water depth [L]
ymax Maximum seasonal sediment level change in
unobstructed nearshore [L]

Standard deviation of monthly average significant wave


heights [L] and grain size distribution [L]

10-18
Chapter 11

SLOPE STABILITY ASSESSMENT

11.1 INTRODUCTION basically a question of whether the anticipated loading will


exceed the resistance available in the sediment mass. References
This chapter provides a background description of the
11-1 and 11-2 describe analytical methods for dealing with slope
major elements and processes influencing offshore slope
stability problems in detail and should be consulted if more
stability. Specific strategies for dealing with field situations are
detail is desired than is presented in this chapter.
proposed, but analytical methods that treat slope stability
Between the terms loading and resistance, however, there
problems in detail are not presented. A structured evaluation
is great uncertainty and the need for much detail. The nature of
procedure applicable to problems of offshore sediment
the most important loading functions is usually uncertain.
instability and based on fundamental principles governing the
Values for loading, for example, are almost always associated
processes is defined. The most important of these principles is
with events different from the present and difficult to predict--
the basic relationship between loading and resistance in a
such as large waves or earthquakes. For designing, probabilistic
sediment mass. These forces are shown schematically in Figure
descriptions are often used as the basis for defining maximum
11.1-1. In any case of potential sediment instability, the
magnitudes for waves and
evaluation is

Figure 11.1-1. Loading and resistance forces on a potential slide mass.

11-1
earthquakes and for how often they occur. Thus, offshore equipment and the large because enough data to define their
sediment stability analysis is often based, in part, on limits may not or cannot be taken.
probabilistic terms (discussed in more detail in Section 11.3.2).
11.2 FORMS OF INSTABILITY
Describing values for resistance to loadings requires an
evaluation of stratigraphy, which is often complex, and core The understanding of offshore sediment instability is
sampling and testing of recovered sediments. The data added by classification of movements into a limited number of
interpretations to determine stratigraphy require considerable types [Ref 11-1, 11-3]. Table 11.2-1 classifies most submarine
knowledge of marine geotechnology and considerable technical movements into four simplified models: translational slides,
interpretative skills. In almost all cases, the amount and quality rotational slides, flow slides, and turbidity currents. This is an
of data upon which the resistance is based are very limited. artificial classification as most actual situations involve more
Another concern is the scale of offshore slides. Mass than one type of movement.
movements measuring a few feet wide or long to several miles The form of a submarine slide that has already occurred is
on a side are found. Both the smaller and larger slides may be almost always defined from the record of an acoustic
difficult to identify: the small because they fall outside the geophysical survey (Section 11.6.3). However, because of
resolution of survey record distortion at the time the data are taken, such

Table 11.2-1. Movement Models for Submarine Slides


in Soft and Loose Sediment
Type of Movement Characteristics
Translational Slides Shear displacement along one or several surfaces or within a
relatively narrow zone. Movement predominantly along planar
or gently undulatory surfaces. Movement frequently controlled
by surfaces of weakness, such as faults, joints, bedding
planes, and variations in shear strength between layers of
bedded deposits.

Rotational Slides Shear displacement along one or several surfaces or within a


(Slumps) relatively narrow zone. Movement due to forces that cause a
turning moment about a point above the center of gravity of
the unit. Surface of rupture concaves upward.

Flow Slides Movement within displaced mass such that the form taken by
Mud, Silt, or Sand moving material or the apparent distribution of velocities and
displacements resemble those of viscous fluids. Slip surfaces
within moving material are usually not visible or are short-
lived. Boundary between moving mass and material in place
may be a sharp surface of differential movement or a zone
distributed shear. Movement ranges from extremely rapid to
extremely slow.

Turbidity Currents Complete fluidization of sediment with introduction of


additional water. Mixing and resuspension of material during
transport. Extremely rapid rates of movement. Transport of
surface material from a large geographical area hundreds of
miles to a new location is possible. A major geological
event.

