Вы находитесь на странице: 1из 560

Geomechanics

IN SOIL, ROCK, AND


ENVIRONMENTAL ENGINEERING
This page intentionally left blank
Geomechanics
IN SOIL, ROCK, AND
ENVIRONMENTAL ENGINEERING

JOHN C. SMALL
The University of Sydney, New South Wales, Australia

Boca Raton London New York

CRC Press is an imprint of the


Taylor & Francis Group, an informa business
CRC Press
Taylor & Francis Group
6000 Broken Sound Parkway NW, Suite 300
Boca Raton, FL 33487-2742
© 2016 by Taylor & Francis Group, LLC
CRC Press is an imprint of Taylor & Francis Group, an Informa business

No claim to original U.S. Government works


Version Date: 20160120

International Standard Book Number-13: 978-1-4987-3930-6 (eBook - PDF)

This book contains information obtained from authentic and highly regarded sources. Reasonable efforts have been
made to publish reliable data and information, but the author and publisher cannot assume responsibility for the valid-
ity of all materials or the consequences of their use. The authors and publishers have attempted to trace the copyright
holders of all material reproduced in this publication and apologize to copyright holders if permission to publish in this
form has not been obtained. If any copyright material has not been acknowledged please write and let us know so we may
rectify in any future reprint.

Except as permitted under U.S. Copyright Law, no part of this book may be reprinted, reproduced, transmitted, or uti-
lized in any form by any electronic, mechanical, or other means, now known or hereafter invented, including photocopy-
ing, microfilming, and recording, or in any information storage or retrieval system, without written permission from the
publishers.

For permission to photocopy or use material electronically from this work, please access www.copyright.com (http://
www.copyright.com/) or contact the Copyright Clearance Center, Inc. (CCC), 222 Rosewood Drive, Danvers, MA 01923,
978-750-8400. CCC is a not-for-profit organization that provides licenses and registration for a variety of users. For
organizations that have been granted a photocopy license by the CCC, a separate system of payment has been arranged.

Trademark Notice: Product or corporate names may be trademarks or registered trademarks, and are used only for
identification and explanation without intent to infringe.
Visit the Taylor & Francis Web site at
http://www.taylorandfrancis.com
and the CRC Press Web site at
http://www.crcpress.com
Contents

Preface xix
Acknowledgments xxi

1 Basic concepts 1
1.1 Introduction 1
1.2 Basic definitions  1
1.2.1 Submerged unit weight  3
1.3 Soil tests  4
1.3.1 Triaxial tests  4
1.3.1.1 Unconfined compression test  4
1.3.1.2 Unconsolidated undrained test  5
1.3.1.3 Consolidated undrained test with
pore pressure measurement  6
1.3.1.4 Consolidated drained test  6
1.3.1.5 Alternative failure plots  7
1.4 Direct shear tests  8
1.5 Consolidation tests  9
1.6 Permeability 12

2 Finite layer methods 15


2.1 Introduction 15
2.1.1 General concepts  15
2.2 Approximation of Fourier coefficients  19
2.3 Formulation 21
2.4 Solution procedure  22
2.5 Three-dimensional problems  23
2.6 Consolidation problems  25
2.7 Fourier transforms  26
2.8 Examples 28

3 Finite element methods 31


3.1 Introduction 31
3.2 Types of elements  31

v
vi Contents

3.2.1 Finding shape functions  32


3.2.2 Isoparametric elements  34
3.2.3 Infinite elements  35
3.2.4 Finite element meshes  36
3.3 Steady state seepage  37
3.3.1 Governing equation  37
3.3.2 Finite element formulation  38
3.3.3 Approximation of total head h 38
3.3.4 Finite element equations  39
3.3.5 Calculation of flows  39
3.3.6 Flow lines  40
3.3.7 Calculation of flow using the stream function  41
3.3.8 Determining the stream function  41
3.3.9 Pumping or extracting fluid  42
3.4 Stress analysis  42
3.5 Consolidation analysis  45
3.5.1 Effective stress analysis  45
3.5.2 Volume balance  47
3.6 Numerical integration  50
3.7 Elastic–perfectly plastic models  52
3.7.1 Formulation 52
3.7.2 Examples for a specific failure surface: Mohr–Coulomb  54
3.8 Work hardening models  55
3.9 Effective stress analysis using Cam Clay type models  56
3.9.1 Normally consolidated clay  58
3.9.2 Overconsolidated clay  59
3.10 Cam Clay type models  60
3.10.1 Cam Clay yield surface  60
3.11 Undrained analysis  63
3.12 Finite element analysis  65
3.12.1 Examples 67
Appendix 3A: Shape and mapping functions for various element types  68
Appendix 3B: Global matrix assembly and boundary conditions  73
Appendix 3C: Boundary conditions  74

4 Site investigation and in situ testing 77


4.1 Introduction 77
4.2 Exploration methods  77
4.3 Site investigation  78
4.4 Object of site investigation  78
4.5 Category of investigation  78
4.6 Planning an investigation  79
4.7 Preparing cost estimates for the work  80
4.8 Detailed exploration  80
4.9 Presentation of information (logs)  80
Contents vii

4.10 Excavation or drilling methods  81


4.10.1 Test pits  81
4.10.2 Excavations 81
4.10.3 Drilling 81
4.10.3.1 Hand augers  81
4.10.3.2 Wash borings  83
4.10.3.3 Rotary drilling  83
4.10.3.4 Auger boring  83
4.11 Sampling methods  86
4.11.1 Thin-walled sampler (or Shelby tube)  87
4.11.2 Split spoon sampler (SPT sampler)  88
4.11.3 Piston sampler  88
4.11.4 Air injection sampler  88
4.11.5 Swedish foil sampler  88
4.12 Rock coring  89
4.13 Field tests  90
4.14 Vane shear test  91
4.15 Standard penetration test  91
4.15.1 Equipment 92
4.15.2 Sampler 92
4.15.3 Drive hammer  93
4.15.4 Rods 93
4.15.5 Test procedure  94
4.15.6 Properties of sands  95
4.15.7 Properties of clays  96
4.15.8 Liquefaction 98
4.16 Pressuremeters 101
4.16.1 Types of pressuremeters  101
4.16.2 Interpreting test results  102
4.17 Dilatometers 103
4.17.1 Type of soil  105
4.17.2 Shear strength of clays  105
4.17.3 Other quantities  105
4.18 Cone penetrometers  106
4.18.1 Equipment 106
4.18.2 CPT equipment  107
4.18.3 CPTu equipment  107
4.18.4 Pushing equipment  109
4.18.5 Calibration 109
4.19 Interpretation of cone data  111
4.19.1 Soil classification  111
4.19.2 Relative density of sands  113
4.19.3 Friction angle of sands  114
4.19.4 Constrained modulus of sands  116
4.19.5 Young’s modulus of sands  116
4.19.6 Undrained shear strength of clays  116
viii Contents

4.19.7 Undrained modulus of clays  117


4.19.8 Permeability 118
4.20 Liquefaction potential  119
4.21 Geophysical methods  121
4.21.1 Seismic surveys  122
4.21.2 Reflection surveys  123
4.21.3 Seismic refraction  124
4.21.4 Rippability of rock  125
4.22 Resistivity 125
4.22.1 Electrical sounding method  128
4.22.2 Push-in resistivity instruments  129
4.23 Magnetic surveying  130
4.24 Ground probing radar  130
4.25 Seismic borehole techniques  130
4.25.1 Down-hole seismic testing  130
4.26 Cross-hole techniques  131
4.27 Other seismic devices  132

5 Shallow foundations 135


5.1 Introduction 135
5.2 Types of shallow foundations  135
5.2.1 Strip footings  135
5.2.2 Pad footings  135
5.2.3 Combined footings  136
5.2.4 Raft or mat foundations  136
5.3 Bearing capacity  136
5.3.1 Uniform soils  138
5.3.1.1 General formulae  143
5.3.1.2 Soil layers of finite depth  147
5.3.2 Non-uniform soils  148
5.3.2.1 Strength increasing with depth  148
5.3.2.2 Fissured clays  151
5.3.2.3 Footings on slopes  151
5.3.2.4 Layered soils  154
5.3.2.5 Working platforms  157
5.4 Numerical analysis  161
5.5 Settlement 162
5.5.1 Limits of settlement  163
5.5.2 Settlement computation  164
5.5.3 Theory of elasticity  164
5.5.4 Rate of settlement  166
5.5.5 Settlement of footings on sand  169
5.5.6 Methods based on settlement and bearing criteria  172
5.6 Numerical approaches  174
5.6.1 Layered soil: Finite layer approaches  174
Contents ix

5.6.2 Finite element methods  175


5.6.3 Estimation of soil parameters  178
5.7 Raft foundations  179
5.7.1 Strip rafts  180
5.7.2 Circular rafts  180
5.7.3 Rectangular rafts  181
5.7.4 Raft foundations of general shape  183
5.8 Reactive soils  184
5.8.1 Pad or strip footings  187
5.8.2 Rafts on reactive soils  189
5.9 Cold climates  189
Appendix 5A  190

6 Deep foundations 191


6.1 Introduction 191
6.2 Types of piles  191
6.2.1 Driven piles  192
6.2.2 Driven and precast piles  192
6.2.3 Jacked piles  192
6.2.4 Bored piles  193
6.2.5 Composite piles  193
6.2.6 Grout injected piles  195
6.3 Installation 195
6.3.1 Types of displacement piles  195
6.3.2 Small displacement piles  195
6.4 Pile driving equipment  195
6.4.1 Piling rigs  196
6.4.2 Piling winches  196
6.4.3 Piling hammers  196
6.4.4 Helmet, driving cap, dolly, and packing  199
6.5 Problems with driven piles  199
6.5.1 Problems from soil displacement  199
6.6 Non-displacement piles  200
6.6.1 Precautions in construction and inspection of bored piles  200
6.6.2 Continuous flight auger piles (or grout injected piles)  201
6.7 Design considerations  201
6.8 Selection of pile type  202
6.9 Designs of piles  202
6.10 Single piles  202
6.10.1 Piles in clay  203
6.10.2 Piles in sand  204
6.10.3 Lambda method  206
6.11 Methods based on field tests  207
6.11.1 Correlations with standard penetration test (SPT) data  207
x Contents

6.11.2 Correlations with cone data  208


6.11.3 Seismic data  209
6.12 Pile groups  210
6.12.1 Piles in clay  211
6.12.2 Piles in sand  212
6.13 Piles in rock  213
6.14 Settlement of single piles  213
6.14.1 Closed form solutions  214
6.14.2 Settlement of single piles  214
6.14.3 Soil modulus increasing with depth  215
6.15 Interaction of piles  216
6.15.1 Use of interaction factors for pile groups  220
6.15.2 Simplified method for pile groups  222
6.16 Assessment of parameters  223
6.17 Lateral resistance of piles  225
6.17.1 Single piles  226
6.17.2 Piles in clay  226
6.17.3 Piles in sand  226
6.18 Laterally loaded pile groups  227
6.19 Displacement of laterally loaded piles  230
6.19.1 Linear elastic solutions (single piles)  230
6.19.2 Constant soil modulus with depth  230
6.19.3 Soil modulus linearly increasing with depth  233
6.19.4 Non-linearity 234
6.20 Deflection of pile groups  236
6.20.1 Interaction methods  237
6.21 Estimation of soil properties  239
6.22 Load testing of piles  240
6.23 Pile load tests  242
6.23.1 Static load tests  242
6.23.2 Types of static load tests  244
6.23.3 O-cell tests  245
6.23.4 Lateral load testing  245
6.23.5 Measurement of deflection  246
6.24 Dynamic pile testing  247
6.24.1 Dynamic pile test  247
6.24.2 Statnamic testing  248
6.25 Pile integrity tests  249
6.25.1 Cross-hole sonic logging  249
6.25.2 Sonic integrity test  249
6.25.3 Gamma logging  251
6.26 Capabilities of pile test procedures  251
6.27 Number of piles tested  252
6.28 Test interpretation  253
6.28.1 Ultimate load capacity  254
6.28.2 Pile stiffness  254
Contents xi

6.28.3 Acceptance criteria  254


6.28.4 Other quantities  255
6.29 Monitoring of piled foundations  256
6.30 Measurement techniques  256
6.30.1 Deflection 256
6.30.2 Pressure cells  257
6.30.3 Strain gauges  258
6.30.4 Piezometers 258
6.30.5 Extensometers and inclinometers  258
6.30.6 Frequency of measurements  258
6.31 Comparison with predicted performance  259
6.31.1 Emirates twin towers, Dubai  259
6.32 Interpretation and portrayal of measurements  264
6.33 Piled rafts  264
6.34 Uses of piled rafts  265
6.35 Design considerations  265
6.35.1 Design process  266
6.36 Bearing capacity of piled rafts  266
6.37 Analysis of piled raft foundations  268
6.37.1 Numerical modelling  268
6.37.2 Finite layer techniques  269
6.37.3 Non-linear behaviour  271
6.38 Example of the finite layer method  272
6.39 Applications 273
6.39.1 Westend Strasse 1 tower  274
6.40 Structural stiffness  277

7 Slope stability 281


7.1 Introduction 281
7.2 Slip circle analysis  282
7.2.1 The method of slices  283
7.2.2 The Swedish, Fellenius, USBR, or Common Method  285
7.2.3 Bishop’s method and simplified method  285
7.2.4 Spencer’s method  285
7.2.5 Finding the critical circle  286
7.2.6 Water pressures  286
7.2.7 Surface loads  288
7.2.8 Computer programs  288
7.2.9 Three-dimensional failure surfaces  288
7.3 Non-circular failure surfaces  289
7.3.1 Morgenstern–Price method  289
7.3.2 Janbu’s method  289
7.3.3 Sarma’s method  290
7.4 Wedge analysis  290
7.5 Plasticity theory  292
xii Contents

7.6 Upper- and lower-bound solutions  292


7.7 Finite element and finite difference solutions  293
7.8 Seismic effects  294
7.9 Factors of safety  295
7.10 Slope stabilisation techniques  296
7.10.1 Control of surface water  296
7.10.2 Horizontal drains  297
7.10.3 Stabilising piles  297
7.10.4 Toe fill  297
7.10.5 Retaining structures  297
7.11 Stability charts  298

8 Excavation 303
8.1 Excavation 303
8.2 Types of excavation support  303
8.2.1 Steel ‘H’ piles and lagging  303
8.2.2 Sheet piles  303
8.2.3 Bored pile walls  305
8.2.4 Diaphragm walls  306
8.3 Stability of excavations  306
8.4 Base heave for cuts in clay  309
8.4.1 Shallow excavations (H/B < 1)  309
8.4.2 Deep excavations (H/B > 1)  310
8.4.3 Excavations of rectangular shape in plans  310
8.4.4 Base failure in sands  311
8.5 Ground settlement caused by excavation  312
8.5.1 Effect of shape of excavation  313
8.6 Forces on braced excavations  314
8.7 Stability of slurry-filled trenches  316
8.7.1 Wedge analysis  317
8.7.2 Purely cohesive soil  317
8.7.3 Cohesionless soil  320
8.8 Numerical analysis  321
8.8.1 Finite element analysis  321
8.8.1.1 Non-linear analysis  323
8.8.2 Finite difference approach  323
8.9 Excavation including groundwater  324
8.9.1 Example excavation problem (no drawdown)  327
8.9.2 Excavation involving drawdown of the water surface  328
8.10 Soil models  329

9 Retaining structures 333


9.1 Introduction 333
9.2 Earth pressure calculation  333
9.2.1 Rankine’s theory  333
Contents xiii

9.2.1.1 Inclined backfill  337


9.2.2 Coulomb’s theory  337
9.2.2.1 Active case  337
9.2.2.2 Passive case  338
9.2.2.3 Surface loads  339
9.2.2.4 Uniform materials  339
9.2.2.5 Active earthquake forces  340
9.2.3 Log spirals  341
9.2.3.1 Passive earthquake forces  341
9.2.4 Upper- and lower-bound solutions  343
9.3 Effect of water  344
9.4 Surface loads  346
9.4.1 Compaction stresses  346
9.5 Sheet pile walls  348
9.6 Anchored walls  350
9.6.1 Anchors 350
9.7 Reinforced earth  352
9.7.1 Sliding 355
9.7.2 Bearing failure  355
9.7.3 Rupture of the reinforcement  356
9.7.4 Pull-out of the reinforcement  357
9.7.5 Overall slip failure  358
9.7.6 Excessive deformation  358
9.8 Computer methods  358
9.8.1 Limit equilibrium methods  359
9.8.2 Finite element methods  361

10 Soil improvement 363


10.1 Introduction 363
10.2 Soft soils  363
10.3 Surcharging and wick drains  364
10.3.1 Surcharging 364
10.3.2 Field observations  368
10.3.3 Sand or prefabricated vertical (wick) drains  370
10.3.4 Vacuum consolidation  375
10.4 Vibroflotation 375
10.5 Vibro-replacement 376
10.5.1 Bearing capacity analysis  377
10.5.1.1 Bearing capacity of single columns  377
10.5.1.2 Bearing capacity of column groups  380
10.5.2 Settlement analysis of column groups  383
10.5.2.1 Flexible foundation  383
10.5.2.2 Rigid foundation  384
10.6 Column-supported embankments  384
10.6.1 Collin beam method  388
10.6.2 BS 8006 method  390
xiv Contents

10.7 Controlled modulus columns  391


10.8 Dynamic compaction  392
10.8.1 Impact rollers  394
10.9 Deep soil mixing  396
10.10  Jet grouting  397
10.11 Grouting  398
10.12 Other methods  400
10.12.1 Ground freezing  400
10.12.2 Electro-osmotic or electro-kinematic stabilisation  401
10.13 Numerical analysis  401
10.13.1 Three-dimensional analysis  402
10.13.2 Equivalent two-dimensional analysis  402

11 Environmental geomechanics 405


11.1 Introduction 405
11.2 Landfills 405
11.2.1 Liners 405
11.2.2 Covers for landfills  407
11.3 Compacted clay liners  409
11.3.1 Compaction of clay  409
11.3.2 Compaction method  411
11.3.3 Compaction control  412
11.3.4 Permeability of clay  413
11.3.5 Measuring permeability of CCLs  414
11.3.5.1 Ring infiltrometer  414
11.3.5.2 Borehole test  415
11.3.5.3 Lysimeters 416
11.3.5.4 Porous probes  416
11.4 Flexible membrane liners  417
11.4.1 Types of geomembranes  418
11.4.1.1 High-density polyethylene  418
11.4.1.2 Very low density polyethylene  418
11.4.1.3 Polyvinyl chloride  419
11.4.1.4 Chlorosulfonated polyethylene  419
11.4.2 Placing geomembranes  419
11.4.3 Seaming 419
11.5 Geosynthetic clay liners  419
11.5.1 Types of GCLs  420
11.5.2 Manufacturing 420
11.5.3 Placement 420
11.5.4 Examples of use  421
11.6 Stability of liners  422
11.6.1 Tension in the membrane  422
11.6.2 Factor of safety  424
11.7 Processes controlling pollutant transfer  425
11.7.1 Advective transport  425
Contents xv

11.7.2 Diffusive transport  426


11.7.3 Dispersive transport  426
11.7.4 Sorption 427
11.7.5 One-dimensional transport  427
11.7.5.1 Ogata–Banks solution  428
11.7.5.2 Booker–Rowe solution  429
11.8 Finite layer solutions  432
11.8.1 Three-dimensional solutions  432
11.8.2 Boundary conditions  434
11.8.2.1 Boundary condition at the base  435
11.8.2.2 Boundary condition at the surface  435
11.8.3 Assembly of finite layer matrices  437
11.8.4 Inversion of transforms  437
11.8.5 Solutions for a three-dimensional problem  437
11.9 Remediation 438
11.9.1 In situ leaching and washing/flushing  439
11.9.2 In situ chemical treatment  440
11.9.2.1 Oxidation 440
11.9.2.2 Chemical reduction  440
11.9.2.3 Polymerisation 440
11.9.3 In situ biological treatment  440
11.9.3.1 Microbial treatment  440
11.9.4 Soil venting  441
11.9.5 Thermal desorption  442
11.9.6 In situ stabilisation/solidification 442
11.9.7 Electro-remediation 443
11.9.8 In situ vitrification  444
11.10 Mining waste  444
11.10.1 Properties of tailings  445
11.10.2 Tailings dam construction  445
11.10.2.1 Upstream method  445
11.10.2.2 Spigotting 445
11.10.2.3 Cycloning 447
11.10.3 Centreline method of construction  448
11.10.4 Downstream method  449
11.10.5 Embankments built entirely of borrowed materials  450
11.10.6 Tailings storages  450
11.10.7 Control of water  451
11.10.7.1 Inflows 452
11.10.7.2 Outflows 452
11.10.8 Stability of embankments  453
11.10.9 Piping 453
11.10.9.1 Filters 454
11.10.10 Cut-offs and barriers  455
11.10.10.1 Controlled placement of tailings  456
11.10.10.2 Clay liners  456
xvi Contents

11.10.10.3 Cut-off trenches  457


11.10.10.4 Slurry cut-off walls  457
11.10.10.5 Grout diaphragm walls  459
11.10.10.6 Grouting 460
11.10.11 Synthetic liners  461
11.10.12 Seepage return systems  462
11.10.12.1 Collector ditches  462
11.10.12.2 Collector wells  462
11.10.12.3 Collection and dilution dams  463
11.10.13 Acid rock drainage  463
11.10.13.1 Factors affecting acid generation  464
11.10.13.2 Control of acid generation  465
11.10.13.3 Control of acid migration  465
11.10.14 Remediation 466
11.10.14.1 Hydrological stability  466
11.10.14.2 Long-term erosion stability  468
11.10.14.3 Vegetation 468
11.10.14.4 Riprap 468
11.10.14.5 Prevention of environmental contamination  468

12 Basic rock mechanics 471


12.1 Introduction 471
12.2 Engineering properties of rocks  471
12.2.1 Point load strength index  471
12.2.2 Unconfined compression test  473
12.2.3 Modulus of rock from unconfined compression test  473
12.2.4 Confined compressive strength  473
12.2.5 Sonic velocity test  474
12.3 Failure criterion for rock  475
12.3.1 Hoek–Brown criterion for intact rock  476
12.3.2 Hoek–Brown criterion for rock mass  476
12.4 Classification of rocks and rock masses  478
12.4.1 Classification on strength  478
12.4.2 Classification by jointing  478
12.4.3 Rock quality designation  479
12.4.4 Classification of individual parameters used
in the NGI tunnelling quality index  480
12.4.5 Rock mass rating method  480
12.5 Planes of weakness  481
12.5.1 Stereographic projections  481
12.5.2 Roughness of joints  483
12.6 Underground excavation  484
12.6.1 Support systems  484
12.6.2 Design process  485
12.6.3 In situ stresses 485
Contents xvii

12.6.4 Stresses around underground openings  487


12.6.4.1 Circular tunnel  487
12.6.4.2 Elliptical tunnel  488
12.6.5 Support design  488
12.6.5.1 Q index method  490
12.6.5.2 RMR method  491
12.6.6 Support types  491
12.6.7 Rock bolts and shotcrete  493
12.7 Rock slopes  493
12.7.1 Planar sliding  494
12.7.2 Wedge failure  496
12.8 Foundations on rock  496
12.8.1 Surface foundations  497
12.8.2 Shafts in rock  497
12.8.2.1 Base resistance  498
12.8.2.2 Shaft resistance  498
12.8.2.3 Lateral capacity  500
12.8.2.4 Uplift capacity  501
12.8.2.5 Piles on sandstones and shales of the Sydney region  502
12.8.3 Deformation of foundations on rock  502
12.8.3.1 Vertical deformation  502
12.8.3.2 Lateral deformation  503
12.9 Vibration through rock  503
12.10 Numerical methods  504
Appendix 12A  506
Appendix 12B  510
Appendix 12C  512
Appendix 12D  515

References 517
Index 537
This page intentionally left blank
Preface

Since the arrival of the twenty-first century, analysis and design of geotechnical facilities
such as foundations, excavations, tunnels, and slopes has advanced rapidly, mainly due to
the availability of computer-based techniques and the ease by which computer software can
be purchased over the Internet.
The use of software makes complex analysis very simple and designs can be improved and
refined through its application. However, there is still a need to look carefully at the results
of any analysis and decide if the results are correct. Careful checking of input parameters
and the output is essential in all design work, and the introduction of advanced graphical
input and output into computer codes has made this process easier and has had the effect of
reducing errors.
However, there is still a strong need for engineering judgment and this still has to be
applied even though the tools for analysis are becoming more and more powerful. Therefore,
in this book, although there are many references to advanced software, and for Internet
locations for obtaining the software, there are many hand-based calculation methods pro-
vided. These enable simple and swift checking of more advanced analysis. If the problem can
be simplified and checked by hand, then this is a useful way of deciding whether to accept
the results of numerical computations.
As many numerical techniques are commonplace today, an introduction is given to finite
element and finite layer techniques in the initial chapters. Then, the chapters that follow deal
with the application of geomechanics to soil and rock mechanics; the use of these numeri-
cal techniques is mentioned where they are applicable. For example, finite layer techniques
can be applied to settlement of foundations, consolidation problems, piled raft analysis, and
pollution migration problems. Finite element techniques are applicable to most geotechni-
cal problems; stress analysis, consolidation, seepage, soil–structure interaction, and rock
mechanics are some applications.
New techniques that involve the use of optimisation are also mentioned where applicable.
These techniques involve finding solutions to stress or velocity fields at discrete points that
minimise a quantity such as the applied load or the power dissipation. These methods can
be used in a whole range of geotechnical problems, and for problems involving collapse,
provide upper and lower bounds to the collapse load. The real solution must lie between
these bounds.
Many classical problems in soil mechanics have only partially been solved, and in recent
years new solutions have become available that either show that the past solutions were
of reasonable accuracy and may still be applicable, or show that these methods should no
longer be used. An example is the Coulomb wedge method for obtaining passive earth pres-
sures. These solutions have been shown to be grossly in error especially for large friction
angles, and so there is no longer any need to use them, as more accurate modern values are
available.

xix
xx Preface

Today, the environment is becoming increasingly important and so one chapter is devoted
to environmental considerations. Municipal and industrial waste is an increasing problem
for large cities, and sites for landfills are rapidly being used up. The problem of ­isolating the
waste so that it does not produce further pollution will become a major issue in the future
and is one that needs to be addressed today.
Mining waste will also become a future problem as it has to be ­disposed of safely. The
practice of dumping waste into rivers and the sea with no treatment or ­containment is no
longer acceptable today, and so this aspect of geomechanics is addressed in Chapter 11.
Rock mechanics is a logical extension to soil mechanics, and often it is overlooked in
texts on geomechanics. Here, an introduction is provided to rock mechanics that may assist
students and professionals who wish to obtain some basic understanding of the design of
­foundations and excavations in rock. More detailed texts are available, and where possible
the reader is encouraged to obtain the papers and textbooks referred to in the chapter on
rock mechanics.
The material contained in this book is information that the author has found useful in his
time as a university professor and as a consulting engineer, and it is hoped that it will serve
as a source of reference for the reader as well.

John C. Small, BSc(Eng) Hons I, PhD, FIEAust, MASCE


Acknowledgments

I wish to express my thanks to my teachers from whom I have learned so much: Professor
Bob Gibson of King’s College London, Professor John Booker, Professor Ted Davis,
Professor Harry Poulos, and Dr. Peter Brown at the University of Sydney.
I also extend my thanks to all of my research students who over the years have gen-
erated much new material, some of which is presented in this book. Finally, I gratefully
­acknowledge the support given to me by my parents throughout my career.

xxi
This page intentionally left blank
Chapter 1

Basic concepts

1.1 INTRODUCTION

In later chapters of this book, various terminologies are used and standard soil mechanics
tests are referred to, therefore, this chapter provides some background on these basic tests
and concepts. More details on testing of soil and rock are provided in the National and
International Standards that are referred to in this and later chapters.

1.2  BASIC DEFINITIONS

Soil is considered to consist of the soil grains and the voids that exist between them. The voids
can contain air and water, and so may be considered fully saturated if the voids are full of
water or partially saturated if air and water fill the voids. Sedimentary rocks also consist of soil
grains, but cementing that is present makes the rock stronger and less deformable than a soil.
Generally, soils subjected to moderate stress levels deform through changes in the void
space in the soil, although calcareous soils will undergo volume change through particle
crushing. At high pressures, soil grains will eventually crush, but in most soil mechanics
applications, the particles are considered non-deforming.
The concept of a soil for engineering purposes is therefore as shown in Figure 1.1.
The void ratio e of a soil is defined as the volume of the voids to the volume of the solid
soil particles and is defined as

Vv V + Vw
e = = a (1.1)
Vs Vs
where Va is the volume of air in the voids, Vw is the volume of water in the voids, Vs is the vol-
ume of solids, and Vv is the volume of voids in a given total volume VTot of soil. Sometimes
it is more convenient to work with the porosity n of a soil where the porosity is defined as

Vv Va + Vw
n = = (1.2)
VTot Va + Vw + Vs

A very useful property of a soil is its water or moisture content that is usually given the
symbol w or m and is usually expressed as a percentage. If the weight of water is w w and the
weight of solids is ws, then we can write

ww
m = × 100% (1.3)
ws

1
2  Geomechanics in soil, rock, and environmental engineering

wA (= 0)

wW (=VW γW)

wS (=VS G γW)
VA

VW
Voids { Air
Water

Vs

Figure 1.1  Three-phase soil model.

The degree of saturation S is a measure of how much of the voids of the soil are filled with
water

Vw
S = × 100% (1.4)
Vv

Another important measure of a soil’s properties is its unit weight. The saturated unit
weight γsat is calculated for a soil with voids fully saturated with water, that is, S = 100%.

w s + ww
γ sat = (1.5)
VTot

The dry unit weight γdry of a soil is calculated for no water in the voids

ws
γ dry = (1.6)
VTot

The soil may be in a state of saturation between the dry and totally saturated case
(0 ≤ S ≤ 100%), where the voids contain some air and some water. In this case, the bulk unit
weight γbulk is calculated from

w s + ww
γ bulk = (1.7)
VTot

Unit weights are expressed in kN/m3 or lbf/ft3 as they are expressed in terms of the weight
as a force. A typical unit weight for a soil may be 19 kN/m3 or 120 lbf/ft3. The unit weight
of water is 9.81 kN/m3 as the acceleration due to gravity is 9.81 m/sec2 .
The density of a soil ρ is sometimes used in calculation and the density is defined as the
mass of a soil per unit volume. For example, the bulk density may be calculated from

ms + mw
ρbulk = (1.8)
VTot
Basic concepts  3

The density of a soil is expressed as the mass of soil per unit volume and has units of kg/m3 or
lb/ft3. A typical value might be 2000 kg/m3 or 125 lb/ft3. The density of fresh water is approxi-
mately 1000 kg/m3 or 62.4 lb/ft3 (as it depends on temperature).
With sands and gravels, often the term ‘relative density’ is used rather than the unit
weight. The relative density Id gives a measure of where the void ratio of the soil is relative
to the minimum emin (densest state) and maximum emax (loosest state) void ratios, that is,

emax − e
Id = (1.9)
emax − emin

It may be seen from the definition, that Id must range between 0 (loose) and 1 (dense).
The specific gravity G of the soil grains is defined as the ratio of the density of the soil
grains ρs to the density of water ρw, that is,

ρs γ
G = = s (1.10)
ρw γw

1.2.1  Submerged unit weight


The concept of submerged unit weight comes from the calculation of stresses in the ground.
According to the effective stress law that was originally proposed by Terzaghi (1923), the
total stress in the ground is equal to the sum of the stress in the soil grains or ‘skeleton’
(called the effective stress) plus the water pressure. This can be expressed as

σ = σ′ + u (1.11)

where σ is the total stress in the soil, σ′ is the effective stress, and u is the water pressure.
This equation is one of the most important in soil mechanics as it explains how soils
behave in both drained and undrained conditions. It governs the strength and deformation
of soils under all drainage conditions as it is the effective stresses in the soil that govern soil
behaviour, and as can be seen by Equation 1.11, the effective stresses depend on the water
pressures that are acting.
If we have a totally saturated layer of uniform soil with a static water table at the surface,
the total vertical stress acting at any depth z is given by

σ v = γ sat z (1.12)

The water pressure will be given by γwz at depth z and so we can calculate the vertical
effective stress as

σ ′v = γ sat z − γ w z = γ ′z (1.13)

where γ  ′ or γsub is equal to the difference in the saturated unit weight and the unit weight of
water, that is, γ  ′ = γsat − γw. We can therefore calculate the effective stress directly by using
the submerged unit weight, and this is why it arises. If there is any doubt in calculating an
effective stress, it is recommended that a return to basics is used, and the water pressure
subtracted from the total stress.
4  Geomechanics in soil, rock, and environmental engineering

1.3  SOIL TESTS

In the following chapters, reference is made to soil properties that are obtained from lab-
oratory tests, and so in this section some of the more common laboratory soil tests are
described. Often the soil properties measured are also found from field testing by using cor-
relations, as this is a faster and cheaper way to obtain the soil property. Because the values
are obtained from correlations, they are generally not as reliable as laboratory values. For
example, the undrained shear strength of a soil may be obtained from triaxial tests, or from
shear vane tests, or static cone tests in the field (see Chapter 4).

1.3.1  Triaxial tests


One of the most valuable and commonly performed laboratory tests is the triaxial test.
A cylindrical shaped sample of the soil is placed into a cell and covered with a latex mem-
brane that is attached to end caps with rubber rings. The cell is pressurised by water and
therefore applies an all-round pressure to the soil specimen called the cell pressure σ3. The
triaxial cell is shown in Figure 1.2.
A loading ram through the top of the cell allows a load to be applied to the top of the soil
specimen through the top cap. This load can be divided by the area of the specimen to calcu-
late the stress being applied in the axial direction σ1–σ3 called the deviator stress. (Because
there is a cell pressure of σ3 acting down on the top cap as well as the deviator stress from
the ram, the total stress acting on the top cap is therefore σ1.)
Provision is made through porous stones placed at the top and bottom ends of the speci-
men for drainage and pore pressure measurement. By knowing the pore water pressure in
the sample, the effective stresses in the sample can be calculated.
There are several types of tests that can be performed with the triaxial cell, and some of
the more common are examined in the following.

1.3.1.1  Unconfined compression test


As the name suggests, in the unconfined compression test, there is no cell pressure applied
and there is no need to use the outer cell or membrane. The test is only applicable to clays as
a clay sample is self-supporting. The unconfined compressive strength su is given by

P (1.14)
su =
2A
Axial load

Loading ram
Bleed cock

Rubber Top cap


Pressure gauge Porous disc
sealing ring

Cell water Soil sample


Rubber Rubber membrane
sealing ring Porous disc
Drainage and pore pressure
Cell pressure measurement

Figure 1.2 Triaxial cell.


Basic concepts  5

where P is the applied axial load at failure and A is the cross-sectional area of the sample
(corrected for increase in diameter after loading; A = A 0/(1 − ε) where A 0 is the original
cross-sectional area and ε is the axial strain in the specimen). In such a test, the initial stress
state of the sample is not controlled and so the strength obtained is for an unknown effec-
tive stress state. However, the test is quick and easily performed and is therefore commonly
used (ASTM D2166 2013).

1.3.1.2  Unconsolidated undrained test


If three similar specimens of the same clay material are tested under three different cell pres-
sures in an undrained state (i.e. they are loaded in the triaxial cell until failure), then the
Mohr’s circles for each of the samples can be plotted in terms of total stress at failure. The
minor total principal stress is the applied cell pressure σ3 and the undrained strength su at
failure is given by the deviator stress

P σ − σ3
su = = 1 (1.15)
2A 2

If the soil sample is not saturated, then consolidation can occur when the cell pressure is
applied even though the sample is undrained. In this case, the undrained strengths will be
different for samples tested at different cell pressures and the envelope to the Mohr’s circles
at failure is often curved. If the failure surface is approximately linear, it can be represented
in terms of total stresses by

τ = cu + σ n tan φ u (1.16)

where cu is the intercept on the vertical axis of the plot and ϕu is the undrained angle of
shearing resistance as shown in Figure 1.3. In applying this failure criterion in design, it
should be noted that total stresses should be used.
If the soil samples tested are totally saturated, then any increase in cell pressure on the
sample causes an equal rise in pore pressure inside the sample and, for each of the tests,
the undrained strength su will be the same (because the effective stress in the sample is the
same) and so the undrained angle of shearing resistance will be ϕu = 0. The test is described
in ASTM D2850-15 (2015).

ϕu = 0 if soil saturated
ϕu
Undrained failure surface

su
cu

σ3 σ1 σ

Figure 1.3 Undrained strength envelope from an unconsolidated undrained test.


6  Geomechanics in soil, rock, and environmental engineering

1.3.1.3  Consolidated undrained test with pore pressure measurement


In this test, soil samples are allowed to consolidate under each cell pressure applied, and
then tested undrained.
If the pore pressures are measured during each stage of the consolidated undrained test, it
is possible to calculate the effective stresses at each stage of the test. Sometimes, three stages
are performed on the same sample, but ideally three separate samples are used. The test is
described in AS 1289.6.4.2 (1998), ISO/TS 17892-9 (2004), and ASTM D4767-11 (2011).
To obtain the effective stresses by subtraction of the pore pressure, the sample needs to
be saturated, and so a test of saturation is performed by shutting the drainage valves and
increasing the cell pressure. If the soil is perfectly saturated, then the increase in pore pres-
sure should be equal to the increase in cell pressure. However, this is rarely achieved because
of air existing in the pore water.
To overcome this problem, de-aired water is used to saturate the specimen and the pore
water in the sample is pressurised under a back pressure to force air into the solution. For
instance, if the back pressure is 200 kPa and the cell pressure is 300 kPa, then the net
consolidation pressure on the sample is 100 kPa. A back pressure of up to 900 kPa may be
required for this to be effective in some cases.
If the increase in the cell pressure is Δσ3 and the increase in pore pressure measured is Δu,
the pore pressure parameter B can be calculated from

∆u
B= (1.17)
∆σ 3

The value of B should be greater than 0.95 before testing; otherwise, the back pressure
should be increased (AS 1289.6.4.2 1998).
As the effective stress is known in this test, both the drained and undrained
strength envelopes can be plotted. The drained strength envelope as plotted in Figure 1.4
is given by

τ = c′ + σ ′ tan φ′ (1.18)

The deviator stress at failure for each stage is the same for the total stresses and the effec-
tive stresses since

P
= σ1 − σ 3 = (σ1′ + u) − (σ ′3 + u) = (σ1′ − σ ′3) (1.19)
A

1.3.1.4  Consolidated drained test


For this test, the sample is consolidated under the action of a cell pressure and water is
allowed to drain from the sample. Back pressure may be used to keep the soil sample satu-
rated as was explained in Section 1.3.1.3. The test is performed at a slow rate of shearing
such that no excess pore pressure is generated in the soil sample and this requires drainage
of the sample at all times during the test.
The test can be performed on sands and clays, although it can be carried out more quickly
on sands because they have relatively high permeability. Back pressure may be required for
dense sands that tend to dilate and cause a drop in the pore pressure. Tests on clays can last
Basic concepts  7

Mohr–Coulomb failure surface ϕ′

c′
σ′3 σ′1 σ′

Figure 1.4 Mohr–Coulomb failure envelope showing the Mohr’s circle at failure for each of the three stages.

for several days, and so the test is not often used for fine-grained soils. Instead, the consoli-
dated undrained test described in Section 1.3.1.3 is used.
The test is performed for three (or more) different cell pressures, and for each, the deviator
stress at failure is found from

P
σ1 − σ 3 = = σ1′ − σ ′3 (1.20)
A

The effective stress σ ′3 at failure is the difference between the cell pressure and the back
pressure ub and may be found from σ ′3 = σ 3 − ub and so the Mohr’s circles for each stage can
be plotted in terms of the effective stresses. The result is the same as shown in Figure 1.4 for
the failure surface found in terms of effective stresses from the consolidated undrained test
with pore pressure measurement. The test is described more fully in ASTM D7181-11 (2011).

1.3.1.5  Alternative failure plots


An alternative form of plot from a triaxial test is a p − q plot. This form of plot is useful as
the stress path plot goes vertically upward when the soil is elastic in an undrained test, and
when the soil yields, the plot deviates towards the failure line (see Figure 3.28). The slope of
the failure line in this type of plot is given by M where

6 sinφ′
M = (1.21)
3 − sinφ′

and so the drained angle of shearing resistance can be found from the plot.

3M
sin φ′ = (1.22)
6+M

More is provided in Section 3.8 on this type of plot, and further information is given in
the books by Atkinson (2007) and Atkinson and Bransby (1978).
8  Geomechanics in soil, rock, and environmental engineering

1.4  DIRECT SHEAR TESTS

The shear strength available on a surface of shearing can be found directly in a shear box test.
The box consists of two halves (Figure 1.5), and the bottom half is held fixed while the upper
half is pushed, thus shearing the soil. The shear force is applied through a curved yoke so that
the force is not applied eccentrically to the upper box. The shear force on the yoke is measured
with a proving ring or load transducer (see Figure 1.6). The outer box into which the shear box
is placed can be filled with water to keep the sample saturated during slow shearing.
For sands and gravels, the shearing rate is not of great importance as they are normally
tested in a dry state. For clays, the test can be performed in the drained or undrained state.
For drained shearing of a clay, the speed of the machine is set to a very slow shearing rate
so that the excess pore pressures that are generated have time to dissipate. When the normal
load is placed onto the lid of the shear box, the settlement of the lid is monitored to see
when the lid stops settling thus indicating that the pore pressures have dissipated (see ASTM
D3080 or AS 1289.6.2.2 1998).
The residual strength of a clay can be found by reversing the shear box. The soil is firstly
sheared forwards and then backwards until the shear force does not change with further
shearing. This type of test may be required for soils that are liable to undergo large shear
displacements where the shear strength drops to the residual value.
To perform an undrained test on a clay, the normal loads are placed on the hanger and the
soil immediately sheared at a fast rate.
The test has the drawback that the pore pressure in the box cannot be controlled as it can
in a triaxial test and so it is not possible to know what pore pressures are being generated
in clay soils.

Lifting screw
Clamping screw
Lifting
handles
Clearance holes

Upper half of box


Gap
Lower half of box
Shear force
applied to yoke

Figure 1.5 Two halves of a shear box.

Water (if necessary)

Worm drive Lid for vertical load


pushing outer box Loading ring dial
gauge

Upper and lower Spacer block (only


shear box halves if box is reversed)

Figure 1.6 A shear box in a loading frame.


Basic concepts  9

τ
Mohr–Coulomb failure surface
τ3 ϕ′

τ2

τ1

c′
σ′n1 σ′n2 σ′n3 σ′

Figure 1.7 A plot of shear strength versus a normal load for a drained shear box test.

For each different normal stress applied to the lid of the box, the shear stress at failure
can be measured and a plot is made as shown in Figure 1.7 (for a drained test). The same
Mohr–Coulomb failure line is obtained as found in the triaxial test, but from the measure-
ment of the shear stress and normal stress on the plane of shearing.

1.5  CONSOLIDATION TESTS

One-dimensional compression tests on clays are performed in an oedometer. The oedom-


eter has a ring that is used to cut a sample of soil so that it is a tight fit inside the ring
(Figure 1.8). Because the soil cannot expand sideways, the compression only occurs verti-
cally (one-dimensional).
Porous disks are placed at the top and bottom of the soil specimen so that it can drain,
and loads are placed on the sample so as to compress it. The time–settlement behaviour is
recorded so as to obtain the rate of consolidation of the soil sample. As well, the amount of
compression of the sample under the applied load is measured.
The vertical load applied to the sample is doubled at each load increment and a plot is
made of the void ratio of the sample versus the effective stress applied (to a logarithmic scale)
as shown in Figure 1.9. This can be used to compute the settlement of soil layers as outlined
in Section 10.3.1 in Chapter 10.
The rate at which consolidation occurs is recorded at each loading stage, and a plot is
made of settlement against either the square root of time, or the logarithm of time. A log
time plot is shown in Figure 1.10.
Various parameters can be found from the tests that are used in the calculation of the
magnitude of settlement or rate of settlement.
In the log time versus settlement plot, it may be seen that the final part of the plot is a
straight line. This is due to creep of the sample after the primary consolidation has finished

Vertical load

Soil
Porous stones
Cutting ring

Figure 1.8 A section through an oedometer.


10  Geomechanics in soil, rock, and environmental engineering

2.50
2.30
2.10

Void ratio
1.90
1.70

1.50
1.30

1.10
1 10 100 1000 10,000
Pressure (kN/m2)

Figure 1.9 Void ratio versus logarithm of effective pressure plot e − log σ′v for an organic silt.

t1 4t1
7.5 15 s 30 s 1 min 2 4 8 16 min 24 h
t50 10 100 1000 10,000
Vertical deformation dial gauge reading (mm)

2.1
y 0%
y
2.0

1.9

50%
1.8

1.7

1.6 ΔHα
100%

1.5

t2 10t2
Elapsed time t

Figure 1.10 A settlement versus log time plot for an oedometer test.

and excess pore pressures are almost zero. Two straight lines are drawn as shown in Figure
1.10 (one through the steepest part of the curve and one through the final straight line part),
and where they intersect is deemed to be 100% primary consolidation. Two points are then
chosen on the initial curved part of the plot: one at a time t and the other at 4t. Because the
first part of the curve is a straight line on a plot against the square root of time, the settle-
ment from 0 to t is the same as from t to 4t. Using this fact, the location of zero settlement
can be found as shown in Figure 1.10. The 50% consolidation point is then midway between
the 0% and 100% consolidation points.
Theoretically, the 50% degree of consolidation takes place at a time factor Tv equal to
0.197, where

cv t
Tv = (1.23)
H2
Basic concepts  11

and t is time, H is the half-depth of the specimen (because of two-way drainage), and cv is
the coefficient of consolidation.
We can therefore solve for the coefficient of consolidation cv that has units of m2/yr or ft2/yr
as shown in Equation 1.24.

Tv H 2 0.197 H 2 (1.24)
cv = =
t50 t50

In the above equation, t50 is the time for 50% consolidation and H is the average half-depth
of the specimen during consolidation.
The rate of creep can also be calculated from the coefficient of secondary compression C α
defined as

∆H
Cα = (1.25)
H0

where ΔH is the change in thickness of the specimen over one log cycle of time (for the last
linear part of the curve), and H 0 is the initial height of the specimen.
From the void ratio versus log of the applied effective pressure plot (Figure 1.9), it can be
seen that there is a distinct kink in the loading part of the curve. This is the point at which the
stress in the specimen reaches the pre-consolidation pressure (or maximum past pressure).
The slope of the curve before this point is called the recompression index C r and is defined as

e0 − ei  p′ 
Cr = = ∆e log  c  (1.26)
log pc′ − log p1′  p1′ 

The soil becomes more compressible after the pre-consolidation pressure is reached, and
for this part of the plot the compression index C c is calculated as

ei − e2  p′ 
Cc = = ∆e log  2  (1.27)
log p2′ − log pc′  pc′ 

Figure 10.3 shows the definitions of effective pressures p′ and void ratios e used in the
above definitions.
Sometimes, the coefficient of volume compressibility mv is calculated from the oedometer
test where

∆ε v
mv = (1.28)
∆σ ′v

and Δεv is the vertical strain that occurs due to a change in vertical stress ∆σ ′v on the sample.
Hence, mv is like an inverse modulus, and if the soil gets stiffer, the value of mv reduces.
This means that mv has to be calculated for the stress range that it is to be used for as it is
not constant, but varies with stress level. We can also write

0.435Cc
mv = (1.29)
(1 + e0 )∆σ ′v
12  Geomechanics in soil, rock, and environmental engineering

In terms of a conventional elastic modulus of the soil, we can write

(1 + ν′)(1 − 2 ν′)
mv = (1.30)
(1 − ν′)E′

where ν′ is Poisson’s ratio of the soil and E′ is its elastic modulus. As well, we can write cv
in terms of elastic constants,

k k(1 − ν′)E′
cv = = (1.31)
γ w mv γ w (1 + ν′)(1 − 2 ν′)

This expression (1.31) is often used to find the permeability k of a clay as a conventional
permeability test cannot be used for very low permeability soils.
The use of the parameters found from the oedometer test is explained more fully in Section
10.3.1. Details of the test are provided in relevant standards such as AS 1289.6.6.1 (1998),
ISO/TS 17892-5 (2004), and ASTM D2435 (2011).

1.6 PERMEABILITY

For sands, the permeability can be found by direct means (for clays use Equation 1.31) by
allowing water to flow through a sample of the soil at the appropriate void ratio. For soils
with permeabilities of between about 10−2 and 10−5 cm/sec, the falling head permeameter
can be used. The apparatus is shown in Figure 1.11, where water in a tube is allowed to drop
as it flows through the soil. The permeability k is then given by

a h 
k = ln  1  (1.32)
At  h2 

dt
h1

h2

Soil sample

Ds

Figure 1.11 A falling head permeameter.


Basic concepts  13

Water

h Overflow

Soil Filter discs



sample

Figure 1.12 A constant head permeameter.

where a is the cross-sectional area of the piezometer tube (a = πdt2 / 4), A is the cross-­
sectional area of the sand sample (A = πDs2 / 4), ℓ is the length of the soil sample, and t is the
time for the total head difference across the sample to drop from h1 to h2 .
For more permeable soils, a constant head permeameter can be used (ASTM D2434-
68 2006). For this type of permeameter, the head difference across the soil sample is
kept constant through a supply of water as shown in Figure 1.12. The permeability k is
given by

Q
k = (1.33)
thA

where an amount of water Q is collected from the outlet in a time t, ℓ is the sample
length, A is the cross-sectional area of the sample, and h is the head difference across
the sample length. The permeability is dependent on temperature and can be corrected
back to a standard value at 20°C (k 20°C = kTμT /μ20°C), and μT is the viscosity of the soil at
temperature T.
These tests can also be conducted using a flexible wall permeameter (which contains the
soil specimen in a rubber membrane such as a triaxial test sample – see ASTM D5084 2010).
Typical permeabilities of soils are shown in Table 1.1.

Table 1.1  Typical permeabilities of soils


Soil type Permeability range (cm/sec)
Gravels >1
Sands −3
10 to 1
Silts 10−6 to 10−3
Clays <10−6
This page intentionally left blank
Chapter 2

Finite layer methods

2.1 INTRODUCTION

There are many problems in the field of geomechanics, where the soil profile may be ide-
alised as consisting of horizontal layers for which the geometry and material properties do
not vary in one or two spatial directions.
Examples include (1) pavements that are constructed by placing horizontal layers of dif-
ferent pavement materials (see Figure 2.1) and (2) sedimentary soils that are horizontally
layered because sediments are deposited under a vertical gravity field (Figure 2.2). In the
latter case, the layering may be in one direction only if the sediments are laid in an ancient
stream bed or valley (Figure 2.3).

2.1.1  General concepts


In order to convey the general concepts of the method, suppose we consider a problem that
is common in geomechanics, that of a strip loading applied to the surface of a layered soil
as shown in Figure 2.4.
A function such as the loading function shown in Figure 2.4, may be represented by a
Fourier series, where the addition of various periodic sine or cosine functions are added
together to approximate the original function. Here, the function is a step function that is
equal to the applied uniform load q between −a and +a if the load is half-width of a as shown
in Figure 2.5 (Cheung 1976).
Because the loading function chosen is an even function of x, we only need use the cosine
part of the Fourier series and to represent the loading function as a sum of cosine terms.

q(x) = ∑Q
n =0
(n )
cos(α n x)

α n = 2nπ /L
L
2 (2.1)
Q (n )
=
L ∫
q(x)cos(α n x)dx
0
n >0

L
1
Q(n) =
L ∫
q(x)dx n =0
0

We are therefore stating that the load function is periodic with a period L (Figure 2.6).

15
16  Geomechanics in soil, rock, and environmental engineering

Tyre load

Basecourse

Sub-base

Figure 2.1 Wheel loading applied to a pavement.

Structure

Clay
Silt Sand
Peat

Clay

Bedro
ck

Figure 2.2 Soil that has been deposited horizontally by sedimentation.

Surface loading

Void

Material interface

Figure 2.3 Soil layering that is horizontal in one direction only.

Figure 2.4 Strip loading applied to a layered soil.


Finite layer methods  17

q(x) = q

–a +a x

Figure 2.5 Strip-loading function.

L L

Figure 2.6 Periodically spaced loadings.

Provided that we make L large enough and that we use enough terms n in the Fourier
series, we will be able to synthesise an isolated strip loading. Figure 2.7 shows the effect of
adding terms to the series used to approximate the strip loading. By the time 50 terms are
used in the series, the sum of the cosine terms is beginning to take the shape of the strip load-
ing. The plot in Figure 2.7 is for a spacing of the loads (the period) of L/a = 20. If the spacing
is made larger, it takes more terms in the Fourier series to get a good approximation of the
step function, so in applying the method in practice, a judicious choice of spacing needs to
be made so that fewer terms can be used in the solution, and the solution time is faster.
We can now make use of the theory of superposition (provided the problem involves linear
elasticity). Suppose that instead of applying the strip loading, we apply the individual cosine
loadings as shown in Figure 2.8.
For each of the applied cosine loads, we will get a periodic response, that is, the displace-
ments in the soil, and the stresses will be distributed with the same period as the load.
If we add up the applied cosine loads (plus the one-dimensional or constant n = 0 term),
we get the applied strip load. If we add up the corresponding displacements, stresses, or
strains, we get the displacements, stresses, and strains for the strip load. We may therefore
write for the vertical uz and horizontal ux displacements

ux = ∑U
n =0
(n )
sin(α n x)


(2.2)
uz = ∑W (n )
cos(α n x)
n =0

Note that the horizontal displacements ux are represented by a sine series, as they will be in
the positive direction (if x is positive) or in the negative direction (if x is negative).
18  Geomechanics in soil, rock, and environmental engineering

1.5
No. of terms
in series
10
50

1.0

5
q(N)

0.5
q

0
a
q
L
= 20
a
L
–0.5
0 0.5 1.0 1.5
x
a

Figure 2.7 Approximation of uniform strip loading by Fourier series.

Periodic loading

uz (same period as
loading)

Figure 2.8 Response of soil to periodic load.


Finite layer methods  19

The stresses also may be represented in this form:


σ xx = ∑H
n =0
(n )
cos(α n x)

σ zz = ∑N
n =0
(n )
cos(α n x) (2.3)

τ xy = ∑T (n )
sin(α n x)
n =0

The direct stresses in the horizontal x and vertical z directions, σxx and σzz, are represented
by a cosine series as they will have the same sign for all x values, but the shear stress τxz is
represented by a sine series as the shear will be of an opposite sign if it is on the negative or
positive side of the x-axis.

2.2  APPROXIMATION OF FOURIER COEFFICIENTS

For a single component in the Fourier series for the displacements, we can write

 ux   U (n) sin(α n x) 
u =   =  (n )  (2.4)
 uz  W cos(α n x)

The strains can then be found since εxx = ∂ux /∂x, εzz = ∂uz /∂z, γxz = (∂ux /∂z) + (∂uz /∂x)
(n )
 ε xx   Ex(n) cos(α n x)
 
ε (n ) =  ε zz  =  Ez(n) cos(α n x) (2.5)
 γ xz   Exz
(n )
sin(α n x) 

where

 
(n ) αn 0 
 Ex    (n )
E (n )
=  Ez  = 0
∂  U  (2.6)
 ∂z  W (n) 
 Exz   
∂ −α n 
 ∂z 

The stresses may be related to the strains since

 ν 
(n )
 1 (1 − ν)
0  (n )
σ xx     ε xx 
σ  E(1 − ν)  ν 
= 1 0   ε zz  (2.7)
 zz  (1 + ν)(1 − 2 ν)  (1 − ν)
 τ xz     γ xz 
 (1 − 2 ν)   
 0 0
2(1 − ν) 

20  Geomechanics in soil, rock, and environmental engineering

or in a more compact form

σ(n) = Dε(n) (2.8)

The Fourier coefficients for the displacements U (n), W(n) will vary with depth z (as the
displacements must vary with depth). In order to obtain a numerical solution, suppose that
we approximate these coefficients by assuming they vary linearly between nodal values (see
Figure 2.9).
We can then write

Wi (zi + 1 − z) + Wi + 1(z − zi )
W (z) = (2.9)
(zi + 1 − zi )

A similar linear interpolation may be carried out for the Fourier coefficient U

Ui (zi + 1 − z) + Ui + 1(z − zi )
U(z) = (2.10)
(zi + 1 − zi )

For the nth Fourier term, we therefore have

(n )
 Ui 
 U (n )   a 0 b 0   Wi 
 (n )  =  (2.11)
W  0 a 0 b  Ui + 1 
 
Wi + 1 

(zi + 1 − z) (z − zi )
a = b =
∆z ∆z

Nodes
Elements

Wi
(i) zi

W(z) Δz
z
Wi+1
(i + 1) zi+1

Figure 2.9 Use of linear interpolation functions.


Finite layer methods  21

or in matrix form

U (n) = Cδ(n)

We can now use Equation 2.6 to obtain

(n )
 aα n 0 bα n 0   Ui 
 0  
1 0 1   Wi 
E(n) =  − (2.12)
 1 ∆z 1 ∆z  Ui + 1 
− − aα n −bα n  Wi + 1 
 ∆z ∆z

Or, in matrix form

E(n) = Bδ(n)

2.3 FORMULATION

For a layered system, we may write the virtual work equation as

∫dε σdV = ∫du T dS


T T
(2.13)
V S

This equation states that if the body is in equilibrium, the work done by the external loads
will be equal to the energy stored in the body under small virtual displacements.
If we take a single Fourier component (the nth) and substitute it into Equation 2.13, we have

∫ d ε(n)T σ(n)dV = In dδ(n)T B(n)T DB(n)δdz



V 0 (2.14)

∫ du (n)T
TdS = Indδ (n)T
T (n )

where

L L

In =
∫ cos2 (α n x) =
∫ sin (α x) 2
n (2.15)
0 0

Substituting the expressions in 2.14 into 2.13, we have (since dδ(n) are arbitrary variations
and not n
­ ecessarily zero)

zi + 1

∫ B( ) DB (2.16)
nT (n ) (n )
δ dz = T (n)
zi
22  Geomechanics in soil, rock, and environmental engineering

or writing the above equation in more compact form

K e(n)δ(n) = Te(n) (2.17)


where

zi + 1

∫ B( ) DB
(n ) nT (n ) (n )
K e = δ dz (2.18)
zi

As B (n) contains terms that are a function of z, we need to carry out the integration of
Equation 2.18. This may be done numerically or analytically.
For all the linear elements, we must sum the effects giving

K (n ) = ∑K e
(n )
e

(2.19)
T (n )
= ∑T e
(n )

As the only node that is loaded in this example problem for a strip footing is the top node,
we will only have one entry in the T (n) vector, corresponding to the top node, for example,

T (n) = (0, Q(n) , 0, 0,… , 0)T (2.20)


where Q (n) is the nth Fourier coefficient in the series for the applied load.

2.4  SOLUTION PROCEDURE

The solution procedure involves solving the set of equations

K (n)δ(n) = T (n) (2.21)

for each of the Fourier components. This involves setting up the T (n) vector for each of the
Fourier coefficients in the series for the applied load (see Equation 2.20) and the correspond-
ing K(n) matrix. This matrix is a function of αn and must be set up for each Fourier term.
Usually, this is a very fast process as we are dealing with only a few linear elements.
Equations 2.21 are then solved to give the Fourier coefficients of the displacements at the
nodes δ(n)
Once these coefficients are known, the actual displacements at the nodes may be found
since (for example, the displacement ux)

ux = ∑U (n )
sin(α n x) (2.22)
n =0

In practice, only a few terms in the series are required to obtain a good approximation
of the displacements. The number of terms depends on the spacing (or period L) of the
Finite layer methods  23

Approximate loads
N
using Fourier series
q=∑ Q(n) cos αnx
n=0 (vertical load only)

K (n) δ(n) = T (n) Form stiffness matrix


α = αn and load vector for
particular harmonic
Q = Q(n)

Do for
n = 0, …., N

δ(n) = K (n)–1 T (n) Solve stiffness


equations
δ(n) = (U1, W1, U2, W2, ....) Tn

N
ux = ∑ U (n) sin αnx Calculate field
n=0
quantities at any
N desired position x
uz = ∑ W (n) cos αnx
n=0

Figure 2.10 The process for solving a problem for each term in the Fourier series.

applied loads, however 20–30 terms would generally be enough. For larger L, more terms
are needed.
The solution procedure is shown schematically in Figure 2.10.

2.5  THREE-DIMENSIONAL PROBLEMS

For three-dimensional problems, we may make use of the double Fourier series. For exam-
ple, we could represent the uniform loading shown in Figure 2.11 by a double cosine series
(Equation 2.23).

∞ ∞

q(x, y) = ∑∑Q mn cos(α n x)cos(βm y) α n = 2nπ /L β m = 2m π /M (2.23)


m=0 n =0

This implies that the loadings are periodic, and hence in the case of general shaped loading,
we would have a series of loads spaced at the centre of L and M as shown in Figure 2.12.
24  Geomechanics in soil, rock, and environmental engineering

x
y

Figure 2.11 Three-dimensional loading.

Periodically spaced loadings

2B = M

y 2A = L

Figure 2.12 Periodic three-dimensional loadings.

We may now express the displacements and stresses as double Fourier series

∞ ∞

uy = ∑∑V
m=0 n =0
mn
cos(α n x)sin(βm y)

∞ ∞

uz = ∑∑W
m=0 n =0
mn
cos(α n x)cos(βm y) (2.24)

∞ ∞

σ yy = ∑∑S mn
yy cos(α n x)cos(βm y)
m=0 n =0

If we again approximate the Fourier coefficients’ variation with depth by using linear
interpolation between the nodal values, exactly the same set of finite layer equations arises
as is shown in Equation 2.21 for the two-dimensional case with the stiffness matrix set up
using ρ = (α n 2 + βm 2 )1 / 2 in place of αn:

K (ρ)δ mn = T mn (2.25)

This means that we may solve the three-dimensional problem by use of a simple one-
dimensional discretisation. It is necessary to set up the stiffness matrix for each value of αn ,
Finite layer methods  25

βm and to solve for all the Fourier coefficients. Again, it is not necessary to carry out the full
summation, but to use only a finite number of terms in the series. If there are N terms in
the αn summation and M terms in the βm summation, we would have to solve the simple set
of equations (M × N) times to obtain all of the Fourier coefficients. Once these have been
found, summations of the type shown in Equation 2.24 may be used to obtain the actual
field values at any point x, y.

2.6  CONSOLIDATION PROBLEMS

The same finite layer techniques may be used for problems involving consolidation of
­ orizontally layered soil deposits. For example, if we take the strip-footing problem again
h
and represent the excess pore pressures p that are induced into the soil as a Fourier series

p = ∑P (n )
cos(α n x) (2.26)
n =0

we obtain a set of equations

 K (n ) − LT (n)   ∆δ(n)   ∆h(n)(t) 


 (n )   =   (2.27)
−L −θ∆tΦ(n)   ∆q(n)   ∆tΦ(n)q(n)(t)

where q(n) is the vector of nodal pore pressure coefficients for Fourier term n, δ(n) is the
­vector of displacement coefficients for Fourier term n, K(n) is the stiffness matrix (for term n),
L(n) is a coupling matrix, Φ(n) is the flow matrix, and h(n) is the force matrix that is completely
analogous to the force matrix in Equation 2.21.
All of the matrices in Equation 2.27 are fully explained in the paper by Small and Booker
(1979).
To obtain a complete solution for pore pressures and displacements at any time, a ‘march-
ing type’ process is used. A time step Δt is taken and the changes Δδq(n), Δq(n) are found.
These rely on the solution for the excess pore pressure coefficients at the start of the time
step q(n)(t) (as may be seen in Equation 2.27 as these form part of the right-hand side of the
equations.
Solutions at a later time may then be calculated from

δ(n)(t + ∆t) = δ(n)(t) + ∆δ(n)


(2.28)
q(n)(t + ∆t) = q(n)(t) + ∆q(n)

The whole solution may be ‘marched’ forward in this fashion for each Fourier component of
the load vector Δh(n). To obtain the solution at any time t, we must carry out a summation
using the appropriate Fourier coefficients (that are now functions of time).
Equation 2.26 may more correctly be written


p(t) = ∑P n =0
(n )
(t) cos α n x

(2.29)
26  Geomechanics in soil, rock, and environmental engineering

2.7  FOURIER TRANSFORMS

If we apply Fourier transforms to the field variables (i.e. the displacements, pore pressures,
and stresses) instead of using Fourier series to represent them, we may again reduce two-
and three-dimensional problems to ones involving only a single spatial dimension.
The Fourier integral in effect replaces the Fourier summation and we are left with a very
similar set of equations to solve.
If we apply a Fourier transform to the loading function shown in Figure 2.5, we would
have
+∞
1
Q(α) =
2π ∫
q(x)cos(αx)dx
−∞
+a
1

(2.30)
= q(x)cos(αx)dx

−a

sin(αa)
= q
πa

In Equation 2.30, the first infinite integral is the Fourier transform of the step function
q(x), and as the strip function is an even function of x, a cosine function is used in the trans-
form. The transforms may also be applied to the displacements and stresses, for example, if
we take a strip-loading problem,
+∞
1
U(α) =
2π ∫ u (x)sin(αx)dx
−∞
x

+∞
(2.31)
1
W (α) =
2π ∫ u (x) cos(αx)dx
z

−∞

For an elastic problem, we would finally have a set of equations

K (α)δ(α) = f (α) (2.32)

The force vector f(α)would contain the transform of the applied load, for example,
T
 q sin(α a) 
f (α) =  0, , 0, 0, , 0 (2.33)
 πa 

and the vector δ(α) would contain the transforms of the displacements,

δ = (U1, W1, U2 , W2 ,)T

As we can write the inverse transforms as say

+∞


ux =
∫U(α)sin(αx)dα
−∞
(2.34)
Finite layer methods  27

+∞

uz =
∫W(α)cos(αx)dα (2.35)
−∞

once the transformed quantities are found from Equation 2.32, we can carry out the inte-
gration to obtain the final results.
This is usually difficult to do, and so numerical integration is used. Gaussian integration
of a function of α may be carried out as follows:

∞ G

∫ f (α)dα = ∑ω f (α )
i =1
i i (2.36)
−∞

where the αi are the values of α at the selected Gauss points, and the ωi are the Gaussian
weights. G is the number of Gauss points. These weights and Gaussian coordinates may be
found from tabulated values in maths books.
Hence, to numerically invert a function of α, we need to know its value at the
Gauss points αi. This means that we need to solve Equation 2.32 at these selected values,
that is,

K (α i )δ(α i ) = f (α i ) (2.37)

Now we may use numerical integration to find all the required values, for example,

ux = ∑ω U(α )sin(α x)
i i i (2.38)
i =1

The Gaussian integration may be carried out in several blocks as shown in Figure 2.13.

+∞ +A

∫ f (α)dα ≈ ∫ f (α)dα (2.39)


−∞ −A

We do not need to integrate all the way from −∞ to +∞, but we can truncate the integral at
some large value A. As long as A is large enough, a good approximation to the integral may
be obtained.

f (α)
f(α)

Block 1 Block 2 Block 3

Figure 2.13 Numerical integration scheme.


28  Geomechanics in soil, rock, and environmental engineering

2.8 EXAMPLES

As an example of finite layer analysis, the program FLEA (Finite Layer Elastic Analysis – Small
and Booker 2014) is used in the following. A simple problem involving a uniform circular load-
ing q = 100 kPa of radius a = 1.0 m applied to a finite layer of uniform soil of thickness h = 2.0 m
is analysed. The elastic modulus is taken as E = 10,000 kPa and Poisson’s ratio as ν = 0.3.
The vertical displacements along the axis of the footing are shown in Figure 2.14, where
it can be seen that the settlement at the surface is about 12 mm.
A more complex problem is one of a square footing 2 m × 2 m in plan. The soil consists of
two layers of soil: the upper layer being 2 m thick and the lower layer 4 m thick. The soil is
anisotropic, and the layers have the properties shown in Table 2.1.
The square footing has unit loads of 1 kPa applied laterally in the x-direction and 1 kPa
applied vertically on it, and the results are calculated throughout the depth of the layer
beneath the point x = 0.5 m, y = 0.5 m.
The stress in the x-direction σxx is shown in Figure 2.15, where it may be seen that the stress
has a discontinuity in it between the layers. A plot of the vertical displacement (Figure 2.16)
shows that the displacement is continuous although there is a kink at the interface of the
two layers (z = 2.0 m).

0.4

0.8
Depth (m)

1.2

1.6

2.0
0 4 8 12
Displacement (mm)

Figure 2.14 Vertical displacement beneath the centre of circular loading.

Table 2.1  P
 roperties of layered soil
Thickness Modulus Modulus Shear modulus Poisson’s Poisson’s
Layer (m) Ex (kPa) Ez (kPa) Gz (kPa) ratio νxx ratio νzx
1 2 1500 1000 450 0.25 0.2
2 4 12,000 4000 2000 0.1 0.3
Finite layer methods  29

2
Depth (m)

6
0 0.4 0.8 1.2
Stress σxx (kPa)

Figure 2.15 Stress in x-direction σxx beneath the square surface loading at x =  0.5 m, y =  0.5 m.

2
Depth (m)

6
0 0.4 0.8 1.2
Displacement (mm)

Figure 2.16 Vertical displacement beneath the square surface loading at x = 0.5 m, y = 0.5 m.
This page intentionally left blank
Chapter 3

Finite element methods

3.1 INTRODUCTION

Because soils are in general complex materials that consist of the soil skeleton, the pore air,
and pore water and exhibit non-linear stress–strain behaviour and perhaps time-dependent
behaviour, it is not always simple to obtain analytic solutions to problems. Analytic solu-
tions are those that contain closed form solutions that may be a simple algebraic formula
that can be evaluated. Some solutions are semi-analytic, for example, the solution may be
written as an integral that is not able to be inverted through analytic means and has to be
evaluated using numerical integration. Finally, there are the numerical solutions that rely on
numerical approximation rather than algebraic manipulation for their solution.
Although there are many very useful analytic solutions available for use in geomechan-
ics, more complex problems often arise that need the power of numerical means to obtain
a solution. One of these numerical methods, the finite element method, is presented in this
chapter.

3.2  TYPES OF ELEMENTS

The basis of the finite element method is that the volume of interest is divided up into dis-
crete elements. The elements may be of different shapes and some of the shapes are shown
in Figure 3.1. As shown in this figure, the elements have ‘nodes’ (shown as black dots), and
different element types may have different numbers of nodes. For example, in Figure 3.1, the
triangular shaped element is shown with three or fifteen nodes.
It is assumed that the field quantities are known at the nodes, but can be found within the
element by mathematical interpolation functions. In geomechanics, the field quantities may
be displacements or water pressure head for instance.
The mathematical functions that are used within the element are called ‘shape functions’
as they depict the shape of the field quantity within the element if plotted. The number of
shape functions used is equal to the number of nodes in the element as the shape function
can be thought of as giving the shape of the field quantity within the element for a unit value
of the field quantity at one of the nodes.
For instance if we give one node a unit displacement, the shape function gives the dis-
placement within the element. Figures 3.2 and 3.3 show how a six-node triangle will deform
for a displacement at node 1 or at node 2, respectively. If all of the nodes displaced a unit
amount, then we superimpose the displacements given by the shape functions for each node.
If the displacement is not unity, we can scale the displacement by multiplying the shape
functions by the actual displacements at each of the nodes and then adding the result.

31
32  Geomechanics in soil, rock, and environmental engineering

3-node triangle 8-node isoparametric 15-node triangle

20-node isoparametric 3D element

Figure 3.1 Different types of finite elements.

Figure 3.2 Shape function for the first node of a six-node triangle.

3.2.1  Finding shape functions


Suppose, as an example, we have an element with six nodes and the values of some quantity
of interest (w), such as displacement, head, and temperature, are known at each of the nodes.
It is assumed that within the element the variation of w at position (x, y) can be approxi-
mated by a polynomial expression:

w(x, y) = a1 + a2 x + a3y + a4 x2 + a5xy + a6 y 2 (3.1)



Finite element methods  33

1
4

Figure 3.3 Shape function for the second node of a six-node triangle.

where ak are polynomial coefficients. Because there are six nodes, we have six unknown
coefficients. The functions of x and y are chosen to give a quadratic variation of the function
in both the x and y directions. If the element were three-dimensional, functions of x, y, and
z would be needed. If the element has m nodes, we would need m terms in the polynomial.
We can therefore write

w(x, y) = aT f (x, y) = f T (x, y) ⋅ a (3.2)


where a = {a1, a2 , …, ak, …, a 6}T and f(x, y) = {1, x, y, x 2 , xy, y 2}T.


Suppose that the element nodes are located at the points (x1, y1), (x 2 , y 2), …, (x6, y6). At
the ‘kth’ node, the value of the quantity w is

wk = a1 + a2 xk + a3yk + a4 xk2 + a5xk yk + a6 yk2 (3.3)


Equation 3.3 holds at each of the six nodes. These equations may be written in matrix
form as follows:

w = Ca (3.4)

where

1 x1 y1 x12 x1y1 y12 


 
1 x2 y2 x22 x2 y2 y22 
1 x3 y3 x32 x3y3 y32 
w = (w1, w2 ,… , w6 )T and C =   (3.5)
1 x4 y4 x42 x4y4 y42 
1 x5 y5 x52 x5y5 y52 
 
1 x6 y6 x62 x6 y6 y62 

The solution of Equation 3.4 is

a = C −1w (3.6)
34  Geomechanics in soil, rock, and environmental engineering

When a from Equation 3.6 is substituted into Equation 3.2, it is found that the quantity
of interest can be expressed in the form of

w(x, y) = f T (x, y) ⋅ C −1w = N ⋅ w (3.7)


The vector N must contain the shape functions as it relates the values at the nodes of the
element w to the value at any point within the element w(x, y).
Hence,

N = (N1, N 2 ,… , N6 )T = f T (x, y) ⋅ C −1 (3.8)


and we can find the shape functions for any element in this manner simply from a knowledge
of the coordinates of the nodes of the element (as they are used to establish the C matrix). The
inverse of the C matrix can be found numerically or if it is possible to find the algebraic values
of the inverse, the shape functions can be written as algebraic expressions. Some shape func-
tions for different types of elements are given in Appendix 3A, but others may be found in spe-
cialized books on finite elements such as Zienkiewicz (1977) or Potts and Zdravkovic (1999).

3.2.2  Isoparametric elements


An isoparametric element is one where the geometry of the element is also determined by a
shape function. The term ‘isoparametric’ comes from the fact that for this particular type of
element, the shape functions used for determining the shape of the element are the same as
those used to interpolate the field quantities within the element. They do not have to be the
same, and could be different, but using the same functions simplifies the analysis.
An eight-node isoparametric element with curved sides is shown in Figure 3.4. By allow-
ing the element to have curved sides, shapes such as circular tunnels and curved geometries
can be more easily discretised.
The shape functions are now written in terms of local parameters ξ, η because the coor-
dinates x, y cannot be used. The local parameters range between −1 and +1, for example,
at node 1, the local coordinates are ξ = −1 and η = −1. The shape functions are given in
Appendix 3A but, for example, the shape function for the first node is given by

1
N1 = (1 − ξ)(1 − η)(−ξ − η − 1) (3.9)
4

(a) (b) 5

7 6 5
6
4
η
3
8 4
ξ
7
8
2
1 2 3
1

Figure 3.4 (a) Local coordinates. (b) Curved shape of an eight-node isoparametric element.
Finite element methods  35

The coordinates within the element are given by

x = N1x1 + N 2 x2 +  + N8x8
(3.10)
y = N1y1 + N 2 y2 +  + N8y8

The problem now is to differentiate the shape functions with respect to x, y because the
actual coordinates are also functions of the local coordinates. Often the shape functions
need to be differentiated to obtain the strains from the displacements or the gradient of the
total head for example. This is done by using the chain rule for differentiation

∂N1 ∂N1 ∂x ∂N1 ∂y


= +
∂ξ ∂x ∂ξ ∂y ∂ξ
(3.11)
∂N1 ∂N1 ∂x ∂N1 ∂y
= +
∂η ∂x ∂η ∂y ∂η

or in matrix form

 ∂N1   ∂x ∂y   ∂N1   ∂N1 


 ∂ξ   ∂ξ ∂ξ   ∂x   ∂x 
  =    = [ J] 
 ∂N   ∂x      (3.12)
∂y ∂ N ∂ N
 1   1 
1

 ∂η   ∂η ∂η   ∂y   ∂y 

The matrix in the above equation [ J] (Equation 3.12) is called the Jacobian, and by inverting
the Jacobian we can obtain the differentials of the shape functions with respect to the actual
coordinates. Hence,

 ∂N1   ∂N1 
 ∂x   ∂ξ 
  = [ J ]−1  
 ∂N1   ∂N  (3.13)
   1
 ∂ y   ∂η 

This can be done for any isoparametric element; if it is three-dimensional then the Jacobian
is a 3 × 3 matrix as there are three axis directions.

3.2.3  Infinite elements


As can be seen with an isoparametric element, the shape of the element can be deter-
mined by using shape functions. If we choose the shape functions (now called mapping
functions) so that the coordinates at one side of the element become infinite, we can cre-
ate an infinite element as shown in Figure 3.5. Infinite elements are useful when the side
boundaries of the problem at hand need to be a long way from the region of interest so
that they do not influence the solution. They can also reduce the size of the mesh required
and therefore save a good deal of computation time, especially for three-dimensional
problems.
36  Geomechanics in soil, rock, and environmental engineering

3
η
ξ
4

ξ = +1 at ∞
5

2
1

Figure 3.5 A five-node infinite element.

For example, a five-node infinite element that can be used with the eight-node isoparamet-
ric element shown in Figure 3.4 can have mapping functions that decay with ξ−1
x = M1x1 + M2 x2 +  + M5x5
ξη(1 − η)
M1 =
(1 − ξ)
0.5(1 + ξ)(1 − η)
M2 =
(1 − ξ)
0.5(1 + ξ)(1 + η)
M3 =
(1 − ξ)
ξη(1 + η)
M4 = −
(1 − ξ)
2ξ(1 − η2 )
M5 = −
(1 − ξ)
(3.14)

It may be seen from the mapping functions that the coordinate in the direction of ξ goes
to infinity as ξ approaches +1. The shape functions relating displacement (say) to nodal dis-
placement are the same as for the eight-node element. To differentiate a shape function with
respect to x and y, the same procedure can be used as for ordinary isoparametric elements,
but with a Jacobian formed by differentiating the mapping functions. It may be noted that
if the infinite element is used at the side of a finite element mesh to extend the mesh laterally
to infinity, then the y coordinate does not go to infinity, only the x coordinate. In this case,
the normal shape function is used for the y coordinate mapping.
For mapping functions that decay as ξ−2 and for three-dimensional infinite elements, see
Appendix 3A.

3.2.4  Finite element meshes


Individual elements are joined together to form meshes covering the field of interest for the
problem. A mesh of elements is shown in Figure 3.6 for a three-dimensional problem involv-
ing a piled raft on a layered soil (1/4 of the mesh is used because of symmetry). The mesh
of Figure 3.6 is made up of different types of elements: (i) solid 20-node elements for the
soil, (ii) 8-node two-dimensional elements for the raft, and (iii) infinite 12-node elements to
extend the soil boundaries to large distances. The elements can be used to model different
behaviours. For example, a raft element will behave differently to a solid soil element. To
Finite element methods  37

Material types

Materials
1
2
3
4
5

y x

Infinite elements

Piled raft (5 × 5 pile group) – Solid elements for piles

Figure 3.6 Example of a finite element mesh made up from different types of elements.

model the behaviour of different materials, we need to formulate the problem and incorpo-
rate laws that best model the behaviour of a material.
Formulation of finite element equations for different problems is discussed in the f­ ollowing
sections.

3.3  STEADY STATE SEEPAGE

Finite element analysis of steady state seepage involves the analysis of flow that is continuous
and for which the total head or permeabilities within the field of flow do not change. If the
location of the water table changes (e.g. when the stored water in a dam drops) or the soil
is being wet by an advancing front, then the seepage is not steady state. Here, steady state
seepage will be examined first, and then non-steady flow problems will be addressed.

3.3.1  Governing equation


The governing equation for steady state flow is established by assuming that for any small
volume in the soil, the volume of water that flows out must be equal to the volume that
38  Geomechanics in soil, rock, and environmental engineering

flows in. This implies that water is not being pumped into or out of the soil at the point, and
that there is continuity of flow.
This leads to the equation

∂  ∂h  ∂  ∂h  ∂  ∂h 
 kxx  +  kyy  +  kzz  = 0 (3.15)
∂x  ∂x  ∂y  ∂y  ∂x  ∂z 

where kxx, kyy, and kzz are the permeabilities of the soil in the three axis directions x, y,
and z. In Equation 3.15, h is the total water pressure head which is made up of the elevation
head hE and the water pressure head hw, that is, h = hE + hw. It is necessary to work in terms
of the total head as flow depends on the elevation of a point in the soil as well as the water
pressure head that exists there.

3.3.2  Finite element formulation


If we approximate the total head h over the volume of interest, then the governing equation
will not be zero everywhere but will contain some error, and this is called the ‘residual’.
To get the best approximation to the real function, we try to make the residual as small
as possible everywhere within the region. This can be done by integrating the residual times
a small change in the approximating function over the region occupying the volume V and
setting the result to zero.

 ∂  ∂h  ∂  ∂h  ∂  ∂h  
∫ δh   kxx  +
 ∂x 

∂x  ∂y 
kyy  + 
∂y  ∂x 
kzz   dV = 0
∂z  
(3.16)
V

This leads to the equation

∫ δi [k]i dV = 0
T
(3.17)
V

where i = (∂h/∂x, ∂h/∂y, ∂h/∂z)T = ∇h is the hydraulic gradient of the pore fluid and

kxx 0 0

[k] =  0 kyy 0 
 0 0 kzz 

is the matrix of permeabilities for the soil.

3.3.3  Approximation of total head h


We can use the usual finite element interpolation functions to approximate the total head h
within an element. Suppose the element has n nodes and Ni are the shape functions.

h = N1h1 + N 2h2 + N3h3 + N 4h4 +  + N nhn (3.18)


Finite element methods  39

Then the value of the hydraulic gradient for the element ie would be

 ∂N1 ∂N 2 ∂N n 
  
 ∂x ∂x ∂x   h1 
 
 ∂N ∂N 2 ∂N n   h2 
ie =  1  (3.19)
∂y ∂y ∂y    
  
 ∂N1 ∂N 2 ∂N n  hn 

 ∂z ∂z ∂z 

or in a more concise form

ie = Eh (3.20)

3.3.4  Finite element equations


Equation 3.17 can now be written for a single element e

∫ δh E [k]Eh dV
T
e
T
e e = 0 (3.21)
Ve

As δhe is an arbitrary variation in head and not necessarily zero, then we must have

∑ ∫ E [k]Eh dV
elem Ve
T
e e = 0 (3.22)

or

Φh = 0 (3.23)

where

Φ = ∑ ∫ E [k]EdV
elem Ve
T
e (3.24)

The matrix is called the global flow matrix for the problem and is assembled for all of the
elements in the finite element mesh. The vector h contains the total head at each node in the
finite element mesh. By solving the set of Equations 3.23 subjected to some boundary condi-
tions, the total head at each node can be found. Assembly of the element stiffness matrices
and the application of boundary conditions is examined in Appendix 3B. Because in flow
problems, there is only one variable per node (i.e. the total head), the set of equations to be
solved is generally not large and the bandwidth of the equations is small.

3.3.5  Calculation of flows


Returning to Equation 3.17 and introducing Darcy’s law (i.e. v = [k]i), we can write
40  Geomechanics in soil, rock, and environmental engineering

Flow in Flow out

Figure 3.7 Confined seepage showing the flow balance.

∫ δi [k]i dV = δh Φh
V
T T

=
∫ (∇δh) v dV
V
T

(3.25)
=

V
δhT ∇T v dV

=
∫ δh v ⋅ n dS
T
by the divvergence theorem
S

Hence, we have

Φh =
∫ v ⋅ n dS (3.26)
S

and so by multiplying the flow matrix by the solution for the heads, we get the flows at the
nodes of the finite element mesh. This is done for each element in the mesh and the flows at
the nodes are summed.
Internal nodes will have no flow, but the boundary nodes will exhibit flows if they are on
a permeable boundary. There should be a flow balance in the finite element mesh with flow
in = flow out as shown in Figure 3.7.

3.3.6  Flow lines


In time dt, the distance travelled by a particle (with flow velocity components vx and v y as
shown in Figure 3.8) of fluid is

In the x-direction = vx ⋅ dt
In the y-direction = vy ⋅ dt

Therefore,

dx v dt v
= x = x
dy vy dt vy (3.27)
vy ⋅ dx − vx ⋅ dy = 0

Finite element methods  41

vx
dy dQ
dx
v
vy

Figure 3.8 Section of a flow line.

If we define a function Ψ(x, y) that is constant along the flow line we would have on
differentiating

∂Ψ ∂Ψ
⋅ dx + ⋅ dy = 0 (3.28)
∂x ∂y
Equations 3.27 and 3.28 suggest that

∂Ψ
vx = −
∂y
(3.29)
∂Ψ
vy =
∂x

3.3.7  Calculation of flow using the stream function


From Figure 3.8, it can be seen that the flow across the face dQ is given by

dQ = vx ⋅ dy − vy ⋅ dx
∂Ψ ∂Ψ (3.30)
= − ⋅ dy − ⋅ dx
∂y ∂x

and so the total flow across a boundary between points A and B can be found by integration.
B

Q =
∫ dQ = Ψ A − ΨB (3.31)
A

This means that the values of the stream function can be found relative to a point on the
­boundary (say A) by summing the flows at the nodes around the edge of the finite element mesh.

3.3.8  Determining the stream function


From Equation 3.29, we can show that

∂h ∂Ψ
vx = −kx = −
∂x ∂y
(3.32)
∂h ∂Ψ
vy = −ky =
∂y ∂x
42  Geomechanics in soil, rock, and environmental engineering

and so the hydraulic gradients are

∂h 1 ∂Ψ
ix = =
∂x kx ∂y
(3.33)
∂h 1 ∂Ψ
iy = = −
∂y ky ∂x

By substituting back into the original equation (Equation 3.17), used to formulate the
problem, we can obtain a set of equations for the stream function. These equations are the
same as the finite element equations for the flow, but with

1
kx →
ky
1 (3.34)
ky →
kx
h→ Ψ

Hence, we can solve the same set of finite element equations as for the total head (Equation
3.23), but using the reciprocal of the permeabilities. The solution will be the stream func-
tion, and these values can be contoured to plot the flow lines (e.g. see Figure 3.7).
To solve the finite element equations, we need some boundary conditions, and these are
the values of the stream function around the edges of the finite element mesh. As mentioned
before (Equation 3.31), the value of the stream function at the nodes around the outside of
the mesh can be computed by summing the flows, and these flows are computed in the first
part of the analysis by solving for the total heads and applying Equation 3.26.

3.3.9  Pumping or extracting fluid


If fluid is being pumped from or injected into the ground, Equation 3.15 does not hold,
and the right-hand side of the finite element equations will not be zero. Equation 3.23 will
become Φh = f where f = (0, 0,…,qn ,…,0)T and qn is the quantity of fluid being injected or
extracted (depending on sign) at node n.

3.4  STRESS ANALYSIS

The finite element equations for problems involving a single-phase elastic material can be
established by using the principle of virtual work. This states that if a body is in equilibrium
under a set of applied forces, then for any virtual displacements applied to the body the
energy stored in the body will be equal to the work done by the external forces and any body
forces (e.g. self-weight of the material).
For soil and rock mechanics, the body considered is the soil mass and applied forces
would be due to loads from structures or embankments for example.
The internal energy W I stored per unit volume due to a virtual displacement to the body
is given by

dWI = σ d ε (3.35)

Finite element methods  43

where σ is the stress at a point in the body and dε is the increment in strain caused by the
virtual displacements. For general three-dimensional problems, the stress vector consists of
the six components of stress at a point, namely the three normal stresses and three shear
stresses. The strain vector also contains six strain components. The energy stored in the
whole body involves the integral of the work per unit volume over the volume of the body,
that is,

WI =
∫ σdε dV (3.36)
V

The work done by the surface tractions (such as applied uniform or point loads) is given by

WF =
∫ qdu dS + ∫ γdu dV + P du (3.37)
S V

where q is a traction (or load) applied over an area of the surface dS and so must be inte-
grated over the area where the loading (which may not be uniform) is applied. P is a point
or concentrated load applied at a point on or in the body. The increment in displacement
at points in the body is du. For general three-dimensional loading, the displacement vector
has three components corresponding to the three displacements in the three axis directions,
as does the point load vector P. The loading traction q also has three components as one
normal and two shear stresses can act on a surface. The body force vector γ has three com-
ponents, as body forces can act in all three-axis directions (these can be due to gravity or
centrifugal forces).
Equating internal and external work, we can write

∫ σdε dV = ∫ qdu dS + ∫ γdu dV + P du (3.38)


V S V

For linear elastic materials, the stress–strain relationship can be found simply from
Hooke’s law which may be written as

σ = DEε (3.39)

where DE is a matrix of elastic constants. For an isotropic elastic material and general three-
dimensional conditions, the DE matrix is a 6 × 6 matrix as shown in Equation 3.40.

λ + 2G λ λ 0 0 0
 λ λ + 2G λ 0 0 0 

 λ λ λ + 2G 0 0 0
DE =   (3.40)
 0 0 0 G 0 0
 0 0 0 0 G 0
 
 0 0 0 0 0 G

44  Geomechanics in soil, rock, and environmental engineering

where E is the elastic modulus of the material and ν is its Poisson’s ratio. G is the shear
modulus relating shear stress to shear strain and λ is Lamé’s parameter defined as


λ =
(1 − ν)(1 − 2 ν)
E
G = (3.41)
2(1 + ν)
σ = (σ xx , σ yy , σ zz , τ xy , τ yz , τ zx )T

ε = (ε xx , ε yy , ε zz , γ xy , γ yz , γ zx )T

In the above equations, σ are the normal stresses to a plane and τ are the shear stresses.
Likewise, ε are the normal strains and γ are the shear strains.
We can now make the finite element approximations of the field quantities (displacements
u = (ux, uy, uz)T) through use of the element shape functions N.

u = Nδ (3.42)

where the displacements at the nodes of the element are contained in the vector δ.
By differentiating the displacements to obtain the strains (see Section 2.2), the strain–dis-
placement relationship can be obtained.

ε = Bδ (3.43)

By substituting the finite element approximations for the strains and the stresses, we have

∑ ∫ dδ B D Bδ dV = ∫ dδ N qdS + ∑ ∫ dδ N
elem Ve
T T
E e
T T

elem Ve
T T
γ dVe + dδ T N T P (3.44)
S

Since the virtual displacements are arbitrary and not necessarily zero, we must have

∑ ∫ B D Bδ dV = ∫ N q dS + ∑ ∫ N
T
E
T T
γ dV + N T P (3.45)
V S V

or in more compact form

K δ = fq + f γ + fP (3.46)

where
K = Σ ∫Ve BT DE B dVe is the stiffness matrix for the body
fq = ∫S NTqdS is the force vector for uniform loads
f γ = Σ ∫Ve N T γ dVe is the force vector for body forces
f P = P is the force vector for point loads

The point load force vector reduces to simply the values of the point forces as the shape
functions are equal to unity at the nodes of an element, and point loads can only be applied
at the nodes.
Finite element methods  45

3.5  CONSOLIDATION ANALYSIS

Consolidation analysis is a combination of stress analysis and flow analysis, as the flow of
water out of the pores of the soil allows the soil to undergo a volume change. Excess pore
pressures caused in the soil due to external loads dissipate during consolidation and the
applied stress is transferred from the pore water pressures to the soil skeleton.

3.5.1  Effective stress analysis


We can set up the finite element equations for the stress analysis part of the equations by
again using Equation 3.36. This time however, use can be made of the effective stress prin-
cipal that states that the total stresses σ are made up of the effective stresses σ′ (the stress in
the soil skeleton) and the pore water pressures p. The pore pressures only add onto the direct
stresses not the shear stresses and so we can write

σ = σ ′ + pj (3.47)

where j = (1, 1, 1, 0, 0, 0)T.


Substituting the total stress into Equation 3.38 gives

∫ (σ′ + pj)d ε dV = ∫ q dudS + ∫ γ du dV + P du (3.48)


V S V

If the finite element approximations of the field quantities are now made, we obtain

∑ ∫ dδ B D Bδ dV + ∑ ∫ dδ B jpdV = ∫ dδ N q dS + ∑ ∫ dδ N
T T
E e
T T
e
T T T T
γ dVe + dδ T N T P

Ve Ve S Ve

(3.49)

where BT j = d a vector that is the sum of the first three rows of the B matrix. It therefore
contains the interpolation functions for the volume strain, as the volume strain is the sum of
the three linear strains, that is, θ = εxx + εyy + εzz. As before, ∑ denotes a summation over all
of the elements in the mesh, and Ve is the volume of an element.
The pore water pressures can be interpolated within the element by using the shape fac-
tors (these are the same shape factors as used for the displacements) that will be contained
in a vector a to distinguish these factors from the displacement shape factors. Some inves-
tigators like to use different interpolation functions for the pore water pressures and the
displacements, but in the author’s experience this is not necessary. We can therefore write

p = aT q
(3.50)
θ = dTδ

If as before we note that the variations in the displacements are arbitrary and not neces-
sarily zero, then we must have

K δ + LT q = fq + f γ + fP (3.51)

46  Geomechanics in soil, rock, and environmental engineering

where the force vectors have the same definitions as in Equation 3.46, and the matrix L has
the definition

LT = ∑ ∫ da T
dVe (3.52)
Ve

In incremental form, the above equation may be written as

K ∆δ + LT ∆q = ∆fq + ∆f γ + ∆fP (3.53)


The pore water pressure may be written in terms of the pressure head as the total head h
is equal to the sum of the elevation head hE and the water pressure head hw, that is,

h = hE + hw (3.54)

The water pressure p may therefore be substituted by

p = γ w hw = γ w (h − hE ) (3.55)

The same vector of interpolation functions can be used for the total head and the water
pressure and so

p = aT q = γ w aT (h − hE ) (3.56)

Any changes in the pore pressure can now be calculated, and since changes in the eleva-
tion head are zero (if we are dealing with small strain analysis), we have

∆q = γ w ∆h (3.57)

This enables us to write the incremental equations in terms of total head as

K ∆δ + γ w LT ∆h = ∆fq + ∆f γ + ∆fP (3.58)


For consolidation problems, we need to take care of the signs of the terms if we want
compression to be positive, so that compressive effective stresses and compressive pore water
pressures are positive.
If a compressive load is downward as shown in Figure 3.9, then it is opposite to the posi-
tive y-axis and is therefore negative. A downward movement will therefore need to yield a
positive stress and so we must write

σ = −Dε (3.59)

Hence, Equation 3.58 will need to have a change in the signs of the force vector and the
stiffness matrix giving finally

K ∆δ − γ w LT ∆h = ∆fq + ∆f γ + ∆fP (3.60)



Finite element methods  47

q
y

Figure 3.9 Sign convention for compression positive.

3.5.2  Volume balance


If it is assumed that the soil is saturated, then any change in volume of the pores of the soil
due to water flowing out of the pores is compensated for by a change in the volume of the
soil that is equal to the volume of water lost from the pores. This is making the assumption
that the pore water is not compressible and that there is no air in the pores of the soil.
In examining the flow of water through the soil, we saw previously that the quantity of
water flowing from an element of soil (Equation 3.15) was balanced by the water flowing
into the element for steady state seepage. The amount of water flowing out of the element of
soil is not zero when consolidation is occurring, and is equal to the rate of the volume strain.
We can therefore write

∂θ
[k]∇ 2h = (3.61)
∂t

where, as before, [k] is the matrix of permeabilities (see Equation 3.17), ∇ is the gradient
vector ∇ = (∂/∂x, ∂/∂y, ∂/∂z)T, and h is the total head. The operator

 ∂2 ∂2 ∂2 
∇2 =  2 + 2
+ 2  = ∇T (∇) (3.62)
 ∂x ∂y ∂z 

As before (Equation 3.16), we can integrate the flows over the volume to minimise the
error in the flows:

∂θ
∫ δh [k]∇ (∇h) dV − ∫ δh ∂t dV = 0
T
(3.63)
V V

∂θ
∫ δi [k]i dV − ∫ δh ∂t dV = 0
T
(3.64)
V V

If we now substitute the finite element approximations for the volume strain and the gra-
dient of the total head, we have

∂δ
∑ ∫ δh E [k]Eh dV − ∑ ∫ δh a d  ∂t  dV
T T
e
T T
e = 0 (3.65)
Ve Ve

48  Geomechanics in soil, rock, and environmental engineering

We can write this equation in more compact form (noting that the increment in total head
δh is not necessarily zero and so the rest of the equation must be zero) as

∂δ
Φh − L = 0 (3.66)
∂t

and the definitions of the matrices are

Φ = ∑ ∫ E [k]E dV
Ve
T
e

(3.67)
L = ∑∫ aT d dVe
Ve

The flow matrix Φ is the same as before for steady state seepage, however, we now have an
extra term in the finite element equations because the flow of water at a point in the soil is
not zero as it was for steady state seepage. Once again, ∑ indicates a summation for all the
elements in the mesh.
Equation 3.66 can now be integrated numerically by noting that if we integrate a function
f(t) with respect to time t we are finding the area under the curve of f(t) versus t. The integral
can be approximated by the following formula

t + ∆t

∫ f (t) = α∆t ⋅ f (t) + (1 − α)∆t ⋅ f (t + ∆t) (3.68)


t

This is illustrated in Figure 3.10, where it can be seen that this is an approximation to the
area under the curve (the areas of the two shaded rectangles), and α is a quantity that can
lie between 0 and 1.
If we apply this integration scheme to Equation 3.66, then we have

α∆tΦht + (1 − α)∆t Φht + ∆t − L∆δ = 0


(3.69)
− (1 − α)∆t Φ∆h − ∆t Φht − L∆δ = 0

f (t)

f (t + Δt)
f (t)

αΔt (1 – α) Δt
t t + Δt

Figure 3.10 Numerical integration scheme.


Finite element methods  49

and so combining the volume balance equation (Equation 3.69) and the effective stress
equation (Equation 3.60), we obtain the final set of equations involving consolidation in an
incremental form.

 K − γ w LT   ∆δ   ∆fq + ∆f γ + ∆fP 
   =   (3.70)
−γ w L −(1 − α)∆tγ w Φ   ∆h   ∆t γ w Φht 

The solution can therefore be ‘marched’ forward, with the solution started off by a knowl-
edge of the total head at time zero. The increments in the displacements and heads are
found, and then the new displacements and heads at time t + Δt are found, that is,

δ t + ∆t = δ t + ∆δ
(3.71)
ht + ∆t = ht + ∆h

The new values of total head are used in the right-hand side of the equations to allow
a new solution to be found and so on. Errors can accumulate with solutions of this kind,
and so it is of interest to know under what circumstances the ‘marching’ scheme is stable.
This has been examined by Booker and Small (1975) where it was found that the scheme
is unconditionally stable if α ≤ 0.5. The scheme can be made stable with values of α
greater than 0.5, but this depends on the eigenvalues of the equations, and so this is not
practical.
An example of a solution for consolidation of a soil layer under a uniform circular load-
ing q is shown in Figure 3.11 where the vertical permeability kv is different to the horizontal
permeability kh. The properties used to obtain the solutions are shown in Table 3.1.
By comparing the solution of the numerical equations of consolidation with an analytic
solution from the program CONTAL (Small 2012), it may be seen that the closest agreement

Time (days)
0.001 0.01 0.1 1 10 100 1000 10,000 100,000
–0.045
1

–0.050

–0.055 kv = 0.1.kh
kv= 0.1.kh
Settlement (m)

–0.060 kv = α = 0.0
kv=10.k
10.khh α=0.0
α=0.5α = 0.5
α=0.0
α = 0.0
–0.065 F.E.F.E.solutions
solutions α=0.5
α = 0.5
0.0

–0.070

–0.075
CONTAL
–0.080

Figure 3.11 Effect of the value of α on the finite element ‘marching’ solution. Comparison with an analytic
CONTAL solution.
50  Geomechanics in soil, rock, and environmental engineering

Table 3.1  Properties used in finite element analysis


Quantity Value
Drained modulus of elasticity 10,000 kPa
Drained Poisson’s ratio 0.35
Radius of load a 8 m
Depth of layer h 16 m
Horizontal permeability kh 0.0001 m/day
Uniform load q 80 kPa

between the solutions is obtained when α = 0, and so this is recommended for numerical
analysis.
The accuracy of the numerical solution also depends on the size of the time step Δt chosen
as for many other forms of time-dependent numerical solutions. Consolidation problems
tend to need small time steps at the beginning of consolidation when pore pressure gradients
are high and consolidation is rapid. As time goes on, the consolidation process slows down
as pore pressures tend to even out. Therefore, it is necessary to increase the size of the time
step to allow a solution to be obtained with a reasonable amount of computational effort.
As can be seen from Equation 3.70, if the time step is changed, the set of equations has
to be set up again and the consolidation matrix has to be refactored (for a Crout–Cholesky
factorisation) or a new Gaussian elimination performed if these kinds of solution methods
are being used. Once this is done a simple back substitution is performed at each time step
until the time step is changed.

3.6  NUMERICAL INTEGRATION

In order to obtain the matrices such as the ones in Equation 3.70, there is a volume integral
to perform. For two-dimensional elements, this reduces to an area integral. In some cases,
this integration can be carried out algebraically, but in most cases, this is complicated and
the integration can be carried out numerically.
As the shape functions used are generally powers of the coordinates x and y, the mul-
tiples are powers of the coordinates as well. Integration of such functions can be performed
exactly using Gaussian numerical integration, and so this is a popular means of obtaining
the integrals when forming the finite element matrices.
Gaussian integration is based on fitting polynomials to the curve of the function being
integrated and finding the area under that curve. If the function being integrated is the same
order of polynomial, the area will be exact, but if it is a different polynomial, say one of
lesser order, then the integration is approximate. Lower order Gaussian integration is some-
times used to obtain a better performance of elements, for instance eight-node isoparametric
elements can sometimes ‘lock’ when used for plasticity problems.
The area A under a function f(x) (or its integral) is given by
n

A = ∑ ω f (g )
i i (3.72)
i =1

where ωi are the Gaussian weights, gi are the Gaussian coordinates, and n is the number
of Gauss points. For higher order schemes, there are more Gauss points used. These values
Finite element methods  51

Table 3.2  G
 aussian weights and coordinates
for a three-point rule
Weight ωi Gauss coordinate gi
−0.77459 66692 0.55555 55555
0.0 0.88888 88888
+0.77459 66692 0.55555 55555

are given in tables that are provided in various books (see Zienkiewicz 1977) for different
integration schemes. The coordinates are generally provided so that they range between
−1 and +1; for instance for a three-point scheme, the values in Table 3.2 may be used.
To integrate between a and b as shown in Figure 3.12, we can make a transformation
xi = (a + b)/2 + (b − a)gi /2.
If we wish to integrate over an area as in a two-dimensional problem we would need to
integrate with respect to x and y, but we can relate the real coordinates to the local coordi-
nates and perform the integration as follows:

+1 +1

I =
∫ ∫ f (ξ, η)|J | dξd η (3.73)
−1 −1

where |J| is the determinate of the Jacobian matrix (Equation 3.12). For three-dimensional
problems, we have to integrate over three local coordinates and the Jacobian is a 3 × 3
matrix.
This is shown in Figure 3.12 for a three-point Gauss scheme. It may be seen that the Gauss
points lie between the integration limits of a and b.
For an eight-node isoparametric element, the integration needs to be performed in both
the x and y directions, and so the Gauss points run in both directions. This is shown in
Figure 3.13 where a 3 × 3 Gaussian integration scheme is used.
For triangular elements, the Gauss points are located at prescribed locations within the
triangle. The Gauss points are normally given in terms of the ‘area coordinates’ of the ele-
ment. There are three area coordinates that define a point within the element, and they each
range from 0 to 1, and the sum of the coordinates is 1. At a corner of the triangle, one area
coordinate will be 1 and the others 0. At the centroid of an element, they are all 1/3.

f (x)

w1 w2 w3

a x1 x2 x3 b

Figure 3.12 A three-point Gaussian integration scheme.


52  Geomechanics in soil, rock, and environmental engineering

x x x

x x x
Nodes

x x x

Gauss points

Figure 3.13 An eight-node isoparametric element with a 3 × 3 Gauss point scheme.

3.7  ELASTIC–PERFECTLY PLASTIC MODELS

When the stresses in a soil or rock reach a combination that will cause failure, the stresses
are said to lie on the failure surface for that material, and the soil/rock will then undergo
plastic behaviour. Within the failure surface, the material is assumed to behave as if elastic,
but once the stress state reaches the failure surface, the stress increments have to stay on the
surface (and run along it) as the stress state cannot go outside the surface.
If unloading occurs, the stress state can move inside the failure surface and the material
will behave like an elastic material once again.

3.7.1 Formulation
When the stress state in the soil is within the failure surface (i.e. f(σ) < 0 where f is the func-
tion for the failure surface), then the material is assumed to behave in an elastic manner and
the stress–strain relationship may be written as

σ = DEε (3.74)

where DE is the stress–strain matrix for an elastic material.


Once the stress state is such that the stresses reach the failure surface, the material becomes
plastic and the strain rate can be expressed as

ε = ε E + ε p (3.75)

The total strain rate ε is made up of the elastic ε E and plastic ε p strain rates where the
elastic strain rate is given by

ε E = D −1σ (3.76)

During plastic flow, the stresses must remain on the failure surface f = 0 and so any change
in the stress state must be such that

df = 0 (3.77)

This means that we can write

∂f
df = dσ = 0 (3.78)
∂σ
Finite element methods  53

or

sT σ = 0 (3.79)

where

 ∂f ∂f ∂f 
sT =  , ,… (3.80)
 ∂σ 1 ∂σ 2 ∂σ n 

In Equation 3.79, the dot product of the two vectors is zero, so this indicates that the vector
 and as the increment of stress is along the
sT is perpendicular to the increment in stress, σ,
failure surface, the vector sT must be perpendicular to the yield surface.
The plastic material is assumed to have a flow rule such that the plastic strain rate vector
is given by

ε p = λr (3.81)

In the above equation, r is a vector giving the direction of the plastic strain increment, and
λ is an unknown scalar multiplier.
Equation 3.75 then becomes

ε = DE−1σ + λr (3.82)

Rearranging gives

σ = DEε − λDEr (3.83)

We can now multiply by the vector sT and noting the result of Equation 3.79, we have

0 = sT DEε − λsT DEr (3.84)

and solving for the unknown λ, we have

sT DEε
λ = (3.85)
sT DEr

Substituting the value of λ into the stress–strain law of Equation 3.83 results in

σ = DI ε (3.86)

where the incremental plasticity matrix DI is given by

αβT
DI = DE − (3.87)
αT s
54  Geomechanics in soil, rock, and environmental engineering

and where

α = DE r
β = DE s

3.7.2  Examples for a specific failure surface: Mohr–Coulomb


In the general derivation of the incremental stress–strain relationship in the previous sec-
tion, the equation for the failure surface f could be any function of the stresses. One failure
criterion that is of interest in soil and rock mechanics is the Mohr–Coulomb failure surface.
For a two-dimensional problem involving plane strain, we can write the failure criterion (for
the failure line as shown in Figure 3.14) simply as

f = (σ x − σ y )2 + 4τ 2xy − sin2 φ(σ x + σ y + 2c cot φ)2 (3.88)


We can therefore differentiate this equation for the failure surface with respect to the
stresses to obtain the s vector.

 ∂f 
 
 ∂σ x  2(σ x − σ y ) − 2 sin2 φ(σ x − σ y + 2c cot φ)
 ∂f   2 
s =   = 2(σ y − σ x ) − 2 sin φ(σ x − σ y + 2c cot φ) (3.89)
 ∂σ y   8τ xy 
 ∂f   
 
 ∂τ xy 

The vector r is the direction of the incremental plastic strain. This vector was traditionally
made perpendicular to the yield surface in the plasticity of metals so that no plastic work
was done. This is called an associated flow rule, as in this case the r vector and the s vectors

Mohr–Coulomb failure φ
envelope

τxy

c σx σ
σy

Figure 3.14 Mohr–Coulomb failure criterion for a 2-D stress state.


Finite element methods  55

Velocity vectors

300.0

250.0

200.0

150.0

Scale
100.0

1.00E–01
50.0

0.0
Phi = 0
Psi = 0
–50.0 0.0 50.0 100.0 150.0 200.0 250.0 300.0 350.0 Uniform soil
Rigid strip footing

Figure 3.15 Velocity vectors computed for strip footing: ϕ = 0, ψ = 0.

are the same. However, for soils, the flow rule can be such that the r vector is not perpen-
dicular to the failure surface, and in this case, we have

 ∂g 
 
 ∂σ x  2(σ x − σ y ) − 2 sin2 ψ(σ x − σ y + 2c cot ψ)
 ∂g   2 
r =   = 2(σ y − σ x ) − 2 sin ψ(σ x − σ y + 2c cot ψ) (3.90)
 ∂σ y   8τ xy 
 ∂g   
 
 ∂τ xy 

where g is called the plastic potential of the soil. This is a surface like the failure surface, and
the vector r will be perpendicular to it. The angle ψ is the dilation angle of the material and
can be seen to be equivalent to ϕ in Equation 3.89. Figure 3.15 shows the velocity vectors
calculated for the case of ϕ = 0, ψ = 0.

3.8  WORK HARDENING MODELS

With simple elastic-perfectly plastic models like those discussed in Section 3.7, the soil is
elastic and then suddenly becomes plastic once the stress state at a point reaches the failure
56  Geomechanics in soil, rock, and environmental engineering

surface. For most soils and rocks, this is not what is observed, as there is a non-linear behav-
iour before final failure.
To account for this, models have been proposed where the soil can yield before it fails, and
these models are generally classed as work hardening (or softening) models.
Unlike Equation 3.77, for a work hardening material there can be a change in the location
of the yield surface as work hardening occurs, that is,

∂f ∂f
df = dσ + dh = 0 (3.91)
∂σ ∂h

where h is the hardening parameter. It could be a change in some parameter such as the
plastic volume strain. In this case,

h = εVp = Hd ε p (3.92)

where H = (1, 1, 1, 0, 0, 0)T for a general three-dimensional case.


If c = ∂f/∂h, Equation 3.91 can be written as

sT σ + ch = 0
(3.93)
sT DEε − λsT DEr + ch = 0

and so solving for λ gives

sT DE ε
λ = T
(3.94)
s DEr − cHr

Hence,

 sT DEε 
σ = DEε −  T ⋅ DEr = DI ε (3.95)
 s DEr − cHr 

and now the incremental plasticity matrix DI is given by

 αβT 
DI = DE −  T (3.96)
 α s − cHr 

The Modified Cam Clay Model is a hardening model that uses the plastic volume strain
as the hardening parameter. This is discussed in the following, Section 3.9.

3.9 EFFECTIVE STRESS ANALYSIS USING


CAM CLAY TYPE MODELS

It is very important in using constitutive models with finite element analysis to under-
stand how they work as this may influence the results. One such model that is worthwhile
Finite element methods  57

examining is the Cam Clay Model which is often used to simulate soft clay behaviour.
Similar models are often implemented in finite element codes.
Soils consist of solids (the soil particles or skeleton), with liquid and air in the voids. If the
soil has a low permeability (i.e. a clay or a silty clay), the fluid cannot flow out of the soil
rapidly when the soil is loaded, and the soil behaves as an undrained material. If the soil is
saturated, it will deform at constant volume, as the water in the pores is incompressible, and
so only shear deformations can occur.
If a soil is loaded slowly, the fluid in the pores has time to flow from regions of high excess
pore pressure to regions of low excess pore pressure, and the soil behaves as a drained material.
At intermediate rates of loading, the behaviour is somewhere between drained and undrained.
The behaviour of a soil therefore is dependent on the loading rate as well as the permeability.
Figure 3.16 shows the results of triaxial tests on a soil consolidated isotropically, that is,
σ1′ = σ ′2 = σ ′3.
The stress path is plotted using the mean effective stress p′ = (σ1′ + σ ′2 + σ ′3)/ 3 and
the deviator stress q′ = σ1′ − σ ′3. During drained loading, the stress path will follow the
straight broken line (which will be at a slope of 3 vert:1 horiz) until it hits the failure
line. If the soil is loaded undrained, it will follow a curved path as shown in Figure 3.16
until it hits the failure line. This is because pore pressure u is generated during loading,
and the difference between the drained path and the undrained path is u. This is because
p = (σ1′ + u + σ ′2 + u + σ ′3 + u)/ 3 = p′ + u.
The failure line in a plot such as the one in Figure 3.16 is the usual Mohr–Coulomb failure
line, but has a slope given by M, where
6 sin φ′
M = (3.97)
3 − sin φ′

and ϕ′ is the angle of shearing resistance of the soil.

Failure line
q′

Partially
drained Drained

Excess pore
2su pressure = u

Undrained
p′
Isotropic
consolidation
σ′1

σ′3

σ′2

Figure 3.16 Triaxial test stress paths.


58  Geomechanics in soil, rock, and environmental engineering

If the soil is partially drained, the stress path will follow the curved broken line shown in
Figure 3.16.
For undrained loading, the point at which the stress path intersects the failure line gives
twice the undrained shear strength of the soil, as the undrained shear strength su is given by

σ1 − σ 3 (σ ′ − σ ′3) q′
su = = 1 = (3.98)
2 2 2

3.9.1  Normally consolidated clay


The effective stress within a soil will increase with depth due to the overburden pressure.
The lateral stress is related to the vertical stress by the earth pressure coefficient K0′ . This
effective stress variation with depth is shown in Figure 3.17, where the vertical effective
stress and the lateral effective stress are seen to increase with depth. At any point in the soil,
we therefore have

σ ′v + 2K0′ σ ′v (1 + 2K0′ )σ ′v
p′ = =
3 3 (3.99)
q′ = σ ′v − K ′ 0 σ ′v = (1 − K ′ 0 )σ ′v

If we plot the initial stress points for soil at two depths A and B, then these plot as points
A′ and B′ as shown in Figure 3.18a. The points lie on what is called the K0 line as it is the
line that the initial stress states lie on under K0 conditions (or one-dimensional consolidation
conditions).
The void ratio e at point B would be less than at point A because the soil is under greater
stress (i.e. it is deeper), and so it would plot as point B′ as shown in Figure 3.18b which is a
plot of the natural logarithm of p′ versus the specific volume υ = 1 + e. This is how the void
ratio versus mean effective stress is conventionally plotted in critical state soil mechanics,
but is equivalent to the conventional e – log10 σ ′v plots made for oedometer tests.
In Figure 3.18b, the isotropic consolidation line (which is the line the stresses in the soil
would follow if the soil is consolidated under vertical and lateral stresses that are all equal)
is shown along with the critical state line (CSL). The CSL is the line showing the void ratio–
mean effective stress relationship at failure.

Effective stress

Point A
Depth z

σ′v

Point B

σ′h = K ′0 σv

Figure 3.17 Effective stress variation with depth in a normally consolidated soil.


Finite element methods  59

(a)
Failure line
q′

B
2suB
K0 line
A
2suA
B′

A′

p′

(b) Isotropic
1+e consolidation line
A A′

B B′

Critical state line

ln(p′)

Figure 3.18 (a) Undrained stress paths to failure at different depths in a soil. (b) Void ratio during undrained
loading.

Upon undrained loading, the stress path at the point A in the soil would be from A′ to A,
and at point B the stress path would be from B′ to B. The void ratio would remain constant
as the soil is undrained (Figure 3.18b) and so the path in this figure is horizontal (because
there is no change in void ratio) until it hits the critical state line at failure.
We can see therefore that the undrained strength depends on the initial stress at a point
in the ground and the path taken to failure. At point B in the ground, the stresses are higher
and so the stress path leads to a higher undrained shear strength at failure at B (suB) than for
point A that follows a path to A (suA). This is what is observed in reality, that the undrained
shear strength in a soil deposit will normally increase with depth.

3.9.2  Overconsolidated clay


If a clay is overconsolidated, which means that it has been loaded in the past to a greater
pressure than that it exists at a point in the ground at present, then the stress path taken by
the soil in the ground will be as shown in Figure 3.19a.
During undrained loading, the volumetric behaviour is as shown in Figure 3.19b, where
once again the void ratio remains constant during loading and the path followed is such that
point A finishes on the critical state line. The path is now such that the CSL is approached
from the left instead of from the right.
This time the undrained shear strength is higher at the same point in ground A, because
of the different stress path taken on loading. The undrained shear strength can be seen to be
different because the stress path hits the failure surface at a greater value of q′.
Hence, the undrained shear strength of a soil depends not only on the initial stress state
but also on whether it is overconsolidated or not (and the degree of overconsolidation).
60  Geomechanics in soil, rock, and environmental engineering

(a)
Failure line
q′

K0 line
A
2suA

A′

p′

(b)
Isotropic
1+e consolidation line
A′ A

Critical state line

ln(p′)

Figure 3.19 (a) Undrained stress path for an overconsolidated soil. (b) Volumetric behaviour of an
­overconsolidated soil during undrained loading.

3.10  CAM CLAY TYPE MODELS

In order to try to simulate the behaviour of real soils described in the previous sections,
models have been developed that can be put into finite element codes, and one such model is
the Modified Cam Clay Model (Roscoe and Burland 1968) so called because it was a modi-
fication of the original Cam Clay Model proposed by Roscoe and Schofield (1963).
The features of this model are that it can simulate the yield of a soil and the increased
compressibility of the soil once it has yielded. The model is based on effective stress analysis,
and therefore the excess pore pressures generated are important as they affect the stress path
as seen in the previous section.

3.10.1  Cam Clay yield surface


With the Modified Cam Clay Model for soft soils, there is both a yield surface and a fail-
ure surface. The yield surface has been noted from triaxial tests on soft clays for which the
model was developed, and can be seen to be an ellipse in Figure 3.20.
The model is a work hardening model such that the ellipse expands when the stress path
reaches the yield ellipse (see the dotted line in Figure 3.20). If the stress path goes back inside
the ellipse, the elliptical surface stays at its maximum size (i.e. the maximum p0′ ). In this
way, the pre-consolidation pressure that is exhibited by soils can be modelled. The stress
path has to reach the new location of the yield surface before it becomes plastic again.
The equation for the yield surface ellipse is
M 2 (p′ 2 − p′ p0′ ) + q2 = 0 (3.100)

Finite element methods  61

q = (σ′1 – σ′3)

M = 6 sinϕ′/(3 – sinϕ′)
q = M . p′

New location of
yield surface

p′0/2
p′ = (σ′1 + 2σ′3)/3
p′0

Figure 3.20 Modified Cam Clay Model showing a yield ellipse and failure line q = M · p′.

Eventually, the stresses will reach the failure surface, which is given by the equation

q = M ⋅ p′ (3.101)

and the stress state must then stay on this failure surface.
For a work hardening plastic material, we can write the incremental stress–strain matrix
as in Equation 3.93, and for the Modified Cam Clay Model, the hardening parameter h is
the plastic volume strain. We can therefore write for a two-dimensional problem

∂f ∂f ∂p0′  1 + e0 
c = = = M 2 p′  p0′ (3.102)
∂h ∂p0′ ∂h  λ − κ 

The parameters λ and κ are the slopes of the normal and recompression lines if we plot the
specific volume υ = (1 + e) versus p′ as shown in Figure 3.21 (e is void ratio).
In order to work out the s vector, the function for the yield surface has to be differenti-
ated with respect to the stresses. For a two-dimensional plane strain problem, there are four
stress components σx, σy, σz, τxy as we have to take into account the stress in the zero strain
direction σz.

Normal
compression line

Slope = λ
υ

Slope = κ

Critical state line

ln p′

Figure 3.21 Volumetric behaviour. Specific volume υ = 1 + e versus ln p′ plot.


62  Geomechanics in soil, rock, and environmental engineering

∂f ∂f ∂p′ ∂f ∂q
s = = +
∂σ ∂p′ ∂σ ∂q ∂σ
 
1
   
3  
 1  3 σ ′x − p′  (3.103)
= M 2 (2 p′ − p0′ ) ⋅   + 2q ⋅  σ ′y − p′ 
 3 2 q  
1  σ ′z − p′ 
3  2τ xy 
 
 
 0 

In the Modified Cam Clay Model, the plastic potential g is the same as the yield surface,
that is, it has an associated flow rule.
While the stresses are inside the yield surface, they deform along the recompression line
that has a slope of κ in the plot of specific volume υ = 1 + e versus ln p′. Once the yield sur-
face has been reached (the ellipse in Figure 3.22a), the soil yields. If we are moving along the
x-axis or along the hydrostatic compression line, once yield occurs, we would move along
a line that has a slope of λ as shown in Figure 3.22b. This is what is observed for real soils
where after yielding, the soil is more compressible. If we were to perform a one-dimensional
compression test such as an oedometer test, the slope of the compression line would also

(a)
q = (σ′1 – σ′3)

q = M . p′
Yield surface at
A′
failure

X
p′0/2
A p′ = (σ 1′ + 2σ′3)/3
p′0

(b)
Normal
compression line

υ=1+e A Slope = λ
X
Slope = κ
A′
Critical state line

ln p′

Figure 3.22 (a) Stress path taken during drained loading showing an expanding yield surface. (b) Volumetric
behaviour during the drained loading path of (a).
Finite element methods  63

be  λ after the pre-consolidation pressure is reached. The relationship between the more
familiar parameters Cc (the compression index) and C R (the recompression index) obtained
from an oedometer test is given by

Cc
λ =
2 .3
(3.104)
C
κ = R
2 .3

Suppose we perform a drained triaxial test on a slightly overconsolidated soil from an


initial stress state at point A in Figure 3.22a. The stress state would hit the yield surface at
the point X and then the yield surface would move with the stress path and expand as load-
ing occurs.
If unloading occurs, the soil returns to being ‘elastic’ inside the yield surface. This allows
overconsolidation to be modelled, as the soil will not now yield on reloading until it once
again hits the (expanded) yield surface. The stress state will eventually hit the failure surface
at point A′.
In terms of the volumetric behaviour, the soil will follow the path shown in Figure 3.22b
where it can be seen to yield at point X before continuing on to failure at point A′ on the
critical state line.
If an undrained analysis is performed on the same clay starting from the same point on
the p′ – q′ plot (i.e. point A), the stress path will move vertically upward hitting the yield
surface at point X. It will then move to the left intersecting the failure surface at point A′ as
shown in Figure 3.23a. The volumetric behaviour is shown in Figure 3.23b where it can be
seen that there is no volume change as the loading takes place and so the path from point A
to A′ is horizontal.
Hence, the success of an effective stress analysis considering pore pressure will depend on

1. Having the correct initial stress state (which depends on K0′ )


2. Choosing the correct permeabilities kv and kh and the correct boundary conditions as
the excess pore pressures control the effective stress path
3. Having the correct loading rate
4. Having the correct overconsolidation (OCR) ratios
5. Obviously, the soil strength ϕ′ parameter (which is used to compute the slope of the
failure line M) and the soil deformation parameters λ, κ play an important role as well

The undrained behaviour of a clay in a numerical analysis will be such that the soil
behaves according to the model, and for a Cam Clay Model, the yield surface is an ellipse
with the top of the ellipse on the critical state line. Other shaped ellipses are possible to get
different soil behaviours.

3.11  UNDRAINED ANALYSIS

In conventional slip circle analysis or in an undrained analysis using a finite element pro-
gram (say), the soil strength is treated as having undrained strength parameters su and ϕu = 0.
As we have seen, the undrained strength su is a function of the initial stress state and the
­overconsolidation ratio.
64  Geomechanics in soil, rock, and environmental engineering

(a)
q = (σ′1 – σ′3 )

q = M . p′
Yield surface at
A′ failure

X
p′0/2
A p′ = (σ′1 + 2σ′3 )/3
p′0

(b)
Normal
compression line

X Slope = λ
A′ A
υ=1+e

Critical state line

ln p′

Figure 3.23 (a) Stress path taken during undrained loading from the K0 line (slightly O/C soil). (b) Volumetric
behaviour during undrained loading from the K0 position.

According to Cam Clay theory (Wroth 1984), the undrained shear strength of a clay
under plane strain conditions (with plain strain angle of friction ϕPS) is given by
Λ
 suPS  sin(φPS )  c 2 + 1 
  = (3.105)
 σ v′  2c  2 

where Λ = (1 − Cr/Cc), Cc is the virgin compression index and C r is the recompression index
of the clay, and c = 1/(1 − 2 sin(ϕPS)). If we take typical soil parameters such as ϕ′ ≈ 26–27°
and Λ in the range 0.7–0.8 for clays of low to medium sensitivity, then this gives
 su  0.8
 σ ′  = 0.27 OCR (3.106)
 v

Data based on back analysis and field observation have led to empirical equations such as

suN /C ≈ 0.22σ v′
suO /C
= (OCR)0.8
suN /C (3.107)
 suO /C  0.8
 σ ′  = 0.22 OCR
v

Finite element methods  65

5
Clays

(su/σ′v0)OC
(su/σ′v0)NC
3

Varved
clay
2

1
1 2 5 10
OCR

Figure 3.24 Ratio of undrained shear strength for an overconsolidated clay to undrained shear strength
for normally consolidated clay versus OCR. (From Ladd C.C. et al. 1977. Proceedings of the 9th
International Conference on Soil Mechanics and Foundation Engineering, Tokyo, Vol. 2, pp. 421–494.)

where the subscript N/C denotes the normally consolidated value of undrained shear strength
and O/C denotes the overconsolidated value. These formulae are similar to the theoretical
equation (Equation 3.106), from the Cam Clay Model that gives a slightly higher value of su.
The su values change depending on soil type, stress paths, and initial stress states, but
the empirical equations (Equation 3.107) apply in many practical situations. A plot of the
undrained shear strength ratios versus overconsolidation ratio OCR is shown in Figure 3.24
(Ladd et al. 1977).

3.12  FINITE ELEMENT ANALYSIS

If we use a finite element analysis (Britto and Gunn 1987) for an undrained problem of
loading of a soft clay, then we would start off with a stress state at point X in Figure 3.25
if the soil is normally consolidated and the stress state would lie on the yield surface. Upon
loading, the stress would follow a path to point A′ on the critical state line. The undrained
shear strength is then half of the q′ value at point A′.
If we take the angle of shearing resistance of the clay as ϕ′ = 27° and assume
K0′ = 1 − sin φ′ = 0.546, then we have an initial stress state (at point X in Figure 3.25) of

q′ = σ v′ (1 − K ′0 ) = 0.454σ v′
(1 − 2K ′0 ) (3.108)
p′ = σ v = 0.697 σ v′
3

6 sin φ′
M = = 1.07 (3.109)
3 − sin φ′
66  Geomechanics in soil, rock, and environmental engineering

q′ = (σ′1 – σ′3)

q′ = M . p′

su = q′/2 Yield surface at


A′ failure
0.454σ′v
X

q′ = 0.454σ′v p′ = (σ′1 + 2σ′3)/3

p′ = 0.697σ′v p′0

Figure 3.25 Stress path taken by normally consolidated clay under undrained loading in a Cam Clay Model.

The actual stress path is shown in the finite element results of Figure 3.26. In this case,
su ≈ 13.6 kPa when σ ′v = 50 kPa or su = 0.27 σ ′v as predicted by the Cam Clay formula of
Equation 3.106. This is slightly higher than the values usually measured for normally con-
solidated clays that are observed to be about 0.22 σ ′v as we saw in the previous section
(ratio = 0.27/0.22 = 1.23).
This may be critical in the analysis of embankments being constructed on soft clays, as
the predicted height to failure may be higher than it should be. In cases such as this, it is best
to use a slip circle analysis with the variation of undrained strength with depth as a check
of the collapse height. This of course can only be done if the distribution of the undrained
strength has been measured. If the embankment loading is such that it is kept constant for
a period of time to allow consolidation, then the effective stresses and the undrained shear
strengths will increase and the initial distribution of undrained shear strengths cannot be
used in a slip circle analysis.

50

40
Deviator stress q (kPa)

Stress path
30

20

10

0 10 20 30 40 50
Mean stress p (kPa)

Figure 3.26 Stress path taken under plane strain conditions.


Finite element methods  67

3.12.1 Examples
Some examples of the finite element implementation of the Cam Clay Model are shown in
Figures 3.27 through 3.29. These show the behaviour of a clay specimen tested in triaxial
compression in either the drained or the undrained condition.
Figure 3.28 shows the behaviour of a lightly overconsolidated soil in an undrained tri-
axial test. The stress path (Figure 3.28b) goes vertically upward until it hits the yield locus
and then it turns towards the critical state line, where failure occurs. The load–deflection
curve (Figure 3.28a) shows the failure at a deviator stress of about 35 kPa.
Figure 3.29 shows the results of an undrained triaxial test on a heavily overconsolidated
clay. The stress path once again goes vertically up until it hits the yield surface (passing
above the critical state line). It then turns towards the CSL and ends on it at failure. Figure
3.29a shows the load–deflection curve for this case, where it may be seen that strain soften-
ing occurs after yielding. The final deviator stress at failure is about 31 kPa.
More on Critical State Soil Mechanics can be found in texts by Atkinson and Bransby
(1978) and Schofield and Wroth (1968).

(a) 100 (b)


80
80
Applied pressure (kPa)
Deviator stress q (kPa)

Critical state
line 60
60

40
40 Unload–reload

20
20

0 0
0 20 40 60 80 100 0 0.4 0.8 1.2 1.6
Mean stress p (kPa) Deflection (m)

(c) 1.85

1.80

1.75
Void ratio

1.70
Unload–reload
1.65

1.60

1.55
50 55 60 65 70 75 80
log (p)

Figure 3.27 A drained triaxial compression test: (a) stress path in p – q space (finite element result);
(b) load–deflection; (c) e – log p′. Cam Clay Model showing load–unload behaviour.
68  Geomechanics in soil, rock, and environmental engineering

(a) 40 (b) 70

60
30

Deviator stress q (kPa)


Applied pressure (kPa)

50

40 Stress path
20
30

20
10
10

0
0 0.04 0.08 0.12 0.16 0 10 20 30 40 50 60 70
Deflection (m) Mean stress p (kPa)

Figure 3.28 An undrained triaxial test on lightly overconsolidated soil: (a) load–deflection; (b) p – q plot.

(a) 70 (b) 35

60 30
Applied pressure (kPa)
Deviator stress q (kPa)

50 25
Stress path
40 20

30 15

20 10

10 5

0 0
0 10 20 30 40 50 60 70 0 0.05 0.1 0.15 0.2 0.25
Mean stress p (kPa) Deflection (m)

Figure 3.29 An undrained triaxial test on heavily overconsolidated soil: (a) load–deflection; (b) p – q plot.