11-2
records often give the incorrect impression that the sliding common as a slope becomes steeper and more irregular
process involves much more rotation than is, in fact, the case. Consequently, rotational slides tend to be more common on
If the vertical to horizontal scale distortion (typically about 10 continental slopes, canyon walls, and similar features than on
times) is eliminated, the actual form of the submarine slide is the more flat shelf or plain areas. Many small-scale slides, such
mere apparent. as collapse of a pipe trench or the slumping of a dredged face,
are of rotational form.
11.2.1 Translational Slides
11.2.3 Flow Slides
Translational slides, such as depicted by the enhanced
profile shown in Figure 11.2-1, are the principal failure mode Flow slides develop when the shearing resistance of the
on the continental shelf and upper continental slope, where sediment is reduced significantly as a consequence of shearing or
most of mans engineering activities take place. Many leading some other process. The two most common sediment
mechanisms, such as earthquakes and waves, can cause rapid characteristics that lead to such strength reductions are high
sediment movement in a translational mode. Also, where they sensitivity and susceptibility to strength reduction under cyclic
occur, slow creep movements probably take place on planar loading. Soil liquefaction is an extreme example of the latter.
surfaces associated with a strength anomaly. This type of Sensitivity is a property generally attributed to clays, while
movement is also classified as a translational slide. high potential for liquefaction is identified with loose sands
The geophysical record of a flow slide is distinguished by
11.2.2 Rotational Shoes
the mixed and chaotic form of the slide mass or debris, as
The second most common slide form is rotational. An illustrated in Figure 11.2-3. Frequently, the flow slide is
example of multiple rotational slides is shown by the seismic associated with much longer movement distances than
profile on Figure 11.2-2. Rotational slides become more translational or rotational forms. It is

Figure 11.2-1. Enhanced seismic profile of a translational slide.

11-3
Figure 11.2-2. Seismic profiles of a rotational slide.

Figure 11.2-3. Seismic profiles of a submarine flow.

11-4
not always apparent where the slide mass originated. Since (2) earthquakes, (3) currents, and (4) structural loads*. Defining
damage can occur anywhere between the source and deposition the load increment magnitude of the first three external
areas of a flow slide, it is important to evaluate the entire extent mechanisms is not easy. Each varies with time, and the critical
of movement in the stability assessment of a region. magnitude of the loading (that used for design) is expected at
some uncertain period in the distant future. Analysis is usually
11.2.4 Turbidity Currents
based on an evaluation of the risk defined by a probabilistic
A turbidity current is a very fluid sediment or sediment- description of the external loading or by a conservative selection
laden water that flows downslope along the seafloor in response of forces that promote instability.
to gravitational forces. Turbidity currents are an active
11.3.2 Probabilistic Prediction of Load
mechanism of submarine sediment transport. The many
references that describe the mechanism and forms of the deposits It is important to recognize that for most sea bottom
are discussed in Reference 11-4. For slope stability assessment, areas a finite probability exists, in which failure or instability
a turbidity current can be treated as an extreme example of flow can occur. In some sea bottom areas where an anchorage or
sliding [Ref 11-2]. There is no definite distinction between the facility is desired, the probability of instability due to natural
slide types or between the deposits they leave. processes is high, and construction of anchoring activities is
likely to increase the risk. Evaluating the risk of inducing a
11.3 LOADING
slope failure can be treated as part of the design process, but it
11.3.1 Loading Mechanisms is usually very complex. A simplified application of a
probabilistic evaluation is often necessary; these are discussed
An accurate assessment of marine sediment stability
below and explained more thoroughly in References 11-5 and
requires that the critical loads, both gravity and external, be
11-6.
evaluated. Instability of offshore sediment is always associated
Knowledge of load and soil resistance variables is
with an increment of external loading added to the influence of
unlikely to be large. To make up for this, probabilistic
gravity on the sediment mass. A reliable estimate of the gravity
distributions can be estimated for these variables. Figure 11.3-
loads depends on an accurate definition of the sea bottom and
1(a) illustrates a simple case in which all the uncertainties in
subsurface geometry. This can usually be accomplished with
the load variables are combined into one curve of loading
high-resolution seismic profiles. The unit weight of sediment is
probabilities. Likewise, all possible levels of resistance and
also required, but reasonable estimates can often be made even if
their probable occurrence are combined into another curve as
samples are not available. Chapters 2 and 3 present data useful
shown in Figure 11.3-1(b). Figure 11.3-1(c) illustrates a
for making these estimates.
common relationship between loads and resistance probabilities
While many external loading mechanisms are possible,
and shows that loading may exceed resistance and that failure, in
four dominate as the initiators of sediment instability; others
this case a slope instability, will occur. This is represented by
need be considered only in special circumstances. The most
the cross-hatched area where the curves overlap. Failure can
common sources of external loads are: (1) waves,
occur ever though a sufficient factor of safety is believed to
____________
have been applied to the forces involved.
*"Man-made" loadings including those from construction and
anchors.

11-5
Figure 11.3-1. Probabilistic description of loading and resistance and risk assessment in
slope stability evaluation.

The methods for combining the separate probabilistic In evaluating the hazard associated with instability of
descriptions of each variable and for mathematically describing marine sediments, the external environmental loads (waves
the relationship of interaction between the loading and resis- currents, and earthquakes) usually dominates any probabilistic-
tance functions are described in References 11-5 and 11-6. These based risk analysis. When one of these factors is considered, it
techniques usually involve complicated mathematics or is necessary to define not only the likelihood of a particular
numerical solutions. level of loading but also the probability of that loading
Collection enough information so that the probabilistic occurring within a particular time interval. In many cases this
descriptions of each variable can be defined is likely to present a time interval is the lifetime of the facility being evaluated or a
serious problem. Frequently, the amount of information critical shorter exposure period. A probabilistic description of
available for many variables is so limited that simplifying environmental loading data is available for many offshore areas
assumptions are necessary. This may severely limit a rigorous that have had historical economic importance. Sources of
probabilistic analysis. The results of such an analysis will be earthquake data, for instance, are listed in Chapter 2, Table 2.2-
highly dependent on the quality of the data used as input. 2.
It is common for one or two of the problems major
11.4 IMPORTANT SOIL PROPERTIES
variables to have a much greater uncertainty than the other
variables. Consequently, the overall probability of failure can 11.4.1 General
be dominated by the influence of one or two factors. For
Data and engineering parameters required for slope
example, sediment weight as a component of the loading may
stability assessment are among those identified in Tables 2.1-1
have a large importance, but the uncertainty may be small
and 2.1-2. Methods for in-situ measurement of these properties
because the volume of slide mass and unit weight are known
and for obtaining samples for laboratory testing are presented in
within narrow limits. On the other hand, probability of a
Chapter 2, while laboratory testing to determine the properties
particular earthquake load occurring in a certain time period is
is discussed in Chapter 3.
likely to involve much uncertainty, and in probabilistic
analysis, this term could dominate. An alternative method, 11.4.2 Special Conditions. Underconsolidated Sediments
therefore, is to assume that uncertainty of the dominant variable
Underconsolidated soils are soils that have not had
dictates the uncertainty of the complete analysis and to do a
sufficient time to consolidate under the
conventional deterministic analysis considering only the
variation of that dominant parameter.