APPENDIX 3A: SHAPE AND MAPPING FUNCTIONS


FOR VARIOUS ELEMENT TYPES

Shape functions for the six-node triangle


t = 1.0
5

t = 0.5
6 4
s=0
s = 0.5 s = 1.0

1 2 3
t=0
Finite element methods  69

Area coordinates are s, t, u = (1 − s − t)

N1 = (1 − 2s − 2t)(u)
N2 = 4(u)s
N3 = (2s − 1)s
N4 = 4st
N5 = (2t − 1)t
N6 = 4(u)t

Shape functions for the eight-node isoparametric element

7 6 5

8 4
ξ

1 2 3

N1 = 0.25(1 − ξ)(1 − η)(− η − ξ − 1)

N 2 = 0.5(1 − ξ 2 )(1 − η)

N3 = 0.25(1 + ξ)(1 − η))(ξ − η − 1)

N 4 = 0.5(1 − η2 )(1 + ξ)

N5 = 0.25(1 + ξ)(1 + η)(η + ξ − 1)

N6 = 0.5(1 − ξ 2 )(1 + η)
N7 = 0.25(1 − ξ)(1 + η)(−ξ + η − 1)

N8 = 0.5(1 − η2 )(1 − ξ)

Mapping functions for five-node infinite element with ξ−2 decay

3
η

ξ
4
ξ = +1 at ∞
5

1 2
70  Geomechanics in soil, rock, and environmental engineering

2ξη(1 − η)
M1 =
(1 − ξ)2
0.5(1 + ξ)2 (1 − η)
M2 =
(1 − ξ)2
0.5(1 + ξ)2 (1 + η)
M3 =
1 − ξ)2
(1
2ξη(1 + η)
M4 = −
(1 − ξ)2
4ξ(1 − η2 )
M5 = −
(1 − ξ)2

Shape functions for 15-node triangle

10 8

15 7
11

12 6
13 14

1 2 3 4 5

Area coordinates are s, t, u = (1 − s − t)

2
c1 = 10
3
c2 = 4c1

c3 = 64
c4 = 128
t1 = s − 0.25
t 2 = s − 0 .5
t3 = s − 0.75
t4 = t − 0.25
t 5 = t − 0 .5
t6 = t − 0.75
t7 = u − 0.25
t8 = u − 0.5

t9 = u − 0.75
Finite element methods  71

N1 = c1 ⋅ s ⋅ t1t2t3
N 2 = c2 ⋅ s ⋅ t ⋅ t1t2
N3 = c3 ⋅ s ⋅ t ⋅ t1t4
N 4 = c2 ⋅ s ⋅ t ⋅ t4t5
N5 = c1 ⋅ t ⋅ t4 ⋅ t5t6
N6 = c2 ⋅ t ⋅ u ⋅ t4t5
N7 = c3 ⋅ t ⋅ u ⋅ t4t7
N8 = c2 ⋅ t ⋅ u ⋅ t7t8
N9 = c1 ⋅ u ⋅ t7t8t9
N10 = c2 ⋅ s ⋅ u ⋅ t7t8
N11 = c3 ⋅ s ⋅ u ⋅ t1t7
N12 = c2 ⋅ s ⋅ u ⋅ t1t2
N13 = c4 ⋅ s ⋅ t ⋅ u ⋅ t1
N14 = c4 ⋅ s ⋅ t ⋅ u ⋅ t4

N15 = c4 ⋅ s ⋅ t ⋅ u ⋅ t7

Shape and mapping functions for a 12-node infinite element


u = +1

t = +1

10
11 12
9
8 ∞
7
3 ∞
6
s = −1
4 2 t = −1
5
1
u = −1

Mapping functions:

−0.5s(1 − u)(1 − t)(−1 − t − u)


M1 =
(1 − s)
0.25(1 − t)(1 − u)(1 + s)
M2 =
(1 − s)
0.25(1 + t)(1 − u)(1 + s)
M3 =
(1 − s)
−0.5s(1 − u)(1 + t)(−1 + t − u)
M4 =
(1 − s)
72  Geomechanics in soil, rock, and environmental engineering

− s(1 − t 2 )(1 − u)
M5 =
(1 − s)
− s(1 − u2 )(1 − t)
M6 =
(1 − s)
− s(1 − u2 )(1 + t)
M7 =
(1 − s)
−0.5s(1 − u)(1 − t)(−1 − t + u)
M8 =
(1 − s)
0.25(1 − t)(1 + u)(1 + s))
M9 =
(1 − s)
0.25(1 + t)(1 + u)(1 + s)
M10 =
(1 − s)

−0.5s(1 + u)(1 + t)(−1 + t + u)
M11 =
(1 − s)
− s(1 − t 2 )(1 + u)
M12 =
(1 − s)

The shape functions are

(i) Mid-side nodes


N 2 = 0.25(1 − s 2 )(1 − t)(1 − u)
N3 = 0.25(1 − s2 )(1 + t)(1 − u)
N5 = 0.25(1 − t 2 )(1 − s)(1 − u)
N6 = 0.25(1 − u2 )(1 − s)(1 − t)
N7 = 0.25(1 − u2 )(1 − s)(1 − t)
N9 = 0.25(1 − s2 )(1 − t)(1 + u)
N10 = 0.25(1 − s 2 )(1 + t)(1 + u)

N12 = 0.25(1 − t 2 )(1 − s)(1 + u)

(ii) Corner nodes

N1 = 0.125(1 − s)(1 − t)(1 − u) − 0.5(N 2 + N 5 + N6 )


N4 = 0.125(1 − s)(1 + t)(1 − u) − 0.5(N3 + N 5 + N7 )
N8 = 0.125(1 − s)(1 − t)(1 + u) − 0.5(N6 + N9 + N12 )
N11 = 0.125(1 − s)(1 + t)(1 − u) − 0.5(N7 + N10 + N12 )

The derivatives of the shape functions are then computed in the usual way (see Equation
3.13) but where derivatives are found for x, y, and z.
∂N i ∂N i ∂s ∂N i ∂t ∂N i ∂u
= + +
∂x ∂s ∂x ∂t ∂x ∂u ∂x
Finite element methods  73

and since

 x1 
x 
x = [M1, M2 ,… , M12 ]  
2

  
 
 x12 

we can find ∂x/∂s, ∂x/∂t, ∂x/∂u, and obtain the inverse derivatives by forming the Jacobian
matrix and inverting it as explained in Section 3.2.3.

APPENDIX 3B: GLOBAL MATRIX ASSEMBLY


AND BOUNDARY CONDITIONS

Once the element stiffness matrix is formed by multiplying element matrices together and
numerically integrating the result, the element matrices need to be assembled into a global
matrix. This is done by assigning a number to each node in the mesh. If the node has only
one variable associated with it, for instance in flow and seepage problems, then the vari-
able can be assigned the same number as the node. For instance, node 22 has unknown h 22
(total head) at the node in a seepage problem. If it has three variables per node as in two-
dimensional consolidation problems (the horizontal u and vertical w displacement and the
total head h or pore pressure), then the node has three unknowns. Again if we take node
22, the variables in the global matrix are 3 × 22 − 2 for u, 3 × 22 − 1 for w, and 3 × 22 for h.
If, for example, we had an element with nodes 6, 13, 22, then for a problem with three
variables per node we would have the following correspondence between the positions in the
element matrix and the global matrix.

Element position Global position


1 16
2 17
3 18
4 37
5 → 38
6 39
7 64
8 65
9 66

So, an entry in position 9,9 in the element matrix becomes entry 66,66 in the global
matrix, and entry 3,9 in the element matrix becomes entry 18,66 in the global matrix.
It can be seen that the further apart the node numbers are for the element, the further
apart the locations in the global matrix are. The bandwidth of the global matrix depends on
how far apart the element numbers are.
From a 12-node infinite element, we can see that the element matrices will add into
the global matrix along the diagonal, and provided the numbering of the mesh is done
74  Geomechanics in soil, rock, and environmental engineering

judiciously, the numbers all fall within a banded region down the diagonal of the global
matrix, and the width of this band is called the bandwidth. The half width of the banded
region is called the half-bandwidth of the global matrix.
Often problems are such that the global matrix is symmetric, and so we only need to
store the upper half of the matrix. In addition, we can store only the upper half of the band
because outside of the band all values are zero. This is often done in finite element work
to increase the efficiency of storage. A more efficient method called ‘skyline’ storage is also
used where each row in the upper half of the band is stored but only to its maximum length.
If zeros exist at the end of a row, they are not stored.
If the problem is not symmetric, the whole bandwidth has to be stored. This occurs for
seepage or consolidation problems where the permeability is anisotropic, or in plasticity
problems where the soil has a non-associated flow rule (see Section 3.7.2).

APPENDIX 3C: BOUNDARY CONDITIONS

To solve the global set of finite element equations for the unknowns at each node, we usually
have to apply some boundary conditions at the nodes of the finite element mesh. A common
condition is that a displacement is zero at a boundary or a total head is prescribed in a flow
problem.
If we know what a prescribed value for a variable is, then it is not an unknown for the
problem and we can multiply it by a column of the global matrix and subtract it from the
right-hand side of the equations.
For example, if we have a set of equations Aδ = f and we know that the 24th variable of
the solution is Q, then we can multiply the 24th column of the matrix by Q and subtract it
from the right-hand vector f as shown in Figure 3B.1. This can be done repeatedly if there

Half-bandwidth

Element matrices

Zeros

Zeros

Global matrix

Figure 3B.1 Placing element matrices into a global stiffness matrix.


Finite element methods  75

δ24 = f – ×Q
A

24

Figure 3B.2 Applying a prescribed boundary condition to a set of equations.

are more than one prescribed value. If the prescribed value is zero, then this obviously does
not need to be done, as zero will be subtracted. All subtractions must be done for each pre-
scribed variable before proceeding to the next step as described in the next paragraph, as the
matrix must not be modified while multiplying the columns by the known values.
However, it should be noticed that this makes the matrix non-symmetric, therefore one
way to make the matrix symmetric again is to set the row and column (in this example row
24 and column 24) equal to zero, and place a 1 on the diagonal for each prescribed bound-
ary condition. The prescribed value is entered at row 24 on the right-hand side so that there
is now an equation that states 1 × δ24 = Q. An alternate method is to delete all the rows and
columns for the prescribed variables, and reduce the size of the global matrix, but this is not
addressed here (Figure 3B.3).

Zeros

24 1 δ24 = Q

24

Figure 3B.3 Replacing row and column with zeros and inserting a unit value on the diagonal.
This page intentionally left blank
Chapter 4

Site investigation and in situ testing

4.1 INTRODUCTION

Most of the analytical and numerical methods that are presented in the following chapters
depend on obtaining the appropriate material properties for use in the analysis. The prop-
erties may be obtained from laboratory tests, but are more often found from field tests as
these can be performed in situ and are often a more cost-effective way of obtaining soil
properties.
The correlation of the field test result with the soil property often depends upon the type
of analysis being performed. For instance, some computer programs may need an initial
modulus, while for others a tangent modulus may be more suitable. Some may require a
small-strain modulus if this is appropriate.
In this chapter therefore, the most common field tests are described, and where applicable,
methods of obtaining the appropriate material property from the field test are presented. As
field tests are performed as part of a site investigation programme, common site investiga-
tion techniques are also described in this chapter.
The most common field tests are the standard penetration test (SPT), the dynamic cone
penetration test (DCP), the static cone penetration test (CPT), the pressuremeter test (PMT),
and the dilatometer test (DMT). Various standards outlining test methods have been devel-
oped in different countries and these should be referred to when performing the test as there
are often slight differences in test procedures among the different standards. The various
field test methods are described in the following sections.

4.2 EXPLORATION METHODS

For some in situ test methods, a hole has to be drilled or formed using wash boring, and the
test is conducted at the base of the borehole. Such tests are the standard penetration test, the
Ménard pressuremeter test, or seismic cross-hole tests.
For SPT tests, the base of the borehole needs to be cleaned (e.g. using a blank bit) so that
loose material that has fallen back into the hole or has been loosened by drilling does not
interfere with the test result.
For other in situ tests, no borehole is needed. For tests such as cone penetration tests, or
dilatometer tests, the device is pushed into the ground using hydraulic pressure. Seismic
refraction tests can be performed by simply striking the ground with a sledge hammer and
recording the seismic waves returning to the ground surface.

77
78  Geomechanics in soil, rock, and environmental engineering

4.3 SITE INVESTIGATION

In order to carry out a geotechnical design, it is necessary to acquire information about the
soil or rock properties pertinent to the type of design to be carried out and to obtain an idea
of the distribution of the various soil and rock types that exist at a particular site. A brief
overview is provided in the following of the most common techniques used to obtain site
and subsoil information.

4.4 OBJECT OF SITE INVESTIGATION

There may be a number of different reasons for performing a site investigation, and it is nec-
essary to establish what the object of the investigation is so that the appropriate information
can be collected when performing the work. Some reasons for performing an investigation
are listed as follows:

1. To determine if the site is suitable for the proposed works.


2. To enable the design of structures, earthworks and excavations.
3. To allow prediction of any difficulties arising during construction.
a. It may be necessary to check the bearing capacity of proposed foundations or to
determine settlements. In this case, it would be necessary to obtain soil samples
which could be tested or to carry out field tests as part of the investigation, so that
soil strengths and compressibilities are known. It would also be necessary to deter-
mine the soil types and their distributions.
b. If cuts are being made, the stability of the slopes would need to be considered.
c. If excavations are to be made, information would be required to enable the support
systems to be designed.
d. Dewatering may be required at the site, and if so, information on materials, their
permeabilities, and their distributions would be required.
4. To locate or determine the availability of construction materials, for example, rockfill,
clay, and filter materials for a dam.
5. To assess the effect of changes brought about by construction, for example, excavation
may affect adjacent sites; construction may alter runoff patterns.
6. To investigate soil/groundwater aggressiveness on buried structures.
7. Environmental and health issues may require an investigation. If the site is known to
be contaminated, the extent of the contamination may need to be determined. Today,
it is common for combined geotechnical/environmental investigations to be carried
out, as use of the same boreholes for both types of investigations leads to obvious cost
savings.

4.5 CATEGORY OF INVESTIGATION

Investigations may need to be performed on ‘green field’ sites (ones that have not been sub-
ject to previous construction or filling) or on ‘brown field’ sites (ones that have had some
form of previous development carried out). Investigations can be categorised as follows:

1. New Work
Examples of new work include
a. Buildings, bridges, roads, railways, airfields, port facilities, industrial plants
Site investigation and in situ testing  79

b. Dams
c. Mines (underground and open cut) and mine infrastructure
2. Existing Facilities
Existing facilities include
a. Pavement which may need repair or upgrade
b. Structures which need extensions or modifications, for example, adding more sto-
ries to an existing building
3. Failures and Remedial Work
Repair or remedial work may need to be carried out for
a. Slope instability or failures
b. Retaining wall collapses
c. Dams (which may suffer collapse or excessive leakage)
d. Structural damage due to swelling or shrinking soils, or subsidence due to mining
or withdrawal of groundwater

4.6 PLANNING AN INVESTIGATION

A good deal of time and money can be saved by making use of pre-existing information
about the site and planning the investigation before going into the field.
Some of the following activities may be useful when planning and executing an
investigation:

1. Define the aim of the investigation or of the problem to be solved.


2. Initial Office Study
a. For large areas, use can be made of topographical and geological maps or aerial
photographs. Internet-based satellite photographs are also commonly used. Stereo
pairs of photographs allow terrain types to be identified, for example, alluvial
flats, shallow slopes, and steep scarps. Large slips can also be identified. Old aerial
photographs of industrial areas can be used to determine previous positions of
buildings, waste storage, or dumping (fill) areas.
b. For smaller sites, information can be obtained from geological maps or from previ-
ous site investigations in the area. Most consulting firms which have been in exis-
tence for any length of time, have comprehensive records of past site investigations,
and investigations could have been previously carried out on adjacent sites or on
the same site. Geographical information systems (GIS) have been set up for some
areas and these contain data such as borehole, cone, and test data for all investiga-
tions in the area covered.
c. Maps of services (water, gas, electricity) need to be obtained from councils or other
government agencies.
  Excavating or drilling through such services will cause costly delays and possible
disruption to surrounding buildings.
3. Site Inspection
a. A preliminary reconnaissance is necessary to confirm features seen on maps and
aerial photographs.
b. At this stage, a more detailed investigation can be planned. The location of drill
holes or test pits can be marked on a map of the site. These are usually along the
line of proposed foundations (dig test pits to one side so as not to create settlement
problems later).
80  Geomechanics in soil, rock, and environmental engineering

  It may be more convenient to use other forms of site investigation than drilling/
excavating. At this stage, the decision may be made to use cone penetrometers or
to use geophysical methods (i.e. seismic refraction or resistivity methods).
c. It may be necessary to locate electricity supplies or water supplies for some equip-
ment which is to be used.
d. Buildings similar to those planned for the site may exist in the area. It is a good
idea to inspect these buildings for signs of damage, for example, cracking due to
swelling soils. Cuts in the area may be at slopes similar to those proposed and their
performance can be assessed.

4.7 PREPARING COST ESTIMATES FOR THE WORK

Some of the main costs in carrying out the site investigation include the following:

1. Cost of establishing plant, for example, may need to drive the drilling rig to country
areas and back
2. Cost of drilling: Drilling with augers and diamond drilling is charged by the metre;
excavators are charged by the hour
3. Cost of supervision by the engineer
4. Cost of laboratory testing
5. Cost of preparing the report

4.8 DETAILED EXPLORATION

1. The detailed exploration depends on the type of structure and the type of material
revealed during the exploration, for example, if soil is uniform, fewer holes need to be
bored.
2. The investigation needs to be taken to a depth where shearing of the soil due to the
applied structural loads is small, for example, to depths of at least 50% greater than
the width of the structure. However, if soft clay exists below this, then the investiga-
tion should be taken to a greater depth. For buildings, usually preliminary borings are
taken to rock (if it is at depth less than the required depth), and if no soft clay layers
are found, holes are taken to 6–10 m. For light industrial buildings, backhoe holes to
depths of 3–4 m are usually sufficient, while for tall buildings with deep pile founda-
tions, some boreholes may need to be taken to depths of 100 m.
  Spacing of boreholes may vary from 10 m in erratic strata to 60 m in uniform
material.

4.9 PRESENTATION OF INFORMATION (LOGS)

The log of a borehole, test pit, or excavation contains information about subsurface condi-
tions with depth. Log layouts vary from company to company, but contain similar infor-
mation about soil types, groundwater conditions, field test, and samples taken. Typical
information included on the logs is

1. The soil type, classification symbol (usually the Unified Classification System is used),
soil colour, plasticity (clays), density (sands), particle size (fine, coarse), and secondary
components
Site investigation and in situ testing  81

2. The position at which samples are taken


3. The depths at which in situ tests are performed
4. The moisture condition of the soil and water levels

An example of a borehole log for a hole drilled in soil is shown in Figure 4.1.

4.10 EXCAVATION OR DRILLING METHODS

Physical methods of investigation of foundations or subsoil may involve either drilling bore-
holes or excavating a hole in the ground so as to expose the soil profile for visual examination.

4.10.1 Test pits
Test pits are an effective method for investigating conditions for shallow foundations or
proposed road alignments. Pits are dug with a backhoe or excavator (see Figure 4.2) and
allow visual inspection of the soil profile. Samples of soil can be cut from the sides of the pit
or undisturbed samples can be taken with a thin-walled sampling tube. Water levels can be
recorded if the pit is left open for a period of time.
The sides of test pits can collapse suddenly, especially for pits dug in sandy soil that is
below the water table. For this reason, various government bodies restrict entry into test pits
unless they are properly shored to prevent collapse.
The soil can be visually classified and a log made of the excavation. In addition, a sketch
of the side of the pit can be made showing the extent of all materials exposed. Test pits are
economical as they can be excavated quickly; however, they are limited in depth to about
3.5 m with a backhoe. Larger excavators can dig deeper holes if they have a long reach
bucket mounted on them.
The test pit should be dug to one side of the foundation alignment so as not to disturb the
soil and cause foundation problems later (i.e. excess settlement). The pit is generally filled in
at the completion of the investigation with the front mounted bucket of the backhoe so as
not to be of danger to humans or animals that may fall into the pit.

4.10.2 Excavations
Excavations can be made on large projects, for example on dam sites, where a large trench
is excavated with a bulldozer. The exposed walls of the trench allow examination of any
faults, fissures, or intrusions in the foundation and their extent. Detailed logs of the founda-
tion cross section can be made by looking at the sides of the trench.

4.10.3 Drilling
Various drilling techniques are available to enable deeper investigation of the subsoil. Unlike
excavation methods, where large cross sections of the foundation are exposed, drilling only
gives an idea of strata encountered in the borehole and interpolation must be made between
boreholes to obtain a geological cross section.

4.10.3.1 Hand augers
Hand augers or post-hole diggers (see Figure 4.3) can be used on small jobs, for example,
investigation of pavement failures or foundations of houses. A borehole log can be made by
82  Geomechanics in soil, rock, and environmental engineering

BOREHOLE LOG
CLIENT: Piling Constructions Pty Ltd SURFACE LEVEL: 3.7m AHD BOREHOLE No: 1A
PROJECT: Shopping Centre EASTING: 333720 PROJECT No: JC123AA
LOCATION: Sydney, Australia NORTHING: 6252303 DATE: 12/3/2015
DIP/AZIMUTH: 90°/-- SHEET: 1 of 1
Degree of Rock strength Fracture Discontinuities Sampling and in situ testing
Description

Graphic log
Depth spacing
weathering

Water
of Test results

Medium
RL

Ex High
(m)

V. High

Rec %
B = Bedding J = Joint

Ex Low

Type

RQD
V. Low

Core
(m) and

%
High
MW
strata

0.01

0.10
0.50
1.00
0.05
Low
HW
S = Shear F = Fault

EW

SW
comments

FR
Fs
CLAY – Yellow brown clay
with ironstone gravel
3

1
2

2 2.0
CLAY – Light grey to grey
clay with ironstone bands
1

3
0

Note: Unless
4.0 otherwise stated,
SANDSTONE – Extremely rock is fractured
–1

low-strength, fine-grained along rough planar


5 sandstone with ironstone bedding dipping at
bands 10°
5.5
SANDSTONE – Medium 5.5–5.58 m: clay band PL(A) = 0.8
–2

then high strength,


5.68 m: B0°, fe
6 moderately weathered and
fresh stained, slightly C 100 91
fractured and unbroken, PL(A) = 2.2
brown then light grey, fine- 6.68 m: B5°, fe, clay, vn
–3

grained sandstone 6.82 & 6.88 m: B5°,


7 –10° fe, clay, vn PL(A) = 1.1
7.4 7.4 m: J20°, pl, ro, fe
LAMINITE – High then very
high strength, fresh, slightly
–4

C 100 100 PL(A) = 2.1


fractured, light grey to grey 7.81 m: 80°, clay, vn
8 laminate with approximately
50% fine-grained sandstone
laminations and bands. PL(A) = 4.4
8.65 8.45–8.65: Very high strength
–5

9
–6

RIG: DT 100 DRILLER: JCS LOGGED: JCS CASING: HW to 4.0 m


TYPE OF BORING: Solid flight auger to 4.0 m; Rotary to 5.5 m; NMLC coring to 8.65 m
WATER OBSERVATIONS: No free groundwater observed whilst augering
REMARKS: Used sucker truck in soil section
SAMPLING and IN SITU TESTING LEGEND

JC JayCee Geotechnics
A Auger sample G Gas sample PID Photo ionisation detector (ppm)
B Bulk sample P Piston sample PL(A) Point load axial test IS (50) (MPa)
BLK Block sample Ux Tube sample (xmm dia) PL(D) Point load diametral test IS (50) (MPa)
C Core drilling Water seepage pp Pocket penetrometer (kPa) Geotechnics/Environment/Groundwater
D Disturbed sample Water level V Shear vane (kPa)

Figure 4.1 Example of a borehole log.


Site investigation and in situ testing  83

Front shovel to fill hole

3–4 m deep

Figure 4.2 Excavation of a test pit using a backhoe.

withdrawing the post-hole digger or auger and noting the soil type adhering to the device
and the depth from which it was recovered.

4.10.3.2 Wash borings
This method is less commonly used than auger boring (see Section 4.10.3.3). Water is forced
through inner rods and it exits from the chopping bit, which is moved up and down and
rotated (see Figure 4.4). The water washes the loosened soil to the surface and enables the
hole to be advanced. A casing is advanced in the wash hole as the chopping bit is advanced.
A change in colour of the wash water at the surface indicates a change of soil type. If a soil
sample is required, the bit can be withdrawn and samples can be taken with thin-walled
samplers.

4.10.3.3 Rotary drilling
With rotary drilling, the drill rod and cutting bit are rotated mechanically and advanced
by hydraulic pressure. Water is forced through the drill rods and it emerges at the cut-
ting bit which may be of the tricone roller type shown in Figure 4.5. The soil cuttings are
flushed to the surface with the wash water or the drilling mud, and changes in soil type
may be noted by observing changes in colour of the wash water. Biodegradable drilling
fluids such as ‘Revert’ can be used for keeping the drilled hole open as well as drilling mud
or casings.

4.10.3.4 Auger boring
Augers are turned mechanically, usually from a truck-mounted rig. They are advanced
hydraulically into the soil so that hard to stiff clays or dense sands can be penetrated.
84  Geomechanics in soil, rock, and environmental engineering

Figure 4.3 Hand augers used for shallow investigations.

Lifting bail
Water hose Water swivel
Water pump Drill rods

Wash tee
Cathead
Wash tub

Casing
Chopping bit
Casing drive shoe

Figure 4.4 Wash boring.


Site investigation and in situ testing  85

Figure 4.5 A tricone roller.

Continuous flight augers consist of separate flights which are generally 100 mm (4 inches)
in diameter and 1.5 m (5 ft) or 1.8 m (6 ft) long. Sections are added to the drill string as
drilling proceeds (Figure 4.6).
Depths of drilling are usually (depending on the power of the motor) 35 m in sand, or
25 m in clay after which the drag on the augers will cause refusal of the drilling motor.

Figure 4.6 Auger used for drilling in soil showing a ‘V’ bit.


86  Geomechanics in soil, rock, and environmental engineering

Types of bits used are

1. Blank bits which are suited to soft soils and cleaning out the base of the drill hole
2. Steel ‘V’ bit which can be used in hard clay or soft rock
3. Tungsten carbide bit which can be used for soft rock
4. Fishtail bit (sands)
5. Drag bit (clay)

To calculate the depth drilled

1. Multiply the number of flights in the ground by the length of each flight. (A good check
is to count the number of flights on the rig before drilling. The number in the ground
can be calculated by subtracting the number left on the rig from the total number.)
2. Add 0.1 m for ‘V’ or tungsten carbide bit.
3. Subtract the distance from the top of the flight (which remains above ground) to the
ground level.

Sampling material using the auger (disturbed samples) involves

1. Using the blank bit, drill in low gear and adjust the speed of penetration so that the
auger ‘winds’ itself into the ground (i.e. spirals do not appear to be going up or down).
2. Stop the flight after about 1 m penetration (stiff clay 0.5 m) and withdraw the auger
without rotation.
3. Cut the contaminated material away and log the material extracted.

As drilling proceeds, the ease of drilling can be noted as a guide to the strength of the soil
(this can range from easy drilling to refusal).
When pulling up the drill string (especially in sands), keep the hole full of water and pull
the drill string up slowly. If the water level drops, the bottom of the hole can blow up and
SPT readings will be erroneous.

4.11 SAMPLING METHODS

As well as performing in situ tests and obtaining soil properties, soil samples may be recov-
ered and taken back to a laboratory for testing.
The site investigation code (AS 1726 1993) recommends that at least one undisturbed
sample should be taken from every different stratum encountered. If the stratum is uniform,
one sample can be taken at every 1.5 m, which is the length of the auger flights.
Samples may be disturbed or undisturbed depending on the type of test to be performed.
Disturbed samples are those for which the soil fabric is disturbed during sampling. Such
samples are normally placed in a plastic bag and sealed so that no moisture is lost. A tag
or label is placed on or inside the bag identifying the soil sample (i.e. date, depth, borehole
number).
Disturbed samples can be used for

1. Compaction tests: 20–30 kg


2. Laboratory or visual classifications, for example, sieve analysis or index tests: 1–5 kg
3. Chemical tests: 1–5 kg
4. Moisture content: 1 kg
Site investigation and in situ testing  87

Undisturbed samples are those for which the soil fabric has been disturbed as little as pos-
sible, although some disturbance will inevitably occur.
Undisturbed samples are required for

1. Oedometer tests
2. Triaxial tests
3. Swell tests
4. Unconfined compression tests

4.11.1 Thin-walled sampler (or Shelby tube)


The tube (see Figure 4.7) is usually 50 mm (2 inches) in diameter (so the specimen is called a
U50 or undisturbed 50 mm diameter specimen). Larger and smaller tube sizes are available
with the larger sizes used if sample defects need to be studied. Tubes are about 0.5 m long.
Use of thin-walled sampling tubes:

1. Ensure that the cutting edge has not been bent (e.g. from previous use in gravelly or
hard soil).
2. Ensure that the tube is not rusty and that there is no old soil or wax in the tube from
previous use. Lubricate the tube with oil or water.
3. Attach the ball valve adaptor.
4. Lower the tube to the required depth. The hole may need to be cleaned out first with
a blank bit, as it may be difficult to distinguish soil which has fallen into the borehole

Figure 4.7 Thin-walled sampling tube.


88  Geomechanics in soil, rock, and environmental engineering

from the sample back in the laboratory. The tube is pushed at a constant rate into the
soil by the hydraulic ram of the drilling rig.
5. Turn the drill rods once to break off the sample and then withdraw the sampler.
6. Record the depth and length of samples on an identification tag and record the sample
on the borehole log, for example, 1.8–2.1 m; 0.3 m recovery.
7. Tubes are sealed at each end with wax, plastic caps, or end seals to prevent moisture
loss before testing.
8. Ensure that the samples are not left in the sun or broken up during transport.

4.11.2 Split spoon sampler (SPT sampler)


Samples are taken as part of the standard penetration test described in Section 4.15.
Because the sampling tube has a thick wall, samples are disturbed more than samples
taken with a thin-walled tube. Samples recovered in this way are suitable for index tests and
identifying soil types.

4.11.3 Piston sampler
The piston sampler has a piston inside the tube which prevents the sample from slipping out
of the tube. The piston also prevents unwanted material entering the tube as it is lowered
(see Figure 4.8b).

4.11.4 Air injection sampler


An air line is provided so that air can be introduced to the underside of the sample facilitat-
ing its retention in the sampling tube (see Figure 4.8c).

4.11.5 Swedish foil sampler


When a tube is pushed into the soil, friction between the walls of the tube and the soil cause
the length of the sample which can be removed to be limited. This also causes disturbance
to the soil samples.

(a) (b) (c)


Drill rods

Sampler head
Ball valve
Fixing screws
Piston
Thin-walled tube Air line

Figure 4.8 Sampling tubes. (a) Simple push sampler, (b) piston sampler, and (c) air injection sampler.
Site investigation and in situ testing  89

A specialised piece of equipment, the Swedish foil sampler, removes the side friction by
means of foil ribbons which line the sides of the sampler. Lengths of sample up to 18 m can
be obtained with this device.

4.12 ROCK CORING

To obtain samples of rock, it is necessary to use diamond drilling equipment. Rock core can
be inspected for fissuring, weathering, and defects or can be used to obtain rock strength by
testing the recovered core.
The core is obtained by using a triple or double core barrel which has a diamond studded
bit attached at one end. A bit being removed is shown in Figure 4.9. Water is passed down
the drill string and it emerges at the bit. This has the dual effect of cooling the bit and of
flushing away the cuttings. However, the wash water will also wash away any weathered or
soft rock or soil and it will be lost.
Core barrels can be of the triple or double type. The triple core barrels have a thin split inner
sheath into which the core slides, and surrounding this two outer core barrels are found. This
arrangement gives better core recovery. In order to retain the core in the core barrel, a core
catcher device is used which can retain the core stopping it from slipping out when an attempt
is made to lift the core barrel. Different sizes of core may be recovered, but the common sizes
are N size multi-leaf coring (triple core barrel) NMLC (52 mm) and H size HMLC (63 mm).
The core, when recovered, is placed in boxes (usually galvanised steel sheet boxes) in a
standardised way. Each section of the box is approximately 1.05 m long and represents l m
of core (50 mm extra is allowed for expansion). The core is placed in the first box and run
from left to right (the top of the core being to the left) and continued onto the next row in
the box (Figure 4.10). Once the first box is full, the core is placed into the second box and so
on. All boxes are labelled to indicate the depths of the core contained in them.
The core is boxed in the following way:

1. Suppose a length of core is retrieved by drilling from 1.4 m depth to 2.6 m depth.
The recovered core is placed in the core box and the depth of the bit at the end of the
run is marked with a piece of wood or polystyrene foam which shows the depth, for
example, 2.6 m.

Figure 4.9 Diamond-drilling bit being removed.


90  Geomechanics in soil, rock, and environmental engineering

Figure 4.10 Method of boxing the cores.

2. The core is removed from the barrel and placed in the box after measuring, say 1.0 m
long.
3. This means that the core loss is 0.2 m (i.e. 2.6–1.4 = 1.2 m should have been obtained)
and this is marked with a red piece of polystyrene (Figure 4.11). Generally, core loss is
shown at the top of the run, but if its true position is known this may be indicated by
a red marker in the correct position. Core loss can be due to cracks in the rock, voids,
or soft material being washed away.

4.13 FIELD TESTS

Various tests may be carried out in the field to obtain soil data rather than taking samples
back to a laboratory for testing. In some cases, field tests can produce better results than
laboratory tests as the soil has been tested in situ and not disturbed. This is especially true
for sands where sampling will change the void ratio of the sand, and unless the material can
be reconstituted to the correct void ratio, results may not be accurate.

Figure 4.11 Fractured core in a box with polystyrene markers showing core loss.
Site investigation and in situ testing  91

4.14 VANE SHEAR TEST

Shear vanes are used in soft clay deposits where sampling of the soil would be difficult or
cause disturbance. The vane consists of two blades at right angles to each other (see Figure
4.12) and is pushed into the soil at the base of a borehole. Vanes come in various sizes and
usually have a height to diameter ratio of 2.
A torque is applied to the vane and increased until the vane shears off a cylindrical volume
of soil (which is contained within the blades). The undrained shear strength of the soil can
then be calculated from the torque at failure T from
T
su = 2
(4.1)
πd (h / 2 + d /6)

where h is the height of the vane, d is the diameter, and su is the undrained shear strength
of the clay.
Corrections can be applied to the field values of shear strength to give values which can be
used for design purposes. The field values are multiplied by a factor λ to give the corrected
design value. Values of λ have been given by Bjerrum (1972) and by Aas et al. (1986) and are
shown in Figure 4.13. The vertical effective stress σ′v, is the present vertical effective pressure
acting at any point in the soil.

4.15 STANDARD PENETRATION TEST

The standard penetration test first originated with the Raymond Pile Company in the United
States in about 1927 and today is widely used in site investigation work. The test originally

Rod

Four-bladed vane

Sheared cylindrical
surface
h

Figure 4.12 Shear vane.


92  Geomechanics in soil, rock, and environmental engineering

(a) 1.2 (b)

1.4
1.0
Correction factor (λ)

Correction factor (λ)


1.0 Normally
consolidated
0.8 clay

0.6
0.6 Overconsolidated
clay
0.2
0 0.2 0.4 0.6 0.8 1.0
0.4
0 20 40 60 80 100 su(field)/σv′
Plasticity index PI

Figure 4.13 Correction factors for vane shear strength. (a: After Bjerrum, L. 1972. Proceedings of the ASCE
Conference on Performance of Earth-Supported Structures, Purdue University, Vol. 2, pp.  1–54.)
(b: After Aas, G. et al. 1986. Proceedings, In Situ 86, American Society of Civil Engineers, pp. 1–30.)

involved hammering a thick-walled sampler into the soil so that a soil sample could be
taken, but it was soon realised that the number of blows that were taken to drive the sampler
into the ground were indicative of the density of sands or the stiffness of clays. The weight
used to drive the sampler and the distance over which the sampler was driven 300 mm
(1 foot) were eventually standardised.
The original test involved manually lifting a weight with a rope, and dropping the weight
onto an anvil connected to the drill rods. More modern equipment uses an automatic trip
hammer that does not have a rope attached and so results from the two types of equipment
vary. International Standards have tried to make the test more repeatable by specifying the
equipment that should be used and the method of performing the test.

4.15.1 Equipment
The equipment consists of a split sampler, rods, and falling weight that is used to drive the
sampler. The equipment specifications are outlined in Australian Standard AS 1289.6.3.1
(2004), American Society for Testing of Materials ASTM D1586-11 (2011), and British
Standards Institution BS EN ISO 22476-3 (2011) documents.
The test is performed at the bottom of a borehole and so auger drilling equipment (usually
a truck-mounted auger) or a wash boring rig is required to drill the hole.

4.15.2 Sampler
The SPT sampler consists of a split cylindrical barrel of outside diameter 50 mm (2 inches)
and inside diameter 35 mm, making the wall thickness about 7.5 mm. Because of the wall
thickness, samples taken will be disturbed, and so are not to be used for tests requiring
undisturbed samples (e.g. triaxial tests). At one end of the sampler, a cutting shoe is attached
and at the other end is a ball valve adaptor. The ball valve consists of a ball bearing that
will allow air and water to be forced out of the sampler as soil moves into the barrel, but
will prevent a return of air or water if the sampler is lifted. This creates a suction above the
Site investigation and in situ testing  93

Driving shoe Sampler head

Four vents

35 mm 50 mm

Split barrel Ball valve

Figure 4.14 Split spoon sampler used in a standard penetration test.

Figure 4.15 Split spoon sampler showing thick-walled split tube and cutting shoe.

sample, and helps break off and lift materials like soft clays that would otherwise slip out of
the sampler. The sampler is shown in Figures 4.14 and 4.15.

4.15.3 Drive hammer
The drive hammer consists of a 63.5 kg (140 lb) weight that can free fall onto an anvil such
that the drop is 760 mm (30 inches). The hammer must be able to fall freely onto the anvil
and not jam on the guide or be restricted by ropes. The hammer is generally dropped by a
self-tripping mechanism so that it does fall freely and delivers most of the energy from the
drop to the sampler (called the ‘efficiency’). A trip hammer commonly used in the United
Kingdom is shown in Figure 4.16.
One way to account for the efficiency of the hammer (and the effect of other factors) is to
standardise the blow count to 60% efficiency. This is discussed in Section 4.15.5.

4.15.4 Rods
The rods that connect the anvil to the sampler should be stiff enough so that there is no loss
of energy through bending and compression of the rods. The Australian Standard requires
94  Geomechanics in soil, rock, and environmental engineering

Lifting assembly

Trip mechanism

63.5 kg weight
(140 lb)

Anvil

Figure 4.16 Drop hammer showing trip mechanism.

AW rod, with holes deeper than 15 m having steadies at intervals of 6 m or stiffer rod. The
rod diameter should not exceed 66.7 mm.

4.15.5 Test procedure
The hole is first cleaned out making sure that in sands the water table is not allowed to fall.
If the water table did fall, water could flow upwards into the borehole and loosen the soil
(especially sand) where the test is to be conducted. The test should also be performed below
any casing that is being used.
Site investigation and in situ testing  95

The sampler and rods are then lowered and three intervals of 150 mm (6 inches) are
marked off along the drill rods. The sampler is driven 150 mm to seat the sampler, and then
two further drives of 150 mm (making 450 mm in all) are performed. The sum of the blow
counts is then made for the last two drives of 150 mm to get the blow count N.
For example, if the blow counts were 10, 13, and 14, then we add the last two to get
N = 27 as the blow count.
If the full 450 mm is not penetrated, the blow count and the penetration achieved over
the last two drives are recorded, for example, 50 blows over 200 mm. If not more than the
first 150 mm is penetrated, the blows and distance are recorded, for example, 60 blows for
initial 100 mm of penetration.
The test is usually stopped once 50 blows are recorded, however when soft rock or very
stiff or dense materials are encountered, blow counts higher than 50 are recorded by some
operators.
In order to account for energy losses due to

• The use of a cathead and winch plus rope


• Loss of energy between the hammer and anvil
• Differences in drop height, especially when a cathead is being used by an operator

the SPT blow count can be corrected to values for a system where 60% of the theoretical free-
fall energy is imparted to the rods. This correlates to the safety hammer with cathead and
winch that is commonly used in the United States and on which many correlations are based.
Correction of the raw SPT blow count N field to 60% efficiency N 60 can be done by use of
the following formula (Equation 4.2).

N60 = (ERm /60)N field (4.2)

where ER m is the efficiency of the hammer in (%). This procedure has been examined by
Kulhawy and Mayne (1990). When using charts and correlations with material properties,
it is necessary to know if the correlation is performed with N 60 or some other value.
To correlate the SPT blow count with some soil properties, it is necessary to correct the
blow count for the effects of overburden pressure as well, for instance when assessing lique-
faction potential of a soil (see Section 4.15.8). A curve giving the correction factor with depth
(or effective stress) needs to be provided and the corrected blow count is then called (N1)60.

4.15.6 Properties of sands
The SPT test is able to indicate the relative density and the drained angle of shearing resis-
tance of sands. The variation of the angle of shearing resistance ϕ′ with the corrected blow
count (N1)60 (see Section 4.15.8) is shown in the plot of Figure 4.17.
The relationship in the figure is given by

φ′ = [15.4(N1)60 ]0.5 + 20 (4.3)


Tests carried out by different investigators have shown that the relative density of sands
depends on the overburden pressure, and so plots can be made of relative density for
sands in terms of the effective vertical stress and the SPT blow count N. This is shown
in Figure 4.18 where results obtained by Gibbs and Holtz (1957) and Bazaraa (1967) are
shown. The curves of Gibbs and Holtz tend to underestimate the relative density and so the
plots of Bazaraa are more generally accepted.
96  Geomechanics in soil, rock, and environmental engineering

55

50 ϕ′ = [15.4(N1)60]0.5+ 20°

Friction angle ϕ ′ (degrees)


45

40

35
Sand (SP and SP–SM)
30 Sand fill (SP to SM)
SM (Piedmont)
25
H&T (1996)

20
0 10 20 30 40 50 60
SPT blow count (N1)60

Figure 4.17 Peak friction angle of sands from SPT resistance. Note: The normalised resistance is
(N1)60 = N 60/(σ′vo/pa) 0.5, where pa = 100 kPa = 1 bar = 1 tsf. (After Hatanaka, M. and Uchida, A.
1996. Soils and Foundations, Vol. 36, No. 4, pp. 1–10.)

0
100

50
Vertical pressure (kPa)

80

100
60

100
150
ex
40

80 Ind
ty
200 60 nsi
De
40 in %
250

0 10 20 30 40 50 60 70
SPT blow count (N1)60
Gibbs and Holtz Bazaraa

Figure 4.18 Relative density of sands as a function of effective vertical stress.

4.15.7 Properties of clays
Because the SPT is a dynamic test, correlations of the blow count with soil properties for
clays are approximate only, since the driving process can set up pore pressures in a clay and
these will affect the blow counts.
Correlations have been made of the blow count with the undrained shear strength su
of clays, but the relationship depends upon the plasticity of the clay. Clays with a lower
Site investigation and in situ testing  97

25
ck
Pe nd n)

d
d

un
i an bou m an

bo
gh per r t
20 za er me y)

r
er I) up Gold (Sch ty cla

we
T i l

) lo
y); m P y
cla rat (s
cla ediu agoangle

PI
d
e r
d y i c oun
er b

h
ld an s (m h S
SPT N value

hig
15 o s C p p
G t( e r I) u ess);

s(
ra ow wP o
gle S s (lo nois (L

r
we
a n ) w e r i
o I l l

So
S R S );
10 SB clay India
n (U
l e r at (
to g via
us San Yugosla
Ho y
5 cla

0
0 0.5 1.0 1.5 2.0
Su/pa

Figure 4.19 Correlation of undrained shear strength of clays su to SPT blow count N (pa = 100 kPa = 14.5
psi = 1 tsf).

Plasticity index PI has a higher undrained shear strength for a given blow count than a high
plasticity clay. Charts like shown in Figure 4.19 (which is based on data from several inves-
tigators) show the large spread in results for clays of different plasticities and so need to be
used with caution. Figure 4.20 also shows the variation of undrained shear strength with
the plasticity of a clay.

10

×
6 ××
su/N (kPa)

×
4

0
0 10 20 30 40 50 60 70
Plasticity index (%)

Boulder clay Kimmeridge clay


Laminated clay Woolwich and Reading clay
Sunnybrook till Upper-Lias clay
London clay × Keuper marl
Bracklesham beds Finz
Oxford clay

Figure 4.20 Ratio of undrained shear strength to SPT blow count versus plasticity index (PI). (After Stroud,
M.A. 1974. Proceedings European Symposium on Penetration Testing, ESOPT, Stockholm 1974.
National Swedish Building Research, Vol. 2.2, pp. 367–375.)
98  Geomechanics in soil, rock, and environmental engineering

Correlations based on local knowledge can give a reasonable correlation once a database
of information is built up.

4.15.8 Liquefaction
If cohesionless soils that are saturated are vibrated very rapidly (e.g. by an earthquake),
excess pore pressures can be generated and this leads to the phenomenon of liquefaction.
This is an important consideration when designing structures in areas prone to earthquake,
as foundation liquefaction has led to widespread damage. The Niigata earthquake of 1964
is an example.
Soils most susceptible to liquefaction are loose soils of uniform grain size. Fluvial depos-
its, and colluvial and aeolian deposits or man-made hydraulic fills when saturated are most
prone to pore pressure build up that will lead to liquefaction.
SPT tests have commonly been used in the United States and many other countries to
determine whether a soil deposit is liable to liquefy. Seed et al. (1985) have presented a plot
of the corrected blow count versus cyclic shear stress ratio for clean sand (Figure 4.21). In
the figure, the cyclic shear stress ratio CSR at which the soil will liquefy under an earth-
quake of magnitude M = 7.5 is indicated by a solid black symbol, and where it does not, by
an open symbol. The data are based on field data from Japan, China, and Pan America and
it can be seen that above the line on the plot, the soil generally liquefied and below the line,
liquefaction did not generally occur.
Because the data were collected from many different countries using different techniques
for performing the SPT, it was necessary to correct the blow counts to an energy ratio of
60% or N60 for use with the charts (see Table 4.1). The corrected blow count must also be
determined for a standard effective overburden pressure of 1 ton/ft 2 (approx. 100 kPa) and
so is called (N1)60. The correction factor C N for determining the overburden correction is
shown in Figure 4.22. The corrected blow count can then be found from the measured blow
count N as follows

N1 = CN N (4.4)

To use the chart, therefore, a knowledge of the corrected SPT blow count (N1)60 and the
cyclic shear stress ratio CSR is required. The cyclic shear stress ratio may be computed at
any depth required from the equation

τ av a σ
= 0.65 max ⋅ v0 ⋅ rd (4.5)
σ ′v0 g σ ′v0

where
amax is the maximum acceleration at the ground surface
g is the acceleration due to gravity
σ v0 is the total overburden pressure at the depth considered
σ ′v0 is the effective overburden pressure at the depth considered
rd is a stress reduction factor that varies as shown in Figure 4.23

For sands containing more than 5% fines, liquefaction is less likely to occur. For this case,
Seed et al. (1985) have presented an alternative chart for determining liquefaction potential
as shown in Figure 4.24.
Site investigation and in situ testing  99

0.6

0.5

0.4
CSRM=7.5

0.3

0.2

Fines content ≤ 5%
Chinese building code (clay content = 0)
Marginal No
0.1 Liquefaction Liquefaction Liquefaction
Pan-American data
Japanese data
Chinese data
0
0 10 20 30 40 50
(N1)60

Figure 4.21 Relationship between cyclic stress ratios causing liquefaction and (N1)60 values for clean sands
in M = 7.5 earthquakes. (After Seed, H.B. et al. 1985. Journal of Geotechnical Engineering, ASCE,
Vol. 111, No. 12, pp. 1425–1445.)

Table 4.1​  ​Energy ratios for different test conditions


Hammer Estimated rod Correction factor for
Country (1) type (2) Hammer release (3) energy (%) (4) 60% rod energy (5)
Japana Donut Free fall 78 78/60 = 1.30
Donutb Rope and pulley with special throw release 67 67/60 = 1.12
United Statesb Safetyb Rope and pulley 60 60/60 = 1.00
Donut Rope and pulley 45 45/60 = 0.75
Argentina Donutb Rope and pulley 45 45/60 = 0.75
China Donutb Free fallc 60 60/60 = 1.00
Donut Rope and pulley 50 50/60 = 0.83
Source: Adapted from Seed, H.B. et al. 1985. Journal of Geotechnical Engineering,  ASCE, V
  ol. 111, No. 12, pp. 1425–1445.
a Japanese SPT results have additional corrections for borehole diameter and frequency effects.
b Prevalent method in the United States today.
c Pilcon type hammers develop an energy ratio of about 60%.
100  Geomechanics in soil, rock, and environmental engineering

100

Effective overburden pressure (kPa)


DR = 40%–60%

200

DR = 60%–80%

300

400

0 0.4 0.8 1.2 1.6


CN

Figure 4.22 Chart for correction values CN.

0 0

5 Average values
20

10
Range for different 40
Depth (m)

Depth (ft)

15 soil profiles

60
20

80
25

30 100
0 0.2 0.4 0.6 0.8 1.0
Stress reduction factor (rd)

Figure 4.23 Reduction factor rd to estimate the variation of cyclic shear stress with depth below level
or gently sloping ground surfaces. (After Seed, H.B. and Idriss, I.M. 1971. Journal of the Soil
Mechanics and Foundations Division, ASCE, Vol. 97, No. SM9, pp. 1249–1273.)
Site investigation and in situ testing  101

0.6

Percent fines = 35 15 ≤5
0.5

0.4
CSRM=7.5

0.3

0.2

Fines content ≥ 5%
Modified Chinese code proposal (clay content = 5%)
0.1 Marginal No
Liquefaction Liquefaction Liquefaction
Pan-American data
Japanese data
Chinese data
0
0 10 20 30 40 50
(N1)60

Figure 4.24 Relationship between cyclic stress ratios causing liquefaction and (N1)60 values for silty sands in
M = 7.5 earthquakes.

4.16 PRESSUREMETERS

Pressuremeters are basically devices which consist of a cylindrical rubber membrane which
can be lowered into a borehole and inflated against the sides of the borehole. Measurements
can be taken of the radial expansion of the membrane as it is inflated by gas pressure (water
or oil is used in some types).

4.16.1 Types of pressuremeters
The types of pressuremeters commonly in use are

1. The Ménard type which is lowered into a pre-formed borehole. The test method is
described in ASTM D4719-07 (2007) or British Standard BS EN ISO 22476-4 (2012).
2. The self-boring pressuremeter (or Camkometer) which has a cutting head to enable
the instrument to be inserted into the ground at the bottom of a borehole caus-
ing very little disturbance. Such a device is shown in Figure 4.25. The name of the
device comes from its original name the Cambridge K0 meter as it was developed at
Cambridge University and could be used to measure the coefficient of earth pressure
at rest K0.
102  Geomechanics in soil, rock, and environmental engineering

Flushing water
Return flow
A Electric/gas
cable
Ax

Feeler

Pore pressure
cell

Total pressure
Rubber
cell
membrane

Water flush

Tapered
Cutter
cutting head

Figure 4.25 The Camkometer.

3. Push in pressuremeters which are pushed into the ground at the bottom of a borehole.
4. The main advantage of the device is that it is able to test soil in a relatively undisturbed
state (especially the Camkometer), however the disadvantage is that the membrane is
being expanded laterally and so the response of the soil is being obtained laterally.

4.16.2 Interpreting test results


A plot is made of the radial strain of the cavity against the pressure inside the membrane.
Such a plot is shown in Figure 4.26 where it can be seen that there is initially very little strain

6
Pressure p (MPa)

0
0 2 4 6 8 10 12 14 16 18 20
Cavity strain (%)

Figure 4.26 Plot of cavity strain versus pressure for pressuremeter showing unload – reload loops.
Site investigation and in situ testing  103

8
7 PL = 7.5 MPa

Pressure p (MPa) 6
5
4

3
2
1

0.001 0.01 0.1 1


ΔV/V

Figure 4.27 Plot of log ΔV/V versus membrane pressure for pressuremeter test in very dense silty sand.

as the pressure increases back to the geostatic lateral stress existing in the ground σh = K0σv
at which point the cavity begins to expand.
Usually, some load and reload loops are carried out so that the reload modulus of the soil
can be obtained as this is generally accepted to be the most appropriate value for design.

a. Shear modulus
The shear modulus G of the soil may be computed from the formula

1  ρ   dp  1  dp 
G = ≈  (4.6)
2  ρ0   d ε c  2  d ε c 

where p is the internal membrane pressure εc is the cavity strain (Δρ/ρ0) and the term
ρ/ρ0 is the ratio of the expanded radius of the cavity to the initial radius and may be
approximated as unity.
  The shear modulus may therefore be found from the plot of cavity strain versus pres-
sure since it is merely 1/2 the slope of the curve as may be seen from Equation 4.6.
b. For clays, the undrained shear strength may be found from Gibson and Anderson’s
(1961) method of analysis. A plot is made of the natural logarithm of the volume strain
ΔV/V of the cavity versus the pressure.
  Since

 ∆V 
p = pL + su ln  (4.7)
 V 

the slope of the plot will give the undrained shear strength of the clay. This is shown
in Figure 4.27 where it can be seen that the theoretical limit pressure pL occurs where
Δ V/V = 1.

4.17 DILATOMETERS

This instrument (which is shown in Figure 4.28) was developed by Marchetti in Italy
(Marchetti 1980). It consists of a flat stainless steel blade 14 mm thick and 95 mm wide
104  Geomechanics in soil, rock, and environmental engineering

Wire

Pneumatic 14 mm
tubing

P0 P1

1.1 m
Flexible
Flexible membrane
membrane
mm
60

95 mm

Figure 4.28 Marchetti dilatometer.

which is pushed into the ground. A 60 mm diameter membrane which is on the side of the
instrument is inflated by gas pressure (see Figure 4.29) after the dilatometer has been pushed
to the required depth. Tests have to be performed at discrete intervals of depth and so tests
are usually conducted at distances of 0.15–1.3 m apart.
The pressure required to inflate the membrane, p 0 and the pressure required to lift the
membrane 1 mm p1 are recorded (and corrections for the instrument are applied).

Figure 4.29 Marchetti dilatometer with pressure gauges.


Site investigation and in situ testing  105

More details for performing the test may be found in British Standard BS EN ISO
22476-5 (2012).
The following quantities are then calculated:

i. The material or deposit index I D


p1 − p0
ID = (4.8)
p0 − u0

where u 0 is the static water pressure.

ii. The horizontal (or lateral) stress index K D


p0 − u0
KD = (4.9)
σ ′v0

where σ′v0 is the in situ vertical effective stress.


iii. Dilatometer modulus

ED = 34.7(p1 − p0 ) (4.10)

4.17.1 Type of soil
The type of soil can be identified by using the chart developed by Marchetti and Crapps
(1981). The chart makes use of the dilatometer modulus E D and the material index ID to
identify the soil type (see Figure 4.30).

4.17.2 Shear strength of clays


The undrained shear strength of a clay su may be calculated from (Marchetti 1980)

su = 0.22σ ′v0 (0.5KD)1.25 (4.11)



If ID ≤ 1.2; if ID > 1.2 the soil is cohesionless.

4.17.3 Other quantities
Other soil properties such as K0 (the coefficient of earth pressure at rest), ϕ′ (the angle of
shearing resistance) can be found from the test (Marchetti 1980).
The formula for computing K0 in clay is given by Equation 4.12.
0.47
K  (4.12)
K0 =  D  − 0 .6
 1 .5 

The angle of shearing resistance of sand can be found from Marchetti’s equation
(Marchetti 1997).

φ′DMT = 28° + 14.6°log KD − 2.1°log 2KD (4.13)



More relationships of soil properties with the dilatometer test results can be found in the
book by Schnaid (2009).
106  Geomechanics in soil, rock, and environmental engineering

SILT SAND 10 )
Clayey sandy Silty (2.
se n
1000 de
ry
Ve )
00
(2.
500 m
CLAY 10
) diu
(2. Me .90)
Silty (1
Dilatometer modulus ED /pa

)
95
200 (1.
rd )
Ha 05) 80
(1.
(2. 0)
(1.
8 ose
100 Lo 70)
90
) (1.
(1. m 70
)
diu (1.
50 Me 80) ) Notes:
(1. (1.
60 a – Number in parenthesis
) is normalised unit
70
(1. weight (γ /γw)
20 ft
So 80) b – If PI > 50, (γ /γw) is
(1.
12 overestimated by about
0.3 0.6 0.9 1.2 1.8 3.3
1 MPa 10 0.1
Very soft
Clay/peat
5
0.1 0.2 0.5 1 2 5 10
Material index ID

Figure 4.30 Chart for identification of soil type using a dilatometer.

4.18 CONE PENETROMETERS

Cone penetrometers can be of two types: mechanical or electrical. For cone penetration test-
ing, it is not necessary to drill a hole as the cone is pushed into the ground using a hydraulic
ram. Cones can be mounted on small trailers that are anchored to the ground by screw-in
anchors, or can be truck mounted using the weight of the truck as a reaction against which
to push the cone into the soil.

4.18.1 Equipment
Most of the modern cone penetrometers consist of a cone which has a 60o apex angle and a
diameter of 35.7 mm (giving a cross-sectional area of 1000 mm 2) as well as a friction sleeve
of 10,000 or 15,000 mm 2 in surface area. Cones also come in other sizes, but these dimen-
sions are the more commonly used.
Cone specifications and the procedures that should be used during testing are outlined
in Australian Standard AS1289.6.5.1-1999 for both the mechanical cones and the elec-
tric cones, in ASTM D 5578-95 for electric friction cones and piezo-cones, or in British
Standard BS EN ISO 22476-1 (2012) for piezo-cones and electric cones.
The older mechanical cones are more rugged than the electric cones, but the advantages
of the electric cone are

• Better accuracy and repeatability of results, particularly in weak soils


• Better delineation of thin strata
• Faster speed of operation
Site investigation and in situ testing  107

• It is possible to measure pore pressure on some devices


• Continuous electronic recording of data

Two types of cones are available, those with pore pressure measurement devices used for
CPTu (cone penetration test with pore pressure measurement) tests and those without pore
pressure measurement used for CPT tests (cone penetration testing).

4.18.2 CPT equipment
The older mechanical cones are pushed into the soil using hydraulic pressure and readings
of the pressure versus the depth are recorded. These devices can have a conical tip that is
advanced by an inner rod while the outer rods are held stationary. Cones can also have a
sleeve (as shown in Figure 4.31). By pushing an inner rod, the cone is firstly advanced, and
then the sleeve is advanced. The sleeve friction is obtained by subtracting the cone pressure
from the combined pressure measured for both the cone and the sleeve.
Electric cones are constructed as shown in Figure 4.32 where it may be seen that the cone
tip is connected to a load cell with strain gauges being used to measure the strain. The output
from the strain gauges can be calibrated so as to give a direct readout of the cone value qc that is
commonly expressed in MPa. A cone with the sleeve and tip removed is shown in Figure 4.33.
The friction sleeve is also connected to a load cell so that the friction on the sleeve of the
penetrometer can be automatically recorded.

4.18.3 CPTu equipment
As mentioned above, some electric cones have the ability to measure pore pressures, and
this provides extra useful information. Such cones are called piezo-cones, and can have
the pore pressure sensors mounted in different locations, at the tip, on the face of the cone,
behind the cone tip, or behind the sleeve of the cone. This is shown in Figure 4.34a,b. The
response to pore pressure is very different depending on the location of the filter and trans-
ducer. The highest pressures are measured when the piezo-element is mounted at the face of
the cone. The pore pressure depends on the type of soil, with highly overconsolidated soils
producing higher pressures at the face.

Figure 4.31 Mechanical cone showing friction sleeve.


108  Geomechanics in soil, rock, and environmental engineering

Signal cable

‘O’ ring seal

Friction sleeve strain


gauge load cell

Friction sleeve

Cone strain gauge load

‘O’ ring seal

Cone

Figure 4.32 Typical electric friction cone penetrometer.

A small pressure sensor is mounted behind the porous element to measure pressure
changes. It should be such that a minimal volume change in fluid passing the filter is required
to operate the transducer to give a fast response time.
In addition (for a fast response time in the pore pressure measuring system), the fluid used
should have a low viscosity and compressibility, and a high permeability filter should be
used. The filter itself should not be compressible, as this leads to filter compression effects,
that is, the compressing filter would cause an increase in pore pressure.
The filter can be made of porous plastic, ceramic, or sintered stainless steel. Ceramic
filters are generally damaged when pushed into dense sands, but polypropylene can survive
being pushed into dense sands and gravelly soils.
Pore pressures measured in the CPTu test rely very strongly on the filters being satu-
rated, and not containing any air. General practice is to saturate the filter elements in the
Site investigation and in situ testing  109

Figure 4.33 Electric cone with friction sleeve and the cone tip removed.

laboratory by placing them under a high vacuum. Some cone users submerge filters in warm
glycerin in an ultrasonic bath under a vacuum. The voids in the cone itself are ­de-aired by
flushing with a suitable fluid (e.g. either glycerin or water) and the cone can be kept de-aired
by placing a latex sheath over it while it is placed on the push rods.
Results of a piezo-cone test are shown in Figure 4.35 where the cone resistance (pressure
on the end of the cone), the friction ratio (see Equation 4.14), and the pore pressure are plot-
ted with depth. The type of soil can be identified from these results as in this plot, the low
friction ratios correspond to sand and silty sand, whereas the high friction ratios (at about
3 m and 7.7 m) correspond to clay layers. At the location of the clay layers, the pore pres-
sures measured also show high values.

4.18.4 Pushing equipment
Truck-mounted rigs are generally specifically built for the equipment, but sometimes anchored
trailer mounted rigs are used. The thrust capacity required is usually between 10 and 20
tonnes, although a capacity of 10 tonnes is enough to carry out most penetration testing to
30 m depth. Trucks ballasted to about 15 tonnes are suitable for this type of testing.
Hydraulic power for pushing the cones comes from the truck engine and the cone must be
pushed at the standard rate of 20 mm/s (AS 1289.6.5.1-1999 allows 10–20 mm/s). Sections
of the rods are added as the cone is pushed deeper, and rigs of this sort can produce up to
250 m of testing in one day.
Friction reducers that consist of an expanded coupling can be used behind the cone to
reduce friction on the rods, or natural or polymer drilling fluid can be pumped down the
drill rods and allowed to flow up the outside of the rods about 1.5 m behind the cone to
reduce friction and therefore pushing force.

4.18.5 Calibration
The cones are normally calibrated by using a load cell before each sounding is made. The
calibration should be done with O-rings and seals in place as would be the case when the
cone is in use.
110  Geomechanics in soil, rock, and environmental engineering

(a)

(b)

Electronics
housing

Piezo-element
behind friction
sleeve
Friction sleeve

Piezo-element
behind tip
Piezo-element on
Cone tip face

Figure 4.34 (a) Location of piezo-elements on piezo-cone. (b) Piezo-cone showing possible locations of
piezo-elements.

Temperature can have an effect on the calibration and this can be overcome by pushing
the cone into the ground about 1 m and leaving it for about half an hour before calibration.
Pore pressure calibration should be done with a pressure chamber that completely encloses
the cone and is sealed at a point above the friction sleeve. Measurement of the tip stress and
friction sleeve stress at applied pore pressures will allow direct determination of unequal
Site investigation and in situ testing  111

Cone resistance Friction ratio Pore pressure


0 0 0

1 1 1
In situ water
level
2 2 2

3 3 3

4 4 4

5 5 5

6 6 6
Depth (m)

Depth (m)

Depth (m)
7 7 7

8 8 8

9 9 9

10 10 10

11 11 11

12 12 12

13 13 13
0 4 8 12 16 0 2 4 6 8 10 0 50 100 150 200
qt (MPa) Rf (%) u (kPa)

Figure 4.35 Pore pressure distributions for a piezo-cone – different soil types.

end effects (i.e. loads caused by pore pressures that may affect the cone readings such as
water pressure at the back of the cone tip).

4.19 INTERPRETATION OF CONE DATA

Cone data can be used for estimating many different soil properties and the type of soil
through which the cone is being pushed. The cone values are affected by the lateral in situ
stress and this should be considered where possible. However, this information is not often
available, and cannot be applied in the interpretation.

4.19.1  Soil classification


Charts such as the ones developed by Douglas and Olsen (1981) or Robertson et al. (1986)
(for a standard CPT) can be used to classify the soil by the use of the friction ratio Rf and the
112  Geomechanics in soil, rock, and environmental engineering

cone resistance qc. The friction ratio is the ratio of the sleeve friction fs to the cone resistance
expressed as a percentage

Rf = (fs /qc )% (4.14)


From such a chart, the soil type can be identified as can be seen from Figure 4.36.
However, the cone resistance depends on the depth of the test and at shallow depths and
depths over 30 m the values measured are affected by overburden stress and soils may not
classify correctly using the chart of Figure 4.36. Therefore, Robertson (1990) proposed a
normalised chart that could be used with both CPT and CPTu tests where a normalised
cone resistance Qt is used.

qt − σ v0
Qt = (4.15)
σv′0

1000
10 12
Drained
11
Cone bearing pressure qc in bars

9
8
100
7
d
ne
d rai 6 Undrained
ally
i 5
Part
4
10
3

2
1
0 1 2 3 4 5 6 7 8
Friction ratio (%) Rf
Zone qc/N Soil behaviour type
(1) 2 Sensitive fine grained
(2) 1 Organic material
(3) 1 Clay
(4) 1.5 Silty clay to clay
(5) 2 Clayey silt to silty clay
(6) 2.5 Sandy silt to clayey silt
(7) 3 Silty sand to sandy silt
(8) 4 Sand to silty sand
(9) 5 Sand
(10) 6 Gravelly sand to sand
(11) 1 Very stiff fine grained*
(12) 2 Sand to clayey sand*

* Overconsolidated or cemented

Figure 4.36 Simplified soil classification chart for a standard electric friction cone. (After Robertson, P.K.
et al. 1986. Proceedings of In Situ 86, ASCE Specialty Conference, Blacksburg, Virginia.)
Site investigation and in situ testing  113

Friction sleeve u

Piezo-element

Figure 4.37 Area acted upon by water pressure at the cone tip.