11-6
weight of soil that lies above. They are weaker than normally Stronger marine soils--heavily overconsolidated silts and
consolidated soils at the same depth and are likely to be more clays, in particular--can also experience strength loss from
susceptible to instability problems. Two mechanisms that can repetitive or dynamic loadings. This usually occurs sometime
produce an underconsolidated condition [Ref 11-7] are rapid after the loading as a result of dissipation of negative pore water
sedimentation and the presence of gas bubbles in the soil mass pressures caused by the loading. At this time water is draining
(bubble-phase gas). An underconsolidated condition can usually into the soil, increasing its water content and lowering its
be identified through measurement of unusually high pore water strength (softening). Softening of stiff marine sediments under
pressure in the sediment or an unusually low undrained shearing dynamic loading car be a problem for facilities such as piles and
resistance. These are most accurately determined by in-situ anchorages in permanent or long-term installations. It may also
testing, but determining either from shallow penetrating core be a problem for slope stability.
samples is also possible--though more difficult [Ref 11-8].
11.5 LEVEL OF ANALYSIS
Bubble-phase gas can be identified without deep cores or
in-situ measurements because a very distinctive acoustic pattern It is common to use simpler analytical models to analyze
of wipe out or white area on the seismic record is indicative slope stability for offshore problems than might be used
of this gas. This occurs because of the distinctly lower acoustic onshore. Two major reasons for this practice are: the relative
velocities associated with these sediments. lack of detailed information necessary for most complex
methods of analysis and the flat shape of many offshore slides.
11.4.3 Repetitive and Dynamic Loading Response of
As a consequence, the infinite slope method of analysis [Ref
Sediments
11-2] has often been used for nonrotational slides. This method
Response to repetitive or dynamic loading is an sums forces acting on the soil layer being analyzed as shown in
important aspect of a marine sediment since many offshore Figure 11.5-1(a) and is usually sufficient for this type of slide.
scenarios involve dynamic loading. Waves and earthquakes, two When slumps involve rotational components, as shown in
of the three dominant natural loading mechanisms, are dynamic Figure 11.5-1(b), a method including moment equilibrium as
types of loading. Similarly, many loads induced by man-made well as force summation is used. More involved analytical
activity are also dynamic (e.g., storm-caused ship movements methods, such as a finite element analysis, may be useful in
ultimately resisted by anchors). offshore soil-structure interaction problems but have little
Response of saturated soils to dynamic and repetitive practical application in assessing the potential for offshore
loading has been studied extensively [Ref 11-9, 11-10]. Most sediment instability.
soft marine silts, cays, and loose sands suffer some reduction in
strength and stiffness under such loading. An extreme example 11.6 SITE INVESTIGATION

is soil liquefaction--an almost total less of strength and 11.6.1 General


subsequent loss of resistance to shearing that can occur in loose
The planning and equipment used for in-situ testing and
sands under certain conditions. Although silts and clays
sample collection or site surveys are discussed in Chapter 2.
generally experience a smaller loss of strength, this loss can
This section discusses information gathered by site survey
still lead to bottom instability [Ref 11-11] and soil mass slides
techniques used for slope stability assessment
or the failure of foundations.

11-7
Figure 11.5-1. Summation of forces and moment for stability analysis.

offshore in relation to land techniques for which most stability stabilized and corrected sensor platform to minimize noise in
assessment techniques have been developed. the record, and side-scan sonar.
Because of the differences in geophysical equipment used Although some analytical methods ignore slide end
for offshore surveys, and the high cost and limitations effects for calculation purposes, for identification the extremes
associated with soil sampling, offshore site investigation of the slide mass should not be ignored. A detailed examination
strategies are very different from those on land. At the present of ends will often disclose the causative mechanism for the
time, an offshore survey, compared to a conventional land slide. Therefore, definition of these areas should always be a
survey, will commonly generate: very significant part of a slope stability assessment.
Some lope instability features may be so large they
(1) More detailed information about the internal structure
cannot be easily or accurately identified using conventional
and stratigraphy in the sediment mass through the use
of high-resolution seismic profiling. sensors and line spacing. The solution to this problem is to
insure that sufficient data are gathered over a wide area and to
(2) Comparable detail about the sediment surface through
the use of bottom profiling surveys and side-scan examine the data from a large, as well as small, perspective. A
sonar.
mosaic of profiles or other techniques that examine groups of
(3) Much less detail about the sediment properties because survey lines together are helpful.
the depth of corer penetration will be limited.
11.6.2 Preliminary Studies
Site investigation methods used for slope stability
Except for bathymetry, most areas of the ocean bottom
assessment must be capable of detecting small, as well as very
are unmapped. Even the bathymetry that exists is usually of
large, surface and subsurface features. However, some features
limited use in evaluating hazards and areas of potential
that are important in evaluating a particular location are too
instability because the scale is too small (i.e., covers a large
small or too shallow to be identified using normal seismic
area) and may be based on interpretation from very limited data.
equipment. Alternative methods that increase the ability to
A few areas of the sea bottom have been mapped
detect these features include: alternative frequencies or higher
specifically to identify slumping or
resolution sensors, a

11-8
other stability hazards. Where this has been done, the overall controls on navigation. The vertical scale accuracy is limited by
objective has been related to offshore petroleum development, knowledge of wave propagation velocities through the water and
such as the study of Mississippi Delta mudslides [Ref 11-12]. sediments. Unfortunately, sediment velocities are not usually
As a consequence, the assessment of offshore sediment well-know, and the vertical scale must be eased on assumed
instability usually requires collection and interpretation of new properties. Sediment sampling and testing can be done to
data for each study. improve the accuracy of these assumptions but is seldom used
because of high cost and lack of adequate ship facilities.
11.6.3 Acoustic Surveys
Consequently, interpretation of seismic records is somewhat
A brief discussion of seismic profiling and side-scan qualitative. Determination of soil acoustic properties is
sonar acoustic survey methods for regional surveying was discussed in Reference 11-14.
presented in Chapter 2, Section 2.3. This section discusses What an individual is able to interpret from a high-
seismic profiling and side-scan sensor techniques with respect to resolution seismic record depends not only on technical skills
their role in assessing slope stability. for pattern recognition but on ability to translate pattern
recognition into a framework of knowledge about geologic
11.6.3.1 Seismic Profiling . In general, the broad-band processes. For example, in Figure 11.6-1, the sea bottom and

systems, known as sparkers and boomers, offer the advantage of relatively horizontal or parallel subbottom reflectors are easily

better resolution at shallow sediment depths (high frequencies) identified. In Figure 11.6-1(a), a large slump feature is obvious

along with deeper penetration (low frequencies). The narrow- because of the break in surface profile and the discontinuous

band sources, such as the 3.5-kHz systems, give very high deeper reflectors. In many other cases, faults and other large

resolution at shallow seafloor depths but are limited as to depth subsurface features are also as easy to identify as this slump.

of penetration. On Figure 11.6-1(b), the slump features are much less clear and

The interpretation of high-resolution seismic records is can only be identified because of the very high resolution

not covered in this report. The seismic interpreter must be capability of the 3.5-khz system and skills of the interpreter. It

highly skilled and experienced to discern real features from is notable that a broad-band seismic survey would likely miss

apparent features, which often result from characteristic this feature completely. These two examples illustrate the high

behavior of wave propagation through individual layers. value of careful seismic profiling in identifying slope failure

Features indicated by side echoes and point source reflections are features. This tool is powerful in evaluating potential for slope

identified by pattern recognition, such as cuts through other instability.

layers or hyperbolic shapes. Pitfalls in seismic interpretation The influence of weather and other disturbances on

are discussed in Reference 11-13. equipment is another consideration in seismic profiling

Assuming that a seismic record has been interpreted so offshore. If the sound source or receiver is supported on the sea

that only the direct reflections are being used, the record must surface, then sea state will have a direct influence on the record.

be related to potential stability problems. The record itself will Some systems compensate for this by towing the seismic

be highly distorted in scale, usually with a vertical exaggeration equipment at a constant depth below the sea surface. Other

of approximately 10 times. The horizontal scale usually is systems not only tow at depth but compensate for any up and

known with a high degree of accuracy because of modern down movement of the towed equipment during the electronic
processing of the signal. These

11-9
Figure 11.6-1. Examples of high resolution seismic records of slides caused by slope instability.

11-10
equipment features result in higher quality records and enable when gravity-piston corers cannot reach much of the failure
more data to be collected in a given amount of time because surface, the samples from these corers can provide useful data
adverse sea conditions have less effect on record quality and will by the following methods and assumptions:
not curtail operations as often. They are, however, more
(1) If the sediment deposit is normally consolidated and of
complex and expensive, and fewer of these are likely to be
uniform composition, the testing of soil from more
available for use. shallow depths can be used to determine soil properties
at the depth of the expected failure surface. These
assumptions should be validated and can be, to some
11.6.3.2 Side-Scan Sonar. Side-scan sonar systems
extent, by a high-resolution seismic survey.
have recently become very popular for assessment of slumping
(2) Where sediments are not uniform, high-resolution
and other potentially hazardous surface features. Figure 2.3-1 seismic records can be used to define surface outcrops
(Chapter 2) illustrates side-scan sonar use. A more complete of the same soil layer that exists along the failure
surfaces buried deep with the slide mass itself. These
discussion of its use in identifying slope failures is given in surface outcrops are sampled by a gravity corer.
Reference 11-15. Side-scan sonar records can be very detailed, Samples from these cores are tested at the state of
stress estimated for the inaccessible material along the
and minor surface features that would be missed by most other failure surface. These test results are then assumed to
systems can be identified with this system. This makes it a apply also to the in accessible failure surface.