In the formula, the cone value is corrected for the effect of pore pressure that acts down-
wards on the cone tip if there is a piezo-element behind the tip (Campanella and Robertson
1988). The correction is given by

qt = qc + (1 − a)u (4.16)

where a is the area ratio of the cone and u is the water pressure acting at the piezo-element.
a = d2 /D 2 where the diameters are shown in Figure 4.37.
Soil type can also be found based on pore pressure measured in the CPTu test by charts
such as the chart presented by Robertson 1990 (see Figure 4.38). In the chart, the pore pres-
sure parameter ratio Bq is defined as

∆u
Bq = (4.17)
qt − σ v0

where
Δu is excess pore pressure measured behind the tip u − u 0
qt is cone resistance corrected for pore pressure effects
σv0 is the total overburden stress

The pore pressure data are not always reliable for classification and so it is recommended
that the pore pressure ratio chart be used in conjunction with the friction ratio chart in
identifying soil type.

4.19.2 Relative density of sands


Relative density of sands is a function of the effective stress state of the sand, and can be
determined from the cone resistance qc. Relationships have been developed from calibra-
tion chamber work for predominantly quartz sands (Jamiolkowski et al. 1985) as shown in
Figure 4.39.
The compressibility of the sand has an effect on the result as can be seen from the figure, and
it should be noted that the results are only applicable to normally consolidated, uncemented,
unaged predominantly quartz sands. The cone resistance and vertical effective stress used
with the chart should be expressed in tonnes/m2 , as the ratio qc /(σ′)0.5 is not non-dimensional.
114  Geomechanics in soil, rock, and environmental engineering

1000 1000
σv0 qt
7 u0
8 7 u
g
in 9
φ′ as e,
re , ag ion
qt – σv0

No
In CR tat u − u0
σv0′

O en Bq =

rm
m qt – σv0
ce

ally
100 100 6
Normalised cone resistance

co
6

n so
lid
5

ate
5
d
4
g
sin ge
10 4 r ea d a 10 OC
c n R
In R a
C 3
O 3
g St
sin
1 c rea ivity
In sit
sen 2
1
1 1
0.1 1 10 0 0.4 0.8 1.2
Normalised fs
× 100% Pore pressure ratio Bq
friction ratio qt – σv0

1. Sensitive, fine-grained soils 6. Sands – Clean sand to silty sand


2. Organic soils – Peats 7. Gravelly sand to sand
3. Clays – Clay to silty clay 8. Very stiff sand to clayey sand*
4. Silt mixtures – Clayey silt to silty clay 9. Very stiff, fine grained*
5. Sand mixtures – Silty sand to sandy silt
* Heavily overconsolidated or cemented

Figure 4.38 Proposed soil behaviour type classification system from CPTu data. (After Robertson, P.K.
1990. Canadian Geotechnical Journal, Vol. 27, No. 1, pp. 151–158.)

For young uncemented silica sands, Kulhawy and Mayne (1990) suggest that the relative
density can be found from the more simple formula

Qtn
Dr2 = (4.18)
350 σ ′v0

and Qtn = (qt /pa)/(σ′v0/pa)0.5 is the normalised cone value and pa = 100 kPa is the atmospheric
pressure used in the non-dimensionalisation.
The value of 350 is closer to 300 for fine sands and 400 for coarse sands.

4.19.3 Friction angle of sands


The friction angle for sands changes depending on the compressibility of the sand and so
no unique relationship between cone resistance and angle of shearing resistance exists.
However, for sands that lie in the zones 7,8,9 of the classification chart of Figure 4.36, the
angle of friction can be estimated by using the chart of Figure 4.40. For overconsolidated
sands, this figure may overestimate the angle of friction by about 2o.
Alternatively, the angle of friction ϕ′ for clean rounded uncemented quartz sands can be
found from the following simple expression (Kulhawy and Mayne 1990).

φ′ = 17.6° + 11 log(Qtn ) (4.19)


where Qtn is defined in Equation 4.18.


Site investigation and in situ testing  115

100
95

85 Moderate compressibility sands


qc
Dr = −98 + 66 log10
[σ v0′ ]0.5
75
Relative density Dr (%)

qc 2
65 σ v0′ expressed in tonnes/m
Probable lower limit
Probable upper limit 2SD (low compressibility
55 (high compressibility sands)
sands)
45 2SD
Edgar sand
Hilton mine sand
35 Hokksund sand
Ottawa 90 sand
Ticino sand
25
SD = standard deviation

15
10 20 40 60 80 100 200 400 600 800 1000
qc
[σ v0′ ]0.5

Figure 4.39 Relationship between relative density and cone resistance of uncemented, normally con-
solidated quartz sands. (After Jamiolkowski, M. et  al. 1985. Theme Lecture, 11th International
Conference on Soil Mechanics and Foundation Engineering, San Francisco, Vol. 1, pp. 57–153.)

0
1 bar = 100 kPa

0.5
φ ′= 48°

1.0
Vertical effective stress σ v0′ bars

46°
1.5

2.0
44°

2.5

3.0

42°
3.5

30° 36° 40°


32° 34° 38°
4.0
0 100 200 300 400 500
Cone pressure qc bars

Figure 4.40 Proposed correlation between cone resistance and peak friction angle for uncemented quartz
sands. (After Robertson, P.K. and Campanella, R.G. 1983. Canadian Geotechnical Journal, Vol. 20,
No. 4, pp. 718–733.)
116  Geomechanics in soil, rock, and environmental engineering

4.19.4 Constrained modulus of sands


The constrained modulus Mt of a sand is the modulus under one-dimensional conditions
and is equal to 1/mv (mv is the coefficient of volume change as found from an oedometer
test – see Section 1.5).
Charts have been presented by Baldi et al. (1981) that enable the constrained modulus to
be found from the cone resistance (noting that 1 bar = 100 kPa) as shown in Figure 4.41.
This chart only applies to uncemented, normally consolidated quartz sands.

4.19.5 Young’s modulus of sands


A relationship similar to that developed for the constrained modulus has been developed for
the secant elastic modulus of a sand (Baldi et al. 1981). The secant modulus depends on the
load level and so the secant modulus is given for load levels of 25% and 50% of the ultimate
load in Figure 4.42.
It may be seen from Figure 4.42 that the assumption that E = 2qc (i.e. the straight line on
the plot) is not a bad average of the data for normally consolidated sands. This is the assump-
tion made by Schmertmann (1970) in his original approach for calculating settlement of
surface footings. He modified this to E = 2.5 or 3.5qc in his later work (Schmertmann et al.
1978) to allow for the shape of footings (i.e. whether square or strip). This is discussed in
Section 5.5.5.

4.19.6 Undrained shear strength of clays


A cone being pushed through a clay is like a deep footing loaded to cause failure of the soil.
We should therefore expect that the cone resistance can be expressed by a formula like the
bearing capacity formula

qu = Nc su + N qqs (4.20)

2000
Baldi et al. (1981)
Normally consolidated Ticino sand
Constrained tangent modulus Mt in bars

Medium dense, Dr = 46% ′ = 8 bars


v0
Dense, Dr = 70%
1500 Very dense, Dr = 90%
4 bars

Mt = qc 2 bars
1000
1 bar
0.5 bar σv′
500 σv′
є
Mt = 3qc єv
Mt
0
0 100 200 300 400 500
Cone bearing pressure qc in bars

Figure 4.41 Relationship between cone resistance and constrained modulus for normally consolidated, unce-
mented quartz sand. (After Baldi, G. et al. 1981. Proceedings of the 10th International Conference
on Soil Mechanics and Foundation Engineering, Stockholm, Vol. 2, pp. 427–432.)
Site investigation and in situ testing  117

600 900
Baldi et al. (1981)
Normally consolidated Ticino sand
Medium dense, Dr = 46% σ v0
′ = 4 bars
500 750

Drained secant Young’s modulus at


Drained secant Young’s modulus at
Dense, Dr = 70%

50% failure stress level E50 (bars)

25% failure stress level E25 (bars)


Very dense, Dr = 90%
400 E25 = 2qc 600

2 bars
300 450
1 bar
200 300
0.5 bar
є25 σDmax
σD є50
100 150
σD 0.5σDmax
єa 0.25σDmax
єa
0 0
0 100 200 300 400 500
Cone bearing pressure qc bars

Figure 4.42 Secant Young’s modulus values for uncemented, normally consolidated quartz sands. (After
Baldi, G. et al. 1981. Proceedings of the 10th International Conference on Soil Mechanics and Foundation
Engineering, Stockholm, Vol. 2, pp. 427–432.)

where
Nc and Nq are the bearing capacity factors (Nq = 1 for a clay)
su is the undrained shear strength of the clay
qs is the surcharge

Here, the surcharge is the total vertical stress σ v0 at cone level and so we can write

qc − σ v0
su = (4.21)
Nk

Nk is a bearing factor like Nc and is generally obtained from empirical correlations.


For normally consolidated marine clays, Nk ranges between 11 and 19 with an average
of 15.
For non-fissured overconsolidated clays, Nk is about 17; for stiff fissured marine clays, it
ranges from 24 to 30; and for glacial clays, it ranges from 14 to 22.
Values for Nk should be developed for individual areas based on local experience and cor-
relations with measured su values. In Sydney, values of Nk in the range 12–15 are used for
normally consolidated clays and values of 15–30 for overconsolidated clays.

4.19.7 Undrained modulus of clays


As for sands, the undrained modulus of a clay E u depends on the stress level and so it may
be estimated at say 25% of the failure load.
One way to obtain the undrained modulus of a clay is to relate the modulus to the und-
rained shear strength su that can be determined from the cone as discussed in Section 4.19.6.
The ratio of Eu /su depends on the overconsolidation ratio and the plasticity index of the clay
and so a knowledge of su is not enough to find the undrained modulus.
118  Geomechanics in soil, rock, and environmental engineering

0.6
Skempton (1957)
Ladd and Foott (1974)
0.4

su/σv0′
0.2 su /σv0′ = 0.11 + 0.0037 Iw

0
0 20 40 60 80 100 120
Plasticity index Iw

Figure 4.43 Statistical relation between su /σ′v0 and plasticity index for normally consolidated clays.

5 Range of data for 7 NC and


OC clays, with recommended
average
4
(su/σv′ 0)NC
su/σv0

1
1 1.5 2 3 4 5 6 7 8 9 10
max. past σ ′
OCR = Overconsolidation ratio present σ vm

v0

Figure 4.44 Ratio of su /σv0 for overconsolidated soil to that for normally consolidated soil versus
­overconsolidation ratio.

The recommended procedure is therefore

1. Obtain the value of su for the clay from the cone resistance as described in Section 4.19.6.
2. Estimate the stress history by computing (su /σ′v0)NC for a normally consolidated soil
from Skempton’s (1957) chart (Figure 4.43).
3. Calculate the actual su /σ′v0 for the overconsolidated soil.
4. Use the relationship of Figure 4.44 to estimate the OCR of the deposit.
5. Use the relationship of Duncan and Buchignani 1976 (Figure 4.45) to obtain the ratio
of Eu /su and hence Eu (that is really a value of Eu at 25% maximum load), because these
values are for working loads that are at about 1/4 of the failure load.

4.19.8 Permeability
The permeability of a soil can be found by performing a dissipation test with a piezo-cone.
The cone is advanced into the soil and this will generate a pore pressure. The cone is then
Site investigation and in situ testing  119

1500

1000
Ip < 30
Eu/su

500
30 < Ip< 50

Ip > 50

0
1 2 3 4 5 6 7 8 9 10
Overconsolidation ratio

Figure 4.45 Ratio of undrained Young’s modulus to shear strength against overconsolidation ratio. (After
Duncan, J.M. and Buchignani, A.L. 1976. An Engineering Manual for Settlement Studies. Department
of Civil Engineering, University of California, Berkeley.)

held stationary and the reduction of the excess pore pressure with time is recorded. A plot
can be made of the percentage dissipation of excess pore pressure with time and the time at
which 50% dissipation occurred is noted (called t50). As the permeability is dependent on
soil stiffness, a chart can be plotted which gives the permeability versus t50 for various nor-
malised cone resistances as shown in Figure 4.46 (Robertson 2010). The chart is for cones
where the piezo-element is located behind the cone tip as shown in the inset to Figure 4.46
(qt is defined in Equation 4.16 and Qt in Equation 4.15).

4.20 LIQUEFACTION POTENTIAL

In a similar fashion to the methods used for evaluating liquefaction potential using SPT
data, cone data may be used. The approach presented by Stark and Olsen (1995) is to com-
pute the corrected cone value qc1 to correspond to the cone value at 100 kPa (1 ton/ft2) by
the equation

qc1 = Cq ⋅ qc (4.22)

The overburden correction factor C q can be calculated from

1 .8
Cq = (4.23)
0.8 + (σ ′v0 /σ ′ref )
120  Geomechanics in soil, rock, and environmental engineering

10–5
qt − σv0
Qtn =
σv0

10–6
σ v0
′ = 100 kPa

10–7 σ v0
′ = 50 kPa

Permeability k (m/s)

10–8

Qtn = 2
10–9

10–10
356 mm (10 cm2)
5
10–11 u2
10
qt Qtn ≥ 14
10–12
0.1 1 10 100 1000 10,000
t50 (min)

Figure 4.46 Permeability as a function of time for 50% consolidation t 50.

where
σ′ref is a reference stress equal to one atmosphere (100 kPa)
σ′v0 is the vertical effective stress at the depth of interest

The corrected cone value can then be used with the chart of Figure 4.47 (for clean sands)
to determine whether the sand deposit is liable to liquefy. To do this, the seismic shear stress
ratio called the SSR needs to be computed. This can be found from

τ av a σ
SSR = = 0.65 max ⋅ v0 ⋅ rd (4.24)
σ 0′ g σ ′v0

where
amax is the maximum acceleration at the ground surface
g is the acceleration due to gravity
σ v0 is the total overburden pressure at the depth considered
σ ′v0 is the effective overburden pressure at the depth considered
rd is a stress reduction factor that may be computed from
rd = 1 − (0.012z) and z is the depth in metres

Soils with SSR values above the line will have the potential to liquefy (in an M = 7.5)
earthquake, whereas those below will not.
For soils that contain some fines, Stark and Olsen (1995) have presented the chart shown
in Figure 4.48. It may be seen from the chart that sands with >5% fines have a greater resis-
tance to liquefaction because it is more difficult for collapse of the sand structure to occur.
Site investigation and in situ testing  121

0.6
Field performance 0.25 < D50 (mm) < 20
Liquefaction F.C. (%) ≤ 5
0.5 No liquefaction
Proposed

Seismic shear stress ratio


0.4 relationship

Liquefaction No liquefaction
0.3

0.2

0.1 M = 7.5

0
0 5 10 15 20 25 30
Corrected CPT tip resistance qc1 (MPa)

Figure 4.47 Relationship between seismic shear stress ratio (SSR) triggering liquefaction and qc1 values for
clean sand and M = 7.5 earthquake.

0.6
Sandy silt Silty sand
D50 (mm) ≤ 0.10 0.10 ≤ D50 (mm) ≤ 0.25
0.5 F.C. (%) ≥ 35 5 < F.C. (%) ≤ 35
Seismic shear stress ratio

0.4
Clean sand 0.25
< D50 (mm) < 2.0
F.C. (%) ≤ 5
0.3

0.2

M = 7.5
0.1

0
0 5 10 15 20 25 30
Corrected CPT tip resistance qc1 (MPa)

Figure 4.48 Relationship between seismic shear stress ratio (SSR) triggering liquefaction and qc1 values for
sands containing fines and an M = 7.5 earthquake.

4.21 GEOPHYSICAL METHODS

Geophysical methods offer a rapid means of determining subsoil information by measuring


some property such as the speed of seismic waves or the electrical resistivity of the soil. As
such, the methods are largely non-destructive, that is, holes do not have to be dug or drilled.
122  Geomechanics in soil, rock, and environmental engineering

Down-hole or cross-hole techniques are the exception, where boreholes are used to deter-
mine what lies between the holes.
Engineering geophysical methods can be used to determine thickness of strata, to map
contamination plumes, find depth to bedrock, to find channels and cavities, and find buried
waste among many other things.
Geophysical surveys can also be used to plan conventional investigations that involve
boreholes or to interpolate data between boreholes.
The type of geophysical method used will depend on the type of problem that is to be inves-
tigated. Table 4.2 shows some of the various geophysical techniques, and their applications.
Offshore geophysical methods are in some ways different to those used on land, and
include echo sounding, the use of pingers and sparkers, and side scan radar.
Waves generated can be compression (P) waves or (S) waves. For a P wave, the motion of
particles is in the direction of propagation of the wave, whereas for a shear wave, the motion
of particles is in a direction perpendicular to the direction of travel of the wave.

4.21.1 Seismic surveys
Seismic surveys involve creating a shock wave in the soil or rock through means such as a
sledge hammer or explosives, and recording the arrival of P waves generated by geophones.
Waves generated by the seismic source, travel through the soil and can be either reflected
or refracted by different soil interfaces.

Table 4.2  S ummary of geophysical methods (on land)


Method Principal characteristics Applicability and limitations
Seismic Refraction of seismic waves at Preliminary investigation for major
refraction interfaces of different materials stratigraphic units and depth to bedrock.
Limited by hard layer over soft layer
High resolution Arrival times of seismic waves Deep bedrock identification. Useful for
reflection reflected from interfaces of adjoining locating groundwater
strata
Vibration Travel time of transverse shear waves Stratigraphy in terms of soil type and
thickness and dynamic properties
Up-hole, Measurement of travel times. Dynamic ground properties at small strain,
down-hole, and Geophones on surface and energy rock quality, cavity detection. Unreliable for
cross-hole source down-hole or vice versa. irregular strata or soft strata with high gravel
Energy in central hole and geophones content
in surrounding holes
Electrical Difference in electrical conductivity or Horizontal extent and variation of strata and
resistivity resistivity of various strata measured depths to 30 m. Granular material searches.
from surface Fresh/salt water boundaries, clay over
bedrock, corrosivity of soils. Polluted regions
Drop in Ratio of potential drops between Similar to resistivity but gives clear definition
potential electrodes as a function of current on vertical or inclined boundaries
imposed
E-logs Difference in resistivity and Correlation of units between boreholes
conductivity measured in boreholes
Magnetic Highly sensitive proton magnetometer Delineation of faults, bedrock, buried utilities,
measures Earth’s magnetic field and steel drums in ground
Gravity Differences in density of subsurface Voids or cavities in bedrock, tracing steeply
materials as indicated by changes in inclined and irregular features such as faults,
gravitational field intrusions, and domes
Site investigation and in situ testing  123

4.21.2 Reflection surveys
Figure 4.49 shows the principle of a reflection survey. The waves from the source travel
down to an interface and are reflected back up to the surface and recorded by the geophones.
The time for arrival at the geophone at a distance x from the source is given by

2SR 2
Tx = = (h12 + x2 / 4) (4.25)
v1 v1

or

Tx2 = T02 + x2 /v12 (4.26)


where

T0 = 2h1 /v1 (4.27)

A plot of T 2 versus x 2 will yield a straight line as shown in Figure 4.50 where the slope is
1/v12 and the intercept is given by 4h12 /v12.

T
Tx′

Tx

T0

0
S G x′ G′
x

R0 R R′

Figure 4.49 Seismic reflection.

T2

Intercept Slope = 1/v12


= 4h12/v12

x2

Figure 4.50 Graphical method of obtaining wave velocity and depth to firm stratum.
124  Geomechanics in soil, rock, and environmental engineering

The slope therefore yields the velocity of the wave and this can be used to obtain the thick-
ness of the layer h1 from the intercept.

4.21.3 Seismic refraction
In the case of seismic refraction, a seismic wave travels down until it hits a stratum with a
different seismic velocity. It will then be refracted at the interface between the two layers,
and then back up to the surface. If the velocity of travel in layer 2 is faster than in layer 1,
the refracted wave will arrive back at the surface first, that is, before the direct surface wave.
This is shown in Figure 4.51 where the time of arrival at the geophones T is plotted
against the distance from the source x. This will also occur for a third layer if the velocity
in that layer is higher, that is, v3 > v2 > v1.
Velocities of travel in different materials are given in Table 4.3.
For a two-layer system, the thickness of the upper layer can be determined from

x1 v2 − v1
h1 = (4.28)
2 v2 + v1

where x1 is the distance to the first crossover point. The velocities are found from the slope
of the plot (see Figure 4.51).

1
v3
1
Ti3 v2

Ti2 1
v1

v1 h1

v2 h2

v3 v3 > v2 > v1

Figure 4.51 Results of seismic refraction survey.

Table 4.3  S eismic velocities in various soil and rock types


Material Velocity (m/s)
Loose sand (above water table) 250–600
Hard clay 600–1200
Soft shale 1200–2100
Soft sandstone 1500–2100
Basalt 2400–4000
Granite 3000–6000
Site investigation and in situ testing  125

T (ms) T (ms)

400 /s
28 m
70 00
m 32
/s

200

s
m/

22
80
Tiu

50
17

m
Tid

/s
x (m)
0
0 500 1000 1500

Figure 4.52 Reversed profile for strata that is slightly dipping.

The method can be generalised for multiple layers as long as the velocity in each of the
layers is increasing with depth.

n −1  hj vn2 − v 2j 
Ti(n) = 2 ∑
j =1

 v j vn


(4.29)

The times Ti(n) are the intercepts on the time axis as shown in Figure 4.51. Equation 4.29
can then be used to progressively calculate the layer thicknesses from the velocities (which
are obtained from the slope of the segments of the plot).
Cross shooting can be used to obtain two sets of data for the soil profile. The seismic
source (i.e. the sledge hammer) is used at either end of the array of geophones. A plot like the
one shown in Figure 4.52 is obtained. This data can be used to identify dipping strata as the
forward and reverse profiles will not be the same in this case (see Sharma 1997).

4.21.4 Rippability of rock
Seismic velocity data can be used for determining when rock can be ripped with a bulldozer
or when drilling and blasting will be necessary. Higher velocities indicate harder rock that
cannot be ripped. Figure 4.53 shows a chart developed by the Caterpillar Tractor Company
that is applicable to a D9 bulldozer.
Rippability of rock also depends on the jointing in the rock, as highly jointed rock con-
sists of small blocks of rock that can be ripped out individually. This needs to be taken into
account as well as seismic velocity.

4.22 RESISTIVITY

The resistivity of different materials is shown in Figure 4.54. Dense rocks have high resistivi-
ties while sands or gravels that contain water will have low resistivities. Salt water has a very
low resistivity, and so this method can be used to locate salt water intrusions.
126  Geomechanics in soil, rock, and environmental engineering

Velocity in feet per second × 1000


0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
TOPSOIL
CLAY
GLACIAL TILL
IGNEOUS ROCKS
Granite
Basalt
Trap rock
SEDIMENTARY ROCKS
Shale
Sandstone
Siltstone
Claystone
Conglomerate
Breccia
Caliche
Limestone
METAMORPHIC ROCKS
Schist
Slate
MINERALS and ORES
Coal
Iron ore

Rippable Marginal Non-rippable

Figure 4.53 Seismic rippability chart for a D9 Caterpillar bulldozer with a No. 9 ripper.

Rocks with fractures and fissures that contain water will also exhibit higher conductivi-
ties. Therefore, low conductivity can be an indication of faulting, shearing, weathering, or
hydrothermal alteration of rocks.
Resistivity of soil and rock can be measured by passing an electrical current through the
ground, and measuring the voltage between two electrodes. Common layouts of the elec-
trode arrays are the Wenner and Schlumberger layouts (see Figure 4.55).

Resistivity (Ohm-m)
0.01 0.1 1 10 100 1000 10,000 100,000

Unweathered
Massive Igneous and rocks
sulphides Graphite metamorphic rocks
Igneous rocks: mafic Felsic Duricrust
Weathered
layer
Metamorphic rocks
Clays Gravel and sand
Glacial
sediments
Tills
Shales Sandstone Conglomerate
Sedimentary
rocks
Lignite, coal Dolomite, limestone
Salt water Fresh water Permafrost Water,
aquifers

Sea ice
100,000 10,000 1000 100 10 1 0.1 0.01
Conductivity (ms/m)

Figure 4.54 Typical resistivities of soil and rock.


Site investigation and in situ testing  127

(a) Battery Current (Amps)

Voltage

a a a

Battery Current (Amps)


(b)

Voltage

2L

Figure 4.55 Wenner (a) and Schlumberger (b) electrode layouts.

i. For the Wenner configuration, the resistivity (ohm m or ohm ft) is given by

V 
ρ = 2πa   (4.30)
I

where I is the current and V is the voltage measured between the electrodes.
ii. For the Schlumberger configuration

πL2 V
ρ = (4.31)
I 2

The resistivity that is calculated is an apparent resistivity as it is a weighted average of the


material within the zone where the current is flowing. The depth of material affected is
roughly the same as the spacing of the electrodes.
With the electrical profiling method, the spacing of the electrodes is kept constant, and
the whole configuration is moved (see Figure 4.56). At each new position, the resistivity is
found and may be plotted as shown in Figure 4.57.

Ground surface
Electrodes

1 2 3 4
Location of lines

Figure 4.56 Electrical profiling method.


128  Geomechanics in soil, rock, and environmental engineering

Position of array (m)


100 200
500

Apparent resistivity ρ (ohm-feet)

Apparent resistivity ρ (ohm-m)


400
a = 20 feet (6.1 m)
300 100

200
50
100 a = 50 feet (15.2 m)

0
100 200 300 400 500 600 700 800 900
Position of centre of electrode array (feet)

Figure 4.57 Resistivity versus position of array (profiling method).

An idea of the change in material type can be obtained from this method and the outline
of each material can be marked on a map.

4.22.1 Electrical sounding method


For this approach, the spacings of the electrodes are increased about the same central posi-
tion. As the spacing is increased, so is the depth of the area through which the current flows.
This is shown in Figure 4.58.
If on spreading the electrodes, the resistivity becomes higher, there must be a layer of
material with high resistivity below a soil with a low resistivity. If there is a material with
low resistivity below one having a high resistivity, then the opposite will occur, that is, the
resistivity will rise as the electrode array is spread. This is shown in Figure 4.59 for a number
of different cases.
The electrical profiling method is therefore capable of detecting softer layers beneath
harder layers, something that the seismic refraction method cannot achieve.

The position of the centre


of the array is fixed

a1 a1 a1

a2 a2 a2

a3 a3 a3

Figure 4.58 The electrical sounding method.


Site investigation and in situ testing  129

Ground surface Ground surface

Higher Layer 1 ρ Lower Layer 1 ρ

Lower Layer 2 Higher Layer 2


a a

Ground surface Ground surface

High Layer 1 Low Layer 1


ρ ρ
Low Layer 2 Medium Layer 2

Medium Layer 3 High Layer 3


a a

Figure 4.59 Resistivity curves for different subsurface conditions.

4.22.2 Push-in resistivity instruments


Resistivity can be measured between the anode and cathode mounted on push-in instruments
such as the piezo-cone and the dilatometer. One such device called an RCPTu (resistivity
cone penetration test with pore pressure measurement) is shown in Figure 4.60 (Campanella
2008). The module can measure the electrical resistance to current flow in the ground on a
continuous basis, and is useful in environmental work as it can be used to detect regions in
the ground where pollution may be present.

Pushed at Cone
2 cm/s

Insulator

10.5 sq cm
350 mm 15 mm ‘isolated’
150 mm resistivity

Electrodes

Accelerometer

U3
650 mm

U2
U1

Figure 4.60 Resistivity module attached behind a piezo-cone.


130  Geomechanics in soil, rock, and environmental engineering

4.23 MAGNETIC SURVEYING

Magnetometers are used for mapping the intensity of the earth’s magnetic field at various
points. A contour map can be made showing the intensity of the magnetic field and areas of
magnetic intensity can be used to identify certain types of rocks, that is, those high in iron
content. Laboratory measurement of magnetic susceptibility of rock types in the area can
be used to interpret results.
Engineering applications include investigations over landfills where buried barrels, pipes,
and domestic rubbish can be identified by more intense magnetic zones.

4.24 GROUND PROBING RADAR

Ground probing radar may be used for imaging at shallow depths. It makes use of electro-
magnetic waves in the frequency band of 10–1000 MHz. The electromagnetic waves are
detected after they are reflected from the different strata in the soil, for example, dry soil,
wet soil, or bedrock.
The method is most effective with low-attenuation media such as ice, sand, crude oil,
bedrock, or fresh water, but is less effective for materials such as wet clay or silt, or salt
water.
The radar pulses are reflected from different materials in much the same way as seismic
reflection signals, and therefore can be processed using similar methods (see Section 4.21.1).

4.25 SEISMIC BOREHOLE TECHNIQUES

Seismic techniques can be used without having to drill a borehole such as those discussed in
Section 4.21.1. However, there are methods that make use of boreholes into which a seismic
detector is lowered and seismic waves generated either at the ground surface or in another
borehole.
The seismic shear wave velocity can be used to compute the small-strain shear modulus of
the soil since the amount of strain generated by the waves is very small. Small-strain mod-
uli can be used in foundation vibration analysis, earthquake analysis, and in the analysis of
retaining structures. The shear modulus can be reduced in magnitude and used for large strain
analysis such as pile settlement. This is discussed further in Section 6.11.3 in Chapter 6.

4.25.1 Down-hole seismic testing


Down-hole seismic testing is performed by drilling a borehole down in which a probe can be
lowered (see Figure 4.61). The vertical seismic shear wave test is performed using pre-drilled
holes (50 mm minimum diameter) which may be cased in PVC and grouted around the out-
side, or uncased. Seismic shear (S) waves are generated at the surface near the top of the hole
using impacts on a wooden plank generally weighted by a vehicle and about 1 m from the
hole. Seismic compressional (P) waves are generated separately by impacting a metal plate.
Both types of waves are detected using an in-hole, geophone probe which is air-packed or
wedged against the hole wall at a pre-selected depth. Individual P and S wave travel times to
the detector are used to compute average P and S wave seismic velocities between the surface
sources and in-hole detector (see ASTM D7400-08 2008).
The seismic velocity of either the shear S wave or the P wave (at two locations down the
hole) is simply calculated by dividing the distance between the two readings in the hole
Site investigation and in situ testing  131

Hammer with impact


Seismograph
switch

Wooden plank
weighted down
by vehicle

Direction of particle
vibration
Borehole
Shear wave
propagates down

Geophone held against


side of borehole

Figure 4.61 Down-hole seismic test.

Δd  by the difference in arrival times Δt of the waves at the two locations, for example,
vs = Δd/Δt. The S wave velocity can be used to compute the small-strain shear modulus as
shown in Equation 4.33.

4.26  CROSS-HOLE TECHNIQUES

For the cross-hole test, two boreholes are drilled typically 3 m apart. By creating a shear
wave in one borehole (by using an impulse rod at the bottom of the hole), and recording the
arrival time in a neighbouring borehole, the shear modulus of the soil can be determined (see
Figure 4.62). Once again, this modulus is a small-strain modulus as the shear waves cause
only small shear strains in the soil.
The velocity of the shear waves can be found from

L
vs = (4.32)
t

and from the shear wave velocity, the shear modulus can be found.

vs2 γ
G = (4.33)
g
132  Geomechanics in soil, rock, and environmental engineering

Seismograph
Cable to
moving
weight in
hammer

Borehole Borehole

Direction of wave
propagation

Geophone Down-hole
Sidewall clamp hammer
Direction of
particle motion

Figure 4.62 Cross-hole seismic survey method.

where
vs is the shear wave velocity
G is the shear modulus of the soil
γ is the unit weight of the soil
g is the acceleration due to gravity
L is the spacing of the boreholes
t is the time for the wave to travel the distance L

Poisson’s ratio ν can be found if the P wave velocity v p is known

(vp /vs )2 − 2
ν = (4.34)
2(vp /vs )2 − 2

The standard for performing this test is ASTM D4428/D4428M-07 (2007) and more
details on the test can also be found in Stokoe and Woods (1972).

4.27 OTHER SEISMIC DEVICES

Seismic sources may also be generated by push-in devices such as the dilatometer or the
cone penetrometer. One type of seismic cone test is very similar to the down-hole seismic
test where the surface of the ground is excited by a sledge hammer, but where the geophone
receiver is attached to the cone (usually a piezo-cone).
Site investigation and in situ testing  133

A seismic dilatometer SDMT is the combination of the standard flat dilatometer (DMT)
with a seismic module. The module is a probe outfitted with two sensors, spaced 0.5 m
apart, for measuring the shear wave velocity vs. Again, the source of seismic waves is a
surface-based source such as a sledge hammer. The seismic velocity is calculated from the
distance between the two sensors and the time difference that it takes for the shear wave to
reach each of the sensors.
This page intentionally left blank
Chapter 5

Shallow foundations

5.1 INTRODUCTION

A foundation is usually termed ‘shallow’ if the base of the foundation is at a depth that is less
than the breadth of the foundation. This type of footing is suitable where structural loads
are not high, or the soil is strong enough to support the applied loads with a foundation of
moderate depth.
The purpose of a shallow foundation is to apply the structural loads to the foundation by
spreading the load from walls or columns over a larger area. The size of the footing can be
selected so that the contact pressure between the footing and the soil is small enough so as
to reduce settlement to acceptable levels and provide an adequate factor of safety against the
possibility of a bearing failure.
Shallow foundations may be of several different types, and the selection of the appropriate
type will depend on the structural loads and the foundation conditions. Some commonly
used shallow foundation types are discussed in Section 5.2.

5.2  TYPES OF SHALLOW FOUNDATIONS

5.2.1  Strip footings


A strip footing is one that is long compared to its width. Such foundations (see Figure 5.1a)
are typically used to support continuous masonry walls.
Strip footings are constructed by excavating a trench with a backhoe (or by hand), placing
a reinforcing cage and pouring concrete into the trench. If the trench is able to stand open
without collapsing (e.g. clayey soils) then formwork is not needed or may only be needed
to form that part of the foundation that is above ground level. For sandy soils, the trench
has to be excavated and fully formed before concrete can be poured as the trench will not
stay open.

5.2.2  Pad footings


For support of isolated loads such as column loads or support of raised slabs or flooring,
pad footings may be used (Figure 5.1c). The process of construction is similar to that for
strip footings as the hole dug for the footing can be filled with concrete without the need
for formwork (if the hole will stay open without collapsing). The hole should be inspected
before placement of concrete to ensure that it does not contain any water or soil that has
fallen in from the sides.

135
136  Geomechanics in soil, rock, and environmental engineering

(a) (b)

Property line

Strip footing Combined footing

(c)
Plan

(d)

Slab fabric

Void or polystyrene

Pad footing Waffle slab

Figure 5.1 Types of shallow foundations.

It is also generally recommended that in sandy areas or on sites that may be subjected to
wind or water erosion, footings should be constructed at least 300 mm below the surface to
prevent any undermining.

5.2.3  Combined footings


A combined footing may be necessary when a column load is close to a property boundary,
and the foundation cannot be made large enough to support the load without encroaching
onto the neighbouring property (Figure 5.1b). In this case, the footing is combined or tied to
a nearby footing that provides additional support.

5.2.4  Raft or mat foundations


For special conditions (i.e. in areas of swelling soils) or where pad footings would need to
be large to support loads, a raft or mat foundation may be used. Rafts can be reinforced
concrete slabs of uniform thickness, which for tall buildings can be several metres thick.
For smaller scale construction, rafts can be stiffened under walls and columns by making
the raft thicker at the point of loading. In areas where reactive soils exist, it is common to
construct a ‘waffle’ slab, so called because it consists of stiffened ribs running perpendicular
to each other under the slab (Figure 5.1d). The upper flat portion of waffle slabs can be cast
on polystyrene blocks so that swelling of the soil does not affect the slab or raft but merely
compresses the polystyrene.

5.3  BEARING CAPACITY

Theoretical solutions for the bearing capacity of shallow foundations are generally based on
the theory of plasticity where the soil is assumed to behave as a rigid plastic material that
fails according to the Mohr–Coulomb failure criterion. Shallow foundations may be defined
Shallow foundations  137

as those which are founded at a depth Df which is less than the full width B of the founda-
tion (B may be taken as the width of a strip, the diameter of a circle or the smaller dimension
of a rectangular footing). Solutions are obtained by establishing the stress characteristic
fields beneath a foundation. The stress characteristics are lines that show the orientation of
the planes on which the stresses in the soil have reached the critical combination τf and σf as
shown in Figure 5.2. The method is described in Davis and Booker (1971) and in the book
by Hill (1983) and in Wu (1976).
Many solutions to bearing capacity problems have been obtained using classical plasticity
theory; however, for more complex problems (i.e. layered soils, eccentric loadings), approxi-
mate methods may need to be used to obtain solutions. Such methods include those based on
the bound theories of plasticity whereby upper and lower bounds to the collapse load may be
found. The true collapse load therefore lies between the bounds that have been established.
An example of the use of this technique is given by Sloan and Yu (1996) who use linear
programming techniques to establish upper and lower bounds to bearing capacity problems.
Other methods include finite element techniques (see Zienkiewicz 1977), finite differ-
ence methods as used in the commercially available computer program FLAC3D (Ver. 5.0)
(ITASCA Consulting Group, Inc. 2014) or optimisation codes (OPTUMG2 2014). Obviously,
a great deal more effort is involved in setting up finite element meshes or finite difference
grids and computing ultimate loads than in using the results of a theoretical solution, and
so such numerical techniques are limited to problems involving complex geometry, loading,
or material properties. A further advantage of numerical techniques is that advanced con-
stitutive laws governing deformation and plastic behaviour of the soil can be incorporated.
Analyses can therefore give predictions of the load-deformation behaviour of a foundation,
rather than just the collapse load.
Solutions have been found to particular types of bearing capacity problems using differ-
ent solution methods. In the following, solutions are presented to various bearing capacity
problems for surface footings, and where applicable, the solution technique is mentioned.

Smooth Qf

45° + ϕ/2 A qs

45° – ϕ/2
| τf | = σf tan ϕ along these lines
Log spiral

Direction of failure planes


τ τ
τf
σf , τf
ϕ ϕ 45° – ϕ/2
45° + ϕ/2
Pole qs Pole
σf σ σ
−τf
−τf

At point A At point B

Figure 5.2 Slip lines for smooth strip footing on cohesionless, weightless soil.
138  Geomechanics in soil, rock, and environmental engineering

5.3.1  Uniform soils


For soil layers that can be considered to be of infinite depth and having uniform strength
with depth (i.e. the cohesion c and angle of shearing resistance ϕ are constant) the well-
known Terzaghi equation may be used (Equation 5.1 – Terzaghi 1943). This equation applies
to strip footings of width B subjected to vertical loading as shown in Figure 5.3a.

1
qu = cNc + γBN γ + qs N q (5.1)
2

where
qu is the ultimate pressure that can be carried by the foundation
qs is the surcharge acting on the surface of the soil
B is the full width of the strip
γ is the unit weight of the soil below the foundation level
c is the cohesive strength of the soil
Nc , N γ , and Nq are bearing capacity factors

The bearing capacity factors depend on the angle of shearing resistance of the soil ϕ.
Values have been presented by Terzaghi (1943) but were approximate. More modern values
of the bearing capacity factors have been computed using slip line theory (Martin 2005) and
are shown in Figure 5.4 and are presented in Table 5.1.
The formula of Equation 5.1 can be seen to consist of three terms, the first arising from
the cohesive strength of the material, the second from the self-weight of the soil beneath
the foundation, and the third from the surcharge considered to be applied at foundation
level (see Figure 5.3a). For foundations that are not on the surface of the soil, the surcharge
is considered to be equal to the pressure exerted by the soil above foundation level. Plastic
failure is therefore not considered in the soil above foundation level and so some error
exists because of this assumption, however, for most practical purposes the error is small.
Figure 5.3b shows the slip lines for the case where the soil above footing level is considered.
Shearing would have to take place through the surcharge in this case, as can be seen from the
figure. Terzaghi used superposition to add together the effects of the three terms as shown
in Equation 5.1. This is an approximation, and improved estimation of the bearing capac-
ity can be obtained by using the theory of plasticity and slip lines, and performing a single
calculation considering all components at the same time. Computer code for performing the
calculation has been provided at the Web site http://www.eng.ox.ac.uk/civil/people/cmm by
C. Martin. An example of the slip lines for a smooth strip footing is shown in Figure 5.5,
where the parameters used are c = 5 kPa, ϕ = 30°, γ = 18 kN/m3, qs = 20 kPa, and B = 2 m.
The calculated bearing capacity for this case is 745.9 kPa for a smooth footing and
936.8 kPa for a rough footing. Note that for this calculation, a greater number of slip lines
was used than shown in Figure 5.5, to improve the accuracy of the computed bearing capacity.

(a) B (b)

qu qs = γD D D
a
d
b
c

Figure 5.3 Failure mode for footing at depth D: (a) Terzaghi’s assumption; (b) actual failure mode.
Shallow foundations  139

(a) 40

35
Nq
30
ϕ (in degrees)

25 Nc
20

15

10

0
0 10 20 30 40 50 60
Nq or Nc

(b) 40

35 Smooth
30 Rough
ϕ (in degrees)

25

20

15

10

0
0 10 20 30 40 50 60 70 80 90

Figure 5.4 Bearing capacity factors for use in Equation 5.1 (i.e. using superposition). (After Martin, C.M. 2004.
Program ABC [Analysis of Bearing Capacity] V1.0. http://www.eng.ox.ac.uk/civil/people/cmm/software.)

The ultimate bearing capacity of a foundation depends upon whether it is loaded rapidly
so that the soil does not drain or whether it is loaded slowly so that all pore pressures have
dissipated.

Undrained Case
For the undrained case, the undrained cohesion su and angle of shearing resistance ϕu
are used with Equation 5.1 (i.e. c = su , ϕ = ϕu). If ϕu = 0, then Nc = 5.14 = (2 + π), N γ = 0, and
Nq = 1 and so the bearing capacity equation becomes
qu = 5.14su + γ bulkD (5.2)

where D is the depth of the foundation and γbulk is the bulk unit weight of the soil above
foundation level.

Drained Case
For the drained loading case, the drained strength parameters are used with Equation 5.1
(i.e. c = c′, ϕ = ϕ′), however the drained case is a little more complex than the undrained case
140  Geomechanics in soil, rock, and environmental engineering

Table 5.1  Bearing capacity factors obtained from plasticity theory


ϕ Nc Nq Nγ
(o) δ/ϕ = any δ/ϕ = any δ/ϕ = 0 δ/ϕ = 1/3 δ/ϕ = 1/2 δ/ϕ = 2/3 δ/ϕ = 1
0 5.14159 1.00000 0.00000 0.00000 0.00000 0.00000 0.0000
1 5.37926 1.09390 0.0106339 0.0110586 0.0112596 0.0114539 0.0118240
2 5.63160 1.19666 0.0242179 0.0257878 0.0265319 0.0272506 0.0286045
3 5.89977 1.30919 0.0408212 0.0443525 0.0460261 0.0476371 0.0506295
4 6.18504 1.43250 0.0607622 0.0672331 0.0702954 0.0732293 0.0785916
5 6.48882 1.56770 0.0844649 0.0950574 0.100057 0.104821 0.113371
6 6.81264 1.71604 0.112443 0.128586 0.13618 0.143368 0.156020
7 7.15820 1.87892 0.145304 0.168723 0.179693 0.190002 0.207770
8 7.52736 2.05790 0.183757 0.216532 0.231807 0.24605 0.270054
9 7.92217 2.25475 0.228629 0.273262 0.293945 0.313066 0.344540
10 8.34493 2.47144 0.280879 0.340379 0.367775 0.392867 0.433164
11 8.79814 2.71019 0.341627 0.419603 0.455253 0.487578 0.538175
12 9.28461 2.97351 0.412173 0.512957 0.558674 0.599689 0.662191
13 9.80746 3.26423 0.494036 0.622817 0.680737 0.732111 0.808259
14 10.3701 3.58556 0.588986 0.751982 0.824616 0.888264 0.979939
15 10.9765 3.94115 0.699096 0.903758 0.994049 1.07216 1.18139
16 11.6309 4.33511 0.826793 1.08205 1.19345 1.28852 1.41748
17 12.3381 4.77215 0.974928 1.29147 1.42803 1.54291 1.69393
18 13.1037 5.25764 1.14685 1.53753 1.70397 1.8419 2.01746
19 13.9336 5.79771 1.34653 1.82673 2.02861 2.19325 2.39600
20 14.8347 6.39939 1.57862 2.16686 2.41065 2.60618 2.83894
21 15.8149 7.07076 1.84869 2.56721 2.8605 3.09162 3.35737
22 16.8829 7.82112 2.16332 3.03892 3.39057 3.6626 3.96449
23 18.0486 8.66119 2.53035 3.59535 4.01573 4.33468 4.67604
24 19.3235 9.60339 2.95919 4.25262 4.75384 5.12649 5.51080
25 20.7205 10.6621 3.46108 5.03017 5.62641 6.06038 6.49131
26 22.2544 11.8542 4.04956 5.95158 6.65942 7.16328 7.64467
27 23.9422 13.1991 4.74097 7.04550 7.88433 8.46773 9.00358
28 25.8033 14.7199 5.55510 8.34686 9.33941 10.0132 10.6076
29 27.8605 16.4433 6.51599 9.89841 11.0713 11.8475 12.5050
30 30.1396 18.4011 7.65300 11.7527 13.1371 14.0294 14.7543
31 32.6711 20.6308 9.00208 13.9744 15.6069 16.6306 17.4275
32 35.4903 23.1768 10.6074 16.6437 18.5673 19.7393 20.6131
33 38.6383 26.092 12.5237 19.8603 22.1254 23.4649 24.4203
34 42.1637 29.4398 14.8188 23.7485 26.4145 27.9427 28.9849
35 46.1236 33.2961 17.5771 28.4643 31.6012 33.3421 34.4761
36 50.5855 37.7525 20.9049 34.2044 37.8947 39.8748 41.1059
37 55.6296 42.9199 24.9357 41.2180 45.5591 47.8083 49.1416
38 61.3518 48.9333 29.8388 49.8224 54.9296 57.4811 58.9219
39 67.8668 55.9575 35.8302 60.4242 66.4339 69.3250 70.8787
40 75.3131 64.1952 43.1866 73.5471 80.6214 83.8937 85.5656
41 83.8583 73.8969 52.2656 89.8703 98.2022 101.902 103.697
42 93.7064 85.3736 63.5316 110.280 120.100 124.280 126.203
(Continued)
Shallow foundations  141

Table 5.1 (Continued)  Bearing capacity factors obtained from plasticity theory
ϕ Nc Nq Nγ
(o) δ/ϕ = any δ/ϕ = any δ/ϕ = 0 δ/ϕ = 1/3 δ/ϕ = 1/2 δ/ϕ = 2/3 δ/ϕ = 1
43 105.107 99.0143 77.5929 135.943 147.525 152.243 154.30
44 118.369 115.308 95.2519 168.401 182.075 187.396 189.592
45 133.874 134.874 117.576 209.715 225.876 231.874 234.213
46 152.098 158.502 145.996 262.657 281.784 288.541 291.026
47 173.640 187.206 182.449 330.993 353.662 361.270 363.907
48 199.259 222.300 229.584 419.882 446.795 455.359 458.150
49 229.924 265.497 291.056 536.469 568.482 578.119 581.067
50 266.882 319.057 371.967 690.752 728.912 739.755 742.863
51 311.752 385.982 479.523 896.883 942.480 954.679 957.947
52 366.660 470.304 624.024 1175.14 1229.77 1243.49 1246.92
53 434.421 577.496 820.392 1554.95 1620.59 1636.04 1639.63
54 518.805 715.074 1090.56 2079.65 2158.80 2176.18 2179.93
55 624.924 893.484 1467.23 2814.00 2909.77 2929.35 2933.25
56 759.793 1127.44 2000.05 3856.36 3972.74 3994.8 3998.85
57 933.170 1437.96 2765.60 5358.81 5500.86 5525.73 5529.93
58 1158.83 1855.52 3884.45 7560.90 7735.20 7763.26 7767.58
59 1456.54 2425.08 5550.24 10847.9 11063.0 11094.7 11099.1
60 1855.10 3214.14 8081.21 15853.6 16120.6 16156.5 16161.0
Source: After Martin, C.M. 2004. Program ABC (Analysis of Bearing Capacity) V1.0. http://www.eng.ox.ac.uk/civil/people/
cmm/software.
Note: δ is the angle of friction between the footing and the soil ϕ is the angle of friction of the soil.

Figure 5.5 Slip lines computed for a smooth strip footing. (After Martin, C.M. 2004. Program ABC [Analysis
of Bearing Capacity] V1.0. http://www.eng.ox.ac.uk/civil/people/cmm/software.)

if the water table is in the vicinity of the foundation. Figure 5.6 shows several different cases
for the level of the water and for each case, the method of calculating the total ultimate pres-
sure that can be applied at foundation level is given in Equations 5.3.

Case 1: qu = c ′N c + 0.5γ bulkBN γ + γ bulkDN q


Case 2: qu = c ′N c + 0.5γ subBN γ + γ bulkDN q
Case 3: qu = c ′N c + 0.5γ subBN γ + [ γ bulk (D − d) + γ subd ] N q + γ w d (5.3)
Case 4: qu = c ′N c + 0.5γ subBN γ + γ subDN q + γ w D
Case 5: qu = c ′N c + 0.5γ subBN γ + γ subDN q + γ w (D + h)

142  Geomechanics in soil, rock, and environmental engineering

B 5
h Ground level
4
qu 3
D d
2
zm
Intermediate case
0 ≤ zm ≤ B
Water level
1 Well below foundation
zm ≥ B

Figure 5.6 Location of a water table beneath a footing.

where
γsub = the submerged unit weight of the soil
γw = the unit weight of water
γbulk = the bulk unit weight of the soil
D, d, h = the distances shown in Figure 5.6

It may be noted that qu is the total stress that can be applied at foundation level and therefore
in Equations 5.3, the effective ultimate bearing pressure has been calculated and the water
pressure added if the water is above foundation level.
If the water level is at a depth of less than one foundation width B beneath the footing, but
below the footing base (i.e. intermediate of Cases 1 and 2), simple empirical corrections are
sometimes made to allow for the changed unit weight of the soil beneath the footing. This
assumes that the water level only begins to affect the unit weight of the soil being pushed
aside by the foundation when it reaches a distance of B beneath it, and that as the water
table rises further, the overall unit weight of the soil can be computed from a simple linear
interpolation, for example,

zm
γ = γ sub + (γ bulk − γ sub )
B (5.4)

where γbulk is taken as the unit weight of soil corresponding to the minimum water con-
tent above the water table and zm is the distance of the water level beneath the footing (see
Figure 5.6).
The unit weight calculated from Equation 5.4 is used in the self-weight term (N γ term) of
the bearing capacity equation.

Effect of Footing Shape

The Terzaghi bearing capacity factors were computed for strip foundations (i.e. the length
of the foundation is large compared to its width) and therefore do not apply to footings of
other shapes. For footings that are square or circular in plan, Terzaghi and Peck (1967)
proposed the following formulae

qu = 1.2cNc + 0.6 γRN γ + qs N q circular


(5.5)
qu = 1.2cNc + 0.4γBN γ + qs N q square

where R is the radius of a circular footing and B the full width of a square footing.
The effect of footing shape is discussed more fully in Section 5.3.1.2.
Shallow foundations  143

Net Bearing Capacity

The value of bearing capacity qu as computed from equations such as Equations 5.1 or
5.3 give the ultimate total pressure that can be applied at foundation level. If a hole is exca-
vated for a foundation, the stress at the foundation level is reduced by an amount equal to
the overburden pressure. Therefore, we should be able to apply a stress at foundation level
equal to the overburden pressure to return to initial conditions. It is only stresses above the
overburden pressure that will contribute to a bearing failure.
This observation leads to the concept of net bearing pressure that is useful in comput-
ing the allowable loads that can be applied to a foundation. The net bearing capacity is
defined as

qnet u = qu − qob (5.6)


where
qnet u  = the ultimate net load
qu = the bearing capacity
qob = the total overburden stress at foundation level

The allowable load qall that can be applied (at foundation level) can be computed from

qnet u
qall = + qob (5.7)
F

where F is the factor of safety.


Equation 5.7 shows that if the factor of safety were infinitely large, then the allowable
load is equal to the overburden pressure. This implies that if a load equal to the excavated
soil were applied, there is no possibility of failure (i.e. F is infinite) since the initial conditions
have been reinstated.

5.3.1.1  General formulae


Loads on foundations may not always be vertical or applied at the central point of the foun-
dation or the foundation may not be on level ground. In order to take into account many
other factors that may affect the bearing capacity, the following general bearing capacity
equation (Equation 5.8) has been proposed by Vesic (1973, 1975).

1
qu = cNcζc + γBN γ ζ γ + qs N qζ q (5.8)
2

where the correction multipliers ζ are obtained from Table 5.2 by multiplying together indi-
vidual correction factors if they are applicable, for example,

ζc = ζcs ζciζ cd ζcβ ζcδ ζcr


ζ γ = ζ γsζ γi ζ γd ζ γβζ γδ ζ γr (5.9)
ζ q = ζ qsζ qi ζ qd ζ qβ ζ qδ ζ qr

144  Geomechanics in soil, rock, and environmental engineering

Table 5.2  Vesic Correction Factors


Factor Cohesion (c) Self-weight (γ) Surcharge (q)
Foundation B ′ Nq B′ B′
shape (s) ζ cs = 1 + ζ γs = 1 − 0.4 ζ qs = 1 + tan φ
L ′ Nc L′ L′
Inclined mH m +1 m
loading (i) ζ ci = 1 − (φ = 0)  H   H 
B ′L ′cNc ζ γ i = 1 −  ζ qi = 1 − 
 V + B ′L ′c cot φ   V + B ′L ′c cot φ 
(1 − ζ qi )
= ζ qi − (φ > 0)
Nc tan φ
Foundation ζ cd = 1 + 0.4ξ (φ = 0) ζγd = 1.0 ζqd = 1 + 2 tan ϕ(1 − sin ϕ)2ξ
depth (d)
(1 − ζ qd )
= ζ qd − (φ > 0)
Nc tan φ
Surface
 2β  ζγβ = (1 − tan β)2 ζγβ = (1 − tan β)2
slope (β) ζ cβ = 1 −  (φ = 0) Note: Nγ = −2 sin β ( φ = 0)a
β < π /4  π + 2 
β is in (1 − ζ qβ )
radians = ζ qβ − (φ > 0)
Nc tan φ
Base tilt (δ)  2δ  ζγδ = (1 − δtan ϕ)2 ζqδ = (1 − δ tan ϕ)2
δ < π /4 ζ cδ = 1 −  (φ = 0)
 π + 2 
δ is in
radians (1 − ζ qδ )
= ζ qδ − (φ > 0)
Nc tan φ
Rigidity (r) ζ γr = ζ qr ζ qr = exp[A+B ]
 B
(see ζ cr = 0.32 − 0.12  
 L 
Appendix  B 
5A for Ir) + 0.60 log10 Ir ( φ = 0) A =  −4.4 + 0.6    tan φ
  L 
1 − ζ qr 3.07sin φ(log10 2Ir )
ζ cr = ζ qr − (φ > 0) B =
Nc tan φ 1 + sin φ

Note: V = vertical load; H = horizontal load; B = foundation width; L = foundation length (L > B); eB = eccentricity parallel
to B; eL = eccentricity parallel to L; B ′ = B − 2eB ; L ′ = L − 2eL ; m = (2 + x)/(1 + x); x = B/L if H parallel to B; x = L/B if H
parallel to L; ξ = D/B if D/B ≤ 1; ξ = tan−1 ( D/B ) if D/B > 1.
a
qu = cNcζ c β + ½γBNγ + qs Nqcos β ( φ = 0 ο ) if only applying surface slope correction.

and the additional subscripts s, i, d, β, δ, and r indicate that the correction factors apply for
foundation shape, inclination, depth, slope of soil surface, slope of foundation base, and soil
rigidity, respectively, (see Figure 5.7) and are given in Table 5.2.
The bearing capacity factors Nc , Nq, and N γ are the factors for a strip footing given in
Table 5.1. The method is approximate, but is accurate enough for practical application
where often large factors of safety (e.g. three or more) are applied to the ultimate bearing
capacities that are computed. For eccentric loads, it is assumed that the load acts at the cen-
tre of a foundation of reduced size. For example, if the eccentric load V shown in Figure 5.7,
were acting at the point X at eccentricities eB, eL , then the footing would be treated as one
having a reduced area of B′ by L′ where the reduced dimensions are given by

B′ = B − 2eB
(5.10)
L′ = L − 2 e L
Shallow foundations  145

MB ML
V eB = eL =
V V
ML MB
H H
eB eL
L X L
B B

V q =
D s γ
Dc
δ osβ
L H
L′ B/2
X B/2 β

B′

Figure 5.7 Footing subjected to vertical and horizontal loads plus moments.

If the footing is subjected to a moment M L in the L direction and a moment M B in the B


direction and has an applied vertical load V an equivalent loading system can be obtained
by placing the vertical load at eccentricities eL and eB (see Figure 5.7) where

ML
eL =
V
(5.11)
MB
eB =
V

The correction multipliers in the general bearing capacity equation (Equation 5.8) can also
be computed from values given by Meyerhof (1953, 1963) or by Hansen (1970). Hansen’s
correction factors include the effects of the slope of the soil surface and the slope of the base
of the foundation. The correction multipliers may be computed by multiplying the correc-
tion factors together as for the Vesic case (Equation 5.9).

EXAMPLE 5.1
A square footing, 1.5 m by 1.5 m in plan, is founded at a depth of 1 m in a deep layer of
clay. The footing is loaded by a vertical load and moment loadings such that the load can
be considered to act eccentrically at a point 0.3 m away from each of the centrelines of
the footing.
The water table is at foundation level and the saturated unit weight of the clay may be
taken as γsat = 17 kN/m3 (as can the bulk unit weight of the clay above water level). If the
magnitude of the inclined load is 100 kN and it is inclined at 60° to the vertical, compute
the bearing capacity of the footing

1. When it is loaded rapidly so that the soil is in an undrained condition


2. When it is loaded very slowly
146  Geomechanics in soil, rock, and environmental engineering

The strength properties of the clay are

su = 60 kPa φu = 0
c′ = 0 φ′ = 28ο

Solution
Because the load is eccentric, we must calculate the reduced length and breadth of the
footing

L′ = L − 2eL = 1.5 − 2 × 0.3 = 0.9 m


B′ = B − 2eB = 0.9 m

The components of the load are H = 100 cos60° = 50 kN; and V = 100 sin60° = 86.6 kN

1.
Undrained Loading

For ϕu = 0, the bearing capacity factors from Figure 5.4 are Nc = 5.14, Nq = 1, N γ = 0.
Using the Vesic Correction Factors of Table 5.2, we have

 B′ N q   mH 
ζc = ζcsζcdζci =  1 + (1 + 0.4ξ)  1 −
 L′ Nc   B′ L′ cNc 
= (1 + 0.194)(1 + 0.267)(1 − 0.3)
= 1.058

ζ q = 1 .0

and so the ultimate bearing capacity qu is given by

qu = su Ncζc + γDN qζ q
= 60 × 5.14 × 1.058 + 17 × 1 × 1 × 1.0
= 343 kPa

2.
Drained Loading
For ϕ′ = 28°, the bearing capacity factors from Figure 5.4 are Nq = 15, N γ = 11.
In this case, we have to compute the factor m that depends upon whether the hori-
zontal load is parallel to side B or L. Because the loading is at an angle to both sides
for this problem, it is necessary to compute m = mB sin2θ + mL cos2θ (as suggested by
Vesic) where θ is the angle of the inclined load to the long side L of the foundation.
As the footing is square, mB = mL = (2 + 1)/(1 + 1) = 1.5 and the angle is 45° so
that m = 1.5.
We now may compute the correction factors from Table 5.2
ζ q = ζ qsζ qdζ qi = (1 + 0.53)(1 + 0.199)(0.274) = 0.503
ζ γ = ζ γsζ γdζ γi = (1 − 0.4)(1.0)(0.116) = 0.0696

and so the bearing capacity for drained conditions is

1
qu = γ subBN γ ζ γ + γDN qζ q
2
= 0.5 × (17 − 9.81) × 1.5 × 11 × 0.0696 + 17 × 1 × 15 × 0.503
= 4.1 + 128.3
= 132 kPa

Therefore, for this particular problem, the drained bearing capacity is the lowest.
Shallow foundations  147

5.3.1.2  Soil layers of finite depth


The solutions for the bearing capacities of foundations mentioned so far have been for soils
that are of infinite depth. It is of interest to determine the effect of the bedrock underlying a
soil layer if the bedrock (or a very stiff layer) is at a finite depth.
Solutions to this problem have been found by Mandel and Salençon (1969) who consid-
ered a strip footing on a uniform layer of soil (i.e. strength parameters are constant) having
a rough base so that the maximum friction has to be mobilised on the base before slip can
occur.
The conditions assumed for the interface between the rigid base and the layer of soil is
important because it influences the collapse load, and Mandel and Salençon have also con-
sidered other base conditions. The rough based solution is more applicable to soils where it
would be expected that there would be some shearing resistance between the soil and the
underlying rock and so only results for the rough base are considered here.
The bearing capacity of strip footings having a full width B may be computed from
Equation 5.12.

1
qu = cNc ⋅ Fc + γBN γ ⋅ Fγ + qs N q ⋅ Fq (5.12)
2

where the bearing capacity coefficients are the same as for the Terzaghi equation, but are
modified by the factors Fc , F γ , and Fq which depend upon the depth h of the soil layer.
Tables 5.3–5.5 give values of the factors F γ , Fc , and Fq, respectively. The values in the
tables may be seen to depend on the angle of shearing resistance and the ratio of the footing
width to the layer depth B/h. It may be noted that the correction factors were computed for
a smooth footing base and by considering the effects of cohesion, self-weight of the soil, and
surcharge separately.

Table 5.3  Values of factor Fγ


ϕ ↓ B/h →  2 3 4 5 6 8 10 15 20 30 40
30° B/h ≤ 1.3 1.20 2.07 4.23 9.90 24.8 178 1450 3.81 × 105 1.3 × 108 1.95 × 1013
Fγ = 1
20° B/h ≤ 2.14 1.07 1.28 1.63 2.20 4.41 9.82 97 340 2.6 × 105 7 × 105
Fγ = 1
10° B/h ≤ 4.07 1.01 1.04 1.12 1.36 2.28 4.33 20 113
Fγ = 1
Fγ = 1 for all ϕ = 0°

Table 5.4  Values of factor Fc


ϕ ↓ B/h →  1 2 3 4 5 6 8 10 15 20 30
30° B/h ≤ 0.63 1.13 2.50 6.36 17.4 50.2 150 1444 1.48 × 104 5.81 × 106
Fc = 1
20° B/h ≤ 0.63 1.01 1.39 2.12 3.29 5.17 8.29 22.0 61.5 905 1.50 × 104
Fc = 1
10° B/h ≤ 1.12 Fc = 1 1.11 1.35 1.62 1.95 2.33 3.34 4.77 11.7 29.4
0° B/h ≤ 1.414 Fc = 1 1.02 1.11 1.21 1.30 1.40 1.59 1.78 2.27 2.75 3.72
148  Geomechanics in soil, rock, and environmental engineering

Table 5.5  Values of factor Fq


ϕ ↓ B/h → 1 2 3 4 5 6 8 10 15 20 30
30° B/h ≤ 0.63 1.12 2.42 6.07 16.5 47.5 142 1370 1.40 × 10 4 5.50 × 10
6

Fq = 1
20° B/h ≤ 0.86 1.01 1.33 1.95 2.93 4.52 7.14 18.7 51.9 763 1.26 × 104
Fq = 1
10° B/h ≤ 1.12 Fq = 1 1.07 1.21 1.37 1.56 1.79 2.39 3.25 7.37 17.9 92.3
Fq = 1 for all ϕ = 0°

5.3.2  Non-uniform soils


Inhomogeneity can be due to the soil strength increasing with depth, or perhaps being stron-
ger at the surface due to overconsolidation. The soil beneath a foundation may also consist
of different layers of material or the soil strength may simply be different at different places
on a site. In the latter case, the footing should be designed for the soil strength at the loca-
tion of the footing, or if this is not known, a conservative approach is to design footings for
the lowest strength.
Solutions have been found for the case where the soil strength varies uniformly with depth
or where the soil is layered, and these solutions are presented in the following sub-sections.

5.3.2.1  Strength increasing with depth


Strip Loading on Infinitely Deep Soil Layer
Solutions to the undrained problem (ϕ = ϕu = 0) where the cohesion of the soil varies linearly
with depth have been produced by Davis and Booker (1973) for the case of a strip footing.
The ultimate load per unit length that can be carried by the foundation is given by

Qu  ρB 
= F (2 + π)c0 + (5.13)
B  4 

where
c 0 is the cohesive strength at the surface of the soil
ρ is the rate of increase of cohesive strength with depth
B is the full width of the footing
F is a factor that is presented in Figure 5.8a

In the figure, the factor F is presented for both rough FR and smooth F S foundations.
Factors were also presented by Davis and Booker for a soil having a crust. The crust is
assumed to have a cohesion c 0 to a depth where it intersects the line c = ρz (see inset to Figure
5.8b). The factor F is presented in Figure 5.8b for both rough and smooth footings, where
the factor is again used with Equation 5.13.
Figure 5.9 shows the stress characteristics or slip lines for the case of a clay, whose
strength increases linearly with depth for both a (a) smooth and (b) rough based strip footing
computed using the program ABC by Martin (2004).

Soil Layers of Finite Depth


The case of an undrained clay (ϕ = ϕu = 0) with a cohesion that varies with depth has been
examined by Matar and Salençon (1977) for the case of a strip footing on a soil layer of
finite depth (overlying a rough, rigid base).
Shallow foundations  149

(a) 2.2
qu co
su
2.0
co + ρz
B

1.8 qu = Fn [(2+π)co + ρB/4] z


Rough FR
1.6
Fn

Smooth FS
1.4

FR/FS
1.2

1.0
0 4 8 12 16 20
ρB/co 0.05 0.04 0.03 0.02 0.01 0
co/ρB

(b) 1.3
qu = Fn [(2+π)co + ρB/4]
gh strip)
1.2 Fn = FRC (rough) F RC (Rou
or FSC (smooth)

1.1
ρB/co
4 8 12 16 20
1.0
Fn

0.05 0.04 0.03 0.02 0.01


co/ρB
co
0.9 su
)
o oth strip
F SC (Sm
0.8
ρz
z
0.7

Figure 5.8 Undrained bearing capacity factor Fn for a strip footing on an infinitely deep layer: (a) shear
strength increasing linearly with depth; (b) with a crust. (After Davis, E.H. and Booker, J.R. 1973.
Géotechnique, Vol. 23, No. 4, pp. 551–563.)

The bearing capacity of the strip is given by

 ρB 
qu = qs + µ cc0  Nc′ + (5.14)
 4c0 

where
qs is the surcharge
N′c is the bearing capacity factor given by Figure 5.10
µc is the correction factor given in Figure 5.10
c 0 is the cohesion at the surface
ρ is the rate of increase in cohesion with depth
150  Geomechanics in soil, rock, and environmental engineering

(a)

(b)

Figure 5.9 Slip line field for a (a) smooth and (b) rough strip footing on a clay whose strength increases with
depth ρB/c0 = 4.

Rough footing – Rough based layer


B/h

30 μc = 1.05 1.10

1.15
B
qs
⇓ 20 1.2

h 1.25
z
1.3
10
s u = co + ρ z 1.4
1.5
1.6
1.7
0.1 1 10 102 103 104
30 25 20 15 10 5 1.715 1.72 ρB/co
N′c π+2 √2

Figure 5.10 Curves of µ c and Nc′ . Strip on finite layer of soil with undrained strength increasing with depth.
(After Matar, M. and Salençon, J. 1977. Sols et Fondations, No. 352, Juillet-Août, pp. 95–107.)
Shallow foundations  151

5.3.2.2  Fissured clays


Some clay deposits, especially overconsolidated clays, can contain fissures which have a
lower strength than the matrix due to shearing and weathering. Some solutions have been
obtained to this problem (Davis 1980, Booker 1991) for the case of a strip footing.
Solutions to the problem have also been presented by Lav et al. (1995) for the case of a
smooth strip footing resting on a weightless, purely cohesive (ϕu = 0) material that has a
regular set of fissures. The problem considered is shown in Figure 5.11. The angle between
the sets of fissures is ωi while the fissures lie at an angle ω to the vertical.
If the cohesive strength of the fissures is cf and the strength of the solid material between
the fissures is cm , then the bearing capacity of the strip footing can be expressed in terms of
cm. Results are shown in Figure 5.12a–d which show the effects of the fissures on the bear-
ing capacity for joint sets with an included angle ωi of 90, 60, 45, and 30 degrees. Where
there is one set of fissures, for example, Figure 5.12a, the inclination ω of that set is used.
Where two sets of fissures exist (Figure 5.12b–d), the mean inclination of the two sets of
fissures is used (ω1 + ω2)/2 to indicate the inclination of the fissure sets. Solutions have been
evaluated for several values of the ratio of the fissure strength to the strength of the solid
material cf /cm.

5.3.2.3  Footings on slopes


The bearing capacity of footings on slopes (rather than on level ground) is of interest to
engineers as abutments to bridges are often founded on valley walls or are at the top of a
slope. Shields et al. (1990) have produced design charts that are based on the centrifuge tests
of Gemperline (1988) and other model tests.
Figure 5.13  shows the locations of footings with respect to the slope and definitions of
the non-dimensional terms λ, η that give the location of the strip footing with respect to the
slope. Design charts for two slopes (1V:2H) and (1V:1.5H) on cohesionless soil are shown in
Figure 5.14a,b. The charts show contours of the percentage capacity P of a footing on level
ground that the footing in a slope can carry.

B
qu
Smooth
interface

ωi
ω
Matrix shear strength, cm
Fissure shear strength, cf

ωi

ω ω ω1
ω2
ω
2 Orthogonal Single set 2 Non-orthogonal
sets sets

Figure 5.11 Strip footing on fissured clay.


152  Geomechanics in soil, rock, and environmental engineering

(a) 6 (b) 6
ωi = 90° or single set ωi = 60°
5 0.9 5 0.9
0.8
0.8
0.7
0.6 0.7
4 4
0.5 0.6
0.4
0.3 0.5

qu/cm
3 3
0.2
qu/cm

0.4
0.1
2 cf /cm = 2 0.3

0.2
1 1
0.1
cf /cm =
0 0
0 10 20 30 40 50 60 70 80 90 0 10 20 30 40 50 60 70 80 90
Fissure slope, ω (degrees) Fissure slope, (ω1 + ω2)/2 (degrees)

(c) 6 (d) 6
ωi = 45° ωi = 30°

5 0.9 5 0.9
0.8 0.8
0.7 0.7
4 4
0.6
0.6
0.5
3 0.5 3
qu/cm

qu /cm

0.4
0.4
2 2 0.3
0.3

0.2 0.2
1 1
0.1 0.1
cf /cm = cf /cm =
0 0
0 10 20 30 40 50 60 70 80 90 0 10 20 30 40 50 60 70 80 90
Fissure slope, (ω1 + ω2)/2 (degrees) Fissure slope, (ω1 + ω2)/2 (degrees)

Figure 5.12 Undrained bearing capacity of strip footing on fissured clay deposit: (a) one or two orthogonal
fissure sets; (b) two fissure sets with included angle of 60°; (c) with included angle of 45°; (d)
with included angle of 30°. (After Lav, M.A., Carter, J.P., and Booker, J.R. 1995. Proceedings of
the 14th Australasian Conference on the Mechanics of Structures and Materials, Hobart, pp. 38–43.)

The bearing capacity of a footing in a slope is therefore given by Equation 5.15 (for a
cohesionless soil)

1
q = γBN γq (5.15)
2

N γqR = (100.1159φ − 2.386 )(100.34φ − 0.2 log10 B) (5.16)


where N γq may be found from N γq = PN γqR (P is from Figure 5.14a,b).


Shallow foundations  153

(i) (ii)

O B
D
D B
β a
a b(–) b(+)

Legend:
Note: β = angle of slope with respect to horizontal
λ = b/B D = depth of footing with respect to the
η = D/B level of the horizontal ground
B = footing width
b = horizontal distance leading edge of
footing is away from crest of slope

Figure 5.13 Definition of terms for a footing on a slope. (i) Footing on slope; (ii) footing at top of slope.

The value of ϕ is in degrees and the value of the footing width B is in inches in the above
formula.
More recent solutions have been provided by Leshchinsky (2015) for c′–ϕ′ materials,
where the footing is at the edge of the slope.

(a) λ
–6 –4 –2 0 2
0 0
2
1
50
1 1
e 70
su rfac 90 0
η

pe 11 30 0
15 170
Slo 1 190
2 2

use 200%
3 3
–6 –4 –2 0 2
Note: Contours give percent capacity, P

(b) λ
–6 –4 –2 0 2
0 0
1.5
1 1
1 30
e 50
fac 70
η

ur 0
190
90

15 170
13 0

es
0
11

2 p 2
Slo

use 200%
3 3
–6 –4 –2 0 2
Note: Contours give percent capacity, P

Figure 5.14 (a) Suggested design factors for 2:1 (26.6°) slope. (After Shields, D., Chandler, N., and Garnier, J.
1990. Journal of Geotechnical Engineering, ASCE, Vol. 116, No. 3, pp. 528–537.) (b) Suggested
design factors for 1.5:1 (33.7°) slope. (After Shields, D., Chandler, N., and Garnier, J. 1990.
Journal of Geotechnical Engineering, ASCE, Vol. 116, No. 3, pp. 528–537.)
154  Geomechanics in soil, rock, and environmental engineering

5.3.2.4  Layered soils


The bearing capacity of footings on layered soils presents a more difficult problem than that
of a footing on a uniform soil, and so solutions are less plentiful and often approximate in
nature. Problems may be divided into two categories, those involving sand and clay layers,
and those involving clay layers of different strengths.

Sand and Clay Layers

If a loading is applied to a foundation that is constructed on a sand layer overlying a clay


layer, then the footing can punch through the sand into the clay layer beneath, especially if
the clay layer is soft. This type of failure has been examined by Meyerhof (1974), Meyerhof
and Hanna (1978), Hanna and Meyerhof (1980), Michalowski and Shi (1995), Burd and
Frydman (1997), and Shiau et al. (2002).
Solutions to this problem should be bounded by the solution for a footing on clay when
the sand layer becomes thin and by the solution for a footing on sand when the sand layer
becomes thick (as the failure would occur through the surface sand layer).
The solutions of Michalowski and Shi (1995) were obtained through the use of a kine-
matic approach that will yield an upper bound to the bearing capacity of a footing, and so
one would expect the collapse loads to be somewhat higher than the true solution. These
solutions are presented in Figure 5.15a–d. The numerical results of Burd and Frydman
(1997) were obtained from finite element- and finite difference-based computer codes, and
so may still contain some error. These results (for an angle of friction of the sand of 40°) are
presented in Figure 5.16.
In these figures, the full width of the foundation is B, the depth of the sand layer is D, the
undrained shear strength of the clay is su , the angle of friction of the sand is ϕ′, and the unit
weight of the sand is γ. The ultimate bearing pressure that can be carried by the footing is qu.
The accuracy of the results can roughly be assessed by looking at the limiting cases as
mentioned above. For deeper sand layers, the value of qu /γB should approach 1/2N γ because
the failure would occur through the sand alone. The rigorous plasticity solutions of Martin
(2004) (see Table 5.1) show that this value should be about 43 for a rough strip footing on
a sand layer having an angle of friction of 40°. The charts of Michalowski and Shi show a
value of 60 for this case while Burd and Frydman’s solution shows a value of about 50.
For thin sand layers, the results should be that for a clay layer alone so that a plot of pu /γB
versus cu /γB should be a straight line with a gradient of 5.14 as in this case qu = Nc · su and
Nc is equal to the classical value of (2 + π). This is certainly true for the Michalowski and
Shi results; however, it is not so easy to see on the plot of Burd and Frydman because their
plot is qu /γB versus su /γD.
Solutions have been found by Shiau et al. (2002) for the problem of a frictional material
(sand or gravel) overlying a clay layer through use of the finite element based upper- and lower-
bound theorems. In the plots, B is the full width of the footing, D is the thickness of the sand
layer beneath the footing, q is the surcharge (at footing level), and γ is the unit weight of the
sand. The plots of Figure 5.17 are based on the average between the upper and lower bounds.
Shiau et al. show that the results of Burd and Frydman, Michalowski and Shi and Hanna
and Meyerhof are slightly overestimating bearing capacities, although if appropriate factors
of safety are used, then the charts should be a reasonable guide for foundation design.

Clay Layers of Different Strengths

For clay layers of different stiffnesses, some solutions are available for the undrained case
where the angle of shearing resistance can be taken as ϕu = 0 (Vesic 1975). For the case where
Shallow foundations  155

(a) 30 (b) 30
ϕ = 30° ϕ = 35°
25 25 3.0 2.0

20 20
t/B = 0 0.5 1.0

qu
15 15
γB

γB
qu

2.0

10 10
1.0
0.5
5 5
t/B = 0

0 0

(c) 80 (d) 200


ϕ = 40° ϕ = 45°
70
5.0
5.0 4.0 160
60
3.0 4.0
50 120
2.0
40
γB

γB
qu

qu

3.0
1.0 80
30
2.0
20
0.5 40
1.0
10
t/B = 0 0.5
t/B = 0
0 0
0 1 2 3 4 5 0 1 2 3 4 5
su su
γB γB

Figure 5.15 Dimensionless limit pressure for strip footing on a sand layer overlying clay. Case of no sur-
charge. (After Michalowski, R.L. and Shi, L. 1995. Journal of the Geotechnical Engineering Division,
ASCE, Vol. 121, No. GT5, pp. 421–428.)

a soft clay layer (having a cohesion su1) overlies a deep stiffer clay layer (having cohesion su2),
the bearing pressure can be found from

qu = su1N m + qs (5.17)

where the coefficient Nm can be found from Figure 5.18 for square or circular footings.
For the case where a stiff clay layer (having undrained shear strength su1) overlies a deep
soft clay layer (having strength su2), then the bearing capacity can be estimated from

1
Nm = + κζc Nc (≤ ζc Nc ) (5.18)
β

where κ = su2 /su1, β = BL/[2(B + L)H], and ζc is the bearing capacity shape factor for the
footing (see Equation 5.9). (Figure 5.18 shows the dimensions B, H. L is the length of the
footing.)
156  Geomechanics in soil, rock, and environmental engineering

60

50

qu/γB 40

30
B/t
1.5
20
1
0.75
10 0.5
0.33

0
0 2 4 6 8 10 12
su/γD

Figure 5.16 Bearing capacity for strip footing on a sand layer overlying a clay layer. ϕ″sand = 40°. (After Burd,
H.J. and Frydman, S. 1997. Canadian Geotechnical Journal, Vol. 34, No. 2, pp. 241–253.)

35 80
D/B = 0·25 70 D/B = 2.0
30 ϕ′ = 50°
Rough base Rough base
q/γ B = 1 ϕ′ = 40° 60 q/γ B = 1
25 ϕ′ = 30°
q/γ B = 0 q/γ B = 0 ϕ′ = 50°
ϕ′ = 50° 50
20 ϕ′ = 40°
40 ϕ′ = 40°
γB

γB
p

15 ϕ′ = 50°
ϕ′ = 30° 30 ϕ′ = 30°
10 ϕ′ = 40°
20
5 10 ϕ′ = 30°
0 0
0 1 2 3 4 5 6 0 1 2 3 4 5 6
su /γ B su /γ B

45 140
40 D/B = 0.5 120 D/B = 2.0
ϕ′ = 50°
35 Rough base Rough base
q/γ B = 1 100 q/γ B = 1
30 ϕ′ = 40° ϕ′ = 50°
q/γ B = 0 q/γ B = 0
ϕ′ = 50° 80
25
ϕ′ = 30° ϕ′ = 50°
γB

γB
p

20 ϕ′ = 40° 60 ϕ′ = 40°
15 ϕ′ = 40°
40
10 ϕ′ = 30° ϕ′ = 30°
5 20
ϕ′ = 30°
0 0
0 1 2 3 4 5 6 0 1 2 3 4 5 6
su /γ B su /γ B

Figure 5.17 Ultimate bearing capacity p for rough strip footings on a frictional material overlying a deep clay layer.
Shallow foundations  157

13

B/H = ∞
B/H = 40 12

qult = su1Nm + qs
qu
qs
11

Bearing capacity factor Nm


H B su1
su2 10
Square or circular
footing
20 9

16
8
12

8 7

≤4 6.17
6
1 2 3 4 5 6 7 8 9 10 ∞
Undrained strength ratio, κ = su2/su1

Figure 5.18 Modified undrained bearing capacity factor Nm for square or circular footings on two layers of
purely cohesive soil. (After Vesic, A.S. 1975. Foundation Engineering Handbook, New York.)

Solutions have been found to the problem of a strip footing on a layer of clay overlying
a deep layer of clay. The solutions were obtained by Merrifield et al. (1999) by using an
optimisation technique that yields an upper and lower bound to the correct solution. Some
solutions to the problem shown in Figure 5.19 for a footing of width B resting on a layer of
clay with undrained shear strength su1 and thickness H which overlies another layer of clay
having strength su2 and which is of great (infinite) depth.
The bearing capacity is given by

qu = Nc*su1 (5.19)

The roughness of the footing only makes about 2% difference in the bearing capacity, and
so the charts can be used for rough and smooth footings.
The solutions are shown in the plots of Figure 5.20a–d.

5.3.2.5  Working platforms


Often engineers are required to design platforms to support tracked machines. For exam-
ple, if piling is being placed in soft soil that cannot support the weight of the piling rig,
a platform of granular material is placed to provide sufficient bearing to support the rig.
Geotextile reinforcement may also be placed at the base of the granular fill to increase the
bearing capacity.
The British Research Establishment (BRE) document 2004 has been widely used for the
design of working platforms. The approach is based on calculating the bearing capacity of
158  Geomechanics in soil, rock, and environmental engineering

r su1 ≠ su2
θ

su1
H
γ=0

su2
γ=0 ∞
Assumed circular failure
mechanism of Chen (1975)

Figure 5.19 Strip footing on a two-layered clay deposit.

a rectangular shaped loading applied to a granular layer (the working platform) overlying
either a cohesive layer (a clay) or a frictional layer (sand). The punching type bearing fail-
ure of the loading through the upper-granular platform is assessed by using the theory of
Meyerhof (1974).
Meyerhof and Hanna (1978) presented an improved method for calculating the bearing
capacity of a granular material overlying a weaker material, and this could be used to cal-
culate the bearing capacity rather than relying on the BRE tables.
The BRE document distinguishes between two load cases: Case 1 – where the crane or
rig  operator is unlikely to be able to aid recovery from an imminent platform failure; or
Case 2 – where the rig or crane operator can control the load safely (e.g. by releasing the load
or reducing power) if there is imminent platform failure. The amounts by which the loads
are factored up when calculating the thickness depend on the loading case. Table 5.6 shows
the factors recommended by the BRE.
The aim of designing a platform is to work out the thickness of a platform required to
support the tracked vehicle. For a cohesive subgrade where the undrained shear strength
of the clay is between 20 and 80 kPa, the thickness of the platform D is given by Equation
5.20, where

D = {W[Fq − su Nc sc − (2T /W )]/[ γ pKp tan δsp ]}


0.5
(5.20)

Note that Equation 5.20 will give a negative value within the square brackets if the soil has
high strength and it will not be possible to take the square root. This generally occurs if no
platform is needed.
If the subgrade is a non-cohesive material such as sand then the platform thickness is
given by Equation 5.21.

D = {W[Fq − ½γ sWN γs sc − (2T /W )]/[ γ pKp tan δsp ]}


0.5
(5.21)

Shallow foundations  159

(a) (b) 7
Finite element upper bound Finite element upper bound
10 Finite element lower bound Finite element lower bound
6
Semi-empirical Meyerhof and Semi-empirical Meyerhof and
8 Hanna (1978), empirical 5 Hanna (1978), empirical
Brown and Meyerhof (1969) Brown and Meyerhof (1969)
Upper bound Chen (1975) Upper bound Chen (1975)
6 4
N*c

N*c
3
4
2 H/B = 0.375
2 H/B = 0.125 1
0 0
0 1.0 2.0 3.0 4.0 5.0 0 1.0 2.0 3.0 4.0 5.0
su1/su2 su1/su2
8 6
Finite element upper bound
Finite element lower bound
Semi-empirical Meyerhof and
5 H/B = 0.5
6 Hanna (1978), empirical
Brown and Meyerhof (1969) 4
Upper bound Chen (1975)
4 3
N*c

N*c

Finite element upper bound


2 Finite element lower bound
2 Semi-empirical Meyerhof and
H/B = 0.25 1 Hanna (1978), empirical
Brown and Meyerhof (1969)
Upper bound Chen (1975)
0 0
0 1.0 2.0 3.0 4.0 5.0 0 1.0 2.0 3.0 4.0 5.0
su1/su2 su1/su2

(c) 6 (d) 6

5 H/B = 0.75 5

4 4

3 3
N*c

N*c

Finite element upper bound Finite element upper bound


2 Finite element lower bound 2 Finite element lower bound
Semi-empirical Meyerhof and Semi-empirical Meyerhof and H/B = 1.5
Hanna (1978), empirical Hanna (1978), empirical
1 Brown and Meyerhof (1969)
1 Brown and Meyerhof (1969)
Upper bound Chen (1975) Upper bound Chen (1975)
0 0
0 1.0 2.0 3.0 4.0 5.0 0 1.0 2.0 3.0 4.0 5.0
su1/su2 su1/su2
6 6

5 H/B = 1.0 5

4 4

3 3
N*c

N*c

Finite element upper bound Finite element upper bound


2 Finite element lower bound 2 Finite element lower bound
Semi-empirical Meyerhof and Semi-empirical Meyerhof and H/B = 2
1 Hanna (1978), empirical 1 Hanna (1978), empirical
Brown and Meyerhof (1969) Brown and Meyerhof (1969)
Upper bound Chen (1975) Upper bound Chen (1975)
0 0
0 1.0 2.0 3.0 4.0 5.0 0 1.0 2.0 3.0 4.0 5.0
su1/su2 su1/su2

Figure 5.20 Variation of bearing capacity factor Nc*: (a) H/B = 0.125 and H/B = 0.25. (b) H/B = 0.375 and
H/B = 0.5. (c) H/B = 0.75 and H/B = 1. (d) H/B = 1.5 and H/B = 2.
160  Geomechanics in soil, rock, and environmental engineering

Table 5.6  Load factors F applied to track


pressure from the vehicle
Platform required
Loading condition Yes No
Case 1 2.0 1.6
Case 2 1.5 1.2

In Equations 5.20 and 5.21,

q is the applied pressure from the tracked vehicle


W is the width of one of the tracks
F is the loading factor from Table 5.6 depending on the loading case
T is the ultimate tensile strength of the geogrid used divided by a factor of 2, for
example, T = Tult /2 (units are force per unit length, e.g. kN/m)
γp is the unit weight of the platform material
sc is a shape factor for a rectangular track sc = (1 + 0.2W/L) and L is the track length
sp is a shape factor for the geometry of the punching shear sp = (1 + W/L)
Kp tan δ is the shearing resistance coefficient for the platform (see Figure 5.21)
Nc is the bearing capacity factor for clay (generally Nc = 5.14)
N γs is the bearing capacity factor for a sand subgrade
γs is the unit weight of a sand subgrade (use submerged unit weight γsub if sand is
submerged)

Several other checks should be performed when designing a platform

• The bearing capacity of the platform material alone should be checked to see if it is
adequate.
• The bearing of the subgrade alone should be checked to asses if a platform is required
at all.

20
18
16
14
12
Kp tan δ

10
8
6
4
2
0
35 40 45 50
ϕ′

Figure 5.21 Design values of K p tan δ. (BRE 2004. Working Platforms for Tracked Plant: Good Practice Guide to
the Design, Installation, Maintenance and Repair of Ground-Supported Working Platforms, British
Research Establishment, Garston, Watford.)
Shallow foundations  161

• The track loading should not come closer than half of the machine width from the edge
of the platform.
• There may be soft spots in the subgrade, for example, backfilled trenches. These should
be eliminated if possible.
• Slopes should ideally be less than 1 in 10 so that plant does not become unstable.

5.4  NUMERICAL ANALYSIS

For difficult problems involving complex material properties, complex geometry, or where
structural elements such as retaining walls or anchors, finite element or finite difference
methods are often used. Generally, care needs to be taken, as the collapse loads calculated
are greater than the exact theoretical loads (see Sloan and Randolph 1982). However, for
engineering purposes, the flexibility of numerical methods is of advantage and these meth-
ods are often used.
Computer programs such as the finite element (FE) codes that are specifically written
for geotechnical applications such as PLAXIS2D (2014) and PLAXIS3D Foundation or
PLAXIS3D Tunnel, CRISP (2013), Phase2 (2013), or the more general FE code ABACUS
(2014) may be used. The finite difference codes FLAC (Ver. 7.0) or FLAC3D (Ver. 5.0) may
also be used (ITASCA Consulting Group 2014).
If a margin of safety against collapse of a geotechnical structure is required, a “c–ϕ” reduc-
tion can be carried out. The process involves reducing the strength of the soil until failure
occurs. The amount by which the strength parameters are reduced gives the margin of safety.
Alternatively, load can be applied and displacement at some point monitored until there
is evidence of large displacement occurring which indicates collapse. An example is shown
in Figure 5.22 for a strip footing with a total width of 2 m resting on a uniform weightless
soil with strength parameters c′ = 10 kPa and ϕ′ = 20°. A non-associated flow rule was used
with a Mohr–Coulomb failure surface where the angle of dilation for the soil was taken
as ψ = 10°, the modulus of elasticity for the soil as E = 30,000 kPa and Poisson’s ratio as
ν = 0.25. The load may be seen to level off at about 153 kPa, which is slightly more than the
rigorous plasticity solution of 148.3 kPa (3% difference).

180

160

140
Average pressure (kPa)

120

100

80

60

40

20

0
0.00 0.01 0.02 0.03 0.04 0.05
Deflection (m)

Figure 5.22 Load–deflection curve for strip footing on weightless soil with c′ = 10 kPa, ϕ′ = 20°, ψ = 10°.
Rigorous plasticity solution for collapse load = 148.3 kPa.
162  Geomechanics in soil, rock, and environmental engineering

Velocity vectors

12.0

10.0

8.0

6.0

Scale
4.0

5.00E–04
2.0

0.0
–0.0002
1
0.0 5.0 10.0 15.0 A
–2.0
Rigid strip footing

Figure 5.23 Velocity vectors at failure for strip footing on weightless soil with c′ = 10 kPa, ϕ′ = 20°, ψ = 10°.
(Half of problem shown due to symmetry.)

For such a model, the numerical analysis has been noted to become unstable if there
are large differences between the angle of friction ϕ and the dilation angle ψ (e.g. ϕ = 40°
and ψ = 0°). This is often the cause of instability in calculation with commercial codes
as well.
The plot of velocity vectors at failure shows a wedge of material moving down beneath
the footing and material being pushed up to the side of the footing as shown in Figure 5.23.
In the figure, the extent of the mesh that is composed of eight-node isoparametric elements
may be seen. Increments of displacement are specified for the nodes beneath the footing and
the loads at the nodes are back figured to obtain the corresponding load.

5.5 SETTLEMENT

If a structure settles substantially with very small differential settlements (settlements are
almost uniform everywhere), then the structure is unlikely to sustain any damage, although
if settlement is too large, the structure will be subject to serviceability problems (i.e. water,
gas, and sewer pipes and other services may be damaged). It is the differential settlements
that cause cracking and damage to structures (i.e. movement of one part of a structure rela-
tive to another). Tilt may also be a problem for crane rails or machinery that has to remain
level. It is therefore necessary to try to limit the differential settlements as well as the overall
vertical movements of a structure.
The magnitude of allowable settlements depends on the type of structure or the purposes
for which the foundation is to be used. Here, only static loadings are considered (i.e. not
cyclic or vibratory loads).
Shallow foundations  163

5.5.1  Limits of settlement


Allowable deflections (or angular distortions) of different structures have been proposed by
Skempton and MacDonald (1956), Bjerrum (1963), and Grant et al. (1974). More recently,
Boone (1995) has proposed a method for determining damage to structures that relies on
many factors such as the ground movement profile, the geometry of the structure, and the
critical strains in the building materials.
Allowable angular distortions for different types of structures and construction materials
have also been presented by Polshin and Tokar (1957) and these are shown in Table 5.7. It
may be seen from this table that in general, stiffer structures such as those constructed from

Table 5.7  A
 llowable settlements and deflections
Subsoil
Sand and clay in Clay in plastic
Item no. Description of standard value hard condition condition
1 Slope of crane way, as well as tracks for bridge crane 0.003 0.003
truck
2 Difference in settlement of civil and industrial
building column foundations:
(a) for steel and reinforced concrete frame 0.002ℓ 0.002ℓ
structures
(b) ​for end rows of columns with brick cladding 0.007ℓ 0.001ℓ
(c) for structures where auxiliary strain does not 0.005ℓ 0.005ℓ
arise during non-uniform settlement of
foundations (ℓ = distance between foundation
centres)
3 Relative deflection of plain brick walls:
(a) for multi-storey dwellings and civil buildings
at /H < 3 0.0003 0.0004
at  /H  5 (ℓ = length of deflected part of wall, 0.0005 0.0007
H = height of wall from foundation footing)
(b)  for one-storey mills 0.0010 0.0010
4 Pitch of solid or ring-shaped foundations of high rigid 0.004 0.004
struc­tures (smoke stacks, water towers, silos, etc.)
at the most unfavourable combination of loads

Item no. Kind of building and type of foundation Average settlement (cm)
1 Buildings with plain brick walls on continuous and
separate foundations with the wall length ℓ to the
wall height H (H counted from the foundation
footing):
 /H >2.5  8
 /H <1.5 10
2 Buildings with brick walls, reinforced with reinforced 15
concrete or reinforced brick belts (not depending
on the ratio of ℓ/H)
3 Framed buildings 10
4 Solid reinforced concrete foundations of blast 30
furnaces, smoke stacks, silos, water towers, etc.
Source: After Polshin, D.E. and Tokar, R.A. 1957. Proceedings of the 4th International Conference on Soil Mechanics and
Foundation Engineering,Vol. 1, pp. 402–405.
164  Geomechanics in soil, rock, and environmental engineering

masonry, are more likely to be damaged than flexible structures such as those having steel
frames with sheet cladding.

5.5.2  Settlement computation


When a load is applied to a footing on a non-cohesive, free draining soil like a sand or a
gravel, the soil will deform immediately and will reach its final settlement (provided there is
no creep settlement). However, when a load is applied to a saturated clay, the load is partly
supported by the soil skeleton and partly by the pore water. There is therefore an increase
in the pore water pressure beneath the foundation and water will begin to flow from areas
of high water pressure to areas of low water pressure, and as it does, the soil beneath the
footing undergoes a volume reduction. The footing will therefore settle with time until all
excess pore pressures have dissipated, at which time the footing will have reached its final
settlement (again if creep is not considered).
When the load is first applied, the footing will undergo an initial settlement with the soil
deforming at constant volume (since the water does not have time to flow through the soil).
This can only occur if the load is not one-dimensional so that shear deformation can take place.
Generally, it is the final settlement of a foundation that is of interest as this is the largest
settlement that will occur and therefore this leads to the largest differential settlements.
However, the load–deflection behaviour may be of importance in some cases.

5.5.3  Theory of elasticity


Working loads applied to foundations are usually selected so that they are well below the
ultimate bearing capacity of the building. It is therefore possible to treat the soil as an elastic
material, provided that the elastic modulus used in any computations is appropriate for the
stress range that occurs in the ground beneath the footing.
For obtaining the initial settlement Si of the footing, the undrained elastic modulus Eu
and Poisson’s ratio νu of the soil are needed. As the soil is assumed to undergo zero volume
change, the value of νu is taken to be 0.5 as this can be shown to be the requirement for zero
volume change in an isotropic material. For computing the final settlement Sf, the drained
elastic modulus E′ and Poisson’s ratio ν′ are used. For free draining sands and gravels, only
the drained parameters will apply.
For one-dimensional loadings (i.e. where the load is large in extent compared to the thick-
ness of the soil layer such as a surcharge), the elastic theory of Terzaghi (1923) may be used.
One-dimensional theory is based on the assumption that the strain will be entirely vertical
and that there will be no lateral strains. Vertical settlement S is then given by

S = ∑m ∆σ′ h
vi vi i (5.22)
i =1

where
mvi is the coefficient of compressibility of layer i of the soil
∆σ ′vi is the increase in vertical effective stress in layer i of the soil
hi is the thickness of the ith layer of soil
N is the number of layers into which the soil is divided

The soil layer is divided into horizontal layers each of thickness hi and the stress increases
are computed at the centre of each layer. The coefficient of compressibility mv is found from
Shallow foundations  165

an oedometer test and is calculated from the strain in the specimen divided by the stress
change (Section 1.5). As the value of mv is dependent on the stress level, the appropriate value
should be used for each layer. A value corresponding to the mean stress (i.e. intermediate of
the initial stress and the stress after application of the load) is appropriate. For strictly one-
dimensional problems, the stress increase ∆σ ′vi should be the same for all layers and equal to
the applied uniform load, however, vertical stress increases are often calculated for rectan-
gular or circular loading shapes using three-dimensional solutions for the stress increases.
This is in effect carrying out a pseudo three-dimensional analysis as stresses are calculated
from three-dimensional solutions and then the one-dimensional equation (Equation 5.22) is
used to compute settlements.
Although Equation 5.22 allows the compressibility of the soil to be stress dependent, it
does not allow for the compressibility to change within a stress increment within each layer.
An alternative method that uses the compression Cc and recompression Cr indices can be
used. It is observed from laboratory testing that the soil becomes more compressible after
the previous pre-compression pressure σ ′p is reached. The pre-compression pressure is the
greatest pressure that the soil has been subjected to in the past and may be due to layers of
soil being eroded away that were once present, or is often due to swelling and shrinking of
the surface layer of the soil that is subjected to wetting and drying. The settlement S is given
by Equation 5.23 when the soil is divided into N horizontal sub-layers in order to perform
the calculation.

N
 Cr  σ ′pi   σ ′fi  
∑ 1 + e
Cc
S = log10   + log10    ⋅ hi (5.23)
0i  σ 0′ i  1 + epi  σ ′pi  
i =1

In Equation 5.23

σ 0i
′ is the initial vertical effective stress in the soil for layer i
σ ′pi is the pre-consolidation pressure in layer i
σ ′fi is the final effective stress in layer i after loading
e 0i is the initial void ratio in layer i
epi is the void ratio in layer i at the pre-consolidation pressure

For problems that are three-dimensional in nature, elastic solutions for the settlement
of foundations of various shapes and stiffnesses and subjected to various types of load-
ings, are available and may be found in books such as the one by Poulos and Davis (1974).
Provided that the appropriate solution can be found, the settlement of the foundation can
be calculated.
The settlement of flexible footings that are rectangular in plan (of length L by breadth
B) and founded at depth D in uniform soil layers of finite depth may be computed from
the charts of Christian and Carrier (1978). These charts are reproduced in Figure 5.24 and
apply for only the undrained case (νu = 0.5). Elasticity solutions are normally presented in
non-dimensional form so that they apply to a wide range of problems. In addition, solutions
for strip or rectangular shaped loadings are often presented for the edge (strip) or corner
(rectangle) because the solutions may then be superimposed. For example, the deflection at
the centre of a rectangle may be obtained by dividing the rectangular loaded area into four
equal rectangular areas and summing the deflections under the corners of each.
Figure 5.25 shows solutions for a rigid circular foundation constructed on a layer of soil
of finite thickness. This plot may be used to determine both initial and final settlements, as
166  Geomechanics in soil, rock, and environmental engineering

1.0
L = Length ν = 0·5
q
ρ = Average settlement D
qB B
ρ =μ0 μ1 H
E
μ0 0.9

0.8
0 5 10 15 20
D/B

2.0
L/B = ∞
L/B = 10
1.5
L/B = 5

1.0 L/B = 2
μ1

Square

0.5 Circle

0
0.1 1 10 100 1000
H/B

Figure 5.24 Chart for evaluating settlement of embedded rectangular footings. (After Christian, J.T. and
Carrier, W.D. 1978. Canadian Geotechnical Journal, Vol. 15, pp. 123–128.)

the solutions have been evaluated for a range of Poisson’s ratios. The solutions apply only to
foundations on the surface of the clay layer and may also be used for square footings since
the settlement of a square footing may be approximated as the settlement of a circular foot-
ing having the same area in plan with very little error.

5.5.4  Rate of settlement


The rate at which footings settle may be determined by carrying out a consolidation analy-
sis. It is generally assumed that the soil is totally saturated, and that as water flows from the
soil, the soil skeleton slowly deforms. The volume change in the soil skeleton is assumed to
be equal to the volume of water that flows from the pores in the soil.
If the problem is one-dimensional, there can be no initial settlement (i.e. upon application
of a load) as the pore water is incompressible, but for three-dimensional problems, shear
deformation of the soil can take place therefore the soil can undergo an initial undrained
deformation at constant volume.
Plots have been produced to allow estimates to be made of the rate of consolidation of
footings, and an example of such a plot is shown in Figure 5.26 (Davis and Poulos 1972).
This plot is for a circular footing of radius a resting on the surface of a clay layer of thick-
ness h, which may drain across its upper surface and through its lower surface. It is assumed
Shallow foundations  167

1.6
P
a
r
1.4
h P
pav =
πa2

1.2
z

Settlement
1.0 p aI
S = av ρ
E

ν=0
0.8

0.2
0.4
0.6 0.5

0.4

0.2

0
0 0.2 0.4 0.6 0.8 1.0
h/a 1.0 0.8 0.6 0.4 0.2 0
a/h

Figure 5.25 Chart for evaluating settlement of rigid circular footing on a soil layer of finite thickness.

0
a

0.2
h

0.4
Up = Us

0 (One-dimensional)
0.6 0.5
Values of h/a
5 2 1
10
0.8 20
50

1.0
10–4 10–3 10–2 10–1 1
ct
Tv = 12
h

Figure 5.26 Degree of settlement versus time factor for circular footing on a soil layer drained at the upper
and lower surfaces. (After Davis, E.H. and Poulos, H.G. 1972. Géotechnique, Vol. 22, No. 1,
pp. 95–114.)
168  Geomechanics in soil, rock, and environmental engineering

that the load is applied to the footing at time t = 0 and then held constant. In the plots, the
degree of consolidation U is plotted versus the time factor Tv where

cv1t
Tv = (5.24)
h2

and cv1 is the one-dimensional coefficient of consolidation that may be measured in an


oedometer test.
If the extent of the load is large enough, it can be seen that the consolidation curve reaches
the one-dimensional curve (h/a = 0).
The degree of consolidation U is defined as

St − Si
U = (5.25)
Sf − Si

where St is the settlement at any time, Sf is the final settlement, and Si is the initial settle-
ment. Initial and final settlements can be calculated using the appropriate elastic moduli and
Poisson’s ratio as described above (Section 5.5.3).
Settlement at any time t can therefore be computed (from Equation 5.25) once the degree
of consolidation is known and the initial and final settlements have been found.

EXAMPLE 5.2
A square footing having a side of 2 m is constructed on a layer of soft clay that is 4 m
thick and has the following properties

E′ = 8 MPa ν′ = 0.3 cv1 = 0.8 m2 /yr


Eu = 12 MPa νu = 0.5

The footing is constructed on a granular base layer of permeable material, so that the
surface of the clay layer may be considered drained. The base of the clay layer is underlain
by sand, therefore drainage can take place at the base of the clay layer also.
If a vertical load of 400 kN is applied to the footing, what would be the expected settle-
ment of the footing after 1 year?

Solution
• The pressure applied to the footing would be pav = 400 kN/(2 m × 2 m) = 100 kPa
• The square footing may be treated as a circular footing of area equal to the square,
therefore πa 2 = 2 × 2 giving a = 1.128 m
• The thickness of the soil layer to the radius of the footing would be

h /a = 4 m/1.128 m = 3.55 or a /h = 0.28


1.
Drained Case

Figure 5.25 gives (for Poisson’s ratio ν′ = 0.3) an influence factor I ρ = 1.1

Therefore, the final settlement Sf is given by

pav aIρ 100 × 1.128 × 1.1


Sf = = = 15.5 mm
E′ 8
Shallow foundations  169

2. Undrained Case (or Initial Settlement)

Figure 5.25 gives an influence factor of I ρ = 0.84 (for νu = 0.5)

The initial settlement Si is given by

pav aIρ 100 × 1.128 × 0.84


Si = = = 7.9 mm
Eu 12

3. At 1 year, the time factor can be computed as

cv t 0.8 × 1
Tv = = = 0.05
h2 16

and so from Figure 5.26, the degree of consolidation U can be found to be 0.82. We
can therefore write

St − Si St − 7.9
U = = = 0.82
Sf − Si 15.5 − 7.9

and solving for St , gives the settlement at 1 year St ≈ 14 mm.

5.5.5  Settlement of footings on sand


Often footing design for sand foundations is based on field-test results (e.g. standard pen-
etration tests [SPT] or cone penetrometer results) because of the difficulty of sampling and
testing sand samples in the laboratory (i.e. testing samples at the correct density). The design
must again be based on both settlement criteria and bearing criteria.
Generally, a small or narrow footing subjected to an increasing pressure will have an
unacceptably small margin of safety against a bearing capacity failure before it settles too
much, whereas a larger or wider foundation will reach a given settlement criterion before
the margin of safety against a bearing capacity failure is exceeded. Therefore, for smaller
foundations, bearing capacity influences design, but for larger foundations, settlement is of
greater importance.
Vibration can cause footings on loose sand to settle unduly as the sand will tend to densify
(or even liquefy if it is saturated); however, this aspect is not considered in the following.
Footings to be constructed on loose sands, which are often the result of hydraulically placed
fills, should be considered in the light of potential vibration or earthquake hazard and den-
sification of the sand deposit may be necessary by means such as vibroflotting.

Methods Based on SPT

D’Appolonia et al. (1970) presented a method based on obtaining the elastic modulus of
the sand from the average SPT blow count taken over one footing width B beneath the foun-
dation. The settlement of a footing can be found once the modulus is obtained, by use of the
theory of elasticity (see Section 5.5.3). The relationship between elastic modulus E′ and the
blow count is shown in Figure 5.27 (ν′ is Poisson’s ratio of the sand).
Burland et al. (1977) presented a compilation of observed data that can be used to deter-
mine the settlement of footings on sands of different densities. The deflection of a foun-
dation per unit applied pressure is presented for different footing widths B as shown in
Figure 5.28. Suggested upper limits for the deflection are shown (as lines) on the plot for
dense, medium dense, and loose sands. The density of the sand can be determined from the
SPT blow count as shown in the inset to Figure 5.28.
170  Geomechanics in soil, rock, and environmental engineering

100
All data for footing foundations on clean
sand or sand and gravel
Pre-loaded sand
75

E′/(1–ν′2) (MN/m2)

50
Normally loaded sand
or sand and gravel
25

0
0 10 20 30 40 50 60 70
Average measured SPT resistance depth B below footing (blows/ft)

Figure 5.27 Correlation between blow count and modulus of sand. Blow count is the average over a depth
B below the footing. (After D’Appolonia, D.J., D’Appolonia, E. and Brissette, R.F. 1970. Journal
of the Soil Mechanics and Foundations Division, ASCE, Vol. 96, No. SM2, pp. 754–762.)

Parry (1977) presented a method based on SPT results in which the likely settlement of
strip or rectangular footings of width B may be computed.
The SPT blow count at a representative depth of 3/4B beneath the base of the footing is
used in the design, and this is not corrected for depth or for the effects of the water table. If
Nm is the SPT blow count at the representative depth, then the settlement of a footing may
be estimated from

qB
ρ = 300 (5.26)
Nm

where q is in MPa, B is in metres, and ρ is the settlement of the footing in mm.


(kg/cm ) 2

L
mm

nd
100 1.0 e sa
Loos
Settlement per unit applied pressure

tative –
mm Ten
Upper limit for
(kN/m 2) medium dense

Upper limit
for dense
10 0.1
Key N
Loose < 10
Medium 10–30
dense
Dense > 30
1.0
0.1 1.0 10.0 100.0
Breadth B (m)

Figure 5.28 Observed settlement of footing on sand of various relative densities. (After Burland, J.B.,
Broms, B.B., and de Mello, V.F.B. 1977. State-of-the-Art Review, IX International Conference on Soil
Mechanics and Foundation Engineering, Tokyo, Vol. 2, pp. 495–546.)
Shallow foundations  171

To be able to use any of these methods for deflection calculation, the load applied to the foot-
ing must be well below the collapse load, therefore a bearing capacity check would also need to
be carried out to verify that the loading level had an adequate factor of safety against collapse.

Method Based on Static Cone Penetrometer

The approach of Schmertmann (1970) and Schmertmann et al. (1978) is based on cone
penetrometer results. The soil is divided up into N depth increments or layers and the
expected settlement ρi is then computed from the formula
N
∆zi
ρi = C1C2 ∆p ∑E si
I zi (5.27)
i =1
where
C1 is a correction to allow for strain relief from embedment
C 2 is a correction for time dependent increase in settlement
Δp is the net applied footing pressure Δp = p − p 0 (see Figure 5.29b)
Δzi is the depth increment
Izi is the influence factor for soil layer i (see Figure 5.29a)
E si is the elastic modulus of soil layer i

(a) Rigid footing vertical strain influence factor = Iz


0 0.2 0.4 0.6 0.8 1.0
0

B/2 For axisymmetrical Δp


Relative depth below footing level z/B

use Es = 2.5 qc } Izp = 0.5 + 0.1 √ σ′vp ........(*)


B

Axisymmetric
L/B = 1
2B
For plane strain
use Es = 3.5 qc Modified strain influence
factor distributions
3B
Plane strain
L/B > 10

4B
(b) B
Δp = p – p′o
p p′o

B/2 (axisym) σ′vp


6B B (pl. str.)
Depth to Izp
Explanation of pressure terms
used in equation (*) above

Figure 5.29 Strain influence factors for use in Schmertmann’s method. (After Schmertmann, J.H., Hartman,
J.P., and Brown, P.R. 1978. Journal of the Geotechnical Engineering Division, ASCE, Vol. 104, No.
GT8, pp. 1131–1135.)
172  Geomechanics in soil, rock, and environmental engineering

Factor C1 allows for strain relief due to embedment of a footing and is computed from the
ratio of the overburden pressure at foundation level p 0 to the net pressure increase at founda-
tion level Δp = p − p 0 (p is the applied pressure).

p 
C1 = 1 − 0.5  0  (5.28)
 ∆p 

Note that C 1 must be greater than or equal to 0.5.


Factor C 2 is used to allow for creep of the foundation and is given by

 t 
C2 = 1 + 0.2 log10  (5.29)
 0.1 

where t is the time in years.


The soil is divided into a number of horizontal layers and the elastic modulus of the soil
is estimated from the cone penetrometer tip resistance qc. The elastic modulus may be found
from the following correlations:

Long or strip footings L /B ≥ 10 Esi = 3.5qc


(5.30)
Axisymmetric footings or L /B = 1 Esi = 2.5qc

L = length and B = breadth of a rectangular footing.

The strain distributions beneath footings are different depending on the footing shape
and vary between the two extremes of the distributions for a strip or for a square footing.
Therefore, different influence factor diagrams are used for square (or circular) and strip-
footings (see Figure 5.29). For intermediate cases, that is, rectangular footings of length L
and breadth B (L/B from 1 to 10), interpolation between the two diagrams may be used. Iz
may be assumed to vary linearly between 0.1 and 0.2 on the Iz axis and z/B to vary from 2
to 4 on the z/B axis.
It may be noted that the peak value on the influence factor plots varies and is given by
0.5
 ∆p 
I zp = 0.5 + 0.1   (5.31)
 σ ′vp 

The effective stress used in computing the peak value of the influence factor σ ′vp is calcu-
lated at the depth shown in the inset to Figure 5.29b.
To design footings using this approach, the field SPT results must be corrected to allow
for the effect of depth. The correction factor C N is shown plotted against vertical effective
stress in Figure 5.30, so that at any depth the corrected blow count N cor can be obtained by
multiplying the field value N field by the correction factor, that is,

N cor = N field × CN (5.32)

5.5.6  Methods based on settlement and bearing criteria


Some methods of footing design (for sands) are based on both settlement and bearing cri-
teria. An example of this is the well-known approach of Peck et al. (1974). This method is
Shallow foundations  173

Ncor
Correction factor CN =
Nfield
0.4 0.8 1.2 1.6 2.0
0

50

Effective vertical overburden pressure, kPa


100

150

200

250

300

350

400

450

Figure 5.30 Chart for correction of N values in sand for influence of overburden pressure (reference value
of effective overburden pressure = 1 ton/sq ft, 96 kPa). (After Peck, R.B., Hanson, W.E., and
Thornburn, T.H. 1974. Foundation Engineering, 2nd ed. Wiley, New York.)

based on SPT results, and allows for a factor of safety of 2 against a bearing capacity failure
or an allowable settlement of 25 mm (1 inch).
The average corrected blow count (using the correction factors of Figure 5.30) over a
depth of one footing width beneath the foundation is then used to design the foundation
using the charts shown in Figure 5.31. This is straightforward if the size of the footing is
known and an allowable pressure is being obtained, but if the applied pressure is known
and the footing size is to be determined, some trial and error is required as the average blow

(a) Df /B = 1 (b) Df /B = 0.5 (c) Df /B = 0.25


600
N= N= N=
50 50 50
500
Soil pressure, kPa

400 40 40 40

300 30 30 30

200 20 20 20
15 15 15
100 10 10 10
5 5 5
0
0 0.4 0.8 1.2 0 0.4 0.8 1.2 0 0.4 0.8 1.2 1.6
Width of footing, B, m

Figure 5.31 Design chart for proportioning shallow footings on sand. (a) Df /B = 1; (b) Df /B = 0.5; (c) Df /B = 0.25.
(After Peck, R.B., Hanson, W.E., and Thornburn, T.H. 1974. Foundation Engineering, 2nd ed.
Copyright Wiley, New York.)
174  Geomechanics in soil, rock, and environmental engineering

count over one footing depth may change with the size of the footing selected. Full details
of the method are given in the book by Peck et al. (1974).
A correction is also made for the effect of the water table by multiplying the allowable
pressure (from Figure 5.31) by the correction factor C w, where

Dw
Cw = 0.5 + 0.5 Dw ≤ D + B
D+B (5.33)
Cw = 1.0 Dw > D + B

In the above equation, B is the footing width, D is the footing depth, and Dw is the depth of
water below the surface. It may be seen that the correction is 0.5 when the water table is at the
surface and 1.0 when greater or equal to one footing width below the base of the foundation –
anywhere between these two extremes, the formula (Equation 5.33) provides a linear interpo-
lation. The reasoning is that the unit weight of the cohesionless material will be approximately
halved when submerged and therefore the bearing capacity will be similarly reduced.

5.6  NUMERICAL APPROACHES

Finding the appropriate analytic solution based on the theory of elasticity may be difficult,
especially if the soil is anisotropic or consists of layers that have different stiffnesses. In such
cases, numerical solutions are often necessary and some of these methods are discussed in
the following sections.

5.6.1  Layered soil: Finite layer approaches


Computer programs can be useful in the situation where the soil is layered (see Chapter 2).
Finite layer programs (Small and Booker 1986, 1996) allow elastic anisotropic layered soils
subjected to multiple surface loadings to be analysed. The surface loadings may be strip,
circular, or rectangular in shape, and the programs may be used to find elastic deflections
only, or the time–settlement behaviour of a loading. One advantage of the formulation is
that solutions may be found for a Poisson’s ratio of 0.5 (i.e. the undrained solution). As for
any solutions based on elasticity, the method can be used for prediction of settlement when
the loads are not near failure but restricted to the initial (approximately linear) part of the
load–deflection curve. Undrained soil parameters Eu and νu = 0.5 are used to compute the
undrained settlements and the drained soil parameters E′ and ν′ are used to compute the
drained settlements.
The approach can be used for estimating time–settlement behaviour as well. In this case,
the soil is treated as a two-phase material (soil skeleton and pore water) and the settlements at
required times can be calculated (Booker and Small 1987, Small 2012). The drained modulus
and Poisson’s ratio and the soil permeability are needed for these calculations. The soil may
have anisotropic permeability and the loads may be time dependent in this kind of analysis.
The finite layer method is particularly easy to use, as the only information required is the
thickness of each layer and the material properties of each layer (Small and Booker 1984,
1986). The only restriction is that the material properties of each material layer do not vary
laterally (but can vary from layer to layer).
A plot is shown in Figure 5.32 of the surface deflection computed beneath a single circular
uniform loading applied to the surface of a soil layer consisting of two isotropic sub-layers
where the modulus E1 of the uppermost layer is greater than that of the bottom layer E 2 .
All the parameters used in the computation and the geometry of the problem are shown on
this plot.
Shallow foundations  175

a/2
0.25 pav
r
Deflection × E1/pav E1, ν1 a

E2, ν2 2a
0.50

0.75
E1 ν1 = 0.3
= 2 ν = 0.5
E2 2

1.00
0 0.5 1.0 1.5 2.0
z/a

Figure 5.32 Vertical surface displacement beneath circular loading obtained from finite layer analysis.

5.6.2  Finite element methods


For soft clays or soils loaded close to failure, the use of simple elastic models will not be
applicable, and a more sophisticated means of analysis is needed to compute settlements.
The most common methods used are finite element techniques or finite difference based
methods. Many commercially available computer codes (see those listed in Section 5.4) can
perform non-linear analyses for problems that are two- or three-dimensional in nature.
The analysis may be linear or non-linear. If the soil is treated as being elastic, the settle-
ments can be computed for the case of rapid or undrained loading (using the undrained soil
parameters Eu and νu) although generally a Poisson’s ratio of 0.5 cannot be used with finite
element programs because they are based on a stiffness formulation. A value close to 0.5
(say, νu = 0.49) can be used instead. For slow loading or loading where there is no pore pres-
sure build up as in sands or gravels, the drained parameters E′ and ν′ may be used.
One advantage of these numerical models is that complex geometries, loadings and mate-
rial distributions can be included in the analysis. The soil can also be treated as a two-phase
material (the pore water and the soil skeleton) and consolidation behaviour can be modelled.
Time-dependent consolidation is of great importance for clays that have a low permeability
so pore water pressures can build up when loads are applied to the clay layer. With time, the
pore water pressures can dissipate as pore water flows from areas of high pore water pres-
sure to areas of low pore water pressure, and as this takes place, the soil will consolidate
and undergo settlement.
If loads are high enough to cause yield or failure of the soil, then non-linear models of soil
behaviour can be used with numerical models. There are two types of analysis that can be
performed:

• Single-phase elasto-plastic analysis where the soil is treated as consisting of a single


elasto-plastic material and the appropriate soil model (or constitutive model) is used.

• For clay there are the undrained and drained cases. For the rapid loading or und-
rained case undrained deformation Eu and νu = 0.49 and strength parameters su
and ϕu = 0 are used. As the undrained strength of clays generally increases lin-
early with depth, the strength increase in su is modelled. The drained case is less
176  Geomechanics in soil, rock, and environmental engineering

common, but can be modelled using the effective strength parameters c′ and ϕ′ in
the analysis and the drained deformation parameters E′ and ν′. A Mohr–Coulomb
failure criterion can be used in this case.

• For permeable soils such as sands and gravels, the soil can be modelled using the
drained soil parameters E′ and ν′ and c′ and ϕ′ with a Mohr–Coulomb soil model.
More sophisticated models of soil deformation involving estimation of more soil
parameters can be used.

• Two-phase elasto-plastic consolidation analysis may be used particularly when the


loading is applied at different rates so that it is sometimes a fast undrained loading
and sometimes a slow loading. An example may be an embankment being constructed
on a soft clay where the embankment is built up to a point where pore pressures are
becoming high, and then left for a period of time to allow the pore pressures to dis-
sipate before loading begins again. Such analyses are mostly applied to clays as they
have low permeabilities and allow pore pressures to build up.

The finite element equations that are solved when considering pore pressures using a
‘marching’ solution are given in Equation 5.34 (and Section 3.5).

K − LT   ∆δ   ∆f 
   =   (5.34)
−L −(1 − α)∆tΦ /γ w   ∆q   ∆t Φqt /γ w 

The stiffness matrix of the soil K is the same as for any finite element stress analysis and
depends on the type of element used. The coupling matrices L arise from the coupling of the
pore water pressures and the deformation behaviour of the soil. Φ is the flow matrix, and
arises from the flow of water through the soil, and γw is the unit weight of water. The term
α arises from integration of the pore pressure over a time step. These matrices are defined as


K = BT DB dV

L = ad T dV

Φ = ET kEdV (5.35)
V V V

The matrix D in Equation 5.35 contains the constants relating stress to strain under con-
ditions of elasticity or if the soil is yielding it is the incremental plasticity matrix, k is the
matrix of permeability, B is the matrix relating strain to nodal displacement within an ele-
ment, a is the vector of shape functions relating pore pressure within the element to its nodal
values, d is the vector relating the volume strain within an element to its nodal displacements,
and E is the matrix relating the gradient of pore pressure to the nodal values of pore pres-
sure. The form of the matrices and vectors B, E, a, d depend on the form of element chosen.
In Equation 5.34, the changes in displacements, and excess pore pressures q (at each node
of the finite element mesh) can be found from the excess pore pressure field at the previous
time step qt. Initially, as the excess pore pressures are zero, the flow term in the right-hand
side is zero, therefore the equations can be solved and new values of pore pressure found at
time t = Δt and the process marched forward.
The values of displacement and excess pore pressure at each time are found by updating
the previous solutions as shown in Equation 5.36.

δ t + ∆t = δ t + ∆δ
(5.36)
qt + ∆t = qt + ∆q
Shallow foundations  177

Time (days)
0.001 0.01 0.1 1 10 100 1000 10,000
–0.045

–0.05

–0.055
Settlement (m)

–0.06

–0.065

–0.07

–0.075

Figure 5.33 Settlement versus time for the central point of circular loading on a soil layer of finite depth –
drained upper and lower surfaces of clay layer.

As with many ‘marching’ processes, errors in the previous solution can be magnified in
the current solution and the process can become unstable. The stability criterion has been
established by Booker and Small (1975) who showed that the process is unconditionally
stable as long as the integration parameter lies in the range 0 ≤ α ≤ 0.5, but that it can be
stable under other conditions that depend on the eigenvalues of the consolidation equations.
It may be necessary for problems where a constant load is applied and then held constant
to increase the time step during calculation as the rate of consolidation slows down with
time. However, when the time step in Equation 5.34 is increased, the set of equations has to
be set up again and resolved.
An example of a finite element calculation using eight-node isoparametric elements to
interpolate both the displacement field and the pore pressure field is given in Figure 5.33.
The problem involves a circular uniform loading q applied to the region 0 ≤ r ≤ a on the
surface of a soil layer of depth h. The entire upper surface of the soil is assumed permeable
and the base is assumed to be underlain by a permeable layer of sand. The parameters used
for the solution are presented in Table 5.8.
The settlement at the central point of the loading versus time is presented in Figure 5.33.
In this figure, the finite element solution is shown plotted over seven log cycles of time. The
time step has been increased by a factor of 5 nine times during the calculation as can be
seen from the increased distance between the plotted circles where the time step is increased.

Table 5.8  Properties used in finite element analysis


Quantity Value
Drained modulus of elasticity 10,000 kPa
Drained Poisson’s ratio 0.35
Radius of load a 4 m
Depth of layer h 16 m
Vertical and horizontal permeability kv = kh 0.0001 m/day
Uniform load q 80 kPa
178  Geomechanics in soil, rock, and environmental engineering

70

60

Critical state line


50
Deviator stress q (kPa)

40 X
Stress path

30

20
su

10

0
0 10 20 30 40 50 60
Mean stress p (kPa)

Figure 5.34 Stress path taken for undrained consolidation analysis of isotropically consolidated clay.

Non-Linear Solutions

It is necessary to use the correct model of soil behaviour in this case, as the model must be
capable of giving the correct behaviour for all rates of loading. Models that contain a yield
and a failure surface are best suited to this kind of analysis as they can correctly simulate
the soil behaviour when undrained. For example, Figure 5.34 shows the effective stress path
that may occur in a soil that was initially in an isotropic stress state. Where the stress path
reaches the failure surface (at point ‘X’) the deviator stress is the undrained shear strength
of the soil (i.e. su = qfailure). The stress path can be seen to deviate to the left after it intersects
the oval-shaped yield surface.
Various soil models may be implemented, one of the most popular, being the Cam Clay
Model that was developed for clays (see Britto and Gunn 1987). Commercial computer
codes that include models that can be used when a clay is treated as a two-phase mate-
rial (i.e. soil skeleton and pore water) are the Modified Cam Clay Model (CRISP, Phase2,
ABACUS, FLAC) or the Soft Soil Model (PLAXIS2D). These models for clays were dis-
cussed previously in Chapter 3.

5.6.3  Estimation of soil parameters


In order to use the analytical methods described previously, it is necessary to estimate the
soil parameters. For many of the more complex non-linear soil models, this may require
special laboratory testing, although for models like the Cam Clay Model, the parameters
can all be determined from standard laboratory triaxial and oedometer tests.
Here, attention will be restricted to determination of elastic soil properties, as these form
the basis of many of the hand-based methods of calculation mentioned previously. The soil
parameters needed Eu , E′, and ν′ are generally determined either from laboratory tests or
from in situ tests. Commonly used field tests are SPT tests, plate loading tests, pressuremeter
Shallow foundations  179

tests, dilatometer tests, static cone penetration tests, and screw plate tests. See Chapter 4 for
more details on field testing and properties of soils.
These field tests often rely upon empirical correlations between the test result and a par-
ticular parameter, for example, correlations between elastic modulus and the SPT or elastic
modulus and the cone penetrometer resistance qc as was seen in the previous Section 5.5.5
on the settlement of footings on sand. As stated earlier, laboratory testing is rarely used for
sands, as it is difficult to keep sand samples in an undisturbed state (at the same density)
as in the field. Correlations between shear modulus of sands and field test results from
pressuremeters, dilatometers, SPT tests, and cone penetration tests have been presented by
Décourt (1994) for use in the computation of settlement of footings.
For clays, the elastic moduli can be found from triaxial testing or from field tests. Triaxial
testing normally involves consolidating undisturbed soil samples back to the in situ stress
state before loading them in an undrained condition (to obtain the undrained parameters)
or in a drained condition (to obtain the drained parameters). Methods of testing have been
described in Davis and Poulos (1968) and in Ladd and Foott (1974). Other methods have
also been discussed in Chapters 1 and 4 of this book.
For clays also, a common procedure is to relate the undrained modulus Eu to the und-
rained shear strength of the clay su. This depends upon the overconsolidation ratio of the
clay, therefore one needs some knowledge of both the undrained shear strength and the
­overconsolidation ratio of the clay to predict undrained moduli. Often this relationship can
be established through experience in certain areas, so it is only necessary to obtain the und-
rained shear strengths to predict moduli. Correlation between the undrained shear strength
of clays and the undrained modulus for various overconsolidation ratios has been presented
by Duncan and Buchignani (1976) (Figure 4.45).

5.7  RAFT FOUNDATIONS

While settlements of pad footings may be calculated by treating the foundation as being
perfectly flexible or perfectly rigid, it is generally necessary to take into account the actual
stiffness of a raft or mat foundation. This is because differential settlements or moments
and shears in the raft usually need to be calculated, therefore the structural rigidity of the
foundation has to be considered.
Often solutions are found by treating the soil beneath the raft as consisting of a series of
springs (a Winkler foundation) as this makes the analysis simpler; however, such methods
are to be used with caution, as they cannot represent true soil behaviour. For example, a
flexible raft carrying a uniform vertical loading (over the whole raft) would be predicted to
settle uniformly with no differential settlement apparent if on a spring foundation, and this
is obviously contrary to observed behaviour. A rigid raft would also make all the springs
compress equally, and hence the load in each spring would be equal. Measurement of con-
tact stress below rigid rafts shows that the contact stress is generally larger at the edges (for
a uniform loading on the raft), so again, the load distribution in the springs would not be
modelling the correct behaviour.
It is therefore desirable to treat the foundation as an elastic continuum, as the observed
behaviour of rafts can be more closely simulated. Full finite element analyses may be neces-
sary if the soil has highly non-linear properties, but generally, the theory of elasticity is suf-
ficient, as rafts are not designed to carry loads that are going to cause yielding of the soil. A
soil modulus representative of the stress levels in the soil can be used in such cases.
Solutions have been found to some problems involving rafts constructed on an elastic
continuum, but the solutions are limited because it is difficult to present results for all raft
180  Geomechanics in soil, rock, and environmental engineering

shapes and for different soil conditions such as layering or modulus increasing with depth.
The book by Selvadurai (1979) contains analytic solutions to problems involving rafts on
elastic continua; however, most solutions have been based on numerical procedures. Some
of the available solutions are presented in the following sections.

5.7.1  Strip rafts


A strip raft may be classified as a raft that has a length L that is much greater than its width
B. Solutions to such a problem have been presented by Brown (1975) for a strip raft, hav-
ing a length to width ratio L/B of 10, resting on an infinitely deep elastic soil. The raft is
considered to be loaded by a point load which is located at a distance sL from one end as is
shown in Figure 5.35. The results are presented for a single point load as the solutions may
be superimposed if more than one point load is acting on the raft.
Shown in Figure 5.36a are the solutions for the deflections in the strip raft while
Figure  5.36b shows the moments in the raft. These particular solutions are for the case
where the relative stiffness of the raft to the soil K has a value of 0.01. The relative stiffness
is defined as

16Er I(1 − ν2s )


K = (5.37)
πEsL4
where Er is the elastic modulus of the raft, I is the second moment of area of the raft
(I = Bt 3/12 for a strip raft of thickness t), E s is the modulus of the soil, and νs is Poisson’s
ratio of the soil.
Solutions for the moment or displacement in the raft are presented in non-dimensional
form at various positions x along the strip raft, for a point load of magnitude P. If several
point loads are applied along the raft, the appropriate values of P (corresponding to each
load) should be used when computing the moments or deflections before superimposing (i.e.
adding) them.

5.7.2  Circular rafts


Brown (1969) has also obtained solutions for circular rafts subjected to uniform vertical
loading q over their entire upper surface. The raft is assumed to be of thickness t and radius
a, and to rest on a layer of soil of depth h.
In this case, the relative stiffness of the raft K is defined as
3
Er (1 − vs2 )  t 
K =  a  (5.38)
Es

sL P
Strip Er, νr

L B

Soil Es, νs
x

Figure 5.35 Strip raft subject to point load.


Shallow foundations  181

x/L
(a) 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
–2
–1

Displacement × LEs/P(1 – ν2s )


0
1 s = 0.5
2
0.4
3
0.3
4
0.2
5
6 0.1 K = 0.01
7 0.01
8
9

x/L
(b) 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
0.10
0.4 0.5
0.3
0.2 K = 0.01
0.05
s = 0.1
Moment/PL

–0.05
s = 0.01

–0.10

Figure 5.36 Solutions for point load on strip foundation: (a) displacement distributions; (b) moment distri-
butions. (After Brown, P.T. 1975. Geotechnical Engineering. Vol. 6, pp. 1–13.)