valuable tool for assessing other failure evidence, such as small- These methods are used in the following example,
scale surface features, indicative of instability problems. Side- illustrated by Figure 11.6-2, to evaluate a slide whose failure
scan sonar can be used effectively to augment records obtained surfaces are too deep for sampling. Figure 11.6-2(a) is
by other high-resolutino seismic sensors that have detailed the developed from a record of continental slope margin slide area.
subsurface stratigraphy. Because the depth of core penetration is much less than the
thickness of the slide mass, samples along the slide failure
11.6.4 Sampling of Sediments
surface is not possible except at the scarp exposed above the
Obtaining a sample of sediment from the sea bottom is a slide. A set of five relatively short cores is used to define the
prerequisite for quantitative evaluation of seafloor stability. engineering properties for analysis of this slide mass. The core
Chapter 2 describes shallow and deep coring equipment required locations are selected to sample the following parts of the
to recover soil samples. Because the cost of obtaining good feature illustrated in Figure 11.6-2:
quality, continuous samples by conventional offshore drilling is
high, much slope stability assessment work is done using a A. The surface sediment at the same water depth as the
central part of the slide, but offset to one side and out
minimum number of samples obtained from less expensive of the slide zone. This is assumed to be undisturbed
gravity-piston corers. sediment of the same type that existed in the slide
mass at that elevation before the slide.
In weak sediments, gravity-piston corers can occasionally
obtain samples as deep as 30 to 45 feet in the seafloor, B. The surface sediment on the slide scarp, above the slide
mass and assumed to be a sliding surface.
although in many types of sediment the corer penetration is
significantly less. Most submarine slides extend to much C. The upper part of the slide mass.

greater depths than this. Therefore, the gravity corers cannot


provide specimens from much of the slide mass, and large parts
of the failure surface are inaccessible. However, even

11-11
Figure 11.6-2. Examples of a shallow coring plan used in stability analysis.

D. The lower part of the intact slide mass. offshore slope stability and slump hazard potential are presented
as follows:
E. The debris fan that flowed from the slide.

(1) A Level I Survey (described in Table 11.7-1) is based


Tests run on these five cores provide insight into the
on use of remotely obtained data only. No sampling or
variability of the sediment type at these locations and identify testing of sediments is done. The basis of the
investigation is definition of previous sliding and
the effects of disturbance on soil physical properties. These data
existing morphology associated with sediment
provided a basis for further testing to identify engineering instability by pattern recognition in survey data. The
Level I Survey may be appropriate for cause-ways,
properties at the failure surface and allow more accurate
anchors, and pipelines, or short-lifetime installations
identification of soil failure resistance along the slide surface. where the time and cost of a more detailed survey is
not justified by the potential cost of a failure.
In a majority of slides, shallow cores obtained by gravity-
piston corers can provide useful information for submarine (2) A Level II Survey (described in Table 11.7-2) includes
slope stability analysis. However, some submarine sediment use of remotely obtained data plus gravity-piston
coring. Shipboard and shipside laboratory measurement
deposits are too complex to be evaluated adequately by such of sediment properties from core samples provides
simplified and relatively inexpensive methods. In these cases, values for quantitative analysis of bottom stability.
The Level II Survey is intended for siting of more
the risk associated with large uncertainties (which cannot be important facilities, such as those where its failure
better evaluated) rises sharply without more extensive analysis. threatens life, loss of very expensive equipment, or
loss of a critical capability when needed.