Maximum moment in the raft occurs at its central point, and this moment is shown plot-
ted in non-dimensional form against raft stiffness K in Figure 5.37. Solutions are presented
for different depth to radius ratios h/a, and for two extremes of Poisson’s ratio ν = 0 and
0.5. In this case, the moments presented in the plot are actually moments per unit length
(i.e. units are kNm/m run). Differential deflections in the raft are shown in Figure 5.38 for
various soil layer depths and Poisson’s ratios. For a circular raft, the differential deflection is
defined as the difference in displacement between the centre and the edge of the raft.
Booker and Small (1983) have extended the charts of Brown to include the effects of the
walls as well as the base raft of a tank. The restraining effect of the walls at the outer cir-
cumference of the base means that there is a moment generated there that is opposite in sign
to the moment at the centre of the base and may be large if the walls are stiff enough to resist
the rotation at the base–wall junction.

5.7.3  Rectangular rafts


Rafts of rectangular shape (in plan) that carry a uniformly distributed load q have been
analysed by Fraser and Wardle (1976). The rafts are assumed to be constructed on a layered
soil profile, and methods are given in the paper to enable an equivalent elastic modulus and
182  Geomechanics in soil, rock, and environmental engineering

0.08

2
2

0.06 1

h/a = 1
0.04
0.5
2
(Moment/unit length) qa
M

0.02
0.2

0.5

–0.02

νs = 0.5
0.2
νs = 0

–0.04
10–2 10–1 1 101 102 103
K

Figure 5.37 Maximum moments in circular rafts of various stiffnesses. (After Brown, P.T. 1969. Géotechnique,
Vol. 19, No. 2, pp. 301–306.)

Poisson’s ratio to be determined for the layered system so that it may be treated as a uniform
material.
The raft of length L and width B is shown as the inset to Figure 5.39. The relative stiffness
K of the raft as defined by Fraser and Wardle is

4Er (1 − ν2s )t 3
K = (5.39)
3Es (1 − ν2r )B3

where the terms have the same meanings as for the strip raft (see Section 5.7.1) and νr is
Poisson’s ratio of the raft. In Figure 5.39, points α, β, γ, and δ are identified on the raft. The
deflections at these points and the differential deflections between them can be calculated
by use of the influence factors shown on the figure. For instance, the curve labelled I α would
be used to compute the settlement at point α, whereas I αβ would be used to compute the dif-
ferential settlement between point α and point β.
To compute the deflection or differential deflection ρsi for an infinitely deep soil layer the
following equation is used

qB(1 − ν2s )
ρsi = ⋅I (5.40)
Es
Shallow foundations  183

0.8

h/a = ∞ νs = 0.5
νs = 0
2
0.6
2
Differential displacement × Es/(1 – ν2s)qa
1

0.4
1

0.5
0.2

0.5

0.2
0
0.2

–0.2
10–2 10–1 1 101 102
K

Figure 5.38 Differential deflections (centre to edge) in circular rafts of various stiffnesses. (After Brown, P.T.
1969. Géotechnique, Vol. 19, No. 2, pp. 301–306.)

where I is the appropriate influence factor as determined from Figure 5.39. To compute the
settlement or differential settlement for a soil layer that is not semi-infinite, corrections need
to be applied to ρsi. The actual displacement ρ is computed from the following expression
where S is a correction factor that is found from charts like those shown in Figure 5.40 (for
a raft with a length to breadth ratio L/B of 2 on a soil layer of depth d).

ρ = Sρsi (5.41)

Similar charts have been provided by Fraser and Wardle to enable moments to be com-
puted in the raft. This involves calculating moments for the case where the soil is infinitely
deep, and then applying a correction for soil layers of finite depth, as was done for the
displacements.

5.7.4  Raft foundations of general shape


When the raft is of general shape and is subjected to general loading (that may include point
loads, moments, or uniform loads), it becomes difficult to provide solutions in the form of
charts. For more complicated cases, it is necessary to use computer programs, and these may
analyse raft behaviour by using finite difference methods or by use of finite element meth-
ods. Analysis of the soil behaviour may be carried out by the use of elasticity theory or by
184  Geomechanics in soil, rock, and environmental engineering

1.6

L/B = 2
1.4

1.2

1.0 Iδ
B

0.8 Iγ
β
I

L α
Iαγ
0.6 δ γ
Iαδ

0.4 Iαβ

0.2

0
10–4 10–2 100 102
K

Figure 5.39 Settlement influence factors for rectangular raft on an infinitely deep layer (L/B = 2).

use of finite element techniques. An early demonstration of the use of finite element methods
for the raft and solutions based on the theory of elasticity for the soil were presented by
Cheung and Zienkiewicz (1965).
For rafts on layered soils, the soil behaviour can be analysed using the finite layer method
(see Chapter 2) and finite element methods can be used for determining raft behaviour. The
raft is divided up into finite elements, for example, in the program FEAR (Finite Element
Analysis of Rafts – Small 2013), eight-node isoparametric shell elements are used for the
raft. This type of element allows the moments and shears per unit length to be computed in
the raft as well as the deflections.
The process of solving the problem is given in Section 6.37.2 (that also includes piles), and
involves obtaining the contact stress distribution between the raft and the soil. The use of finite
elements for the raft allows point loads, moments and distributed loads to be applied to the raft
and the raft can be of any shape and have different thicknesses in different parts of the raft.
The soil response can be calculated using finite layer techniques, which makes the analysis
quick and simple to use, although the soil response could be calculated using finite element
methods. Finite layer techniques allow layered soils to be used for the foundation although
this means that the soil must be considered to have an elastic response.
An example of this is shown in Figure 5.41 where the computed displacement contours in
the raft are shown. The centre of the raft can be seen to be deflecting more than the edges
in Figure 5.41b.

5.8  REACTIVE SOILS

Soils such as plastic clays that contain clay minerals like montmorillonites or smectites
tend to shrink when their moisture content decreases or to swell as their moisture content
Shallow foundations  185

(a) 1.0
Sαγ
0.8
Sαβ

0.6 α β
S

0.4
Sαδ δ γ
Flexible
0.2 K = 0.5
νs = 0, L/B = 2 Rigid
0
(b) 1.0

0.8 Sαγ

Sαβ
0.6

S

0.4
Sαδ
0.2
νs = 0.3, L/B = 2

0
(c) 1.0

0.8 Sαγ

0.6 Sαβ
S


0.4

0.2 Sαδ
νs = 0.5, L/B = 2

0
0 0.2 0.4 0.6 0.8 1.0 0.8 0.6 0.4 0.2 0
D/B B/D

Figure 5.40 Settlement correction factor (L/B = 2).

increases. Soil suctions exist in the uppermost partially saturated ‘active zone’ in the soil and
it is changes in these suctions (due to moisture content variations) that cause shrink–swell
behaviour.
Swelling can cause damage to structures especially if differential movements occur in
the structure. For example, moisture changes may be small beneath the centre of a raft or
mat foundation, thus little swelling or shrinking will occur. At the edges of the foundation,
moisture changes are larger, therefore swell–shrink movements are large. This creates dif-
ferential movements in the foundation, which are seasonal, and this can cause damage if the
raft is not stiff enough to resist the movements.
186  Geomechanics in soil, rock, and environmental engineering

(a)
Point and distributed loads

Uniform loads
1.877E+03
2.110E+03
Mx5
Mx4
My4 Mx3
2.334E+03
My5 My3
Mx2 2.512E+03
My2
Mx6
My6 2.171E+03
Mx1
My1 3.184E+03
Mx7
My7
Mx24
My24
Mx8
My8
Mx23
My23
Mx9
My9
My26 Mx22
My22
Mx10 My25 My28
My10
My27 Mx21
My21
Mx11
My11
Mx20
My20
Mx12
My12
Mx19
Mx13 My19
My13
Mx18
Mx14
My18
My14
Mx15 Mx17
Mx16 My17 Loads and pressures in (kN) and (kPa)
My15
My16
Raft – 800 mm/1200 mm thick
2 layers, 3 and 20 m
LC155 – full-size model

Foundation for wind farm

(b)
Contours of vertical displacement

Contour legend
0.000
0.005
0.010
0.015
0.020
0.025
0.030
0.035
0.040
0.045
0.050

Displacement in (m)
Raft – 800 mm/1200 mm thick
2 layers, 3 and 20 m
LC155 – full-size model

Foundation for wind farm

Figure 5.41 (a) Finite element mesh and (b) displacement contours for raft on a layered soil. (From the
Program FEAR [Finite Element Analysis of Rafts]).
Shallow foundations  187

5.8.1  Pad or strip footings


A fairly simple means of computing swell movements has been reported by O’Neill and
Poormoayed (1980). The amount of surface heave y can be calculated by summing the swell
occurring in a number of sub-layers in the active soil zone as shown in Equation 5.42.
N

y = ∑s(%)h /100
i (5.42)
i =1

where s (%) is the swell as a percentage (found from a swell test) in layer i. The thickness of
each layer is hi and there are N layers in the active zone.
To obtain the swell s (%), undisturbed samples of soil taken from representative depths
are placed in an oedometer under a pressure equal to the overburden pressure plus the
anticipated surcharge. The sample is then inundated and sufficient extra load applied in
small increments to prevent swelling. Finally, the sample is unloaded back to the initial pres-
sure (overburden plus surcharge) in decrements and the swell measured (as a percentage of
original sample height). O’Neill and Poormoayed (1980) have stated that samples should
be taken during the construction period, so that the method will predict the expected swell
from the time of construction if the soil is totally saturated (i.e. the worst case).
Other methods used for computing swell that are used in North America are reported in
Settlement Analysis (1994) adapted from the U.S. Army Corps of Engineers, and are based
on oedometer swell tests.
A method that is commonly used in Australia (Residential Slabs and Footings 2011, see also
Mitchell and Avalle 1984) to compute the likely heave of foundations is based on determining
the changes in soil suction beneath a foundation and relating these to soil movements. Soil
suction occurring in partially saturated soils is commonly expressed in pF units where

pF = log10 (suction in kPa)+1 (5.43)


Soil suctions can be measured by using commercially available psychrometers or by using


filter paper techniques, although this is not necessary if the simplified methods outlined
below are used.
An instability index Ipt is used to relate changes in vertical strain in the soil to changes
in the soil suction. Soil suction changes are the result of wetting and drying of the soil in
the zone above that part of the soil which is permanently saturated. The instability index is
related to the swell–shrink index Iss of a soil and this index may be found from experiment
by allowing the soil to swell when inundated with water and to shrink when dried.
The swell test is carried out in an oedometer (or consolidation cell) by loading a sample
at natural moisture content with a pressure of 25kPa and allowing the sample to come to
equilibrium under the load. The sample is then inundated with water and allowed to swell
for a minimum of 24 h or until movements are only 5% of the total swelling up to the time
of reading. The swelling strain ϵsw (in percent) is then computed from

(R1 − R2 ) × 100%
ε sw = (5.44)
H0

where
R1 is the final dial gauge reading in mm
R 2 is the initial dial gauge reading (before inundating) in mm
H 0 is the initial specimen thickness in mm
188  Geomechanics in soil, rock, and environmental engineering

The shrinkage test is performed on a cylindrical soil sample (at natural moisture content)
of length 1.5–2 diameters. A drawing pin is pushed into each end of the sample, and the
distance D 0 between the rounded heads of the sample measured. The sample is then dried
in an oven, and the final length is measured between the pins Dd. The shrinkage strain ϵsh
(percent) is then calculated from

(D0 − Dd ) × 100%
ε sh = (5.45)
H0

where H 0 is the average initial length of the sample (not including the drawing pins).
The shrink–swell index Iss is then computed from

 ε sw 
 2  + ε sh (5.46)
I ss =
1.8

The swell strain is divided by 2 to allow for the effect of lateral restraint in the swelling test.
The factor of 1.8 is the change in pF value assumed to take place over the range of moisture
contents experienced in the test, and can be replaced by an actual measured value if desired.
The swell–shrink index is therefore a measure of the percentage strain per pF change in
suction.
The instability index is related to the swell–shrink index by Ipt = α · Iss where α is intro-
duced to take account of whether the soil is in a cracked (unrestrained) zone or in an
uncracked zone.

1. In a cracked zone

α = 1.0

2. In an uncracked zone (where the soil is restrained laterally by surrounding soil and
vertically by soil weight)

α = 2.0 − z/5

where z is the depth (in metres) from the ground level to the point under consideration
in the uncracked zone. As suction changes generally only occur over 4 m or less, α will
have a positive value.

The design surface movement ys is then calculated from

H
1
ys =
100 ∫
I pt ∆udh (5.47)
0

where
Δu is the suction change at depth h (in pF units)
Ipt is the instability index
H is the depth of the active zone
Shallow foundations  189

Soil surface
Δu Δu Δu
level

H
Bedrock Water table
H H level

Consistent soil case Effect of bedrock Effect of groundwater

Figure 5.42 Change in soil suction Δu with depth.

The integral of Equation 5.47 can be treated as a summation if so desired.


The method therefore depends on being able to predict the soil suction changes with depth
for use in Equation 5.47, as well as a knowledge of the cracked zone. It is often assumed that
soil suction Δu varies linearly with depth as shown in Figure 5.42 and the values of suction
change at the surface and the depth of influence H can be established for certain areas (for
example, in Adelaide Δu = 1.2 pF and H = 4 m unless the water table is higher in which case
H = depth to water table). In Sydney, the values are Δu = 1.5 pF and H = 1.5 m. The depth
of cracks can also be established on a regional basis (for example in Adelaide the depth can
be taken as 0.75H).

5.8.2  Rafts on reactive soils


The swelling and shrinking of reactive soils can produce differential movements in raft
foundations, and this is usually due to soil movements at the edges of the raft where mois-
ture changes are most likely to occur. Rafts constructed in areas where soils are reactive are
generally stiffened with ribs (i.e. the waffle slabs as shown in Figure 5.1d).
If swelling and shrinkage movements take place around the edges of a raft foundation,
the raft will undergo uplift at the edges (swelling) or will be supported on a mound of soil if
the soil shrinks away at the edges. Analysis of this type of behaviour has been examined by
Sinha and Poulos (1996) using finite element analysis of the raft, and by Li et al. (1996) who
have used finite element methods to analyse moisture flow and soil–structure interaction for
a stiffened raft foundation.

5.9  COLD CLIMATES

In cold climates, when the ground begins to freeze, water can be drawn upward towards the
surface by capillary action (from the water table below) and can form ice lenses. The lenses
usually grow perpendicular to the direction of heat flow, which for most soils is parallel to
the ground surface. The increase in volume of the ice will cause the surface of the soil to
heave, and these upward movements can be very large (i.e. 300 mm or more) causing struc-
tural damage.
Frost-susceptible soils are those that have pores small enough to produce ice lenses at the
freezing front (i.e. the surface separating frozen and non-frozen soil). Silty materials, fine
190  Geomechanics in soil, rock, and environmental engineering

silty sands with more than 15% of particles finer than 0.02 mm and lean clays having a
plasticity index <12, varved clays and chalks are most susceptible to frost heave while grav-
els and clean sands are not because of their high permeability. Clays are less susceptible to
heave because their low permeability prevents the flow of water (and hence the formation
of ice lenses).
For a soil to exhibit frost heave, there must be sufficiently low temperatures, water avail-
able in the soil, and the soil must be of a type (as mentioned above) that is frost susceptible.
Freezing tests may be carried out on soil samples to determine frost heave potential.
Freezing of a foundation may be due to sub-zero climatic conditions, or it may be due to
man-made causes, for example freezer rooms or ice rinks (see Chapuis 1988). In the former
case, resistance to frost heave may be engineered by draining the soil, founding footings
below the depth of frost penetration or replacing frost-susceptible soils by non-susceptible
materials such as gravels or sands. For man-made sources of low temperature, these meth-
ods may also be used or good insulation provided between the soil and the structure.
Prediction of frost heave is not generally a simple matter, as any analysis involves com-
putation of the heat flow in the soil, the thermal properties of the soil, the cooling rate and
suction at the frost front. Such analyses have been reported by Konrad and Morgenstern
(1982) and Nixon (1991). Field measurements and numerical predictions of the measured
heaves have also been reported in Hayhoe and Balchin (1990).

APPENDIX 5A

For soils that are compressible, the bearing capacity solutions need to be modified to take
account of punching failure. The classical bearing capacity equations are for rigid plastic
materials, and therefore do not apply to softer or looser soils.
To deal with loose or compressible soils, Vesic (1975) introduced a rigidity index Ir defined as

G
Ir = (5A.1)
c + q tan φ

where
G is the elastic shear modulus of the soil
c, ϕ are the soil strength parameters
q is the vertical overburden pressure evaluated at a depth of B/2 below foundation level
(B is footing width)

The value of Ir is compared to a critical rigidity index Irc where

1   B   φ
Irc = exp  3.30 − 0.45    cot  45o −   (5A.2)
2   
L   2 

When Ir > Irc the soil behaves like a rigid plastic material and there is no need to use the
modifying factors (i.e. they can all be considered to be 1). If Ir < Irc the soil is considered
compressible and punching shear may occur. The correction factors in Table 5.2 may then
be used.
Chapter 6

Deep foundations

6.1 INTRODUCTION

Deep foundations are defined as those where the depth of the foundation exceeds the breadth.
Such foundations are generally caissons, barrettes, or piles. The most common type of deep
foundations is piles, but barrettes (which are deep foundations of rectangular cross section)
or caissons (which are large hollow foundations) can be regarded as special types of piles.
Piles are generally used to transfer the loads of structures down to rock or to deeper stiffer
layers. This improves the bearing capacity of a foundation or has the effect of reducing
settlements. However, there are other reasons for the use of piles:

1. As anchors to resist uplift forces


2. To resist against lateral forces, for example, bridge piers, wharves
3. To provide a deep foundation in cases where scour may remove upper soil layers
4. As foundations where shrinking and swelling soils exist. The piles transfer loads down
below the active surface zone, where water content changes occur

6.2 TYPES OF PILES

Piles can be classified in different ways:

1. Classification by material from which the pile is made:


• Steel: A steel pile may be an ‘H’ section or hollow circular section. Sheet piles have
sections that may be interlocked with the adjacent sheet pile
• Concrete: Concrete piles are generally reinforced with steel or may be pre-stressed
• Timber: Timber piles may be treated with chemicals to prolong their life
• Composite: Composite piles may be combinations of steel and concrete such as
concrete-filled steel pipe piles
2. Classification by the effect of installation:
• Displacement piles (i.e. piles that displace the soil when driven such as a large
closed end pipe pile)
• Low displacement piles (such as ‘H’ section piles)
• Non-displacement piles (such as drilled piles)
3. Classification by method of installation:
• Driven
• Driven tube filled with concrete
• Bored (drilled) piles
• Composite piles

191
192  Geomechanics in soil, rock, and environmental engineering

• Screwed (Atlas)
• Pushed
• Vibrated

6.2.1 Driven piles
Steel, concrete, or timber piles can be driven into the soil by a hammer. Examples are

• Timber piles: Treated with chemicals so as to resist insects and rotting. Such piles
can be used in soils with pH ranging from 2 to 11.5. Piles can be spliced, and cut off
to length. Treated piles can be expected to last in excess of 100 years. In fact, the first
masonry London Bridge built in 1176 stood on untreated elm piles and lasted 600
years. Marine hardwood piles have to withstand attacks from borers and therefore
are impregnated with chemicals. They are vacuum/pressure impregnated twice. The
first treatment is with CCA (copper chrome arsenic) preservative and after re-drying
of the wood, the second treatment is conducted with PEC (pigmented emulsified
creosote).
• Precast concrete piles: Such piles can be solid or hollow and are reinforced with steel
cages. Precast concrete piles are best where there are soft materials overlying rock or
hard strata. The pile sections can be spliced if desired to extend their length. A precast
pile is shown in Figure 6.1.
• Pre-stressed concrete piles: These can be solid or hollow sections and are pre-stressed
in a production yard.
• Steel piles: Solid steel piles can be ‘H’ section, box, or tube (also interlocking sheet
piling).

6.2.2 Driven and precast piles


A steel tube with a closed end is driven into the soil and the tube filled with concrete. The
steel tube may be withdrawn or it may be left in place. There are variations where a steel
shell is driven by a withdrawable mandrel and the shell is concreted.

6.2.3 Jacked piles
In this case, the pile is jacked into the ground against a large reaction. The rigs that are used
to do this can be very heavy and may be difficult to get into areas of soft soil. They may also
cause movement of existing foundations. Steel or concrete piles can be jacked.

(a)

(b)

Steel shoe Spiral steel cage

Figure 6.1 Precast reinforced concrete piles: (a) solid square; (b) hexagonal.
Deep foundations  193

Figure 6.2 Steel cages ready for placing inside a drilled shaft.

6.2.4 Bored piles
There are various types of bored piles, but most are formed by drilling a hole, and then plac-
ing the reinforcing cage before concreting. This is shown in Figures 6.2 and 6.3. Advantages
of bored piles are

• The base of the hole can be under-reamed to provide an enlarged base.


• A continuous flight auger (CFA) can be used (see Figure 6.4).
• It is possible to drill into rock using a bucket with tungsten tipped teeth.
• There is less vibration and noise than for a driven pile.
• Soil is not displaced (such as for a large cross-sectioned driven pile), so there is less
chance of damaging nearby structures.
• Piles can be of large diameter to support very heavy loads.

6.2.5 Composite piles
Piles can be combinations of materials, for example, steel piles placed into a concreted pile
or an ‘H’ section steel pile encased in concrete.

Figure 6.3 Drilled shaft pile being installed. Concrete is tremied into the shaft displacing water.
194  Geomechanics in soil, rock, and environmental engineering

Figure 6.4 Continuous flight auger rig.


Deep foundations  195

6.2.6 Grout injected piles


Jet grouted piles are formed by drilling to the required depth and then injecting high pres-
sure grout that cuts away the soil and forms a mix of soil and cement slurry.

6.3 INSTALLATION

Piles can be classified according to installation methods, whether they are large displace-
ment piles or piles that cause very little displacement.
If large diameter piles are driven into the soil, they will cause the soil to be pushed side-
ways as the pile head advances, therefore the soil must undergo lateral compaction. Such
piles are called ‘displacement piles’. If a hole is bored initially, then concreted, the soil under-
goes minimal disturbance and the pile is termed a ‘non-displacement pile’.

6.3.1 Types of displacement piles


Types of displacement piles are

1. Timber piles that are often used for marine and smaller scale structures.
2. Steel tubes, that can be readily extended, but may corrode and are generally more
expensive.
3. Precast concrete piles: Common lengths are from 12 to 15 m and have the advantage
of lower cost. They are not suited to hard driving, and can be spliced, but not as easily
as steel piles.
4. Proprietary types: Many use steel casings that are withdrawn as concrete is placed.

Common problems are

• Vibration from driving


• Excess pore pressures generated during driving
• Lateral soil movements occur
• Access for driving rigs may be difficult

6.3.2 Small displacement piles


Types of small displacement piles are

1. ‘H’ sections and rolled steel sections: Such piles are useful for punching through thin
hard layers. They can have a steel point to assist in driving. However, thin sections
may bend about the weakest axis of the pile.
2. Steel tube piles: The plug of soil that is pushed up inside the pile can be removed after
driving. Tube piles are often filled with concrete once cleaned inside.
3. Pre-drilled piles: Such piles are usually pre-bored over part of their depth, and then
driven. They are useful if there are hard layers near the surface.

6.4 PILE DRIVING EQUIPMENT

If piles are to be driven into the ground, they are usually driven using a heavy drop hammer,
although vibration can also be used. Piles can be driven vertically or at an angle (called a
‘raked pile’).
196  Geomechanics in soil, rock, and environmental engineering

6.4.1 Piling rigs
A piling rig usually consists of a set of leaders mounted on a standard crane base. Leaders
consist of a stiff box or tubular member which carries and guides the pile as it is driven into
the ground. Leaders can be raked backwards or forwards to install raked piles. A piling rig
is shown in Figures 6.5 and 6.7.

6.4.2 Piling winches
The piling winch is used for raising the pile and (if necessary) the hammer. It may be pow-
ered hydraulically or by steam, diesel, or petrol engines.

6.4.3 Piling hammers
There are several different types of hammers that are used to drive a pile into the ground.
Some hammers like drop hammers, simply rely on their weight to drive the pile, while others

Figure 6.5 Rig for driving piles.


Deep foundations  197

Hammer at
top of stroke

Diesel fuel
injected as
hammer falls

Figure 6.6 Diesel hammer.

like the diesel hammers provide the hammer with an extra force to accelerate the hammer
onto the pile head. Types of hammers are

1.
Drop hammers
The drop hammer is simply a weight that is dropped onto the pile head. A typical ham-
mer mass is 1–5 tonnes.
2.
Single acting steam or compressed air hammer
Steam or compressed air is used to raise the hammer, and once at the required height,
the hammer is allowed to fall under the action of gravity.
3.
Double acting hammers
These hammers are used mainly for sheet piling. They can impart a rapid succession of
blows (more than 100 blows per minute) and rely on steam or air to raise and acceler-
ate the hammer downward. They do not need a guide but can be attached to the head
of the sheet pile by leg guides.
4.
Diesel hammers
The falling ram compresses air in a cylinder into which diesel fuel is injected. The die-
sel fuel ignites, forcing the hammer back up the cylinder and the reaction imparts an
additional force to the pile head (Figure 6.6). Diesel hammers may not be effective in
all soil types, for example, soft soils and may cause breakage of concrete piles if a hard
layer is encountered.
5.
Hydraulic hammers
The hammer is raised by hydraulic fluid and then hydraulic pressure is used to accel-
erate the hammer onto the pile (Figure 6.8). Hydraulic hammers are less noisy than
diesel hammers and can impart a wide range of driving energies.
198  Geomechanics in soil, rock, and environmental engineering

Figure 6.7 Precast concrete piles being driven.

Actuator
Lift cylinder

Segmented ram
weight

Shock adsorber
Pile sleeve

Figure 6.8 Hydraulic hammer.


Deep foundations  199

6.
Vibratory hammers
A vibrating head is attached to the top of the pile and vibration and downward force
causes the pile to be driven into the soil. This type of pile driver is best suited to sandy
or gravelly soils. The technique is also useful for extracting piles such as sheet piles or
the casing of bored piles.

6.4.4 Helmet, driving cap, dolly, and packing


Various components are placed on the pile head to stop it from suffering damage due to
driving. These are

1.
Helmet: This is placed over the top of a concrete pile. It holds the dolly and packing
between the hammer and pile to prevent damage to the pile head.
2.
Driving cap: The cap is used to protect the heads of steel piles, and is shaped for the
particular type of pile being driven. The cap is fitted with a recess for a hardwood or
plastic dolly.
3.
Dolly: The dolly is placed in a recess in the top of the helmet. It can be made of timber
or plastic. Under moderately hard driving conditions, plastic dollies can last for several
hundred piles.
4.
Packing: This is placed between the helmet and pile top to cushion the blow between
the two. Materials used include hessian, paper sacking, thin timber sheets, coconut
matting, and wallboards.

6.5 PROBLEMS WITH DRIVEN PILES

Driving piles into hard soil with a heavy hammer can cause damage to the pile and as the
pile may not be visible above ground, this may not be detected. Types of damage can be

1.
Steel piles: The top or shaft of the pile can be damaged by driving. Tubular piles may
collapse or the pile toe may be damaged. Base plates may rise relative to the casing, or
plugs and shoe may be lost in a cased pile.
2.
Concrete piles: The head, shaft, or toe can be damaged.
3.
Timber piles: The head or toe can be damaged.

6.5.1 Problems from soil displacement


Driving large displacement piles can cause problems because of the soil movements both
vertically and laterally. Some of the problems caused by soil movements are as follows:

Vertical Soil Movements

1. Uplift causing squeezing, necking, or cracking


2. Uplift resulting in shaft lifting off base
3. Uplift resulting in loss of end bearing capacity
4. Ground heave lifting pile
5. Ground heave separating pile segments or units or extra tensile forces being placed on
joints
200  Geomechanics in soil, rock, and environmental engineering

Lateral Soil Displacements

1. Squeezing or waisting of piles


2. Inclusions of soil forced into pile
3. Shearing of piles, or bends in joints
4. Collapse of casing prior to concreting
5. Movement and damage to adjacent structures

6.6 NON-DISPLACEMENT PILES

Non-displacement piles are commonly used where large loads are to be carried or vibration
and noise from driving are not wanted. Common installation methods include drilling a dry
hole, drilling the hole supported by a slurry or by drilling a hole supported by a steel casing.
Holes when drilled should be concreted preferably within the same day, as leaving the
hole open can lead to softening of the soil along the shaft with resultant loss of bearing
capacity. Slurries such as bentonite can form a cake along the sides of the drilled hole and
this too can lead to loss of bearing capacity. Some of the advantages and disadvantages of
drilled shaft foundations are listed below:

1. Advantages
• Absence of ground heave
• Absence of excessive noise and vibration
• Piles can be installed where headroom is limited
• Length and diameter can easily be varied
• Base can be enlarged
• Can obtain high capacities
• Can inspect prior to concreting

2. Disadvantages
• Loosening of sandy soils
• Softening of clays
• Waisting or necking of pile may occur in loose or soft soil if support from casing
or slurry not used
• Inflow of water can damage concrete
• Belled bases may be difficult to construct in sandy soils or clays containing silt or
sand lenses
• Concrete should be placed as soon as possible after drilling

6.6.1 Precautions in construction and inspection of bored piles


Bored piles rely on construction methods being able to produce a defect free pile, and some
of the precautions that need to be applied when constructing bored piles are summarised
below:

1. The pile should be supported by casing through soft or loose soils to prevent shaft
collapse.
2. A casing should be provided to seal off water-bearing soil layers.
3. Where drilling fluid is used, strict control of fluid density should be maintained and
cleaning and de-sanding carried out prior to re-use.
Deep foundations  201

4. Compare soil or rock cuttings removed from pile with descriptions from site
investigation.
5. Shear strength tests on soil from bottom of selected piles can be performed to check
against design assumptions.
6. Plumb deep boreholes to bottom immediately before concreting. Compare plumbed
depth to depth measured at end of drilling.
7. Proper measures should be taken for base cleaning, with video or visual inspection
where possible.
8. Safety procedures must be followed strictly.
9. Time interval between completion of boring and placing of concrete should be as short
as possible, and no longer than 6 h.

6.6.2 Continuous flight auger piles (or grout injected piles)

1. The flight auger usually has a central hollow stem closed by a plug at the bottom (see
Figure 6.4).
2. The borehole walls are supported at all times by soil rising within the flights.
3. On reaching the required depth, grout is injected down the hollow stem. This pushes
out the bottom plug, and the pile shaft is concreted by raising the auger (with or with-
out rotation).
4. A fairly fluid grout with plasticiser to improve pumpability and an expanding agent to
counter shrinkage during setting and hardening is used.
5. After removing the auger, a reinforcing cage is placed while the concrete is still fluid.
Cage lengths up to 12 m long are usually achievable.
6. Strict control of construction is necessary, especially when end bearing resistance is
required.
7. A check can be made of shaft soundness by recording the amount of concrete injected
as the auger is withdrawn.
8. An indication of soil strength can be obtained by measuring the torque on the drill
stem over the full depth of drilling.
9. Size limits: The diameter is restricted to less than 1.5 m but more commonly is less
than 1 m. The length of the pile is governed by the length of the auger, and is com-
monly less than 35 m.

6.7 DESIGN CONSIDERATIONS

The designs of piles is discussed in following sections, but may involve some of the following
considerations:

1. Selection of pile type and method of installation.


2. Estimating the size and number of piles so as to obtain an adequate factor of safety.
3. Computing the settlement and differential settlement of piles or groups of piles to
check if they are within specified tolerances.
4. Effects of lateral loading including lateral load capacity and lateral deflection may
need to be calculated.
5. The effects of soil movements may need to be taken into account in design. Movements
may be due to nearby excavation, slope movement, or soil consolidation (causing
downdrag).
202  Geomechanics in soil, rock, and environmental engineering

6. Evaluation of pile performance through load testing is often carried out, and designs
may need to be refined once this is done.

6.8 SELECTION OF PILE TYPE

As mentioned previously (Section 6.2), there are many different types of piles that the
designer can choose from. Some of the factors which can influence the type of pile to use
are listed below:

1. Location and type of structure


2. Loads being applied by the structure
3. Availability of piling equipment, and access of equipment to the site
4. Ground conditions
5. Durability requirements
6. Effects of installation on adjacent piles and structures
7. Relative costs

6.9 DESIGNS OF PILES

The ultimate load that a pile can carry can be estimated from a dynamic approach, that is,
pile driving formulae or the wave equation. However, the load capacity can also be com-
puted from knowledge of the soil strength. These methods are termed ‘static methods’, and
there are several different ways of carrying out the computations ranging from empirical
approaches to semi-analytical methods.
Soil properties can be found from laboratory tests or from field tests and correlations to
field tests. The design methods also depend on the geotechnical conditions, as there are dif-
ferent approaches taken for clays, sands, and piles driven to rock.

6.10 SINGLE PILES

The ultimate load Pu that can be carried by a pile is the combination of the base load and the
shaft friction minus the weight of the pile itself, that is,

Pu = Pus + Pub − W (6.1)

where
Pus is the ultimate shaft load
Pub is the ultimate base load
W is the weight of the pile

The shaft resistance can be obtained by integrating the available shear strength along
the shaft, while the base resistance can be found from the bearing capacity for a deep
foundation.
As for shallow foundations, there are two cases that can be considered, the drained and
undrained cases. For clays, both undrained (short term) and drained (long term) analyses
may be applicable, while for free draining materials like sands and gravels, only drained
conditions apply.
Deep foundations  203

The general formula for computing the bearing capacity of a single pile can therefore be
written as


Pu = C(ca + σ v Ks tan φ a )dz + Ab (cNc + σ vb N q + 0.5γ dN γ ) − W (6.2)
0

where
C is the shaft circumference
c a is the adhesion between the shaft and soil
ϕa is the angle of friction between the shaft and soil
σv is the vertical overburden stress at depth z
σvb is the vertical overburden stress at pile base level
K s is the coefficient of lateral pressure
Ab is the area of the base of the pile
d is the diameter of the pile
γ is the unit weight of the soil
Nc , Nq, and N γ are the bearing capacity factors for a deep foundation
W is the weight of the pile

6.10.1 Piles in clay
For piles in clay under undrained conditions, the angles of friction ϕa , ϕu can be taken as
zero, therefore the general formula (Equation 6.2) becomes


Pu = Cca dz + Ab (su Nc + σ vb ) − W (6.3)
0

Because the weight of the pile and the term for the overburden pressure are approximately
equal, that is,

W = γ concLAb ≈ σ vb Ab = γLAb

these terms cancel.


The final ultimate load for a pile in clay is therefore


Pu = Cca dz + Ab su Nc (6.4)
0

For piles that are usually considered to have an L/d ratio high enough to be classified as a
deep foundation, Nc = 9 and su is the undrained shear strength of the clay below base level.
The adhesion c a between the soil and the pile shaft, is related to the undrained shear
strength of the clay su(z) at a depth z along the pile shaft. The relationship is given by

ca = αsu (6.5)

The method is sometimes called the ‘alpha’ method because of this. Relationships between
α and the undrained shear strength s u are given in charts such as the one shown in
204  Geomechanics in soil, rock, and environmental engineering

1.2
Tomlinson (1957) Shafts in uplift
Data group 1
(concrete piles) Data group 2
1.0 Data group 3
Shafts in uplift
Data group 1
Data group 2
0.8 Data group 3

Adhesion factor, α
65U and 41C
load tests
0.6

α = 0.21 + 0.26 pa/su (≤1)


0.4

0.2

0
0 50 100 150 200 250 300
Undrained shear strength, su (kN/m2)

Figure 6.9 Variation of α = ca /su against shear strength of clay su.

Figure 6.9. It should be noted that there is quite a large scatter of the data that this chart
is based upon.
For a pile with a constant diameter, the integral in Equation 6.4 is the circumference of
the pile times the area under a plot of the adhesion versus depth.
In the long term, the effective stresses and drained soil parameters should be used to
compute the shaft and base loads of a pile. In this case, the skin friction f is given by
f = βσ ′v where β = Ks tan φ′a and is therefore often termed the ‘beta’ method. In the general
equation (Equation 6.2) therefore, the shaft load may be computed by using ca′ = 0 and
Ks tan φ′a .
For normally consolidated clays, it has been suggested that

Ks = (1 − sin φ′R) (6.6)


while for overconsolidated clays

Ks = (1 − sin φ′R) OCR (6.7)


where
φ′R is the remoulded angle of friction of the soil and
OCR is the overconsolidation ratio of the clay

6.10.2 Piles in sand
Piles constructed in sand can be designed by using the effective stresses (a drained analysis)
with the general equation (Equation 6.2). The equation then becomes (if φ′a = δ )


Pu = Cσ ′v Ks tan δ dz + Ab (σ ′vb N q + 0.5γ dN γ ) − W (6.8)
0
Deep foundations  205

If it is assumed that the pile self-weight and the term in N γ cancel (as they are of the same
magnitude), then


Pu = Cσ ′v Ks tan δ dz + Abσ ′vb N q (6.9)
0

For a pile with a uniform diameter and a constant angle of friction δ between the pile and
soil, the formula becomes


Pu = CKs tanδ σ ′v dz + Abσ ′vb N q (6.10)
0

The integral in this equation is just the area under a plot of the effective vertical stress dia-
gram versus depth, and thus may be easily determined geometrically.
However, it has been noticed that the skin friction acting on the sides of piles in sand
does not keep on increasing linearly with depth as Equation 6.10 may suggest. It has been
noticed from pile test data that the skin friction increases up to a certain ‘critical depth’ and
then does not increase in value a great deal beyond that depth (Vesic 1967). This is shown
in Figure 6.10 where the skin friction variation depends on the density of the sand as well as
the depth from the top of the sand layer.

Average skin resistance (lb/in2)


0 1 2 3 4 5 6
0

20

40 Dense sand (G-3)


Pile length (in)

60
Field tests
loose, moist,
sand (G-4)
80
Medium dense
sand (G-2)
100

Loose sand (G-1)

120

140
0.1 0.2 0.3 0.4
Average skin resistance (tons/ft2)

Figure 6.10 Variation of pile skin friction in sands of different densities.


206  Geomechanics in soil, rock, and environmental engineering

W.T.
zc
σ′
vc

Figure 6.11 Effective stress variation used to compute skin friction for piles in sand.

Therefore to design piles in sand, the critical depth is determined (this depends on the
density of the sand) and then it is assumed that the effective stress is constant below this
depth. This pseudo-effective stress variation is then used to compute the skin friction. By
doing so, the skin friction on the pile remains constant past the critical depth. This is shown
in Figure 6.11 where it may be noted that the location of the water table is taken into account
in computing the vertical effective stress variation.
We also need values of the coefficient of lateral earth pressure K s to compute the skin fric-
tion. Values of K s tan δ vary for different sand densities (where δ = φ′a is the angle of friction
between the pile shaft and the sand), and therefore values have been tabulated for use with
different sand densities as shown in Table 6.1.
The base bearing capacity of piles in sand also requires a knowledge of the bearing capac-
ity factors, and some values of Nq have been suggested by Vesic (1967) as well. The value of
σ ′vb in this case is found from the truncated value of σ ′vc in Figure 6.11.

6.10.3 Lambda method
A method mainly used for the design of offshore pipe piling is called the ‘lambda’ method
reported by Vijayvergia and Focht (1972). The shaft friction for the pile is calculated from

Pus = λ(σ ′m + 2sum )As (6.11)

where
σ ′m is the mean effective vertical stress along the pile shaft
sum is the mean undrained shear strength along the pile shaft

The value of lambda depends on the pile penetration and is shown plotted in Figure 6.12. The
method applies to fairly uniform clay deposits and does not apply if sand layers are present.

Table 6.1  Parameters for calculating pile capacity in sand


F = Ks tan δ Nq
Bored or Bored or
Condition of soil zc/d Driven cast in situ Driven cast in situ
Loose (R.D. = 0.2–0.4) 6 0.8 0.3 60 25
Med. Dense (R.D. = 0.4–0.75) 8 1.0 0.5 100 60
Dense (R.D. = 0.75–0.9) 15 1.5 0.8 180 100
Deep foundations  207

λ
0 0.1 0.2 0.3 0.4 0.5
0

25

Psu
50 λ=
(σ′m + 2cm)As

75
Pile penetration (ft)

Location Symbol Source


100 Detroit Housel
Morganza Mansur
Cleveland Peck
125 Drayton Peck
North Sea Fox
Lemoore Woodward
150 Stanmore Tomlinson
New Orleans Blessey
Venice McClelland
175 Alliance McClelland
Donaldsonville Darragh
MSC Houston McClelland
San Francisco Seed
200 British Columbia McCammon
Burnside Peck
225

Figure 6.12 Frictional coefficient λ versus pile penetration.

6.11 METHODS BASED ON FIELD TESTS

In many cases, the parameters such as shear strength and angle of shearing resistance between
the pile shaft and the soil are not determined through laboratory tests, but field tests are used
and base bearing and skin friction are found directly by correlation with test results.
Field tests of various types have been discussed in Chapter 4.

6.11.1 Correlations with standard penetration test (SPT) data


Empirical correlations with SPT data take the form:

fs = An + Bn N (kPa) (6.12)

where
An and Bn are empirical numbers, and depend on the units used
N is the SPT blow count

The base load (pressure) is computed from

fb = Cn Nb (MPa) (6.13)

where
Cn is an empirical factor
Nb is the average SPT below the pile base (typically 1–3 pile base diameters)
208  Geomechanics in soil, rock, and environmental engineering

Table 6.2  Factor Cn for base resistance


Soil type Displacement piles Non-displacement piles
Sand 0.325 0.165
Sandy silt 0.205 0.115
Clayey silt 0.165 0.100
Clay 0.100 0.080

The most widely used correlations are those due to Meyerhof (1956) for piles driven in
sand, where

An = 0
Bn  = 2 for displacement piles
   = 1 for small displacement piles
C n = 0.3

Limiting values of fs of 100 kPa are recommended for displacement piles and 50 kPa for
small displacement piles.
Décourt (1995) has developed more recent and extensive correlations between fs and N
which take into account both the pile type and method of installation. For displacement
piles An = 10 and Bn = 2.8, while for non-displacement piles An = 5 − 6 and Bn = 1.4 − 1.7.
For the base, values of C n are given in Table 6.2.

6.11.2 Correlations with cone data


Empirical correlations with cone data often take the form:

fs = Aqqc (6.14)

where
Aq is an empirical number
qc is the cone penetration resistance at a particular depth along the shaft

The base pressure is given by

fb = Cqqcb (6.15)

where
Cq is an empirical number
qcb is the average cone penetration resistance below the pile tip

Figures 6.13 and 6.14 give values of shaft friction fs correlated to cone values for the dif-
ferent soil types given in Table 6.3.
A useful adaptation of the method of Bustamante and Gianeselli (1982) has been sum-
marised by Frank and Magnan (1995). The ultimate shaft friction and base capacity are given by

qc
fs = ≤ fsl
b (6.16)
fb = kcqc
Deep foundations  209

Curve Application pile types Note:


no. (see Table 6.3) Lower limit applies for unreliable construction
300
1L Lower limit for IB control; Upper limit applies for very careful
1U Upper limit for IB construction control
2 IIB
250 3L Lower limit for IA and IIA
3U Upper limit for IA and IIA
4 IIIA
Shaft resistance fs (kN/m2)
5 IIIB
200

150
5 3U
4 3L
100 3U
3
2 1L
50
1

0
0 5 10 15 20 25 30
Cone resistance qc (MN/m2)

Figure 6.13 Shaft resistance values for sand.

Tables 6.4 and 6.5 give values of the factors b and the limiting skin friction fsl as well as kc.

6.11.3 Seismic data
For design of piles, the soil modulus E can be found from the small-strain modulus E s-strain as
found from cross-hole seismic tests. The suggested value of modulus is given by

E = 0.3Es-strain (6.17)

160
Curve Application pile types
Note:
140 no. (see Table 6.3)
Lower limit applies for unreliable construction
1 IIB, lower limit for IA, IB, and IIA control; Upper limit applies for very careful
2 Upper limit for IB construction control
120 3 IIIA, upper limit for IA, IIA, and IIIB
Shaft resistance fs (kPa)

4 IIIB
100
4
80 3
2
60

40 1

20

0
0 2 4 6 8 10 12 14 16
Cone resistance qc (MPa)

Figure 6.14 Shaft resistance for piles in clay.


210  Geomechanics in soil, rock, and environmental engineering

Table 6.3  Classification of pile types


Pile category Description
1A Plain bored piles, mud bored piles, hollow auger bored piles, cast screwed piles
Type I micropiles, piers, barrettes
1B Cased bored piles
Driven cast piles
IIA Driven precast piles
Pre-stressed tubular piles
Jacked concrete piles
IIB Driven steel piles
Jacked steel piles
IIIA Driven grouted piles
Driven rammed piles
IIIB High pressure grouted piles (d > 0.25 m)
Type II micropiles
Source: Adapted from Bustamante, M. and Gianeselli, L. 1982. Proceedings of ESOPT II, Amsterdam,Vol. 2, pp. 492–500.

6.12 PILE GROUPS

Piles are usually used in groups rather than alone, and when loaded the piles will often act
as a single block as the soil trapped between the piles will act in unison with the piles in the
group. If the piles are closely spaced, this group action is more likely than if the piles are
spaced more widely apart. Therefore, in calculating the bearing capacity of a group of piles,

Table 6.4  Ultimate shaft friction correlation factors for a CPT test
Clay and silt Sand and gravel Chalk
Pile type Soft Stiff Hard Loose Medium Dense Soft Weathered
Drilled
b 75a – 200 200 200 125  80
fsl, kPa 15 40 80a 40 80a – 120 40 120
Drilled; removed casing
b – 100 100b – 100b 250 250 300 125 100
fsl, kPa 15 40 60b 40 80b – 40 120 40  80
Steel; driven
closed-ended
b – 120 150 300 300 300
fsl, kPa 15 40 80 120 c

Driven; concrete
b 75 150 150 150
fsl, kPa 15 80 80 120 c

Source: Adapted from MELT 1993. Règles techniques de conception et de calcule des fondations de ouvrages de genie civil.
CCTG, Fascicule No. 62, Titre V, Min. de L’Équipement du Lodgement et de Transport, Paris.
a Trimmed and grooved at the end of drilling.
b Dry excavation, no rotation of casing.
c In chalk, f can be very low for some types of piles; a specific study is needed.
Deep foundations  211

Table 6.5  B
 ase capacity factors for a CPT
kc
Soil type qc (MPa) Non-displacement pile Displacement pile
Clay silt A  Soft <3
B  Stiff 3–6 0.40 0.55
C  Hard(clay) >6
Sand gravel A  Loose <5
B  Medium 8–15 0.15 0.50
C  Dense >20
Chalk A  Soft <5 0.20 0.30
B  Weathered >5 0.30 0.45
Source: Adapted from MELT 1993. Règles techniques de conception et de calcule des fondations de ouvrages de genie civil.
CCTG, Fascicule No. 62, Titre V, Min. de L’Équipement du Lodgement et de Transport, Paris.

it is usual to compute the capacity of the group acting as a block, as well as the sum of the
individual capacities of the piles, and the lesser of the two taken as the group capacity.

6.12.1 Piles in clay
For piles in clay, the group is firstly analysed as consisting of a number of individual piles
and the bearing capacity estimated from

Pgi = nPu (6.18)


where
n is the number of piles in the group
Pu is the ultimate load of a single pile

The group behaviour is then calculated for the piles acting as a single block using
L


PuB = 2(B + W ) ca dz + BWNc sub (6.19)
0

where
B is the breadth of the group in plan
W is the length of the group in plan
L is the length of the piles
Nc is the bearing capacity factor for a shallow foundation with an L/B ratio equal to
that of the group
sub is the undrained shear strength below the base of the group

Note that if the pile group has a non-rectangular shape in plan, that the plan perimeter of
the group is used in Equation 6.19 instead of 2(B + W) and the base area instead of BW.
The adhesion c a acting on the side of the group can be taken as the undrained shear
strength of the soil su if most of the shearing is soil to soil. A refined calculation can be made
if the adhesion between the piles and soil around the perimeter of the group is included in
the calculation for that part of the perimeter where piles exist, and the undrained shear
strength used where soil exists.
212  Geomechanics in soil, rock, and environmental engineering

W/B = 1
8
2

3
7

N*c

4
0 1 2 3 4 5
L/B

Figure 6.15 Bearing capacity factors for use with pile groups in clay.

Note that the integral in Equation 6.19 is just the area under a plot of the adhesion versus
depth. The integral can be evaluated simply by estimating the area of the plot and this can be
done by using simple trapezoidal or rectangular areas to approximate the area under the plot.
Values of Nc need to be evaluated and a plot of the type shown in Figure 6.15 can be used
for this purpose (where the group is W by B in plan and L deep).

6.12.2 Piles in sand
Where pile groups are constructed in sand, the capacity of the group can be greater than the
individual capacities of the piles. This can be due to the sand densifying as driving proceeds.
Because of this, it is always desirable to drive the piles at the centre of the group first and
work outwards, driving the perimeter piles last.
In general, the capacity of the group can be found from the sum of the individual pile
capacities, as the group capacity may be 1.3–2 times the group capacity calculated from the
individual pile capacities.

EXAMPLE 6.1
A storage tank, 25 m in diameter and having a total weight of 100 MN, is to be sup-
ported on 400 equally spaced hollow steel pipe piles driven 30 m into a deep bed of clay
with a saturated unit weight of 18 kN/m3. The piles are 0.4 m in diameter and weigh
0.77 kN/m run. Estimate the bearing capacity of the foundation and the factor of safety
that it has against a bearing failure.

Solution
In this case, the weight of the piles is taken into account as the piles are hollow and the
weight of the pile may not compensate for the surcharge term. The ultimate bearing
capacity of a single pile is (from Equation 6.3)

30


Pu = C ca dz + Ab (su Nc + σ vb N q ) − Wpile (6.20)
0
Deep foundations  213

For a pile in clay, the bearing capacity factors are calculated for ϕu = 0, so that Nc = 9
and Nq = 1.
The adhesion on the shaft of the pile is found from Figure 6.9, where for su = 30 kPa the
alpha factor can be seen to be 1 and so c a = αsu = 30 kPa.
Hence, substituting these values gives

Pu = π0.4 × 30 × 30 + π0.22 (30 × 9 + 18 × 30) − 0.77 × 30


= 1130.9 + 0.125 × 810 − 23.1
= 1209.1 kN

For the group of piles, the bearing capacity is firstly calculated for the 400 individual
piles acting independently, that is,

Pug = 400 × 1209.1 = 483, 640 kN


The capacity is then calculated for the piles acting as a block that is 25 m in diameter
and 30 m deep. For such a foundation, the L/B ratio (length to diameter) is 30/25 = 1.2
and so from Figure 6.15 the bearing capacity factor (for W/B = 1) is about 7.9. Generally,
the skin friction acting around the perimeter of the group is taken as  su since most of
the shearing is between soil and soil rather than soil and pile. The weight of the block is
approximately the weight of soil in the block if the weight of the piles is considered small
in comparison, therefore

Publock = π25 × 30 × 30 + π12.52 (30 × 7.9 + 18 × 30) − π12.52 × 30 × 18


= 70685.8 + 381408.9 − 265071.9
= 187, 022.8 kN

Hence, the lower of the two values is the block value so the bearing capacity of the
group is taken as about 187 MN, and hence the factor of safety is 187/100 = 1.87.

6.13 PILES IN ROCK

Piles that are socketed into rock can also be designed by calculating shaft friction and end
bearing. For shaft friction, the roughness of the shaft affects the values used, and often the
shaft is roughened up with a grooving tool to increase the values of the shaft friction. The
design of rock-socketed piles and piers is discussed more fully Chapter 12, “Basic Rock
Mechanics” rock (Section 12.8.2).

6.14 SETTLEMENT OF SINGLE PILES

Often it is the displacements or differential displacements that are of importance in the


design of piled foundations rather than bearing capacity. Piles may be used singly or in
groups, and if used in groups, they tend to interact with one another, so that if one pile is
loaded, then it causes the surrounding piles to settle. Therefore, in order to compute the
settlement of groups of piles, it is necessary to be able to estimate the ‘interaction’ between
one pile and another.
214  Geomechanics in soil, rock, and environmental engineering

This section deals with vertical loading only on piles and pile groups, and horizontal load-
ing is left to a later section.
For single piles, there are several approaches that can be taken ranging from empirical
methods to closed form solutions and charts. Generally, the solutions are based on the
assumption that the pile deflection is linear which at working loads is a reasonable assump-
tion. However, at higher loads the pile deflection will become non-linear, and methods such
as hyperbolic functions, load transfer (t − z) methods, or finite element methods may be
required to predict the non-linear behaviour.

6.14.1 Closed form solutions


Often solutions based on the theory of elasticity can be expressed as an analytic expression
such as (Randolph and Wroth 1978)

4P(1 + νs )  1 + (8η/(πλ(1 − νs )ξ)) (tanh(µL)/µL)(L /d) 


s = (6.21)
EsLd  (4 /(1 − νs ))(η/ξ) + (4πρ/ζ)(tanh(µL)/µL)(L /d) 

where
P is the load on the pile
E sL is the modulus of the soil at pile base level
νs is Poisson’s ratio of the soil
η is the ratio of the pile tip diameter to shaft diameter = db /d
ξ is given by E sL /Eb where Eb is the modulus of any hard layer below the tip of the pile
ρ is the ratio E/EsL and E is the average value of modulus along the shaft

and

 2L 
ζ = ln  {0.25 + [2.5ρ(1 − νs ) − 0.25]ξ}
 d 
Ep
λ = 2(1 + νs ) (6.22)
EsL
2  L
µL = 2  d 
ζ

6.14.2 Settlement of single piles


Charts have been developed by Poulos and Davis (1980) for single piles in a deep uniform
elastic soil, or an elastic soil with a rigid underlying layer. The solutions are based on bound-
ary element methods, and the problem considered is shown in Figure 6.16. In the figure, the
following quantities are shown:

d is the shaft diameter


L is the pile length
h is the depth to the hard stratum
P is the load carried by the pile
db is the base diameter
Ep is the modulus of the pile
E s is the modulus of the soil
Deep foundations  215

P P
(a) (b)

d Soil, Young’s
modulus Es
Pile modulus L Poisson’s ratio L
Ep νs d

h
db db

Soil, Young’s modulus Es


Poisson’s ratio νs Stiffer stratum, Young’s
modulus Eb
Rough rigid base
Ep
Pile stiffness factor K = RA
Es

Figure 6.16 Definition of the problem: (a) floating or friction pile; (b) end bearing pile.

Eb is the modulus of the firm stratum


R A is the area ratio of the pile = area of pile cross-section/total area of pile

The settlement for a single pile is computed from

1. Floating pile

P
S = I1RK RhRν (6.23)
Esd
where I1 is an influence factor that is presented in Figure 6.17, and R K , R h, and Rν are
correction factors for the effects of pile stiffness, layer depth, and Poisson’s ratio and
are shown in Figures 6.18, 6.19, and 6.21.
2. End bearing piles

P
S = I1RK RbRν (6.24)
Esd
For end bearing piles, the correction factor Rb is used in Equation 6.24 to correct for
the effect of a stiffer underlying layer, and this factor is presented in Figure 6.20.

6.14.3 Soil modulus increasing with depth


Where the soil modulus is increasing with depth, the head deflection of a pile can be esti-
mated by using the plot shown in Figure 6.22. Several sets of curves are shown, for different
rates of increase of modulus with depth Nv. Settlements can be calculated from the influence
factor I ρ from

P
s = Iρ (6.25)
N v dL
216  Geomechanics in soil, rock, and environmental engineering

1.0

0.8

0.6

0.4

0.2

Values of db/d
I1

1
0.10 3 2

0.08

0.06

0.04
For L/d = 100
I1 = 0.0254
For 3 ≥ db/d ≥ 1

0.02
0 10 20 30 40 50
L/d

Figure 6.17 Pile settlement influence factor I1.

6.15 INTERACTION OF PILES

As mentioned in the chapter introduction, piles will interact with each other when used in
groups, as if one pile is loaded, it will cause the surrounding piles to settle.
Therefore, it is useful to be able to calculate the interactions of piles in a group, and this
can be carried out conveniently by looking up plots giving interaction factors for either float-
ing piles or piles driven to a firm base. Some examples of interaction factor plots for two
piles spaced at a distance s (see Figure 6.23) are given in Figures 6.24, 6.25, and 6.26 where
the interaction factor αF is defined as

Increase in settlement of pile 1 due to pile 2


αF = (6.26)
Settlementt of pile 1 under its own load

Interaction factors can be corrected using the factors in Figure 6.27 when the soil layer
has a finite depth. Further plots of interaction factors can be found in the book by Poulos
and Davis (1980).
Deep foundations  217

Values of L
d
100
50
25

2 10
RK

1
1
10 102 103 104
K

Figure 6.18 Correction factor for the effect of relative pile stiffness.

1.0

Values of L
d

0.8 50
25
10
5
0.6
2
Rh

0.4 1

0.2

0
1 2
h/L 0.5 0
L/h

Figure 6.19 Correction factor for the effect of layer depth.


218  Geomechanics in soil, rock, and environmental engineering

(a) 1.0 (b) 1.0


100 100
500
0.8 1000 0.8 500

1000
0.6 0.6
5000
Rb

Rb
0.4 Values of K 0.4 Values of K 5000
≥20,000 ≥20,000
0.2 0.2
L L
= 75 = 50
d d
0 0
1 10 100 1000 1 10 100 1000
Eb/Es Eb/Es

(c) 1.0 (d) 1.0


Values of K
100
0.8 0.8

Values of K
0.6 0.6
500
100
Rb

Rb

0.4 1000 0.4

0.2 0.2 500


L L
= 25 ≥2000 = 10 1000
d d
≥10,000
0 0
1 10 100 1000 1 10 100 1000
Eb/Es Eb/Es

(e) 1.0

0.8
Values of K

0.6
5000 100
Rb

20,000 500
0.4 1000
5000
L
0.2 =5
d
20,000
0
1 10 100 1000
Eb/Es

Figure 6.20 Correction factors for the effect of a stiffer underlying layer. (a) L/d = 75; (b) L/d = 50; (c) L/d =
25; (d) L/d = 10; (e) L/d = 5.
Deep foundations  219

1.00

0.95

K = 100
0.90
500

0.85 1000
2000

0.80

0.75
0 0.1 0.2 0.3 0.4 0.5
νs

Figure 6.21 Correction factor for the effect of Poisson’s ratio of the soil.

0.4

K = Ep/Nv d = 15,000
K = Ep/Nv d = 4000
P

0.3 z
L Es = Nvz
d νs = 0.3

Values of Eb/Nv L
0.2

0.1 5

10

100 P
S= Iρ
NvdL

0
5 7 10 15 20 30 40 50 70 100
L/d

Figure 6.22 Settlement of a single pile in soil where the modulus increases with depth.
220  Geomechanics in soil, rock, and environmental engineering

s
P P

p p

d d

Pile 1 Pile 2

Figure 6.23 Interaction between two piles.

6.15.1 Use of interaction factors for pile groups


The settlement of one pile i due to another loaded pile j is given by (see Figure 6.28)

sij = ρ1Pjα ij (6.27)



where ρ1 is the settlement of a single pile under a unit load, and Pj is the load on pile j.
Therefore, for a group of n piles where all the piles are the same, the settlement of a pile
k due to all the other loaded piles is given by
n

sk = ρ1 ∑P α
j =1
j kj (6.28)

1.0

0.8 L = 10
d
Values of K νs = 0.5
0.6
1000
αF


0.4 500
100
10
0.2

0
0 1 2 3 4 5
s/d 0.2 0.15 0.1 0.05 0
d/s

Figure 6.24 Interaction factors for floating piles L/d = 10.


Deep foundations  221

1.0

0.8 L = 100
Values of K
d
∞ νs = 0.5
0.6
100 10 5000

αF

500 5000
0.4
1000
1000
500
0.2
100 10
0
0 1 2 3 4 5
s/d 0.2 0.15 0.1 0.05 0
d/s

Figure 6.25 Interaction factors for floating piles L/d = 100.

For all of the piles in the group, the settlements can be written in matrix form where A is
the matrix of the alpha factors:

{s} = ρ1[ A]{P} (6.29)


This is suitable for piles carrying different loads applied to the head of each pile. If the
pile cap is rigid, the load carried by each pile is not known, but for symmetric vertical load-
ing where the cap does not rotate, it is known that (1) the settlements of all piles are equal
and (2) the sum of the loads on the piles is equal to the total load on the group P Tot. The set

1.0

0.8 L = 100
d
Values of K νs = 0.5
0.6 10
50
αF

100
200
0.4
500
1000

0.2
5000


0
0 1 2 3 4 5
s/d 0.2 0.15 0.1 0.05 0
d/s

Figure 6.26 Interaction factors for an end bearing pile L/d = 100.


222  Geomechanics in soil, rock, and environmental engineering

1.0

Values of h/L
5

0.8 2.5

0.6 1.5
Nh

0.4
1.2

0.2

0
1 5
s/d 0.2 0
d/s

Figure 6.27 Correction factors Nh to interaction factors for the effect of the finite layer depth.

of equations in Equation 6.30 can be set up and solved for all of the pile loads P and the
unknown deflection of the group Δ.

ρ1[ A] −a P   0 
 − aT = (6.30)
 0   ∆   −PTot 

The vector aT = (1, 1, 1, 1,… , 1). This of course is assuming that there is only a small reaction
between the pile cap and the soil, and that the piles carry most of the load. Piled raft analysis
is examined more fully in Section 6.37.

6.15.2 Simplified method for pile groups


For rapid estimation of pile group behaviour without recourse to a computer, the following
simple formula due to Randolph (see Fleming et al. 1992) can be used:

SG = RS Sav
(6.31)
Rs ≈ nω
where SG is the settlement of the group, S av is the settlement of a single pile at the average
load of piles in the group, n is the number of piles in the group, and ω is an exponent depend-
ing on the pile spacing, pile proportions, relative pile stiffness, and variation of soil modulus
with depth.
Poulos has suggested approximate values of ω as 0.5 for piles in clay, and 0.33 for piles
in sand.
Deep foundations  223

EXAMPLE 6.2
A reinforced concrete pile, 16 m long and 0.3 m in diameter is driven through a clay layer
to a dense gravel stratum. Calculate the final settlement of the pile under a working load
of 800 kN.
What would be the settlement if a 12 m long floating pile was used instead?
The elastic modulus of the soil, pile, and gravel may be taken as

Soil E s = 7 MPa
Pile Ep = 15 × 103 MPa
Gravel Eg = 140 MPa

Solution
The length to diameter ratio of the pile is L/d = 16/0.3 = 53.3, and the ratio of the base
modulus to the soil modulus is Eb /E s = 140/7 = 20. Because the pile is solid, its area ratio
R A is 1.0, hence the pile stiffness ratio is given by

 Ep  15, 000
Kp =   RA = × 1 = 2143
 Es  7

For an end bearing pile, settlement is given by Equation 6.24

P
S = I1RK RbRν
Es d

where I1 = 0.043 (Figure 6.17), R K = 1.19 (Figure 6.18), Rb = 0.7 (Figure 6.20), and Rν = 1
(Figure 6.21) if it is assumed that the undrained Poisson’s ratio of the soil νu is 0.5.
Hence,

800
S = × 0.043 × 1.19 × 0.7 × 1.0 = 0.0136 m
0.3 × 7000

or 13.6 mm.
If the pile is a floating pile, then Equation 6.23 is applicable:

P
S = I1RK RhRν
Es d

Figure 6.19 gives the value of R h as 0.62 for a value of h/L = 16/12 = 1.33. The length
to diameter ratio of the pile is now L/d = 12/0.3 = 40 and so I1 = 0.052 and Rk = 1.15.
Hence, the settlement is given by

800
S = × 0.052 × 1.15 × 0.62 × 1.0 = 0.0141 m or 14.1 mm
0.3 × 7000

As this solution is based on the assumption that the layer beneath the base is rigid, the
settlement may be slightly more than 14.1 mm.

6.16 ASSESSMENT OF PARAMETERS

Soil modulus is the key parameter that needs to be estimated when predicting pile displace-
ment. The soil modulus (or shear modulus) is a very variable quantity and depends on a number
of factors including soil type, method of pile installation, stress level imposed by the pile or pile
group, and whether long or short term analysis is required (i.e. drained or undrained analysis).
224  Geomechanics in soil, rock, and environmental engineering

The best way to assess the soil modulus is to carry out a pile load test on a prototype pile
and to back figure the modulus of the soil. As this is not possible in the preliminary stages
of design, it is more common to estimate the soil properties through field tests such as the
SPT or CPT (cone penetration test).
Four different values of modulus can be distinguished for pile analysis:

1. The value for the soil next to the shaft E s


2. The value immediately below the pile Eb
3. The initial tangent value of the soil between the piles Ei. This will affect the interaction
between the piles
4. The value of E sl for the soil well below the tip; the influence of soft layers at depth will
influence pile group behaviour and the group effects will reach deeper as the pile group
gets bigger

Table 6.6 summarises some values of E s, E si, and E sl that can be correlated to SPT data
and to cone data. There is not a great deal of information for values of modulus that should
be used below the base of a pile. Therefore, for clays the value of E sb can be taken as equal
to Es and for sands, the value can be taken as 3–5 times E s.
It should also be noted that the values in Table 6.6 are secant values at typical load levels
of one-third to one-half of the ultimate load.
Correlations can also be made to typical back-figured moduli from pile load tests such as
the relationships shown in Figure 6.29.