11.7 EVALUATION PROCEDURES


Figure 11.7-1 illustrates key features used in
Two levels of investigation procedures for conducting a interpretation of seismic data for these investigations.
site investigation to evaluate

11-12
Table 11.7-1. Level I Slope Stability Survey

11-13
Table 11.7-2. Level II Slope Stability Survey

11-14
Figure 11.7-1. Features on seismic profiles indicative of bottom instability problems.

11.8 REFERENCES 11-5. A. Ang and W. Tank. Probability concepts in engineering


planning and design. New York, N.Y., John Wiley and Sons,
11-1. R.L. Schuster (ed. ). Landslides: Analysis and control,
1975.
National Academy of Sciences/Engineering, Transportation
Research Board. 1978.
11-6. J.R. Benjamin and C.A. Cornell. Probability, statistics,
and decision for civil engineering. New York, N.Y., McGraw-
11.2 N.R. Morgenstern. Submarine slumping and the
Hill Book Co., 1970.
initiation of turbidity currents, in Marine Geotechnique, A. F
Richards (ed. ). Urbana, Ill., University of Illinois Press, 1967,
11-7. D.A. Sangrey. "Marine geotechnology-state of the art,"
pp 189-220.
Marine Geotechnology, vol 2. 1976, pp 45-80.

11-3 D.J. Varnes. Landslides - Chapter 2: Slope movement


11-8. D.A. Sangrey, E.C. Clukey, and B.F. Molnia.
types and processes, Transportation Research Board, Special
"Geotechnical engineering analysis of underconsolidated
Report No. 176. 1978, pp 11-33.
sediments from Alaska," Eleventh Offshore Technology
Conference, Houston, Tex., vol 1, 1979, pp 677-682.
11-4. G.V. Middleton and A.H. Bouma, Turbidity and deep
water sedimentation, Society of Economic Paleontologists and
11-9. D.A. Sangrey et al. "Cyclic loading of sands, silts, and
Mineralogists. Los Angeles, Calif., 1973.
clays," American Society of Civil Engineers Specialty
Conference on Earth-

11-15
quake Engineering and Soil Dynamics, vol 2, 1978, pp 836- 11-15. D.B. Prior and J.M. Coleman. "Disintegrating
851. retrogressive landslides on very-low-angle subaqueous slopes,"
Mississippi Delta, Marine Geotechnology, vol 3, no. 1, 1978,
11-10. H.B. Seed. "Evaluation of soil liquefaction effects on
pp 37-60.
level ground during earthquakes," Journal of the Geotechnical
Engineering Division, American Society of Civil Engineers, 11-16. N. R. Morgenstern and V.E. Price. "The analysis of the
vol 105, GT2, 1979, p. 201. stability of general slip surfaces," Geotechnique, vol 15, no. 1,
1965, pp 79-93.
11-11. B.C. Heezen and M. Ewing. "Turbidity currents and
submarine slumps and the Grand Banks earthquake," American 11-17. D.J. Henkel. "The role of waves in causing submarine
Journal Science, vol 250, 1952, pp 849-873. landslides," Geotechnique, vol 20, 1970, pp 75-80.

11-12. Environmental information on hurricanes, deep water 11-18. M.A. Hampton et al. "Quantitative study of slope
technology, and Mississippi Delta mudslides in the Gulf of instability in the Gulf of Alaska," Tenth Offshore Technology
Mexico, Bureau of Land Management, U S. Department of the Conference, vol 4, 1978, p 2307.
Interior, BLM Open File Report 80-02. New Orleans, La.,
1980.
11.9 SYMBOLS
11-13. P.M. Tucker and J.J. Yorston. Pitfalls in seismic
_
interpretation, Society of Exploration Geophysicists, c Effective soil cohesion [F/L2]
_
Monograph Series, vol 2. 1973.
Effective soil friction angle [deg]

11-14. E.L. Hamilton. "Prediction of in situ acoustic and


elastic properties of marine sediments," Geophysics, vol 36,
1971, pp 266-284.

11-16

Vous aimerez peut-être aussi