Table 6.6  Summary of some correlations for drained Young’s modulus for pile settlement analysis
Small-strain modulus, Modulus well below
Near-shaft modulus, Es (MPa) Esi (MPa) pile tips, Es (MPa)
CLAYS (2.5 ± 0.5)N (Décourt et al. 1989) 14N (Hirayama 1991) (0.5 ± 0.2)N (Stroud 1974)
Driven piles (15 ± 5)su (Poulos 1989) 49.4qc0.695⋅e0−1.13 (7.5 ± 2.5)qc
(Mayne and Rix 1993)
(500 ± 5)qc (Callanan and 1500su (Hirayama 1991)
Kulhaway 1985)
Bored piles (150–400)su (Poulos and Davis (150 ± 50)su
1980)
10qc (Christoulas and Frank 1991) (0.5–0.7)M
SILICA (2.5 ± 0.5)N (Décourt et al. 1989) 16.9N0.9 (Ohsaki and 7N0.5 (Denver 1982)
SANDS Iwasaki 1973)
Driven piles (7.5 ± 2.5)qc (Poulos 1989) 53qc0.61 (Imai and Tonouchi (7 ± 4)qc (Jamiolkowski
1982) et al. 1988)
Bored piles (3 ± 0.5)qc (Poulos 1993)
Source: Adapted from Poulos, H.G. 2001. Geotechnical and Geoenvironmental Engineering Handbook, Chapter 10, Ed. Rowe,

R.K., Pile Foundations.


Notes:
1. Values of Es and Esi for sands are for single isolated pile. In a group, the values may be increased, depending on pile spacing
and initial density.
2. Below pile tip, Esb can be taken as equal to Es for clays and bored piles in sands; and 3–5 times Es for driven piles in sands.
3. Above values of Es and Esb are for use in an elastic analysis. Higher values are appropriate for non-linear analyses (e.g. the
initial tangent values for a hyperbolic model should be 1.4–1.6 times the values in this table).
4. N is the SPT value (blows per 300 mm), and should be corrected to a rod energy of 60%.
5. qc = cone penetrometer resistance in MPa; su = undrained shear strength in MPa; eo = initial void ratio; M = constrained
modulus.
Deep foundations  225

Pile i

s
( ) Pile j
d ij

Settlement of pile i due to pile j


= Sij = Pj . αij . S1

Figure 6.28 Interaction of piles in a group.

6.17 LATERAL RESISTANCE OF PILES

The lateral resistance of single piles or pile groups is of importance when piles are used as
anchors or in offshore structures. The methods used for computing the ultimate resistance
of piles under lateral load generally rely on computing the forces acting on the pile or pile
group at the point of failure.

105

Average for
driven piles
E′ (kPa)

104 Average for


bored piles

Driven piles
Bored piles
Bored piles in London clay

103

0 50 100 150 200 250 300


su (kPa)

Figure 6.29 Back-figured soil modulus E′ for piles in clay.


226  Geomechanics in soil, rock, and environmental engineering

6.17.1 Single piles
As for vertically loaded piles, the ultimate load that a single pile can carry is different for piles
in clay, or piles in sand. However, in both cases, two types of failure need to be considered.

• Failure of the soil


• Failure or yielding of the pile itself (i.e. structural failure)

6.17.2 Piles in clay
For piles in clay, solutions by Broms (1964a) may be used. He assumed that the pile would
rotate about a point and that the lateral soil pressure py in front of the pile and at the back
of the pile would eventually reach a maximum value of

py = 9su (6.32)

The pressure in front of the pile is not allowed to act over the full shaft length, but a zone
of 1.5 pile diameters at the top of the pile is assumed to apply no pressure due to disturbance
during installation (see Figure 6.30).
Results for the loading, obtained by Broms are shown in Figure 6.31, for the ‘short pile’
failure mode where the soil yields rather than the pile. Figure 6.32 shows the ultimate load
for a ‘long pile’ where the pile will yield before the soil fails. Both modes of failure should be
checked out, and the one giving the lowest lateral failure load should be selected.
In these figures, e is the eccentricity of the lateral load (i.e. its distance above the soil sur-
face) and Myield is the yield moment of the pile.

6.17.3 Piles in sand
For cohesionless soils, the ultimate pressure acting in front of the pile (where it is advancing
into the soil) is assumed by Broms (1964b) to be equal to a function of the passive pressure
of the soil (Figure 6.33). It is therefore a function of the depth below the ground surface and
will increase with depth. Behind the pile (where it is moving away from the soil), the active
pressure is assumed to be small and neglected.
For piles of rectangular or circular cross section therefore, Broms suggests that the lateral
pressure be taken as

py = 3pp (6.33)

where pp is the lateral passive earth pressure of the soil computed from

pp = Kpσ ′v (6.34)

and

1 + sin φ′
Kp =
1 − sin φ′

Again the pile can fail due to yield of the soil, or due to yield of the pile itself, thus there are
two plots shown in Figures 6.34 and 6.35 for the ‘short pile’ and ‘long pile’ modes, respectively.
Deep foundations  227

Mmax
Hu
Mmax

1.5d
L

d
9sud
Myield
Hu
Myield
1.5d

9sud Mmax
Myield
Hu Myield Myield

1.5d

9sud
Soil reaction Bending moment

Figure 6.30 Failure modes for laterally loaded piles in clay.

In Figures 6.34 and 6.35, γ is the unit weight of the soil, and D is the diameter of the pile.
For loads of different eccentricities e, the ultimate lateral resistance P can be estimated for
each of the failure modes, and the lowest one is selected.

6.18 LATERALLY LOADED PILE GROUPS

As for vertical loading on pile groups, under lateral loading, the whole group can behave
as a single block if the piles are closely spaced, and the soil is trapped between the piles.
Therefore, the group capacity under lateral loading is taken as the lesser of

1. The sum of the individual pile capacities


2. The capacity of a block containing the piles and the soil

For single pile analysis, it is still necessary to consider both ‘short pile’ and ‘long pile’
modes when calculating the group capacity.
228  Geomechanics in soil, rock, and environmental engineering

60

e/D = 0

50 P
1
e

Ultimate lateral resistance P/su D2


2

40 L 4

D 8
30

16
20

10
Restrained
Free headed
0
0 4 8 12 16 20
Embedment length L/D

Figure 6.31 Ultimate lateral resistance of ‘short piles’ in cohesive soil (e.g. clay).

However, for block failure, it is only necessary to consider ‘short pile’ failure. For this
mode, beware of direct use of Broms’ charts for piles in clay, as these charts assume that
there is a ‘dead zone’ of 1.5 pile diameters (in this case the diameter of the block) and this is
not realistic for a large block. It is suggested that a dead zone of 1.5 times single pile diam-
eter be used in this case. Also, the Broms’ theory does not account for the shear on the two
sides and base of a block moving forward, and it would only apply to a group of piles where
the depth to width ratio is high.
For groups with low depth to length ratios that will move forward rather than rotate, the
side and base shears should be estimated, and included in the calculation. The resistance

100

60 Restrained
Ultimate lateral resistance P/su D2

40 Free headed

20 e/D= 0 1
2 4
10 8
16 P
6 e
4
L
2
D
1
3 4 6 10 20 40 60 100 300 600
Yield moment Myield/su D3

Figure 6.32 Ultimate lateral resistance of ‘long piles’ in cohesive soil (e.g. clay).
Deep foundations  229

Hu

f
L
g

3γdLKp Mmax

Hu

Myield

Soil reaction Bending moment

Figure 6.33 Free-head piles in cohesionless soil.

200
Restrained
Free headed
e/L = 0
160 P
Ultimate lateral resistance P/KpD3 γ

L
120 0.2

D 0.4

80 0.8

1.5
40
3.0

0
0 4 8 12 16 20
Embedment length L/D

Figure 6.34 Ultimate lateral capacity of ‘short piles’ in cohesionless soils (e.g. sands).
230  Geomechanics in soil, rock, and environmental engineering

103
P

Ultimate lateral resistance P/KpD3 γ


e Restrained
Free headed
L
102

D
10

e/D = 0 1
24
8 16
32
1
10–1 1 10 102 103 104
Yield moment Myield/D4 γKp

Figure 6.35 Ultimate lateral capacity of ‘long piles’ in cohesionless soils (e.g. sands).

at the front of the block (for piles in clay) can be taken as 9su down the front of the group
(allowing for a 1.5 pile diameter dead zone).

6.19 DISPLACEMENT OF LATERALLY LOADED PILES

There are several approaches to computing the lateral deflection of piles. A common analysis
is the p–y analysis (Reese et al. 1974) in which the pile is represented as an elastic beam
supported by non-linear springs for which the load–stiffness relationship is known. The p–y
relationship can be different for each spring down the pile if desired.
The p–y relationships are generally found from pile load tests, and typical forms of these
relationships have been presented by Reese et al. (1974) for sands, and Matlock (1970) for clays.
Other methods include solutions based on linear elasticity, and non-linear modifications
applied to these elastic solutions.

6.19.1 Linear elastic solutions (single piles)


Solutions for the lateral deflections of single piles have been obtained by Poulos (1973) and are
presented in the form of charts. The problem basically involves a pile as shown in Figure 6.36 of
length L and carrying a moment M and lateral load H. A load at any eccentricity e can always
be replaced with an equivalent moment and lateral force at the pile head, that is, M = He.

6.19.2 Constant soil modulus with depth


Where the soil modulus is reasonably constant with depth, the ground line deflection ρ of a
free-head pile can be expressed as

H M
ρ = ⋅ IρH + ⋅ IρM (6.35)
EsL EsL2

where the factors I ρH and I ρM are presented in the charts shown in Figures 6.37 and 6.38.
Deep foundations  231

Pile diameter
L
or width = d

Figure 6.36 Laterally loaded single pile.

The rotation of the pile head θ is given by

H M
θ = IθH + IθM (6.36)
EsL2 EsL3

In this case, the influence factors I θH and I θM are given in Figures 6.38 and 6.39.
For a pile that has a fixed head (i.e. cast into a foundation or cap so that it does not rotate),
the ground line deflection is given by

H
ρ = IρF (6.37)
EsL

where I ρF is given in Figure 6.40.

50
νs = 0.5

Values of L/d
20
100
50
25

10 10
IρH

1
10–6 10–5 10–4 10–3 10–2 10–1 1 10
KR

Figure 6.37 Values of I ρH for a free-head pile in soil with a constant modulus.


232  Geomechanics in soil, rock, and environmental engineering

1000
νs = 0.5

Values of L/d

100

100 50
10 25
IρM and IθH

10

1
10–6 10–5 10–4 10–3 10–2 10–1 1 10
KR

Figure 6.38 Values of I ρM and I θH for a free-head pile in soil with a constant modulus.

Values of L/d
104 νs = 0.5
100
50
25
10

103
IθM

102

10

1
10–6 10–5 10–4 10–3 10–2 10–1 1 10
KR

Figure 6.39 Values of I θM for a free-head pile in soil with a constant modulus.


Deep foundations  233

50

νs = 0.5

20
Values of L/d
100
50
10
25
IρF

10
5

1
10–6 10–5 10–4 10–3 10–2 10–1 1 10
KR

Figure 6.40 Values of I ρF for a fixed-head pile in soil with a constant modulus.

To use these charts, it is also necessary to compute the dimensionless flexibility factor K R
where

EpI p
KR = (6.38)
EsL4

where
Ip is the second moment of inertia of the pile cross section
Ep is the modulus of the pile

6.19.3 Soil modulus linearly increasing with depth


Where the soil modulus increases linearly with depth, the ground line deflection and rota-
tion are given by

H M
ρ = I +
2 ρ′ H
Iρ′ M
N hL N hL3
(6.39)
H M
θ = I +
3 θ′ H
Iθ′M
N hL N hL4
where Nh is the rate of increase of modulus with depth. In this case, the pile flexibility factor
K N is given by

EpI p
KN = (6.40)
N hL5

The influence factors in this case can be found in Poulos and Davis (1980).
234  Geomechanics in soil, rock, and environmental engineering

6.19.4 Non-linearity
For laterally loaded piles, non-linearity is more important than for vertical loading as the
soil resistance can be small at the top of the pile, and the pile will tend to behave in a non-
linear fashion. One way of dealing with this is to use p–y analysis as mentioned before.
Another approach is to compute the deflections from the theory of elasticity and to mod-
ify these by dividing by a ‘yield factor’, that is,
ρel
ρ =

(6.41)
θel
θ =

where F ρ and F θ are the yield factors for deflection and rotation, respectively, and the
subscript ‘el’ denotes the deflection or rotation for an elasticity solution. The correction
­factors are shown in Figures 6.41 and 6.42 for free-head floating piles in a uniform soil,
and Figures 6.43 and 6.44 for free-head floating piles in soil where modulus and strength
increase linearly with depth. The correction depends on the load level H/Hu where Hu is the
ultimate lateral resistance of a rigid pile.
For fixed-head piles, the pile head fixing moment M F is of interest because there is no rota-
tion. The correction is then given by
MFE
MF = (6.42)
FM

where M FE is the elastic moment at the pile head, and FM is the yield moment factor. Values
of this factor and the displacement factor are shown in Figures 6.45 and 6.46.
Estimates of the moment at the head of the fixed-head pile can be found from Figure 6.47.

1.0
0 Values of e/L
.25
0 2.0
1.0 .25
2.0

0
.25
2.0
1.0

0.1

0
.25
1.0
2.0

KR = 10–2
KR = 10–3
KR = 10–4
KR = 10–5
0.01
0 0.2 0.4 0.6 0.8 1.0
H/Hu

Figure 6.41 Yield correction factor Fρ for a free-head floating pile in uniform soil.
Deep foundations  235

1.0 0
.25 2
1
1
2 .25
2 0
1
.25
0 2
.25 0
1
2
1

.25

0.1

Values of e/L 0 KR = 10–2


KR = 10–3
KR = 10–4
KR = 10–5

0.01
0 0.2 0.4 0.6 0.8 1.0
H/Hu

Figure 6.42 Yield correction factor Fθ for a free-head floating pile in uniform soil.

1.0 0 .25
2 2
1
0 0
.25
1
0 2
.25 1
2

Values of e/L
0.1
Fρ′

KR ≥ 10–2
KR = 10–3
KR = 10–4
KR = 10–5

0.01
0 0.2 0.4 0.6 0.8 1.0
H/Hu

Figure 6.43 Yield displacement factor Fρ′ for a free-head floating pile in soil where the modulus and strength
vary linearly with depth.
236  Geomechanics in soil, rock, and environmental engineering

1.0
2
2 0
0
.25
2 1
0 1
.25
2
1
.25
Values of e/L 0

0.1
Fθ′

KR = 10–5
KR = 10–4
KR = 10–3
KR = 10–2

0.01
0 0.2 0.4 0.6 0.8 1.0
H/Hu

Figure 6.44 Yield correction factor Fθ′ for a free-head floating pile in soil where the modulus and strength
vary linearly with depth.

6.20 DEFLECTION OF PILE GROUPS

Deflection of pile groups under lateral loading is similar to that of pile groups under vertical
loading in that the piles interact, and the deflections become bigger for a group than for an
equivalent single pile under the average load of a pile in the group.
Simple methods for estimating the group behaviour exist, one being given by the formula:

ρG = ρ1av nωL (6.43)


where
n is the number of piles in the group
ωL is an exponent given in Figure 6.48
ρ1av is the deflection of a single pile under the average load of the group

The exponent can be seen to depend on the critical length of the piles and the pile spacing.
For piles in uniform and Gibson (modulus increasing linearly with depth) soils, the critical
lengths are given by

Lc
= 4.44 KR1 / 4 (uniform soil)
L
(6.44)
Lc 1/ 5
= 3.30 KN (Gibson soil)
L
Deep foundations  237

(a) 1.0

Values of Lc/d

( )
EpIp 1/4
0.8 Lc = π√2
Es
Es = 500su 20
10
0.6 2cu 5
py
2.5
Fu

0.4
9cu
3.5d

0.2
z

0
0.4 0.6 0.8 1.0 2.0 4.0 6.0 8.0 10.0
H
sudLc

(b) 1.0

0.8
Es = 500su

0.6 2cu 20
py 10
5
FM

2.5
0.4 Values of Lc/d
9cu

( ) EpIp 1/4
3.5d

0.2 Lc = π√2
Es
z

0
0.4 0.6 0.8 1.0 2.0 4.0 6.0 8.0 10.0
H
sudLc

Figure 6.45 Non-linear correction factors for a flexible fixed-headed pile in stiff clay: (a) deflection correc-
tion factor Fu; (b) fixing moment correction factor FM .

6.20.1 Interaction methods
Interaction factors can be used to compute the effect of one loaded pile on another. In the
case of laterally loaded piles, this depends on where the unloaded pile lies relative to the
loaded pile as shown in Figure 6.49.
The interaction factors have been computed for

αρH: Interaction factor for head deflection due to horizontal load


αρM: Interaction factor for head deflection due to moment
αθH: Interaction factor for head rotation due to horizontal load = αρM
αθM: Interaction factor for head rotation due to moment
238  Geomechanics in soil, rock, and environmental engineering

(a) 1.0

0.8 Es = 500su
su = nz

py
0.6
2.5
Fu

5
0.4 9cu
10
Values of Lc/d 20
z
(
0.2
)
EpIp 1/5
Lc = 4π4
Es
0
0.004 0.006 0.008 0.1 0.2 0.4 0.6 0.8 1.0
H
ndL2c

(b) 1.0

0.8
Es = 500su
su = nz 2.5
0.6 5
py 10
FM

20
Values of Lc/d
0.4
9cu
(
Lc = 4π4
EpIp
Es )
1/5

0.2
z

0
0.04 0.06 0.08 0.1 0.2 0.4 0.6 0.8 1.0
H
ndL2c

Figure 6.46 Non-linear correction factors for a flexible fixed-head pile in soft clay: (a) deflection correction
factor; (b) fixing moment correction factor.

An example of the interaction factors for horizontal loading is given in Figure 6.50. More
plots can be found in the book by Poulos and Davis (1980).
To compute the deflection (say) of the head of pile j due to a horizontal load Hi on pile i,
the following formula would be used:

ρ j = Hiρ1α ij(ρH ) (6.45)



Deep foundations  239

–0.6
vs = 0.5
–0.5
Fixing moment at head
of fixed-head pile
–0.4

HL
Mf

–0.3
L
Values of 50
d
–0.2 10
2

–0.1

0
10–6 10–5 10–4 10–3 10–2 10–1 1 10
KR

Figure 6.47 Values of the moment at the top of a fixed-head pile in a uniform soil.

6.21 ESTIMATION OF SOIL PROPERTIES

Soil properties can be back figured from field load tests and fitting the observations to the
theory, but in the absence of such data, the value of the quantities such as the soil modulus
can be estimated from correlation with field tests.
Tables 6.7 and 6.8 show correlations between field tests and soil properties for use with
laterally loaded piles. Correlations are shown for both clays and sands.

0.6

Pile spacing s/d 2


0.5

0.4
4

0.3
ωL

0.2
Homogeneous soil
Pinned head piles
0.1 Rr = nωL

0
0 5 10 15 20 25
Lc/d

Figure 6.48 Group factor exponent ωL for laterally loaded groups in uniform soil.
240  Geomechanics in soil, rock, and environmental engineering

H, M
β

H, M

Figure 6.49 Angle between a loaded and unloaded pile.

1.0 1.0
Values of L/d νs = 0.5 νs = 0.5
Values of L/d
0.8 0.8
β=0 β=0
β = 90 β = 90
100
0.6 0.6
KR = 10–5 100 KR = 10–3
αρH

25 αρH 25

0.4 0.4
10
10
100
25
0.2 10
25 100 0.2 10

0 0
0 1 2 3 4 5 0 1 2 3 4 5
s/d .2 .15 .1 .05 0 s/d .2 .15 .1 .05 0
d/s d/s

1.0 1.0
Values of L/d νs = 0.5 Values of L/d νs = 0.5

0.8 0.8 β=0


β=0
β = 90
100 β = 90 100
25 25
0.6 0.6 KR = 10
10 KR = 10–1 10
αρH

αρH

100
10 25 100
0.4 0.4 25
10

0.2 0.2

0 0
0 1 2 3 4 5 0 1 2 3 4 5
s/d .2 .15 .1 .05 0 s/d .2 .15 .1 .05 0
d/s d/s

Figure 6.50 Interaction factors αρH for laterally loaded free-head piles.

6.22 LOAD TESTING OF PILES

Pile tests carried out in the field can be broadly grouped into two types of tests:

1. Tests performed to assess the integrity of a pile. Such tests are aimed at assessing if
there are any defects in the pile such as voids or cavities in the concrete, or if the pile
has been constructed to the designed dimensions with no ‘necking’ or thinner sections
in the shaft for example.
Deep foundations  241

2. Load tests that are designed to assess the load that can be carried by a pile or to
obtain the load–deflection behaviour of a pile. Generally, tests are performed on
single piles, although there have been instances where small groups of piles have
been tested.

For most large structures, pile testing is almost always carried out as there are many fac-
tors including construction techniques and soil variability that may change the pile behav-
iour from that predicted. In addition, pile integrity also needs to be confirmed and selected
piles are tested to check if pile installation methods are resulting in piles that are free from
defects.

Table 6.7  Empirical correlations for Young’s modulus in clays (laterally loaded piles)
Relationship Theory Reference Remarks
Esi/su
300–600 Non-linear subgrade Jamiolkowski Initial tangent modulus for
reaction and Garassino driven piles in soft clays
(1977)
180–450 Non-linear Poulos (1973) Tangent modulus from
boundary element model tests on jacked
piles
280–400 Non-linear subgrade Kishida and Tangent modulus
reaction Nakai (1977)
100–180 Linear boundary Banerjee (1978) Secant value
element
Nhi su, kPa Non-linear subgrade Sullivan et al. Tangent values of rate of
MPa m−1 reaction (1979) modulus increase Esi = Nhiz
0.8 12–25
2.7 25–50
8 50–100
27 100–200
80 200–400

Table 6.8  E mpirical correlations for Young’s modulus in sands (laterally loaded piles)
Relationship Theory Reference Remarks
Non-linear subgrade Jamiolkowski and Tangent value for
Nhi = 0.19I1D.16 , MPa m−1
reaction Garassino (1977) driven piles in
where ID = density index (%)
saturated sands
Condition Nhi, MPa ​ ​m−1 Non-linear subgrade Reese et al. (1974) Tangent value for
Loose 5.4 reaction driven piles in
Medium ​15.3 submerged sands
Dense ​34.0
Nh = 8–19 MPa m−1 Linear boundary element Banerjee (1978) Secant value Is
(av. 10.9)
Esi = 1.6 N MPa Non-linear subgrade Kishida and Nakai Tangent value
where N = SPT value reaction (1977)
242  Geomechanics in soil, rock, and environmental engineering

6.23 PILE LOAD TESTS

Once piles have been designed, using some of the techniques mentioned in Sections 6.9 to
6.20, pile load tests are often carried out to assess if a pile is behaving as predicted. The pile
load tests can be used to

1. Act as a proof load test, where the pile is taken above the working or serviceability
load to ensure that the design load is adequate. The maximum load applied to the pile
head is a multiple of the working load.
2. To allow load–deflection behaviour of the pile to be assessed, to see if it is similar
to predicted load–deflection behaviour. If the measured response is not as predicted,
the parameters used in design can be revised to give better fits to the observed load–
deflection behaviour.

Pile load tests can be static or dynamic tests. In a static test, the load is applied slowly to
the pile whereas in a dynamic test, the load is applied rapidly through a dropped weight or
an explosive charge. The load may be applied to the pile:

1. Vertically (either in compression or in tension). This is the most common type of load
test where the direction of loading is in the direction of the pile shaft
2. Laterally by applying load horizontally to the pile head
3. Using a torsional load, although this is not common

6.23.1 Static load tests


Static tests involve loading the pile against a reaction. The reaction can be supplied by

1. Dead weight (called ‘kentledge’), as shown in Figures 6.51 and 6.53


2. Reaction piles as shown in Figure 6.52
3. Ground anchors as shown in Figure 6.54
4. The pile shaft in an Osterberg cell test. In the Osterberg or ‘O-cell’ test, a hydraulic
jack is placed within the pile shaft, and the load is applied by the jack to the upper and
lower halves of the pile shaft. In some cases, two jacks can be placed within the pile
shaft (see Figure 6.55)

Kentledge

Universal beam

Universal beams

Support blocks

Dial gauges Reference beam


Hydraulic jacks Test pile

Figure 6.51 Reaction provided by kentledge.


Deep foundations  243

Reaction beam

Tension
Dial gauge or
Hydraulic jack transducer cables

Test pile
Reference beam

Anchor piles

Figure 6.52 Reaction provided by anchor piles.

Figure 6.53 Kentledge being used as the reaction system.

Figure 6.54 Anchor cables being used as a reaction for a pile load test.
244  Geomechanics in soil, rock, and environmental engineering

Tell-tales

Casing

Strain gauges

LVWDT (transducers)
Break in pile
‘O’-cells (2 × 405 mm)

LVWDT (transducers)
Break in pile
‘O’-cells (2 × 405 mm)
Pile toe tell-tales

Figure 6.55 O-cell test using two levels of hydraulic cells.

Load is applied to the pile head by one or more hydraulic jacks and is applied according
to the type of test (see Section 6.23.2).
The reaction system should not be too close to the pile that is being loaded or it will
interact with the pile. Various codes specify minimum distances for the reaction system
from the pile.

1.
Anchorages: The Australian code AS 2159-2009 specifies that no part of the anchor
shall be closer than 3 times the shaft diameter of the pile; the American code ASTM
D1143-81 (1994) specifies a minimum distance of 5 times the pile diameter or 2 m.
2.
Anchor piles: AS 2159-2009 specifies that the distance of the anchors from the loaded
pile should be the greater of 5 times the pile diameter or 2.5 m; ASTM D1143-81
(1994) specifies a minimum distance of 5 times the pile diameter or 2 m.
3.
Kentledge: AS 2159 specifies that no part of the kentledge support system shall be
closer than 2.5 times the diameter of the pile head; ASTM D1143-81 specifies the sup-
ports for the kentledge should be at least 1.5 m from the pile being loaded.

6.23.2 Types of static load tests


The main types of static load test are (1) the incremental sustained load test (ISL test) and
(2) the constant rate of penetration test (CRP test).
Deep foundations  245

1.
Incremental Sustained Load Test
In this test the load is applied in stages and maintained on the pile for a period of time
during the loading and unloading phases. The ASTM Standard Loading Procedure is
an example of this type of test.
  The load is applied to a single pile up to 200% of the anticipated working or service-
ability load, applying the load in increments of 25% of the estimated working load. Each
load increment is maintained until the rate of settlement is less than 0.25 mm/h but is held
constant for a maximum of 2 h. At the maximum load, the load is held constant for 12 h
and then removed if the settlement is less than 0.25 mm over a 1 h period. The maximum
load is removed in any case if the load has been maintained for 24 h. The load is removed
in decrements of 25% of the total test load with 1 h allowed between decrements.
  In other loading procedures, the load is applied in increments up to a maximum, then
the pile is unloaded in increments to zero load, and then reloaded before unloading again
(i.e. two load–unload cycles). The load is maintained for a set period of time at each load
step and readings of deflection taken during this time. With O-cell tests, the load is often
applied in this manner with the load being applied over two load–unload cycles.
  Codes or standards from different countries vary in how a test is conducted, there-
fore, it is important to reference the regulations in the country where the test is being
performed. This type of test is the most commonly performed test for piles supporting
tall buildings.

2.
Constant Rate of Penetration Tests
As the name suggests, the pile is jacked into the ground at a constant rate and the
load–deflection behaviour of the pile monitored. Jacking rates specified by ASTM
D1143-81 are between 0.25 and 1.25 mm/min for cohesive soil and between 0.75 and
2.5 mm/min for granular soils.

3.
Other Types of Tests
There are several other test methods where load is applied in increments and held for
a constant length of time with no rate of movement criterion, or cyclic loading tests.
These types of test are less common than (1) and (2) type tests.

6.23.3 O-cell tests
The Osterberg cell test (Osterberg 1989) is commonly used for proof testing piles as it does
not need kentledge for the reaction (which can be a safety concern as it has been known to
topple over), and it is capable of providing additional information about the pile behaviour.
A typical O-cell test set up is shown in Figure 6.55 where two levels of hydraulic jack
(O-cells) are used. In this case, the base of the lower O-cell assembly is 4 m from the toe of
the pile and the upper assembly is 21 m from the toe.
Strain gauges are provided at 12 different levels within the pile to provide information on
pile compression throughout its length. From this information, pile skin friction with depth
can be back figured. Two tell-tales are installed to monitor pile toe movement and upper pile
compression (end of tell-tale just above the upper O-cell).

6.23.4 Lateral load testing


In some instances, lateral load tests are performed if behaviour of the piles under lateral
loads that can be applied by wind or earthquake are to be assessed. The lateral load test is
described in ASTM D3966-90 (1995), and is generally performed by jacking one pile against
246  Geomechanics in soil, rock, and environmental engineering

another so that both piles are loaded horizontally. Dial gauges are mounted so that the lat-
eral deflection is measured.
For the standard type of loading test, the load is taken up to 200% of the working load
and then reduced back to zero. The load is maintained for a set amount of time at each load
increment (or decrement) and readings taken at the start and end of each time period.

6.23.5 Measurement of deflection
For piles loaded at the head, the deflection is measured by dial gauges, or electronic trans-
ducers attached to a reference beam. Movements can also be measured with precise levels
or laser beams placed at some distance from the pile head. A dial gauge used for measure-
ment of deflection with reference to a support beam is shown in Figure 6.57. Generally, four
gauges are placed at equal intervals around the pile head. They should be accurate enough
to measure pile head deflection to about 0.25 mm and have 50 mm of travel.
Measurements are affected by temperature and this can be quite pronounced in regions
where the early morning and midday temperatures vary widely. Care should be taken to
shield the measuring equipment and the pile head from the sun in such circumstances.
When a pile is loaded, it causes the ground around it to deflect as well as the pile, and as
a result, the location of the supports for the reference beam will be affected. For this rea-
son, the supports for the reaction beam should be as far as possible from the loaded pile.
Standards from different countries specify different distances that the reference beam sup-
ports should be from the pile head. ASTM D1143-81 specifies that the supports should be
at least 2.5 m from the pile.
If the soil properties or the pile head stiffness is to be back figured from the pile load test,
the relative movement of the pile head and the reference beam supports can become critical.
Erroneous values of soil modulus can be calculated if this is not taken into account.
Methods of correcting for interaction effects have been presented by Poulos and Davis
(1980) for various pile reaction systems, and one such system is shown in Figure 6.56. The
true settlement of the pile can be calculated from the measured settlement by multiplying by

2.5

2.0
P
L/d =
1.5 10
25 r
Fc

100
1.0
L

K = 1000
0.5 νs = 0.5

True Measured
0 settlement = Fc ∙ settlement
0 0.25 0.5 0.75 1.0
r/L

Figure 6.56 Correction factor for the effect of movement of reference beam supports.
Deep foundations  247

Figure 6.57 Dial gauge used to measure pile head deflection.

a correction factor Fc. In Figure 6.56, the correction factor is plotted against the r/L factor
where r is the distance of the support from the pile, and L is the pile length.
For O-cell tests, measurement of movement is performed using tell-tale rods that are
placed inside casings within the pile. The end of the tell-tale rod can be placed at any level,
and the movement recorded at that level. This allows movement of the base and top of the
pile as well as at intermediate points, therefore more data can be collected on pile behaviour.
The movements are still measured relative to a reference beam and so the supports of the
beam are subjected to movement as they are in conventional tests.
O-cell tests also involve placing strain gauges at several locations within the pile shaft as
mentioned previously (see Figure 6.55). Strains can therefore be measured over most of the
pile shaft and load in the pile shaft back figured along with shaft friction on the pile.

6.24 DYNAMIC PILE TESTING

Dynamic tests are sometimes performed instead of static tests when it is not desirable to use
large amounts of kentledge or to construct reaction systems.

6.24.1 Dynamic pile test


A hammer having sufficient energy to mobilise the pile resistance is necessary if a pile is to
be tested dynamically. In this case, the energy of the blow applied to the pile should be large
enough to mobilise the equivalent of at least 150% of the working pile load or in terms of
limit state design 150% of the design action effect.
The pile head is instrumented with accelerometers and strain gauges as shown in Figure 6.58,
and from the recorded values, a plot is made of force versus time and of velocity versus time.
The parameters for the soil and pile are adjusted using dynamic analysis software until
a good fit to the measured force and velocity plots is obtained. There are several alterna-
tive approaches available for doing this, but commonly used procedures are the ‘CAPWAP’
(Rausche et al. 1985) and the TNO procedures (Middendorp and van Weele 1986).
The rate of loading in a dynamic test is obviously much higher than the loads applied to
piles beneath a tall structure but correction can be made for the rate of loading in the analy-
sis to give estimates of the ultimate static pile load.
248  Geomechanics in soil, rock, and environmental engineering

(a)

Counterweight
releasing device
Ram

Guide tube

Measurement equipment
Cushion Accelerometers
and strain equipment

(b)

Force

t (ms)
10
2L/c Velocity

Figure 6.58 Dynamic pile load test: (a) typical equipment; (b) typical force and velocity records.

6.24.2 Statnamic testing
The Statnamic test was developed in Canada and the Netherlands (Middendorp et al. 1992,
Birmingham et al. 1994). The principle of the test is illustrated in Figure 6.59.
Fast burning fuel is detonated in a chamber accelerating a mass upwards and the recoil
applies a load to the head of the pile. Accelerations of about 20 g can be achieved, therefore

(a) (b)
Reaction mass Gravel container

–Fstn Gravel
Reaction masses
Silencer
Pressure Cylinder
chamber
Platform
Laser sensor
+Fstn Laser
Laser beam
Foundation pile

⇓ Piston
Load cell
Pile to be tested

Figure 6.59 Statnamic test setup: (a) principle of test; (b) test setup.
Deep foundations  249

the reaction mass, which generally consists of rings of concrete or steel only needs to be
about 5% of the load to be applied to the pile head.
During the test, a load cell and a laser sensor are used with a high speed laptop computer
to take about 400 readings per second.
Some of the advantages of the test are

1. The test is quick and easily mobilised.


2. High loading capacity is available with small reaction loads.
3. It can be adapted to lateral loading.
4. The test is quasi-static (i.e. the load is applied more slowly than for a dynamic test).

During the test, the load on the pile head increases until it reaches a maximum, and then
unloads back to zero. The reaction load is caught by gravel pouring in underneath the ram
so that the weights do not fall back onto the pile head.

6.25 PILE INTEGRITY TESTS

There are a number of different tests that can be carried out on bored concrete piles in order
to assess if the piles have been constructed correctly without defects. Common defects can
be cavities in the pile shaft or inclusions caused by material falling from the sides of the
drilled shaft. Pile integrity tests include

1. Drilling cores
2. Sonic tests
3. Radiometric logging
4. Vibration testing

6.25.1 Cross-hole sonic logging


Cross-hole sonic systems involve lowering two piezo-electric probes down parallel access
tubes (steel tubes are preferred) inside the pile (see Figure 6.60). The tubes are filled with
water prior to the test to ensure good acoustic coupling (ASTM D6760-02). One of the
probes is an emitter, and the other probe is a receiver. The system is restricted to bored
piles, and tests the integrity of the concrete between the tubes by measuring its effect on the
propagation of the sonic wave between the emitter and receiver (Stain and Williams 1991).
Sound concrete shows good transmission characteristics, but the presence of voids, soil,
or other foreign material affects the transmission of the signal. Generally, two tubes are
installed in the pile, but three tube or four tube (placed at the corners of a square) layouts
can be used.

6.25.2 Sonic integrity test


A simple and common test that does not require the pre-installation of equipment is the
sonic integrity test. The test involves striking the head of the pile with a plastic mallet, and
the measurement of the time interval for the reflected wave to return to a transducer con-
nected to the pile head (Tchepak 1998). Details of the test are given in ASTM D5882-07.
If a pile is sound, the waves will travel to the base of the pile and will be reflected back to
the pile head, with the travel time dependent on the wave velocity and the length of the pile.
In Figure 6.61a, the wave can be seen being reflected at the pile toe which is at 14 m. If there
250  Geomechanics in soil, rock, and environmental engineering

Thermal
Signal Zero Signal printer
generator processing
Digital
oscilloscope

Electrical Received Sonic profile


impulse signal printout

Voltage proportional to
the depth of the test
Winch
with
sensor

Concrete pile

Transmitter Receiver

Figure 6.60 Elements of a cross-hole sonic logging system.

(a)
Stroke = 72% Vel = 3916 m/s
Pile = 13.9 m

2 4 6 8 10 12 16
14

0
Pile = 14.0 m Fil = 0.0 m Vel = 3900 m/s Exp = 20 ×

(b)
Stroke = 57%
Pile = 7.4 m
Vel = 2200 m/s

1 2 3 4 5 6 7 8 9

0
Pile = 7.4 m Fil = 0.0 m Vel = 2200 m/s Exp = 10 ×

Figure 6.61 Typical results for sonic integrity tests on bored piles: (a) a sound pile; (b) an unsound pile.
Deep foundations  251

is a defect in the pile, the reflected wave will return earlier as shown in Figure 6.61b where
there is a change in impedance at about 2.7 m depth.
The test can identify:

1. Reductions in section
2. Increases in section
3. Shaft cracks
4. Zones of poor quality concrete
5. Large inclusions
6. Soil restraint
7. Toe level of pile

6.25.3 Gamma logging
Another technique that can be used for pile integrity testing is gamma logging, which uses
a radioactive source and a detector that can be used to measure variations in the density of
concrete in a drilled pile.
The radioactive source is generally Cesium 137 that emits gamma radiation. The detector
is a Geiger–Mueller probe, and the source and detector are placed into different PVC pipes
(50 mm diameter) that are cast into the pile in the same way as is done for cross-hole sonic
tests. The PVC pipes must be free of any water, and so are sealed to prevent water ingress.
The pipes should be inspected to make sure that they are free from water and other obstruc-
tions before testing.

6.26 CAPABILITIES OF PILE TEST PROCEDURES

Based on the comments made above in relation to the various types of test, Tables 6.9 and
6.10 summarise the perceived capabilities of the various tests to satisfy the needs of the

Table 6.9  Summary of capabilities of various pile load tests with respect to the results obtained
Ult. Struct.
Ult. axial lateral capacity
geot. geot. Load– Lateral Group and Special Ground
Test procedure capacity capacity settlment deflection effects integrity loadings movements
Static 3 0 3 0 1 1 1 0
uninstrumented
Static 3 0 3 0 2 2 2 2
instrumented
Static lateral 0 3 0 3 1 2 2 0
Dynamic (PDA) 3 0 2 0 0 3 1 0
Osterberg cell 3 0 2 0 0 1 1 0
Statnamic 3 2 2 2 2 1–2 1–2 0
(uninstrumented)
Statnamic 2 2 2 2 2 2–3 2 1
(instrumented)
3 = Very suitable
2 = May be suitable under some circumstances
1 = Possible but unlikely to be suitable
0 = Not suitable or not applicable
252  Geomechanics in soil, rock, and environmental engineering

Table 6.10  Summary of capabilities of various pile load tests with respect to the accuracy and relevance of
the results
Additional Accuracy of Accuracy of Similar duration
Pile loaded in stress changes movement load of loading to
Test procedure same way? (side effects) measurement measurement prototype?
Static uninstrumented 3 2 2 3 3
Static instrumented 3 2 2 3 3
Static lateral 3 2 2 3 3
Dynamic (PDA) 3 2 1 1 1
Osterberg cell 2 2 2 3 3
Statnamic 3 3 3 3 2
3 = Good
2 = May be adequate
1 = Generally not good

designer. It will be seen that no single test can satisfactorily supply all the information which
the designer may require, and that the static load test, which is usually considered to be the
‘benchmark’ test, usually provides only single pile capacity and stiffness. In addition, no test
can provide the ‘perfect’ load test without ‘side effects’, and as discussed in Section 6.23.5,
the interpretation of the static test should allow for the interaction between the test pile and
the reaction system.
The testing system chosen will depend on the information that is required from the test,
the cost of the test and the availability of equipment to perform the test.

6.27 NUMBER OF PILES TESTED

For tall buildings that may be supported by large numbers of piles, the question arises as
to the number of piles that should be tested so that the test results are representative of the
whole pile group.
The Australian piling code AS 2159-2009 specifies the percentage of piles to be tested for
serviceability conditions. This depends on the Risk Rating which is a number calculated
from Risk Factors such as the variability of the geology and the extent of the site investiga-
tion, experience with design in similar conditions, the extent of soil testing and the quality
of construction supervision. From tables in the code, an average risk rating (ARR) can be
calculated (higher risk has a higher ARR) and from this the number of piles to be tested
estimated as shown in Table 6.11. Testing of piles is only specified if the strength reduction
factor applied in design (which is an Ultimate Limit State Design) is greater than 0.4 (i.e. the
soil strengths are factored down by a value >0.4).
AS 2159-2009 also gives guidance for integrity testing. The amount of testing depends on
the pile type (e.g. precast or cast in place). Lower percentages of piles are specified for testing
if the pile design load is governed by soil strength rather than pile structural capacity. For

Table 6.11  Pile testing requirements for serviceability (AS 2159-2009)


Average risk rating (ARR) 2.50–2.99 3.00–3.49 3.50–3.99 4.00–4.49 ≥4.5
Percentage of piles to be tested for serviceability 1 2 3 5 10
Deep foundations  253

Table 6.12  Pile testing requirements according to risk


Characteristics of the piling works Risk level Pile testing strategy
Complex or unknown ground conditions. High Both preliminary and working pile tests essential.
No previous pile test data. 1 preliminary pile test per 250 piles.
New piling technique or very limited 1 working pile test per 100 piles.
relevant experience.
Consistent ground conditions. Medium Pile tests essential.
No previous pile test data. Either preliminary and/or working pile tests can be
used.
Limited experience of piling in similar 1 preliminary pile test per 500 piles.
ground. 1 working pile test per 100 piles.
Consistent ground conditions. Low Pile tests not essential.
Previous pile test data is available. If using pile tests either preliminary and/or working
tests can be used.
Extensive experience of piling in similar 1 preliminary pile test per 500 piles.
ground. 1 working pile test per 100 piles.
Source: Federation of Piling Specialists, Handbook on Pile Load Testing, 2006.

bored piles, the percentage of pile integrity testing depends on how carefully drilling fluid,
base cleaning and concrete tremie pouring is monitored.
For example, for a bored pile constructed using a casing or drilling fluid with good con-
struction monitoring and the design load governed by geotechnical capacity, 5%–15% of
piles should be tested. If the design load is governed by pile shaft structural capacity and
there is minimal construction monitoring, 15%–25% of piles need to be tested.
The Federation of Piling Specialists (United Kingdom) Handbook on Pile Load Testing
(2006) gives guidelines for the number of piles tested, and these are shown in Table 6.12.
The amount of testing depends on the amount of risk associated with the project.
The number of tests performed can also be estimated on a cost basis as described by Kay
(1976). More tests mean that the pile sizes can be refined saving money, but too many tests
will raise the cost due to the cost of testing. The formula of Equation 6.46 can be used to
calculate cost C.

XFm
C = (6.46)
F0 + mY

where F m is the factor of safety for m load tests, F 0 is the original factor of safety for no pile
load tests, X is the cost of the total number of piles, and Y is the cost of a single load test.
For tall buildings where the geological conditions are uniform and construction control is
good, generally one or two vertical pile load tests are performed, with perhaps a lateral load
test. Integrity testing may be performed on 10–15 piles and sonic integrity tests on 20–30
piles. A tension test may be required if some of the piles are subjected to uplift forces.

6.28 TEST INTERPRETATION

Information obtained from pile testing may be

1. The ultimate load capacity of a single pile


2. The load–settlement behaviour of a pile
254  Geomechanics in soil, rock, and environmental engineering

3. The acceptability of the performance of a pile, as-constructed, according to specified


acceptance criteria
4. The structural integrity of a pile, as constructed

Such information may be used in a number of ways, including

1. Construction and quality control


2. As a means of verification of design assumptions
3. As a means of obtaining design data on pile performance which may allow for a more
effective and confident design of the piles

6.28.1 Ultimate load capacity


Once a loading test has been carried out, the load–deflection curve for the pile may be plotted.
From this curve, it may be difficult to estimate where the pile reaches its ultimate load, as the
deflection curve may continue to climb with the loading and not show any clear-cut failure. In
this case, it is more usual to define the failure load as the load for a specific displacement. For
example, for conventional compression load tests, Eurocode 7 (2004) defines the failure (or
‘limit’) load as that causing a gross settlement of 10% of the equivalent base diameter.
In the case where the load is well below the failure load and deflections are small, a fre-
quently used approach is that of Chin (1970) which, in effect, assumes that the load–settle-
ment curve is hyperbolic, and extrapolates the load–settlement data on this basis. It has
been found commonly that Chin’s method tends to over-estimate the failure load, and it is
occasionally modified so that the failure load is taken as a proportion (typically 90%) of the
value derived from Chin’s construction. It is also possible to adopt a consistent approach and
extrapolate the load–settlement curve via Chin’s approach, but to define the failure load as
the value at a settlement of 10% of the diameter.

6.28.2 Pile stiffness
The deflection of a pile under load may be found from a pile load test and used to refine pre-
dictions of pile group or piled raft behaviour. If the deflection at a working load is required,
it may be adequate to back figure a secant pile stiffness, and to assume the pile has a linear
load–deflection behaviour over the range of loads anticipated.
At higher loads, the load–deflection behaviour measured for the pile will be non-linear.
Often, a hyperbolic relationship is used to model the pile stiffness in this case. Parameters
for the hyperbolic relationship can be changed until a good fit to the measured load–deflec-
tion behaviour of the pile is found.
In obtaining the pile stiffness whether linear or non-linear, it is necessary to allow for
the interaction of the pile with the datum for the measuring system as discussed in Section
6.23.5. Misleading values of pile head stiffness may be obtained if this is not done.
Correction is also needed for interaction with the reaction system. Kitiyodom et al. (2004)
have presented charts to allow correction of pile head stiffness found from pile load tests.
The charts are for the case where anchor piles are used. Poulos and Davis (1980) also pres-
ent charts to correct for the effects of the reaction system.

6.28.3 Acceptance criteria
In many cases, acceptance criteria are specified for quality control purposes, and are taken
from a code, without necessarily being related directly to the design. Typical criteria as
Deep foundations  255

Table 6.13  Acceptance criteria for vertical pile load tests (AS2159-2009)
Maximum settlement (mm)
Load Static load test
Serviceability load Ps PsL/AE + 0.01da
After removing serviceability load Max(0.01d,5)
Factored up pile load Pg PgL/AE + 10 + 0.05d
After removing the factored up pile load 10 + 0.05d
Notes:
1. The movement is to include no more than 2 mm creep over 3 h 45 min (after load has been
in place for 15 min).
2. Ps is the pile serviceability load, Pg is the load comprised of the factored up load combinations,
d is pile diameter, L is pile length, A is the pile cross-sectional area.
3. Loads are for no downdrag (or negative friction).

specified in the Australian Piling Code (AS2159-2009), for example, are shown in Table
6.13. The design may be affected by the load testing in that, if the piles are deemed to be
unacceptable, a decision then needs to be made on the future course of action by

• Re-design the pile foundation using more appropriate assumptions


• Replacement of the piles which have shown inadequate performance
• Addition of extra piles to compensate for the piles which have performed inadequately
• Re-analysis of the proposed foundation with the inadequate piles carefully to assess
whether the performance of the foundation system as a whole will perform
adequately

While there may be circumstances in which one of the first three options is inevitable,
there may also be instances where the group action may allow re-distribution of some of
the loads from the inadequate piles to the other piles, without causing unacceptable conse-
quences to the group performance.

6.28.4 Other quantities
Other information can be obtained from pile load tests as well as the usual ultimate pile load
and the load–deflection behaviour (or pile stiffness). With specially instrumented piles, the
load in the pile as well as the skin friction along the pile shaft may be obtained.
Strain gauges can be attached to the reinforcing cage of a pile or concrete type gauges can
be cast into the concrete. The strain gauges allow strains and therefore stresses to be calcu-
lated at various depths within the pile shaft, and therefore the stress and the load in the pile
shaft at that location. Tell-tales may also be used that consists of steel rods placed inside a
casing. The end of the rod is cast into the pile at a chosen depth, and the movement of the
head of the tell-tale monitored.
Extensometers may also be used to measure vertical movement within a pile shaft.
Vibrating wire or DCDT displacement transducers are installed inside 51 mm steel or
PVC sonic testing pipe. In one type of gauge, anchors at the top and bottom of the gauge
can be expanded using compressed air to attach the gauge to the sides of the pipe. The
relative movement between the top and bottom anchor of each gauge can be read to
assess strains. The gauges can be retrieved after testing by releasing the air pressure in
the anchors.
256  Geomechanics in soil, rock, and environmental engineering

6.29 MONITORING OF PILED FOUNDATIONS

Monitoring of the performance of piled foundations can be carried out so as to confirm


predictions of performance (i.e. settlements, pile loads), for reasons of safety or to provide
a warning of impending problems. The data obtained from monitoring is invaluable as it
may be used to refine soil models and analysis techniques for use in the design of similar
structures in the same area or similar soil profiles.
For research purposes, very comprehensive instrumentation may be installed in the piles
and beneath the raft of a foundation, and even in the surrounding soil. This may be per-
formed when constructing tall buildings in an area for the first time, or when foundation
conditions are very different across a site. For smaller structures or structures where there
is experience with similar site conditions and the risk is low, monitoring may simply involve
settlement measurements to confirm design predictions.

6.30 MEASUREMENT TECHNIQUES

There are many types of instruments that can be placed beneath or in piled founda-
tion ­systems, and the extent of the instrumentation depends upon the purpose of the
monitoring.
For example, the foundations of the Messe Turm in Frankfurt (which is supported by a
piled raft) contained 13 contact pressure cells, 1 piezometer, 3 multi-point borehole exten-
someters, and 12 instrumented piles. The piles contained strain gauges and load cells so
that the pile loads in the shaft and at the base of the piles could be measured. Details are
contained in Chapter 13 by Katzenbach et al. (2000) in the book by Hemsley (2000).
Katzenbach et al. (1995) describe the instrumentation placed beneath the Commerzbank
Tower in Frankfurt (Figure 6.62) that included 300 strain gauges placed inside 30 piles, 15
piles had load cells at the pile toe, 5 also had load cells at the pile head, and 6 piles had small
concrete load cells. Thirteen extensometers measured ground deformation down to a depth
of 95 m below the raft level, and 13 contact pressure cells and 4 piezometers were installed
beneath the raft.
These structures were the subject of extensive research as they were some of the first tall
buildings constructed on piled rafts in the Frankfurt clay. Some of the types of instruments
that may be used are discussed in the following sections.

6.30.1 Deflection
Probably, the most common and important measurements taken are of the movement of the
foundation. As extra stories are added to the structure, the foundation will compress, and
there may be immediate, consolidation or creep settlements of the soil, and elastic compres-
sion of the pile that take place. In addition, the foundation settlement may not be uniform
and the structure may rotate. Rotation can be serious for tall buildings since a small rotation
at the foundation level may mean large lateral movements at the top of the structure. Often,
corrections to the verticality of a tall structure are made as more stories are added if the
building is moving away from the vertical.
Deflection measurements may be taken with accurate levels onto a measurement marker
placed on the foundation. A benchmark that is not affected by the settlement of the struc-
ture needs to be used as the datum for the measurements. Total station theodolites may be
used to obtain both vertical and lateral movements of markers.
Deep foundations  257

Figure 6.62 Commerzbank Tower (Frankfurt, Germany).

Other equipment such as lasers and electronic inclinometers can be used to record lateral
movements and tilt of buildings. Dynamic behaviour of structures has been measured using
GPS techniques that are capable of measuring distances to sub-centimetre accuracy and col-
lecting data at 10 Hz (Luo et al. 2000).

6.30.2 Pressure cells
Pressure cells may be used to monitor the pressure beneath raft foundations or the load in
piles. This may be of interest if the load sharing between the raft and the piles is to be mea-
sured and compared with design estimates.
For measurement of pressure beneath a raft, Glötzl-type cells may be used as reported
in Hemsley (2000). The Glötzl cell has a thin sealed chamber containing oil or fluid that
causes a membrane to deflect when the fluid is pressurised. A fluid pressure is then applied
to the membrane to return it to its null position, and that pressure is taken as the pressure
applied to the cell.
258  Geomechanics in soil, rock, and environmental engineering

Load cells for measuring pressure at the base of piles are available and consist of
a ­fluid-filled  cell between two plates. The pressure of the fluid is measured by pressure
transducers.
In O-cell testing, the hydraulic pressure in the cell that is performing the loading within
the pile shaft is used to back figure the load being applied.

6.30.3 Strain gauges
Strain gauges have been mentioned previously in the section on pile testing (see Section
6.24). The gauges allow loads in the pile shaft to be calculated. These types of gauges are
generally used with bored piles where the gauge can be attached to the steel reinforcing cage,
cast into the concrete, or extensometers can be placed in tubes within the pile shaft. For steel
piles, strain gauges can be welded to the pile shaft.

6.30.4 Piezometers
Piezometers may be installed beneath rafts or piled rafts to monitor excess pore water pres-
sures generated during loading. In the case where groundwater has been lowered to allow
excavation of a basement, the total groundwater pressures will fall and rise again as the
pumping is ceased. The water pressures need to be suppressed by pumping until the weight
of the structure can counter the water uplift pressure, therefore water pressure monitoring
is important.
Various types of piezometer may be used (see Dunicliffe 1993) including standpipe,
hydraulic, and vibrating wire devices. The advantage of vibrating wire piezometers is that
they are connected by electric wires to the readout location and have a short lag time (i.e.
can register the pore pressure quickly).

6.30.5 Extensometers and inclinometers


Extensometers are sometimes placed beneath foundations to obtain the settlement
of the  foundation soils with depth. There are many different kinds of extensometers
­available  commercially, but most involve a hollow tube that can telescope and move
with the ground. Either magnets or steel rings are placed around the tube, and the posi-
tion of these rings is detected with a probe lowered into the tube. The probe can accurately
locate the position of the rings, so the soil movement at the locations of the rings can be
found.
Inclinometers may be used to measure lateral soil movement. A plastic casing is placed
into a borehole and grouted in place. The casing has grooves in the sides (generally two sets
at right angles) in which the wheels of a probe can run. As the probe is lowered down the
tube, an accelerometer takes readings of the inclination of the probe, and from these read-
ings, the lateral movements of the casing may be found. Some inclinometers can have a dual
role as an extensometer and an inclinometer.
Different types of extensometers and inclinometers are discussed in the book on instru-
mentation by Dunnicliffe (1993).

6.30.6 Frequency of measurements
Measurements of displacements, pile loads, etc. need to be taken as the structure is increas-
ing in height as the increased loads cause changes in the measurements of all instruments.
Once the construction is complete, the structure may continue to settle due to consolidation
Deep foundations  259

and creep of the foundation. Measurements may need to be taken for many years to assess
if the rate of settlement is slowing down. During this period, there may also be changes in
pile loads and raft moments.
It is therefore necessary to take several measurements for each storey that is constructed
of a tall building. For the One Shell Plaza building constructed in Houston, Texas, for exam-
ple (Focht et al. 1978), readings of instruments were taken at 2–4 month intervals during
construction. After the structural frame was completed, observations were made every 4–6
months. Two years after completion of the structure, readings were taken at yearly intervals,
up to 10 years post-construction.

6.31 COMPARISON WITH PREDICTED PERFORMANCE

Measurements from instruments installed on or in structural elements such as the piles and
raft or in the soil beneath or beside the foundation of a tall building, can provide valuable
information for use in predictive models.
The design and performance of the foundation can be carried out using finite element
methods (PLAXIS 3D, ABACUS), combined finite element and boundary element methods
like GARP (Small and Poulos 2007) or other simple hand techniques. Comparison of the
predicted and measured behaviour of a foundation may be used to assess whether the foun-
dation is behaving as predicted by the numerical models.
If the monitoring shows that the structure is not behaving as predicted (i.e. deflections are
suddenly becoming larger than expected, or excessive tilting is occurring or pile loads are
excessive) then remedial action may need to be taken. This may take the form of strength-
ening the foundation by adding more piles or trying to re-analyse the foundation to assess
whether the observed performance will result in an acceptable foundation for the particular
structure.
Comparison of numerical predictions and measured performance allows design parame-
ters to be refined and for less conservative designs to be carried out for similar soil c­ onditions.
Correlations used between results of field tests (such as SPT, CPT, s­ eismic tests, and pres-
suremeter tests) and pile design parameters (such as skin friction and end bearing pressures)
can also be refined through comparisons with field performance data.

6.31.1 Emirates twin towers, Dubai


The Emirates Project is a twin tower development in Dubai (Figure 6.63), one of the United
Arab Emirates. The towers are triangular in plan with a face dimension of approximately
50–54 m. The taller, the Office Tower, has 52 floors and rises 355 m above ground level,
while the shorter, Hotel Tower, is 305 m tall (see Poulos 2009).
A comprehensive series of in situ tests was carried out. In addition to standard SPT tests
and permeability tests, pressuremeter tests, vertical seismic shear wave testing, and site uni-
formity borehole seismic tests were carried out.
Conventional laboratory testing was undertaken, consisting of conventional testing,
including classification tests, chemical tests, unconfined compressive tests, point load
index tests, drained direct shear tests, and oedometer consolidation tests. In addition,
a considerable amount of more advanced laboratory testing was undertaken, including
stress path triaxial tests for settlement analysis of the deeper layers, constant normal
stiffness (CNS) direct shear tests for pile skin friction under both static and cyclic load-
ing, resonant column testing for small-strain shear modulus and damping of the founda-
tion materials, and undrained static and cyclic triaxial shear tests to assess the possible
260  Geomechanics in soil, rock, and environmental engineering

Figure 6.63 Emirates twin towers with construction almost complete (Dubai, United Arab Emirates).

influence of cyclic loading on strength, and to investigate the variation of soil stiffness and
damping with axial strain.
The geotechnical model for foundation design under static loading conditions was based
on the relevant available in situ and laboratory test data, and is shown in Figure 6.64. The
ultimate skin friction values were based largely on the CNS data, while the ultimate end
bearing values for the piles were assessed on the basis of correlations with unconfined com-
pressive strength (UCS) data (Reese and O’Neill 1988) and also previous experience with
similar cemented carbonate deposits (Poulos 1988).
Using the geotechnical data shown in Figure 6.64, predictions were made for pile load
tests, and once the tests were complete, the predictions and measured pile responses were
compared. Compression tests were performed using anchor cables as the reaction system.
Other tests that were performed were tension tests, lateral load tests, and cyclic load tests.
Four main types of instrumentation were used in the test piles:

• Strain gauges (concrete embedment vibrating wire type) – To allow measurement of


strains along the pile shafts, and hence estimation of the axial load distribution.
Deep foundations  261

Eu E′ fs fb pu
ν′ Unit
MPa MPa kPa MPa MPa
0
SILTY SAND, some 40 30 0.2 18 0.15 0.1 1
calcarenite bands
As above 125 100 0.2 73 1.5 1.5 2
10

CALCAREOUS 700 500 0.1 200 2.3 2.3 3


SANDSTONE
20

30 SILTY SAND 125 100 0.2 150 1.9 1.9 4


Depth (m)

40

CALCISILTITE 500 400 0.2 450 2.7 2.7 5

50

60
As above 90 80 0.3 200 2.0 2.0 6

70

As above 700 600 0.3 450 2.7 2.7 7

80

Figure 6.64 Geotechnical model adopted for design.

• Rod extensometers – To provide additional information on axial load distribution


with depth.
• Inclinometers – The piles for the lateral load tests had a pair of inclinometers, at 180°,
to enable measurement of rotation with depth, and hence assessment of lateral dis-
placement with depth.
• Displacement transducers – To measure vertical and lateral displacements.

Comparisons for one of the compression loading tests that were performed are shown in
Figure 6.65 (load–deflection behaviour) and in Figure 6.66 (for pile axial load with depth).
The predictions for this pile and for the other pile tests that were performed were considered
to be reasonable.
Predictions were then made for the foundation of the towers. In the final design, the
piles were primarily 1.2 m diameter, and extended 40 or 45 m below the base of the raft.
In general, the piles were located directly below 4.5 m deep walls which spanned between
the raft and the first level floor slab. These walls acted as ‘webs’ which forced the raft
and the slab to act as the flanges of a deep box structure. This deep box structure created
a relatively stiff base to the tower superstructure, although the raft itself was only 1.5 m
thick.
262  Geomechanics in soil, rock, and environmental engineering

30
Predicted
25 Measured

Applied load (MN)


20

15

10

0
0 10 20 30 40
Settlement (mm)

Figure 6.65 Predicted and measured load–settlement behaviour for Pile P3 (Hotel Tower).

Load (MN)
0 5 10 15 20 25 30
2
0
–2
–4
–6
–8
–10
–12
–14
Level DMD (m)

–16
–18
–20
–22
–24
–26
–28
–30
–32
–34
Measured (15,000 kN)
–36
Measured (23,000 kN)
–38 Predicted
–40

Figure 6.66 Predicted and measured axial load distribution for Pile P3 (Hotel Tower).
Deep foundations  263

Time (months)
1998

0 1 2 3 4 5 6 7 8 9 10 11 12
0

T4

10 Measured
T15
Settlement (mm)

20 Predicted

30

40

50

Figure 6.67 Measured and predicted time–settlement behaviour for the Hotel Tower.

Predictions of settlement were made using various computer programs, and these were
compared with the measured vertical settlement of the towers. A comparison of the pre-
dicted settlement and the measured settlement for the Hotel Tower is shown in Figure 6.67.
As can be seen from the plot in Figure 6.67, the predicted and measured settlements of the
Hotel Tower were not in close agreement even though the predictions for the pile load tests
were similar to those measured.
Two prime reasons for the larger prediction of settlement for the pile group were thought
to be (1) the interaction of the piles and (2) the stiffness adopted for the ground below
RL-53m. The calculation of interaction among the piles can have a large influence on group
settlements when there are a large number of piles, so the estimates of the interaction factors
were re-assessed. Pile groups stress the ground to greater depths than single piles, and if the
soil modulus at depth is different than assumed, this can lead to inaccuracies. Lower strain
levels in the ground at depth mean that a higher modulus should be used as the stiffness of
the ground is strain dependent. These two reasons may explain why single pile predictions
are reasonable, but the pile group predictions are not.
This experience was used when estimating the settlements of the Burj Khalifa where
stiffer layers at depth were used and interaction factors were calculated assuming stiffer
material existed between piles (and therefore the interaction was less). This gave reasonable
predictions of the long-term settlements of about 74 mm. Measured settlements of the foun-
dation before the tower was completed had reached about 42 mm, so the final settlement
may be close to that predicted.
The monitoring programme was therefore very useful in indicating that the modelling pro-
cedure for the piled raft had some shortcomings, and the experience gained from the Emirates
project could be used to make better predictions for the foundations of the Burj Khalifa
(Figure 6.81) which was founded in similar materials. This example demonstrates the value of
monitoring programmes, especially when designing in new or unfamiliar ground conditions.
264  Geomechanics in soil, rock, and environmental engineering

6.32 INTERPRETATION AND PORTRAYAL OF MEASUREMENTS

Measurements are often portrayed as a function of time. For instance, the settlement of a
structure can be plotted against time thus showing how it increases with height of a struc-
ture and how it continues to increase after construction is complete due to consolidation
and creep.
Loads in piles may also be plotted against time to monitor increases with applied load
increases. Loads in the pile shaft can be found from strain gauges, and this can be done for
conventional top loaded piles or when performing O-cell tests. The load in the pile shaft can
be used to make estimates of pile shaft skin friction and pile base load as is shown in Figure
6.68 for monitored piles of the Messe Turm building in Frankfurt.
Measurement deflections may be used to provide warning if the structure is not behaving
as predicted. For example, the tilt of the building may be seen to exceed the allowable value
or pile loads may exceed allowable values. In such cases, some remedial work may need to be
carried out or decisions need to be made as to whether the measured values can be tolerated
by the structure. For example, if some piles are overloaded, this may be acceptable if the
deflections are not excessive, since the load can be shed to other piles in the group.

6.33 PILED RAFTS

If a surface foundation is not adequate to carry structural loads without excessive differen-
tial deflections, piles may be needed. Both the raft and the piles then transfer load to the soil,
and the interaction problem involves both the raft and the piles. In some cases, the piles are
only placed beneath the raft to provide differential settlement control and are allowed to fail
under load (Hansbo and Källström 1983).
It is important to realise that piles do not need to be uniformly placed over a foundation,
but can be judiciously placed so as to carry the larger loads or to limit the differential deflec-
tions. In this regard, it is useful to have a quick and simple computer program or simple
design method that can be used in the design stage to determine the best layout of the piles

Pile load (MN) Skin friction (kPa)


I M O 0 5 10 15 0 50 100 150 200
0
0

I
10

O
Depth (m)

M
20
–26.9 m
O I
–30.9 m 30
–34.9 m M
O Outer pile-circle
40
M Middle pile-circle
I Inner pile-circle

Figure 6.68 Distributions of the pile load and skin friction from monitoring of the piles beneath the Messe
Turm building in Frankfurt.
Deep foundations  265

beneath the foundation. For example, Horikoshi and Randolph (1997) have shown that the
optimum design of a piled raft carrying a uniform load would involve piles placed under the
central 16%–25% of the raft area.
A piled raft foundation therefore, combines a raft with piles so the overall performance
of the raft can be improved.

• Piled rafts are useful where settlement or differential settlement of the raft is inadequate,
even though the raft may have an adequate factor of safety against a bearing failure.
• Piles are used as settlement controllers or reducers.
• There is a combined action of the raft and the piles – the piles do not carry all of the load.

6.34 USES OF PILED RAFTS

Rafts with piles placed so as to control settlements may seem an attractive alternative to
piles with individual caps that carry column loads. However, piled rafts are not the best
solution in every case, and the following sections discuss when piled rafts are favoured as
foundations and when they are not.

1.
Favourable circumstances
• Where load capacity of a conventional piled foundation is adequate, but settlement
or differential settlement is not
• There are relatively stiff clay profiles
• Dense sand profiles exist
• The foundation consists of layered profiles with no soft layers below pile tip level
• Where soil movements due to external causes do not occur

2.
Unfavourable circumstances
• Where soft clays exist near the surface
• Where loose sands exist near the surface
• Where consolidation settlements may occur due to external causes
• Where swelling movements may occur

3.
Alternate strategies for piled raft design
• Where piles operate at their normal load levels (a factor of safety of typically 2–3),
there is not a great potential for savings in design but if piles are used as settlement
inhibitors and are allowed to yield (factor of safety reaches 1), there is potential for
large savings

6.35 DESIGN CONSIDERATIONS

When designing a piled raft, the following design factors need to be taken into account:

• Bearing capacity of the piles and raft


• Maximum settlement of the foundation
• Differential settlement of the foundation
• Raft moments and shears
• Loads carried by the piles

There may also be circumstances where lateral loading due to wind loads or earthquake
loads need to be taken into account.
266  Geomechanics in soil, rock, and environmental engineering

6.35.1 Design process
The preliminary design process can involve estimating the number of piles required to sat-
isfy overall bearing capacity and settlement requirements.
Then the detailed design can be used to determine

• The optimum pile locations (to limit differential settlement)


• The differential settlements
• The shears and moments in the raft

6.36 BEARING CAPACITY OF PILED RAFTS

Work by de Sanctis and Mandolini (2006) (based on 3D finite element analysis) suggests
that the Factor of Safety for a piled raft can be calculated from the factor ξPR defined as

QPR, ult FSPR


ξ PR = =
QUR, ult + QG, ult FSUR + FSG
(6.47)

where
Q PR,ult is the ultimate load that the piled raft can carry
Q UR,ult is the ultimate load that the raft alone can carry
Q G,ult is the ultimate load that the pile group (with no raft) can carry
FS PR is the factor of safety of the piled raft = Q PR,ult /Q
FS UR is the factor of safety of the raft alone = Q UR,ult /Q
FSG is the factor of safety of the pile group (with no raft) = Q G,ult /Q
Q is the applied (working) load

Numerical analyses were carried out for different pile layouts (as shown in Figure 6.69)
beneath piled rafts using a three-dimensional finite element program. The results of the
analyses are shown in Table 6.14. In the table, H is the depth of the soil stratum, BR is the
full width of the raft, s is the centre-to-centre pile spacing, L is the pile length, and d is the
pile diameter.
It can be seen that the value of the ratio ξPR varies from a minimum of 0.82 up to a maxi-
mum of 1.0. It was therefore recommended by de Sanctis and Mandolini to take the value of
0.80 for estimating the factor of safety of piled rafts in bearing, that is,

QPR ,ult = 0.8 (QUR , ult + QG, ult )


(6.48)

or dividing both sides of the equation by the applied load Q

FSPR = 0.8 (FSUR + FSG) (6.49)


Hence, the suggested ultimate load that can be carried by a piled raft is 0.8 times the sum
of the capacity of the raft alone and the capacity of the piles alone.
As much greater movement of the cap is needed to mobilise the bearing capacity of the
cap than for the piles, it is suggested that the bearing pressure of the cap be taken as that
of a strip footing with a width equal to the distance between the edge of the cap and the
outer pile.
Deep foundations  267

BR = 28 d

n = 9, 49
AG
L/d = 20, 40
s/d = 4, 8
H/L = 2

BR = 20 d

AG n = 9, 25
L/d = 20, 40
s/d = 4, 8
H/L = 2

BR = 12 d
n=9
AG
L/d = 20, 40
s/d = 4
H/L = 2

Figure 6.69 Pile groups examined by de Sanctis and Mandolini (2006).

Table 6.14  Ratio ξPR for piled rafts analysed by de Sanctis and Mandolini (2006)
Case L/d n s/d BR/d FSUR FSG FSPR ξPR
1 40 49 4 28 1.95 6.46 7.76 0.92
2 40 9 4 28 1.95 1.19 2.89 0.92
3 40 9 8 28 1.95 1.19 2.99 0.95
4 20 49 4 28 1.95 2.15 3.35 0.82
5 20 9 4 28 1.95 0.40 2.28 0.97
6 20 9 8 28 1.95 0.40 2.32 0.99
7 40 25 4 20 2.11 6.98 8.84 0.97
8 40 9 4 20 2.11 2.51 4.29 0.93
9 40 9 8 20 2.11 2.51 4.51 0.98
10 20 25 4 20 2.11 2.33 3.64 0.82
11 20 9 4 20 2.11 0.84 2.54 0.86
12 20 9 8 20 2.11 1.84 2.70 0.92
13 20 9 4 12 2.26 2.49 4.06 0.86
14 40 9 4 12 2.26 7.47 9.71 1.00
268  Geomechanics in soil, rock, and environmental engineering

6.37 ANALYSIS OF PILED RAFT FOUNDATIONS

In the past, piles were treated as groups that were rigidly joined at the head or carried equal
loads, and the flexibility of the raft that joined the pile heads was ignored. The book by
Poulos and Davis (1980) includes many of the methods for computing the settlement of piles
or pile groups when the raft is assumed to be totally rigid or totally flexible (i.e. raft flex-
ibility is one of two extremes). These solutions are based on treating the shear forces acting
down the pile shaft as a series of uniform shear stresses acting over sections of the pile shaft.
Mindilin’s (1936) equation for a sub-surface point load is integrated over the section of pile
to obtain the solution for the effect of the uniform shear stress on deflections of the soil at
other sections of pile for the pile itself or for other piles. Interaction between piles can there-
fore be found using this technique often called a ‘boundary element’ technique.
Many different means of analysing piled raft foundations have been developed over the
years (comprehensive reviews have been provided by Randolph 1994 and Hemsley 2000).
Some of the methods that have been used to analyse piled rafts can be conveniently divided
into the following groups:

1.
Simple Plate on Spring Approaches
These methods treat the piles as springs with the raft treated as a plate, and include the
methods of Clancy and Randolph (1993), Poulos (1994), and Viggiani (1998).
2.
Boundary Element Methods
These methods employ the technique described above and include solutions obtained
by Butterfield and Banerjee (1971), Brown and Weisner (1975), Hain and Lee (1978),
Kuwabara (1989), and Chow (1986).
3.
Finite Layer Techniques
Ta and Small (1996) used finite layer techniques (see Chapter 2) to compute the behav-
iour of piled rafts, where the piles were driven into layered soils. Cheung et al. (1988)
had previously used series to analyse the behaviour of pile groups in layered soils, and
the method can be extended to piled rafts. Zhang and Small (2000) have extended
these techniques to horizontal loading of a piled raft.
4.
Simplified Finite Element Analyses
Analyses can be carried out by approximating the piles as a two-dimensional or axi-
symmetric body and assigning ‘smeared’ material properties to the piles in order to
approximate the actual three-dimensional behaviour. That is the solid continuous
‘pile’ in an axi-symmetric or 2D analysis is given a lower modulus to make it compress
the same amount as the actual individual piles. Analyses of this sort include those of
Desai et al. (1974) and Hooper (1973). Lin et al. (1999) have used a finite difference
technique to compute the behaviour of the soil beneath a piled raft, and applied the
theory to piled rafts in Bangkok clay using a two-dimensional finite difference grid.
5.
Three-Dimensional Finite Element Analyses
As computer storage has increased, full 3D analyses of piled rafts have been carried
out and early examples of this are given by Zhuang et al. (1991), Katzenbach and Reul
(1997), Katzenbach et al. (1997), and Ottaviani (1975).

6.37.1 Numerical modelling
In the previous section, many different methods of piled raft analysis were listed, and the
model chosen for a particular application would depend on the degree of sophistication
required in the analysis. Spring models that treat the soil and piles as springs, with no
interaction between the springs are inaccurate and it is recommended that they not be used
Deep foundations  269

because the effects of interaction are large. Most of the simple techniques involve treating
the soil as being a uniform elastic material so that there is interaction between the piles and
the raft, and this leads to much improved solutions. In some cases, loads on the piles are
limited by using a load ‘cut-off’ to limit pile loads to a maximum value, thus simulating yield
of the piles.
Early analytic solutions such as those by Hain and Lee (1978) made use of the theory of
elasticity to compute the settlement of a pile under load and its effect on other piles and
on the ground surface. To compute the behaviour of a piled raft, they treated the contact
stress beneath the raft as a series of blocks of uniform load, a technique originally used
by Zhemochkin and Sinitsyn (1962). It is therefore also necessary to be able to compute
the interaction of surface loads and the piles, and surface loads with each other. Hain and
Lee did this using the theory of elasticity as well, and combined all of these interactions to
compute the piled raft behaviour. For the analysis of the raft, they used the finite element
technique. The loads were considered to be uniform over each of the rectangular elements in
the raft, so an equal and opposite block of rectangular pressure was applied to the soil. The
interactions required are shown schematically in Figure 6.70.
The solution involves calculating the deflections at each pile and block of load due to
loadings on the piles or due to other surface blocks of load. This, along with the equations
of equilibrium (i.e. the total upward contact force is equal to the total downward force of
the applied loads and moment equilibrium) provides enough equations to solve for the mag-
nitudes of all the contact stress blocks. Once these are known, the deflections and moments
in the raft can be obtained. This approach can be extended to include both horizontal and
vertical loading (Zhang and Small 2000).
Similar approaches have been used by Clancy and Randolph (1993), Ta and Small (1996),
and Poulos (1994), with different means of calculating the interactions between the piles
and the surface loadings from the raft being used to allow non-linearity or layered materials
to be taken into account.
It is desirable to know the effects of assumptions made in the different types of analyses,
therefore in the following sections, a limited examination is made of some aspects of the
analyses listed.

6.37.2 Finite layer techniques


Finite layer techniques have been discussed in Chapter 2, and can be used to calculate the
deflection of the surface of a horizontally layered soil due to a rectangular uniform load.
This can be used to find the effect of the surface load on deflections of other surface patches
of load or on locations down the pile shaft.
If the shear loads acting on the pile shaft can be treated as a series of ring loads, then the
finite layer method can be used to compute the deflections at the locations of these ring loads
and all other ring loads as well as at the location of the rectangular surface loads (generally
the centre of the load is chosen).
This can also be carried out for horizontal ring loads down the pile shaft and for lateral
loads over rectangular regions on the surface. This then enables both vertical and lateral
loading to be applied to the piled raft as shown in Figure 6.71.
The method is similar to that originally used by Hain and Lee (1978) except that the finite
layer method means the analysis of the piled raft can be performed for layered soils where
each layer has different properties. Soil layers of finite depth can also be analysed.
The various ring and rectangular loads are shown in Figure 6.72 for general three-dimen-
sional loading. Loads are also applied back to the raft and the piles (in the same locations as
the loads on the soil) and the deflections of the structure are calculated.
270  Geomechanics in soil, rock, and environmental engineering

(a) P

d
S

(b) q

a
L

d
S

(c) P

d
S

(d) q

a
S

Figure 6.70 Interactions among piles and surface loads: (a) pile-to-pile interaction effect; (b) surface-to-pile
interaction effect; (c) pile-to-surface interaction effect; and (d) surface-to-surface interaction
effect.

By matching the deflection of the structure and the soil, plus using the equations of equi-
librium for the piled raft, there are enough equations to solve for the magnitudes of all the
unknown forces applied. These forces can be re-applied to the structure to calculate deflec-
tions, moments, and shears in the raft and moment and deflections in the piles.
The raft can be analysed by the use of finite element techniques. Often it is treated as
being a thin plate, so that thin plate theory (Timoshenko and Woinowski-Krieger 1959) can
be used in the analysis. For rafts that are 2–3 m thick, this may lead to inaccuracy, although
it is the width to thickness ratio of the raft that determines its flexibility.
Slip of the piles can be modelled by limiting the ring loads to a maximum value. In addi-
tion, failure of the soil under the raft can be modelled by limiting the pressure to a maxi-
mum value. Liftoff of the raft may be modelled by not allowing the contact stress to become
negative. If it does become negative, it is set to zero.
Deep foundations  271

Mz

Pz My
Mx
Py
Px

Figure 6.71 Three-dimensional set of loads applied to a piled raft.

6.37.3 Non-linear behaviour
If piles are designed to reach their maximum load, or are designed to carry a high propor-
tion of their maximum load, then slip of the piles becomes important and non-linear behav-
iour of the piles should be taken into account. This type of behaviour becomes important if
piles are used to control differential deflections and are designed to yield or fail.
Full three-dimensional finite element analyses have been used to allow non-linear behav-
iour of piled raft foundations (Katzenbach et al. 1997), but simpler techniques have been
developed. Clancy and Randolph (1993) have presented a non-linear analysis for piles and
this is incorporated into the computer program HyPR (Hybrid Piled Raft Analysis), and
Bilotta et al. (1991) have presented two methods for computing the behaviour of piled rafts

Pz External forces
Mx
Pile Px
tr

{Pr}
Interface forces
Interface forces between the raft
between the and soil
piles and soil

Interface forces transferred from piled raft to the soil

z
Soil
Ring loads
acting on soil
nodes

Circular loads
acting on pile
base

Figure 6.72 Loads applied to the soil and to the piles and raft.
272  Geomechanics in soil, rock, and environmental engineering

Computed (piles elastic)


6
Computed (piles skin friction = 80 kPa)
Measured
Settlement (mm)

100

% Pile load
4 Lp = 225 mm
8-pile group

50
2

0 0
0 5 10 15 20 0 5 10 15 20
Pile diameter (mm) Pile diameter (mm)

Figure 6.73 Centrifuge results compared with computed results.

where the pile may have a non-linear load–displacement relationship. The latter authors
stress the point that if piles are designed to yield, then a non-linear analysis of the piles is
essential.
Poulos (1994) has demonstrated the need to take pile non-linearity into account in analy-
sis of centrifuge tests on piled rafts (Thaher and Jessberger 1991). Figure 6.73 shows the
settlement and pile load predictions made for a piled raft with eight piles. If the pile is
not allowed to fail, then the results of the analysis do not match the observed behaviour.
However, if the skin friction on the piles is limited to 80 kPa so that the piles can yield, then
the predicted values are much closer to the measured values for both settlement and percent-
age load carried by the piles.

6.38 EXAMPLE OF THE FINITE LAYER METHOD

In order to test the accuracy of the Finite Layer Method, solutions were obtained from a
three-dimensional finite element program, and from a finite layer program (APRAF – Zhang
and Small 2000) for a piled raft with a horizontally applied loading.
The raft is shown in Figure 6.74 and consists of a 3 × 3 pile group with a raft in contact
with the ground surface. The raft overhangs the piles by one pile diameter (around the
perimeter). The finite element mesh used to model this raft is shown in Figure 6.75 where it
may be seen that one quarter of the raft is modelled because of symmetry. The mesh extends
further in the x-direction because loading is to be applied to the raft in that direction, and
the boundary should not affect the results by being too close.
All of the properties of the piled raft are given in Table 6.15, and dimensions are shown
on Figure 6.74.
Two horizontal point loads were applied to the heads of each pile (18 loads in all) making
a total horizontal loading of 18 MN. For the purposes of comparison, no slip was allowed
between the raft and the soil, or the piles and the soil. The deflection of the raft can be calcu-
lated from the finite layer method, and a section (A–B in Figure 6.74) through the deformed
raft is shown in Figure 6.76. In the figure it can be seen that the raft rotates under the hori-
zontal loading and at its centre (x = 0) does not undergo vertical movement. The computed
results from the finite layer and finite element methods can be seen to be in reasonably close
agreement.
Deep foundations  273

Br

3
y
Pxi
A 2 5 x B Lr
D S

1 4

Overhang

t
x

z Soil
L
Pile

Figure 6.74 Piled raft analysed using either finite element or finite layer methods.

The moments in the piles may also be computed, and are shown in Figure 6.77. In this
figure, moments down the pile shaft are shown for pile 1 (the corner pile) and pile 5 (the
centre pile), and it may be seen that there is very close agreement between the finite element
and finite layer values.

6.39 APPLICATIONS

The application of various computer programs that have been developed for the analysis
of piled raft foundations are examined in the following when applied to case histories.

x
y

Figure 6.75 Finite element mesh used for piled raft analysis.


274  Geomechanics in soil, rock, and environmental engineering

Table 6.15  Properties of a piled raft (3 × 3 group)


Quantity Value
Pile diameter d 0.5 m
Pile length L 10 m
Depth of soil 15 m
Pile spacing y; Raft width Lr s/d = 3; 9 m
Pile spacing x; Raft breadth Br s/d = 3; 9 m
Overhang of raft 0.5 m
Raft thickness 0.25 m
Soil modulus 10 MPa
Soil Poisson’s ratio 0.3
Raft modulus 30,000 MPa
Raft Poisson’s ratio 0.3

Full three-dimensional finite element programs can be used to model piled rafts, but there
is considerably more effort involved than using programs based on continuum theory and
soil–structure interaction as described in Section 6.38.

6.39.1 Westend Strasse 1 tower


A second example of the application of some of the analytic techniques mentioned in Section
6.37 is that of the Westend Strasse 1 tower in Frankfurt, Germany (see Figure 6.78). The
building is 51 stories high (208 m) and has been described by Franke et  al. (1994) and
Franke (1991). The foundation for the building was a piled raft with 40 piles that were 30 m
long as shown in Figure 6.79.

–60
APRAF
Finite
–40 element

–20
Deflection (mm)

–0.50 –0.40 –0.30 –0.20 –0.10 0.00 0.10 0.20 0.30 0.40 0.50
0
Normalised distance x/Br

20

40
Piled raft under point
horizontal loadings s/d = 3
60

Figure 6.76 Deflection of the raft along centreline A–B.


Deep foundations  275

Bending moment in pile (MN m)


–3 –2 –1 0 1 2
0.0

0.2 Pile 1, finite element


Pile 1, APRAF
Pile 5, finite element
Pile 5, APRAF
Normalised depth

0.4

0.6

Piled raft under point


horizontal loadings s/d = 3

0.8

1.0

Figure 6.77 Moment variation with depth for piles beneath laterally loaded piled raft.

The foundation was constructed in a deep deposit of the Frankfurt clay 120 m thick,
and using pressuremeter tests reported by Franke et al. (1994), the modulus of the clay was
assessed to be 62.4 MPa.
The ultimate load capacity of each pile was computed to be 16 MN and a total load of
968 MN was assumed to be applied to the foundation (this is greater than the ultimate
capacity of the individual piles).
Six methods were used to predict the performance of the piled raft foundation:

1. The boundary element approach of Poulos and Davis (1980)


2. Randolph’s method (1983)
3. The strip on springs approach using the program GASP Poulos (1991)
4. The raft on springs approach using the program GARP Poulos (1994)
5. The finite element and finite layer method of Ta and Small (1996)
6. The finite element and boundary element method of Sinha (1997)

Measured values were available for the settlement of the foundation, the percentage of
load carried by the piles, the maximum load carried by a pile in the group and the minimum
load carried by a pile in the group. The results of the six different analysis methods are
shown in the bar chart of Figure 6.80 compared with the measured values and the values
reported by Franke et al. (1994).
276  Geomechanics in soil, rock, and environmental engineering

Figure 6.78 Westend Strasse 1 tower building in Frankfurt.

(a)

(b)

Main Side building raft


tower

Side
building
208 m

Main tower
40 piles raft

15 m
30 m

Figure 6.79 Layout of Westend Strasse 1 tower in Frankfurt, Germany. (a) Elevation; (b) plan. (Franke,
E., Lutz, B., and El-Mossallamy, Y. 1994. Vertical and Horizontal Deformation of Foundations and
Embankments, ASCE, Geotechnical Special Publication No. 40, Vol. 2, pp. 1325–1336.)
Deep foundations  277

Central settlement Proportion of pile load


200 80

150 60
Settlement: mm

% Pile load
Poulos and Davis
100 Ta and Small 40

Ta and Small
Franke et al.

Franke et al.
Measured

Measured
Randolph

Randolph
50 20
GARP

GARP
GASP

GASP
Sinha

Sinha
0 0
Method Method

Maximum pile load Minimum pile load


20 20

15 15
Pile load: MN

Pile load: MN
10 10
Ta and Small

Ta and Small
Franke et al.

Franke et al.
Measured

Measured
5 5
GARP

GARP
GASP

GASP
Sinha

Sinha
0 0
Method Method

Figure 6.80 Comparison of results from different analysis methods. Westend Strasse 1 tower in Frankfurt,
Germany.

From Figure 6.80, it may be seen that

1. Most of the methods over-predicted the settlement of the foundation. However, this
depends on the soil modulus chosen, and it can only be concluded that most of the
methods gave a reasonable estimate of the settlement for the adopted soil stiffness.
2. Most of the methods over predicted the percentage of load carried by the piles, although
the calculated values are acceptable from a design point of view.
3. All of the methods that are able to give a prediction of pile load suggest that the most
heavily loaded pile is almost at its ultimate capacity, and this is in agreement with the
measured value.
4. For the minimum pile load, there is a considerable variation in the calculated results,
with three of the methods indicating a much larger value than was measured.

These results show that when some of the piles are carrying loads close to their capacity,
there can be significant variability in the computed results, especially for simple methods
and methods based on the theory of elasticity.

6.40 STRUCTURAL STIFFNESS

When analysing rafts, or piled rafts, inclusion of the stiffness of the superstructure will
reduce the differential deflections in the raft, and this aspect may need to be addressed. The
relative stiffness of the superstructure will determine the effect on deflections, but for very
flexible structures, the raft alone can be analysed without great error. Neglecting the stiff-
ness of the superstructure can be conservative, but it may be of importance where there are
278  Geomechanics in soil, rock, and environmental engineering

Figure 6.81 The Burj Khalifa, Dubai (United Arab Emirates).

(a) (b)

Figure 6.82 (a) Raft Model 2. (b) Raft Model 3.


Deep foundations  279

Distance along wing (m) NAPRA Model 1


0 20 40 60 80
0 NAPRA Model 2
10
NAPRA Model 3
20
Settlement (mm)

NAPRA – Using
30
average pile loads
40 NAPRA – Model 3
modified
50
Original design
60 profile
70 Measured (18 February
2008)
80

Figure 6.83 Measured and computed settlements along Wing ‘C’ of the Burj Khalifa.

stiff shear walls. These walls can be treated as stiff raft elements in some cases, thus model-
ling the restraining effect of the stiff structural elements.
An example of this was an analysis carried out for the Burj Khalifa in Dubai (Russo et al.
2012). In order to investigate the effect on the computed settlement and differential settle-
ment, and to try and obtain as accurate an estimate of the pattern of settlement, the stiffen-
ing effect exerted by the superstructure on the raft was taken into account by increasing the
bending stiffness of the raft in each wing (estimated by the structural designers to be equiva-
lent to an increase of 25,200 kNm 2 per wing). Three alternative methods of incorporating
this increased bending stiffness were adopted:

1. Increasing the thickness of the whole raft (Model 1)


2. Increasing the raft thickness over the central part of the wings and on the core, as
shown in Figure 6.82a; this is denoted as Model 2
3. Increasing the raft thickness only below the shear walls (see Figure 6.82b), denoted as
Model 3

In addition, the actual pattern of loading via the columns and walls was applied. The pro-
gram NAPRA was used to carry out the analysis (the details are given in Russo et al. 2012)
and the deflections along one of the ‘wings’ (see the three wings in Figure 6.81) calculated.
As can be seen in Figure 6.83, the deflections computed were similar in magnitude to the
values measured up until 2008. Further settlement is expected to occur in the future as was
predicted in the original design profile (Figure 6.83).
This page intentionally left blank
Chapter 7

Slope stability

7.1 INTRODUCTION

The stability of slopes is an important part of geotechnical engineering as some of the great-
est damage to property and loss of life has occurred through landslips. Slips can be driven
by gravity alone but the effects of earthquakes can also be a factor in causing a slip to occur.
Most importantly, it is the water that exists in a slope that can cause instability, and it
is often after rainfall events that slope instability occurs. An example of a slope failure
caused by the introduction of water is the slip that occurred in 1997 in the ski resort town
of Thredbo in the Australian Alps (see Figure 7.1). The slip took place at 11.30 P.M. on the
30th July when skiers were asleep in their ski lodges. It demolished Carinya Lodge killing
one person, before the earth flow pushed the wreckage of Carinya Lodge into Bimbadeen
Lodge that was further downslope, killing another 17 people. There was only one survivor
from those who were in the two lodges that night.
The ski lodges in that part of Thredbo village where the slip occurred were located on a
fairly steep slope beneath a road embankment for the Alpine Way that had been constructed
some 42 years earlier (in 1955) by pushing soil into place with a bulldozer. The road was
only constructed as an access road for the Snowy Mountains hydro-electric scheme, and
contained logs from trees that had been pushed over by the bulldozer. The fill was not prop-
erly compacted, as it would be for a modern highway.
Carinya Lodge was constructed on a slope of about 30° that became much steeper (esti-
mated to be about 45° similar to other slopes in the area) where the Alpine Way embank-
ment sloped up towards the roadway. A cross section through the site is shown in Figure 7.2.
What caused the road to slip after so many years after its construction was the subject of
an intense investigation, and the findings of the investigation were that the slope above the
lodges had been moving slowly (creeping) for some time. The movement eventually became
large enough to pull apart the coupling in a pipe that was buried in the downhill slope of the
roadway. Once this occurred, water gushed into the slope making it totally unstable, and a
rapid slope failure took place. The pipe coupling is shown in Figure 7.3.
This case illustrates two features of slope stability: (1) water has a profound effect on the
stability of a slope and (2) slope movements can be very slow (creep) and can take place
over several years before the stability of the slope or the cumulative movement of the slope
becomes a problem.
The strength of the soil in the slope resists movement of the slope through shearing,
and this is also of importance. Slopes that have been in place for a period of time behave
in a drained fashion, and the effective strength of the soil and the pore water pressures
are used in analysis. This is the most common type of analysis performed; however, there
are slopes that have been cut rapidly for which an undrained analysis is more appropriate.

281
282  Geomechanics in soil, rock, and environmental engineering

Figure 7.1 Destroyed ski lodges after the 1997 Thredbo landslip.

Various forms of analysis that may be applied to slope stability are discussed in the follow-
ing sections.

7.2  SLIP CIRCLE ANALYSIS

It is often observed that when a slope fails it does so by undergoing a rotational failure with
the soil shearing along a roughly circular failure surface. If there are planes of weakness
in the soil lying on the slope, then the failure surface is likely to pass through one or more of
the weak planes and the surface will not be circular but have some other shape. However
for uniform soils, the circular failure surface may be considered, and this has been used in
many simple methods of analysis including the method of slices. This is assuming that the
slip is two-dimensional in nature, when in reality it will be a three-dimensional bowl shaped
failure surface. Three-dimensional failure surfaces are addressed in Section 7.2.8.

1440 m
Alpine Way
1435 m
1996 1430 m
1946
1425 m
1420 m
1415 m
Carinya
1410 m
Section E
1405 m
1400 m
1395 m
1390 m
0 5 10
Metres

Figure 7.2 The Thredbo landslip. The slip occurred below Alpine Way and above the Carinya Lodge.
Slope stability  283

Figure 7.3 Fibre cement pipe at the top of the slope, thought to have leaked prior to the slide.

7.2.1  The method of slices


This method involves dividing the region above the supposed circular slip surface into a
series of slices that can be of different thicknesses as shown in Figure 7.4. In some methods,
the slices do not need to have vertical sides, but this is not addressed here.
The forces acting on each slice are then considered. As is known from testing of soils in
shear (say in a direct shear box), the shear strength τ on a shear surface at failure is given
by Equation 7.1 where c is the cohesive strength and σn is the normal stress acting on that
surface.

τ = c + σ n tan(φ) (7.1)

If the problem is one involving drained conditions, the angle of shearing resistance is ϕ′,
and the cohesion is c′ while the effective stress σ′n is used for the normal stress. If the prob-
lem involves undrained conditions then ϕ = ϕu is the undrained angle of shearing resistance,
the cohesion c = cu and the total stress σn are used.

Surface
loading

Piezometric surface
May include
tension crack

Circular slip surface

Slices do not need to be


of equal thickness

Figure 7.4 The method of slices used for a circular failure surface.


284  Geomechanics in soil, rock, and environmental engineering

bn Tn
c′bn sec α

an ϕ
N n′ t F
F

Xn+1 θ1
Zn+1 N n′ Wn
θn+1 Force
En+1 polygon
En
Wn U Q
θn
Resultant
yn Zn
Zn Zn+1
Xn
Tn
αn Nn = Nn′ + U
Nn′ Tn = (Nn′ tan ϕ′+ c′bn sec α)/F
ℓn tan θ1 = tan ϕ′/F
U

Figure 7.5 Forces acting on a single slice.

If the soil is not at the point of failure then the full soil strength is not mobilised, and the
shear τf operating on a shear plane will be

c tan(φ)
τf = + σn (7.2)
F F

This is assuming that the factor of safety F applies equally to the cohesive strength c com-
ponent and the frictional strength tan(ϕ). This may not be the case, but it is a simplifying
assumption. In addition, the same factor of safety may not apply all along a potential slip
surface, and this too is an assumption.
Hence, the angle of mobilised friction θ is given by tan θ = tan(ϕ)/F. The resultant forces
due to friction therefore act at an angle of θ to a face of the slice, as the full angle of friction
is not operating if the slope has not slipped. This is shown in Figure 7.5, which shows forces
acting on a single slice in the drained case. In the figure, Wn is the weight of the nth slice,
U is the water force, N′n is the normal effective force on the base of the slice, and Tn is the
shear force acting on the base of the slice. The forces acting on the sides of the slice are the
normal En and the shear force Xn such that the resultant force is Zn acting at the angle θn to
the normal. The locations at which the force at the base of the slice ℓn and at the sides of the
slice yn act are not known, but various assumptions can be made to overcome this problem
as will be seen later.
Various methods have been proposed by several researchers for finding the factor of
safety of the overall assembly of slices above the slip circle, and some of these are listed
below:

• The Swedish method for circular failure surfaces (Fellenius 1927)


• Bishop’s simplified method; circular failure surfaces (Bishop 1955)
• Spencer’s method; circular surfaces (Spencer 1967)
Slope stability  285

7.2.2  The Swedish, Fellenius, USBR, or Common Method


This method (which is referred to by several names – most commonly the Swedish method)
is the simplest method in which the assumption is made that the side forces on the slices
cancel each other (i.e. are equal and opposite).
Then if moments are taken about the centre of the circle, and the overturning and resist-
ing moments are assumed to be equal (for a static situation) then the factor of safety can be
found (in terms of drained strength parameters) as shown in Equation 7.3.


n
[ci′∆ i + (Wi cos θi − Ui )tan φ′i ]
F = i =1
(7.3)

n
Wi sin θi
i =1

and Δℓi = bi /cos θi is the length of the base of the slice.

7.2.3  Bishop’s method and simplified method


Bishop (1955) developed a method that did not involve as many assumptions as the simple
Swedish method. The simplified Bishop method makes the assumption that the vertical
forces on the sides of the slices cancel, and the normal force at the base of the slice N′ni can
be found by resolving forces vertically. The resulting expression for the factor of safety (in
terms of effective strength) is given in Equation 7.4.


n
[ci′bi + (Wi − uibi )tan φ′i ]/Mi (θ)
F = i =1
(7.4)

n
Wi sin θi
i =1

 tan θi tan φi 
Mi (θ) = cos θi  1 + 
 F

It can be seen from the equation that the factor of safety is involved in both sides of the equa-
tion as it is included in the Mi(θ) term. Therefore, an iterative approach needs to be taken,
whereby an initial value of the factor of safety is chosen and used to calculate M i(θ) and
then a new value of F is found from Equation 7.4. The new value is then used to get a better
estimate of Mi(θ), and the process is repeated until convergence is obtained.
In practice, the Swedish method is generally used to obtain the first estimate of F and then
this is used as the starting value for the iterative process. If this is done, an accurate value of
the factor of safety can be obtained in about three iterations.
Bishop’s full method requires values of the side forces Xn to be found so as to obtain force
equilibrium of the slice; however, the error involved in omitting this in the simplified method
is small.

7.2.4  Spencer’s method


Spencer (1967) made the assumption that all of the side forces were inclined at the same
angle θi to the sides of the slice. He then found a factor of safety Ff that would give force
equilibrium for various values of the angle θi and a factor of safety F m that would give
286  Geomechanics in soil, rock, and environmental engineering

1.10

Fi = 1.070

1.05 Fm
Fmo = 1.039
F

1.00

Ff
0.95

θi = 22.5°
0.90
0 5 10 15 20 25
θ (degrees)

Figure 7.6 Variation of factors of safety for moment equilibrium and force equilibrium with θ.

overall moment equilibrium for different values of θi. The value of θi for which Ff = F m is the
required angle and the factor of safety for the slope is the corresponding value of F = Ff = F m
as shown in Figure 7.6.

7.2.5  Finding the critical circle


To find the slip circle with the lowest factor of safety, many different circles must be trialled,
and the one with the lowest factor of safety found. This can be done fairly quickly with com-
puter programs, and different search algorithms have been developed to aid in the search.
One method is to create a grid above the slope on which the centres of the circles lie. At
each grid point, several circles of different radii are trialed before moving on to the next grid
point (Figure 7.7).
Another method is to select points on the lines defining the top of the slope and allow
circles to pass through these points as shown in Figure 7.8. There are other search options
that have been developed to guide the search for the critical circle such as the algorithms by
Nguyen (1985) and Arai and Tagyo (1985), among others.

7.2.6  Water pressures


The factor of safety may be seen to depend on the water pressure in drained analyses, as
the water force can be seen in Equations 7.3 and 7.4 for example. The correct estimation
of the water pressure is very critical, as the factor of safety is fairly sensitive to the vales
used.
There are several ways that the water pressures can be specified and some are listed below:

1.
Using a phreatic surface
A free water surface may be specified to the computer program as a series of straight
lines, and then the distance from the free surface to the base of the slice (generally the
Slope stability  287

Contours of
minimum factor Centres of circles
of safety at each lie on grid points
centre may be
plotted

Different radii used


at each grid point

Circles tangent to firm


strata should be tested

Figure 7.7 Grid of centres used for locating the critical slip circle.

central point) calculated. The water pressure can be computed as the distance to the
base of the slice times the unit weight of water.
  Some computer programs allow several different phreatic surfaces to be input, and
water pressures in different materials to be computed from the free surface associated
with that material. This allows for perched water table calculations.
2.
Pore pressure ratio
For problems involving the stability of embankments such as earthfill dam embank-
ments, it is convenient to use the pore pressure ratio r u which gives the ratio of the
pore  water pressure to the total vertical stress at a point in the soil as defined in
Equation 7.5.

u
ru = (7.5)
γh

  The total unit weight of the soil is γ and the depth of the point beneath the surface
of the soil is h.

Circle centres lie along


this perpendicular

Selected points
on slope

Figure 7.8 Search method using points on the surface of a slope.


288  Geomechanics in soil, rock, and environmental engineering

  This allows for the fact that in earth fills, the water pressure generated is propor-
tional to the depth of the soil above a point in the fill. The fill may not be saturated,
therefore the value of r u will reflect this by being less than 1. Values of r u may be mea-
sured in triaxial tests using the appropriate total stress increases applied to the sample
that has been compacted at the appropriate moisture content.
3.
Pore pressure grid
For problems involving seepage (i.e. through embankments), the water pressures can
be found from a seepage analysis and then transferred to a grid of points. The grid of
points and the water pressure at each point can then be used with a slip circle analysis
by using interpolation between the grid point values. This is commonly done with
finite element analysis of seepage.
  Pore pressures from other analyses such as consolidation analysis can be transferred
in the same way.

7.2.7  Surface loads


If there are loads applied to the surface of a slope as may happen if there are traffic loads,
then the load can be added to the weight of the slice that it acts upon. The increased weight
is then used in the formula for the factor of safety such as Equations 7.3 or 7.4.
In total stress (undrained) analysis of slopes that are submerged, the water can be treated
as a force acting on the face of the slope and the water not considered as a separate material.
Water can also be considered (in an undrained analysis) by introducing it as a material that
has self-weight but no strength.

7.2.8  Computer programs


The summations shown in equations Equations 7.3 and 7.4, for example, can be carried out
by hand, but today there are many commercially available computer codes that enable the
calculations to be done rapidly with simple input of slope geometry and material properties.
Different search algorithms are available for locating the critical slip circle (i.e. the one with
the lowest factor of safety).
Generally, commercial programs allow different methods to be used such as the Bishop
method, Spencer’s method, or the full method. They also allow non-circular failure surfaces
to be used. Non-circular failure surfaces are discussed in Section 7.3.
Commonly used codes are SLOPE/W (2012), SVSLOPE (2014), and XSLOPE (2014).
Several other commercial codes are available on the Internet.

7.2.9  Three-dimensional failure surfaces


As was mentioned previously, failure surfaces are in reality bowl shaped and not two-
dimensional as assumed in the slip circle analyses mentioned in previous sections. The slip
circle method can be extended to the three-dimensional case where the slices become three-
dimensional columns. Early work on such analysis was reported by Chen and Chameau
(1983) and Leshchinsky et al. (1985). These papers conclude that the factor of safety that
is calculated using three-dimensional methods is generally higher than that from 2-D
methods.
As three-dimensional analysis requires a great deal of computation, it is necessary to use
a computer program to perform the analysis. Codes such as SVSLOPE/3D (2014) are able
to perform the analysis.
Slope stability  289

7.3  NON-CIRCULAR FAILURE SURFACES

If there is some reason for the shape of the slip surface to be non-circular, for example, there is
a weak plane along which failure can occur, then the non-circular analyses can be performed.
As for the circular surfaces, the region above the slip surface is divided up into strips. Trial
surfaces are used to locate the one with the lowest factor of safety, and generally, this is done
with input of the surface shape using graphical techniques (using a mouse) and a computer
program.
There are several methods that are used for non-circular surfaces, and these are examined
in the following sections.

7.3.1  Morgenstern–Price method


This method was described by Morgenstern and Price (1965). It satisfies all static equilibrium
requirements, but the solution obtained must be checked for physical admissibility. The prob-
lem is made determinate by assuming a relationship between the interslice shear force Xn and
the interface normal force En. The form of the relationship is given in Equation 7.6.

Xi = λf (x)Ei (7.6)

The function f(x) determines the angle of the interslice forces, while λ determines their mag-
nitudes. Various functions can be tried and there may be many functions that give admis-
sible solutions.
The line of thrust (which is the line showing where the interslice forces act) and the nor-
mal stress on the base of the slice are obtained as part of the solution. The line of thrust
should lie within the slice and the stress at the base of the slice should be compressive. As
well, the shear strength between the slices should not be exceeded.

7.3.2  Janbu’s method


In Janbu’s method, the line of thrust is assumed and then the equations of equilibrium are
solved. Janbu (1973) states that the factor of safety so calculated is relatively insensitive to
the location of the line of thrust as long as it is reasonable. The line of thrust (see Figure 7.9)

P Distributed load

External point loads


Xa
Q
Ea
Xn+1 a
Water table Crack
En+1 Xn En

Slice
Xb Tn N′n
Eb
U
b Line of thrust

Non-circular shear surface

Figure 7.9 Non-circular slip surface showing the line of thrust.


290  Geomechanics in soil, rock, and environmental engineering

should be at about one-third of the height of the slice from the base for cohesionless soils,
and it should be below this height in the active zone and above it in the passive zone for
cohesive soils.
Details of the method are given in Hirschfeld and Poulos (1972).

7.3.3  Sarma’s method


Sarma (1973) originally proposed a method of analysis whereby the stability of the slope
was measured by the horizontal acceleration required to bring the soil mass into limiting
equilibrium. If a conventional factor of safety was required, the solution could be found by
successive reduction of the shear strength parameters until zero acceleration was required
for equilibrium. The conventional factor of safety was the reduction of strength required.
The original method (1973) was based on vertical slices, and as the forces at the base of
the slice were assumed to act at the centre of the slice, the result depended on the number
of slices used. Sarma later presented a method (Sarma 1979) in which the factor of safety
could be calculated for slices that do not have vertical sides, and the slices can be as large as
possible as dictated by the geometry of the slip surface. Again, the method involves finding
the horizontal acceleration on the slope to produce equilibrium.

7.4  WEDGE ANALYSIS

Hand analysis of slope failures may be undertaken if the failure mode is a wedge type of
failure. This can occur where there is a weak plane due to a soft soil layer or a fracture or
fissure. This approach may also be used when analysing the upstream slope of a sloping core
rockfill dam (Sultan and Seed 1967, Seed and Sultan 1967).
A simple form of analysis is to assume that there are two wedges such as those shown in
Figure 7.10. The upper wedge is the active wedge that is pushing down on the lower passive
wedge that is sliding sideways.
The following assumptions are made (see Figure 7.10):

1. The critical failure surface bc passes through the slope.


2. On planes a−b, b−c the friction and cohesion are only partly mobilised, for example,
on b−c

tan φ1′
tan θ1 = (7.7)
F

c1′ 1
S1 =
F

3. The shear strength which is mobilised between the two wedges may not be mobilised
to the same degree as the shear strength on the base of the wedges. The angle of fric-
tion δ could therefore lie between the limits of 0° and

 tan φ′ 
δ = tan−1  (7.8)
 F 

while the inter-wedge cohesion could lie between values of zero and c′/F.
Slope stability  291

Wedge 1
a
d
Wedge 2 P12 W1 ϕ′1, c′1 apply
S21 on ab (= ℓ1)
W2 δ U12 U1
P21
U21 S1 θ1
b S12 P1 ϕ′3, c′3 apply
c S2 on bd (= ℓ3)
U2 Trial slip surface
ϕ′2, c′2 apply P2
θ2
on bc (= ℓ2)
U21
P12 tan θ1 = tan ϕ′1/F S21
S1 = c′1ℓ1/F P21
δ δ
S12 P1
tan θ2 = tan ϕ′2/F
U12 W1 1
S2 = c′2ℓ2/F θ1
P2
W2 θ2 S1

Force polygons U1
2

U2 S2

Figure 7.10 Two sliding wedges and force polygons for each wedge.

One approach, therefore, is to assume that the degree of mobilisation is midway between
the two extremes and to assume that both cohesive and frictional forces have this median
value throughout the analysis.
If δ is assumed to be zero, the calculated factor of safety will be ≈25% on the conserva-
tive side.
The procedure which may be used to find the factor of safety is as follows:

1. Assume two values of the factor of safety which are likely to span the true value, say
F 1, F 2 .
2. Determine for each block the mobilised angles of friction θ1, θ2 and the mobilised shear
forces S1, S 2 (see Equations 7.7) using the first factor of safety F 1.
3. Draw the force polygon for each block, as shown in Figure 7.10. The magnitudes of
forces P1, P2 , P12 , and P21 are unknown but their directions are known so that the
polygons may be completed.
4. Values of P12 , P21 may thus be found from the force diagram.
5. The correct solution is the one for which P12 = P21. If the two values are not equal,
another solution for P12 and P21 is then found using F 2 . Hence, the values of P may
be plotted against the factor of safety (say F). Results are as shown in Figure 7.11 and
linear interpolation may then be used to find the required value of F.
6. The positions of the lines a−b, b−c may be altered and corresponding factors of safety
found. The lowest value is chosen.

This kind of analysis can be rapidly performed using non-circular slice methods, and
therefore is not often performed by hand these days.
292  Geomechanics in soil, rock, and environmental engineering

P21
P12

P21

P12

F1 F2 F

Required value

Figure 7.11 Factor of safety plot.

7.5  PLASTICITY THEORY

Solutions to the stability of slopes have been found using rigorous plasticity theory (Booker and
Davis 1972). The lines along which slip is occurring may be seen in Figure 7.12 compared with
a slip circle solution for the case where the soil is undrained (ϕu = 0) and the undrained strength
varies with depth. In the plasticity solution, it may be seen that there is a wedge of material
below OA with sides at 45° to the horizontal, and a curved section below the slope OB.
Booker and Davis showed that the slip circle factor of safety was less than 10% in error over
a practical range of slope angles (angle from the horizontal >5°) for this particular problem.

7.6  UPPER- AND LOWER-BOUND SOLUTIONS

Solutions can be obtained to the slope stability problem by use of the upper- and lower-bound
theorems (Sloan 2013). The stress field is interpolated from values at the nodes of triangu-
lar elements using linear interpolation, and the failure criterion is represented as a series of

O A

Critical circle
from slip circle
analysis

Figure 7.12 V
 elocity field for a slope failure in undrained soil with strength increasing with depth.
Slope stability  293

Figure 7.13 Upper-bound solution for slope stability from program OPTUMG2 showing the location of the
failure surface.

linear functions. Optimisation (linear programming) is then used to obtain the stress field
that gives the lowest unit weight to cause failure, giving an upper bound.
The velocity field is then represented at the nodes of a triangular mesh (this can be the same
mesh as for the stress field) and the velocities are optimised to give the lowest power dissipation.
This gives the upper bound to the problem. The true solution then lies between the two bounds.
This type of solution does not depend on trial and error as do the limit equilibrium meth-
ods, so the location of the failure surface is found directly. If there are weak layers, then the
failure surface will be found to run along the weak layer as part of the analysis. Such a solution
is shown in Figure 7.13, which was found by using the computer code OPTUMG2 (2014).
Solutions may also be found (using the upper- and lower-bound approach) where water
pressures such as those caused by seepage in the slope are involved. Joints may also be
included in the analysis.

7.7  FINITE ELEMENT AND FINITE DIFFERENCE SOLUTIONS

The stability of slopes can be found by using conventional finite element or finite difference
(FLAC) solutions and a ‘strength reduction’ technique. This involves reducing the strength
of an elasto-plastic material until collapse occurs, and then the reduction that is applied
gives the factor of safety for the slope. The method has the advantage that different reduc-
tion factors can be applied to the cohesive strength and the frictional strength if desired.
This is not a rigorous method of finding the factor of safety as the result depends on the
mesh fineness, but it can give an indication of the failure mechanism as the failure surface is
indicated by the region of high plastic shear strain.
An example is given by Griffiths and Lane (1999) for a slope in a two-layered undrained
soil as shown in the upper diagram in Figure 7.14. The solutions show that there is a transi-
tion in the failure mechanism as the ratio of the shear strengths in the two layers changes as
can be seen in the lower plots of Figure 7.14.
294  Geomechanics in soil, rock, and environmental engineering

ϕu = 0
2H 2H

2
1
H 2H
cu1

cu2
H

(a)

(b)

(c)

Figure 7.14 Deformed meshes for a slope with two layers of material with different strengths: (a) su2/su1 = 0.6;
(b) su2/su1 = 1.5; (c) su2/su1 = 2.0.

Finite element methods are easily extended to three-dimensions (Griffiths and Marquez
2007) and probabilistic methods can be applied to the soil properties. Different soil strengths
can be applied to each element in the mesh according to statistical methods, and the cor-
responding factors of safety found for the slope.
An early example of the use of finite difference methods for slope stability is that of
Hammet, and this is reproduced in the paper by Cundall (1976).
Commercial programs such as PLAXIS (2014) and FLAC2D (ITASCA Consulting Group
2013) are capable of performing the strength reduction analyses. Stabilising features such as
soil nails can be incorporated into such analyses.

7.8  SEISMIC EFFECTS

The simplest way to incorporate seismic effects into a slope stability analysis is to add an extra
force vertically or horizontally to a slice to simulate the force due to ground acceleration.
Slope stability  295

ℓ1 b c
R

kW
Arc length
a W ac = L
ℓ2
su

Figure 7.15 Static force representing earthquake force.

The most common way to estimate the force is to make the force a multiple of the weight of
the slice as shown in Equation 7.9.

Fh = khWi
(7.9)
Fv = kvWi

The vertical or horizontal earthquake forces are multiples of the weight of the slice Wi where
the horizontal and vertical seismic coefficients are kh and kv, respectively. This is suitable if
the soil beneath or in the slope is not subject to liquefaction.
The horizontal force acting on a soil mass is shown in Figure 7.15 for a simple undrained
analysis. By taking moments about the centre of the slip circle, the factor of safety against a
rotational slide can be found if the undrained shear strength used is su /F.
Values of the seismic coefficient can be found for various regions from seismic maps. For
example, the Australian standard AS1170.4 gives acceleration coefficients for different areas
based on historical data. For Sydney, the coefficient is kh = 0.08. However, for seismically
active areas such as Japan or the west coast of the United States, the accelerations are much
higher. Seed (1979) recommended (for large slopes) kh = 0.1 for sites near faults capable of
generating magnitude 6.5 earthquakes, and the acceptable pseudo-static factor of safety is
1.15 or greater, while kh = 0.15 is appropriate for sites near faults capable of generating mag-
nitude 8.5 earthquakes (for an acceptable pseudo-static factor of safety of 1.15 or greater).
Full dynamic analysis can be performed by finite difference codes such as FLAC (ITASCA
Consulting Group 2014), and these can be used for slope stability analyses of dam slopes.
The onset of liquefaction in slopes of hydraulically filled dam walls such as tailings dams
can also be analysed by these methods.

7.9  FACTORS OF SAFETY

If slope stability analysis shows that the slope does not have the required factor of safety
against a slip, then remedial work may be necessary to improve the stability. Factors of
safety that are chosen depend upon the consequences of failure and need to be chosen to
reflect the risk. They also reflect how reliable the soil strength data is, and whether a mean
value or a lower-bound value has been selected for design.
Hence, factors of safety are largely selected by engineers according to the perceived risk.
However, some guidance is provided in the Geotechnical Manual for Slopes (Geotechnical
Control Office 2011) as shown in Table 7.1. These values were developed for use in Hong Kong.
296  Geomechanics in soil, rock, and environmental engineering

Table 7.1  Recommended factors of safety for new slopes for a 10-year
return period of rainfalla
Risk to life
Economic risk Negligible Low Highb
Negligible  > 1.0 1.2 1.4
Low  1.2 1.2 1.4
High  1.4 1.4 1.4
a The factors of safety in this table are recommended values. Higher or lower factors of
safety might be warranted in particular situations with respect to economic loss.
b In addition to a factor of safety of 1.4 for a 10-year return period of rainfall, a slope in
the high risk to life category should have a factor of safety of 1.1 for the predicted
worst groundwater conditions.

Table 7.2  Factors of safety for slopes of tailings dams


Slope Condition Required factor of safety
Upstream Rapid drawdown 1.2–1.4
Upstream Rapid drawdown with earthquake 1.0–1.2
Upstream End of construction 1.1–1.3
Downstream End of construction 1.1–1.3
Downstream Steady seepage 1.5–1.7
Downstream Steady seepage plus earthquake 1.3–1.5

D’Andrea and Sangray (1982) quote a number of sources that give values of the conven-
tional factor of safety lying between 1.25 and 1.5 depending on the particular conditions
for the slope.
For tailings dams, the CANMET (1977) manual recommends the factors of safety given
in Table 7.2. The factor chosen depends on the condition and should be checked for all of
the different conditions that are likely to affect the dam.

7.10  SLOPE STABILISATION TECHNIQUES

Precautions can be taken to mitigate slope instability especially in the long term. It may
be necessary also to modify a slope so as to increase the factor of safety that it has against
failure. Generally, removing sources of water infiltration into a slope or lowering the water
table are methods that can be applied to increase stability, but other methods can be used as
presented in the following sections.

7.10.1  Control of surface water


Collection of water from above the slope can be carried out with a surface drain constructed
at the top of the slope. Water collected by the drain must be channelled away from the slope
and not allowed to infiltrate into the soil. Therefore, the drain should be lined, either with
concrete or with other materials (e.g. plastic or steel half-pipe). Water caught in the drain
can be drained off to either side of the slope or down to the foot of the slope in pipes.
Joints in the drain could open if the slope moves after installation and this could be a
source of water influx into the slope.
Slope stability  297

7.10.2  Horizontal drains


Horizontal drains are drilled into the slope at a slight angle to the horizontal (between −5°
and +5°) so that water intercepted by the drain can flow out. The type of drain depends on
the soil or rock type. In fractured rock, it may be adequate to drill open holes, but in soil
slopes, drainage pipes are necessary.
Drains are normally 100–120 mm in diameter and can be lined with slotted PVC pipes
wrapped in a filter geotextile to prevent ingress of soil. In some cases, the pipe is not slotted
or is grouted near the exit to prevent blockage by tree roots and erosion near the exit.
Drains should be checked to determine if they are operative, and if not can be cleared with
high pressure water.

7.10.3  Stabilising piles


Stabilising piles have been used to resist slope movements, and can be effective in supplying
lateral force to retain the soil. Several methods have been developed for designing the piles,
including those of Ito and Matsui (1975), Ito et al. (1982), Viggiani (1981), and Poulos (1995).
The Ito and Matsui results should be used with caution, as the loads tend to approach
infinity when the pile spacing tends to zero (see Beer and Carpentier 1977).

7.10.4  Toe fill


In some circumstances, it is possible to remove the soil at the base of a slope, and to replace it
with rockfill. The effect of the rockfill is to provide a free draining material with a high angle
of shearing resistance that is much more stable than the original soil material on the slope.
This approach was used to stabilise the Hue Hue Road landslip (Fell et  al. 1987) that
occurred on the main highway leading north of Sydney. Features of the reconstruction were
the use of filter fabric between the natural soil and the rockfill, catch drains at the top of the
slope and a drainage trench at the base of the rockfill as shown in Figure 7.16.

7.10.5  Retaining structures


All types of retaining structures mentioned in Chapter 9 may be used to retain a slope.
Gravity walls, reinforced earth walls, crib-block walls, soil nails, and anchored walls can
be used.

L
C
Pavement Topsoil and vegetation
Catch drain
1 m ripped sandstone
Catch drain

L
C
Filter fabric 1.5
1 8m
Gravel layer
around filter fabric Rockfill
1 Roadway
1 2
1 in 1
0 2 4 6 8 10
Scale 13 m Drainage trench

Figure 7.16 Stabilisation of a slope with rockfill at the toe.


298  Geomechanics in soil, rock, and environmental engineering

If the slide is large in extent, walls cannot be used because of the large forces involved that
would make the wall too large to be economical. In this case, water control methods are a
more effective means of stabilising the slope.

7.11  STABILITY CHARTS

Stability charts for slopes have been developed by many authors, but one that is commonly
used is that of Hoek and Bray (1981). Although, it is a simple matter to use slip circle computer
programs for calculating Factors of Safety today, use of such charts is still quick and easy.

Groundwater flow conditions Chart number

Fully drained slope

Surface water 8 × slope


height behind toe of slope

Surface water 4 × slope


height behind toe of slope

Surface water 2 × slope


height behind toe of slope

Saturated slope subjected


to heavy surface recharge

Figure 7.17 Groundwater conditions: Cases 1–5.


Slope stability  299

2 1

c
tan ϕ γH tan ϕ
3
F

c
γHF

Figure 7.18 Method of using charts.

0
.01

.02

.03

.04

2.0
.05

.06

Circular failure chart number 1


.07
.08
.09

1.8
.10
.11
.12
.13

c/γH · tan ϕ
.14

1.6
.1 5
.1
.1 6
7

8
.1 9
1.4 .1 0
.2

1.2 .25
tanϕ/F

.30
1.0
90
.35
(°)
le .40
ng
0.8
p ea
.45
Slo .50
80
0.6 .60
70 .70
60 .80
0.4 .90
50 1.0
40
30 1.5
0.2 20 2.0
10 4.0
0 ∞
0.

0.

0.

0.

0.

0.

0.

0.

0.

0.

0.

0.

0.

0.

0.

0.

0.

0.
00

02

04

06

08

10

12

14

16

18

20

22

24

26

28

30

32

34

c/γHF

Figure 7.19 Stability charts for Cases 1–5. (Continued)


300  Geomechanics in soil, rock, and environmental engineering

.01

.02

.03

.04
2.0

.05

.06
Circular failure chart number 2

.07
.08
.09
1.8

.10
.11
.12
.13
c/γH · tan ϕ

.14
1.6

.15
.1 6
.1
7
8
.1 9
1.4 .1 0
.2

1.2 .25
tanϕ/F

.30
1.0 90
.35
(°)
le .40
0.8 ng
pea .45
Slo .50
80
0.6 .60
70 .70
60 .80
0.4 50 .90
40 1.0
30
20 1.5
0.2 10 2.0
4.0
0 ∞
0.

0.

0.

0.

0.

0.

0.

0.

0.

0.

0.

0.

0.

0.

0.

0.

0.

0.
00

02

04

06

08

10

12

14

16

18

20

22

24

26

28

30

32

34
c/γHF

0
.01

.02

.03

.04

2.0
.05

.06

Circular failure chart number 3


.07
.08
.09

1.8
.10
.11
.12
.13

c/γH · tan ϕ
.14

1.6
.15
.1 6
.1
7

8
.1 9
1.4 .1 0
.2

1.2 .25
tanϕ/F

90 .30
1.0
(°) .35
gle
0.8 e an .40
p
Slo .45
.50
80
0.6 .60
70
.70
60 .80
0.4 50 .90
40 1.0
30
20 1.5
0.2 2.0
4.0
0 ∞
0.

0.

0.

0.

0.

0.

0.

0.

0.

0.

0.

0.

0.

0.

0.

0.

0.

0.
00

02

04

06

08

10

12

14

16

18

20

22

24

26

28

30

32

34

c/γHF

Figure 7.19 (Continued ) Stability charts for Cases 1–5.


Slope stability  301

.01

.02

.03

.04
2.0

.05
.06
Circular failure chart number 4

.07
.08
.09
1.8

.10
.11
.12
.13
1.6 c/γH · tan ϕ

.14
.1 5
.1
.1 6
7
8
.1 9
1.4 .1 0
.2

1.2 .25
90
tanϕ/F

.30
1.0 (°)
gle
e an .35
p
0.8 Slo .40
.45
80 .50
0.6 .60
70
60 .70
50 .80
0.4 40 .90
1.0
1.5
0.2
2.0
4.0
0 ∞
0.

0.

0.

0.
0.

0.

0.

0.

0.

0.

0.

0.

0.

0.

0.

0.

0.

0.
00

02

20

22

24

26

28

30

32

34
04

06

08

10

12

14

16

18

c/γHF
0
.01

.02

.03

.04

2.0
.05
.06

Circular failure chart number 5


.07
.08
.09

1.8
.10
.11
.12
.13

c/γH · tan ϕ
.14

1.6
.15
.1 6
.1
7

8
.1 9
1.4 .1 0
.2

1.2 .25

(°)
tanϕ/F

.30
1.0
n gle
p ea .35
Slo .40
0.8
80 .45
.50
0.6 70
.60
60
50 .70
.80
0.4 40 .90
30 1.0
20 1.5
0.2 2.0
10
4.0
0 ∞
0.
0.

0.

0.

0.
0.

0.

0.

0.

0.

0.

0.

0.

0.

0.

0.

0.

0.

30
24

26

28

32

34
00

02

04

06

08

10

12

14

16

18

20

22

c/γHF

Figure 7.19 (Continued ) Stability charts for Cases 1–5.


302  Geomechanics in soil, rock, and environmental engineering

A slope with different water conditions ranging from a dry slope (Case 1) to a fully satu-
rated slope (Case 5) is shown in Figure 7.17. A choice is made of the condition closest to that
existing in the field and then the appropriate circular failure chart selected.
The method of using the charts is shown in Figure 7.18. Firstly, a value of c/γH tanϕ is
calculated, and then this value is found on the outside circular part of the chart. Then by
moving down the radius of the circular chart, the slope angle of the slope is reached. The
value of either tan ϕ/F or c/γHF can be read from the chart, and either can be used to cal-
culate the factor of safety F of the slope. (The unit weight of material in the slope is γ, the
height of the slope H and the cohesion and angle of shearing resistance of the slope material
c and ϕ, respectively.)
The circular charts for each of the five groundwater cases are shown in Figure 7.19.
Other specialist charts that may be useful are those of Morgenstern (1967) for rapid
drawdown (such as occurs if water in a dam drops); Bishop and Morgenstern (1960); and
for three-dimensional surfaces, Cheng and Yip (2007).
Chapter 8

Excavation

8.1 EXCAVATION

Excavations may be unsupported or may require some kind of supporting structure that is
designed to prevent the sides of the excavation from collapse. The base of an excavation may
also fail through heave or piping and so must also be designed to resist failure.
If an excavation does not require support, the stability can be checked with slip circle
techniques (see Chapter 7). If it does require support, the supporting structure can range
from simple block work or brick retaining walls to concrete or steel pile walls held in place
by anchors or props.
Support systems must be designed to reduce the load from groundwater, as this can be con-
siderable. Walls are generally provided with weep holes (that need to be periodically checked
to see that they are working) or drains that can channel any groundwater away from the back
of the wall. If groundwater cannot be drained away from the wall (as may be the case for
some basements), then the wall must be designed to withstand the water pressures.

8.2  TYPES OF EXCAVATION SUPPORT

When vertical cuts are made in soils, it is often necessary to brace or support the sides of the
cut in some way so that collapse of the sides does not occur. Several techniques can be used
for bracing the sides of cuts as described in the following sections.

8.2.1  Steel ‘H’ piles and lagging


For shallow cuts less than about 4 m in depth, vertical wooden planks can be used. The
planks are braced by struts that run from one side of the cut to the other and bear onto
wales that run along the face of the planks. Alternatively, prefabricated steel panels with
steel bracing can be lowered into excavations or trenches that are not very wide (Figure 8.1).
For deeper cuts, sheet piles or ‘H’ piles with timber lagging that slots in between the piles
may be used (see Figure 8.2). The ‘H’ section piles are driven first and the lagging is added
as the excavation proceeds. Anchors may be needed or bracing may be used to increase
stability if required. Anchors passing through horizontal cross-beams or wales are shown
supporting a wall in Figure 8.3.

8.2.2  Sheet piles


Sheet pile walls are also a common means of support for retaining walls. The steel sections
are driven or vibrated into the ground before excavation, each section interlocking with the

303
304  Geomechanics in soil, rock, and environmental engineering

Figure 8.1 Prefabricated wall and bracing system.

Figure 8.2 Wall constructed in sand using ‘H’ piles with timber lagging.
Excavation 305

Figure 8.3 Excavation supported by anchors.

adjacent pile. The soil is then excavated leaving the sheet piles supporting the sides of the
excavation. If additional support is required, anchors or bracing can be used as for ‘H’ pile
walls.

8.2.3  Bored pile walls


Other construction techniques involve soldier pile, secant pile, or contiguous pile walls,
where piles are constructed in a row. Piles are drilled and concreted before excavation and
after excavation form the wall of the excavation. The different methods of construction are

1.
Secant piles: In this case, the piles actually overlap each other. First, a series of ‘soft’
piles (made from a weak concrete) are drilled with a space between them and then
a hole is drilled between two of the soft piles cutting into each of them. This pile is
306  Geomechanics in soil, rock, and environmental engineering

Overlapping hard piles

Soft piles constructed first

Figure 8.4 A secant pile wall.

concreted with a stronger mix and is therefore called a ‘hard’ pile. The result is a wall
consisting of overlapping piles that is reasonably watertight (see Figure 8.4).
2.
Soldier pile or king pile walls: For these walls, the piles are drilled and concreted at
some distance apart, and as excavation proceeds the soil exposed between the piles
can be left unfaced or can be shotcreted to prevent weathering.
3.
Contiguous piles: Contiguous pile walls are constructed with the piles touching or
nearly touching each other but not overlapping as for a secant pile wall. There are
some small gaps left between piles, as it is difficult to drill precisely vertical shafts.

With all of these types of walls, it is possible to provide further support through anchors
or bracing.

8.2.4  Diaphragm walls


Diaphragm or slurry walls are constructed by excavating a trench generally filled with benton-
ite slurry. Bentonite is a clay that has a high liquid limit and forms a thick slurry when mixed
with water. The slurry is able to support the sides of the trench and prevent collapse before
concreting can take place. Two concrete guide walls are constructed at the top of the trench to
guide the clamshell bucket that is used to excavate the soil. This helps keep the trench vertical
(see Figure 8.5a,b). The bentonite slurry is circulated through centrifuges that can separate the
heavier soil contaminants such as sand from the bentonite so that it can be used again.
The wall is constructed in sections with a vertical steel stop placed at the end of the new
section so as to form up the end of the section. A plastic water stop (Figure 8.6b) can be
placed between the previous section and the new section so that the joint is watertight.
Concrete is tremied into the base of the wall section using a long pipe (Figure 8.6a), displac-
ing the bentonite slurry upwards. Once the concrete has hardened, the excavation can take
place exposing the concrete wall.
Anchors can be added as excavation proceeds, and concrete slabs can be keyed into the
exposed concrete wall.

8.3  STABILITY OF EXCAVATIONS

There are a number of ways that failure of the support systems for an excavation can occur,
and some of the failure modes are specific to the type of supporting structure. Some of these
modes of failure are shown in Figure 8.7.

1. For gravity type walls made from concrete or brickwork, the wall can overturn due to
excessive earth and water pressures at the back of the wall. Excessive bearing pressure
may occur at the front of the wall (the toe) and bearing failure may take place. Sliding
Excavation 307

(a) (b)

Figure 8.5 (a) Slurry in trench excavated between two guide walls. (b) Clamshell bucket used for excavating
a trench.

(a) (b)

Figure 8.6 (a) Tremie being used to place concrete. (b) Plastic water stop.
308  Geomechanics in soil, rock, and environmental engineering

(a) Overturning (b) Sliding (c) Overstress

Overstress

Toe
Check moment equilibrium Check force equilibrium Check bending and shear stresses

(d) General stability (e) Granular soil, associated with


excessive upward seepage

Carry out overall (slope) stability analysis

(f ) Failure of anchor system (g) Bottom of piles move outward (h) Settlement behind the wall
(passive resistance not sufficient)

(i) Bottom heave (j) Buckled struts (k) Overstress of foundations

Upward and inward


movement of soil

Figure 8.7 Failure modes for support structures.

forward of the wall can be another failure mode. Overall stability due to a slip failure
beneath the wall can be checked using a slip circle analysis. All of these modes can be
checked simply by hand calculation.
2. For sheet pile walls and diaphragm walls, the moments and shears in the wall at vari-
ous depths need to be calculated and checked against allowable values. The capacity of
the anchors and an outward failure of the toe of the wall due to a passive failure also
need to be checked. Settlement behind the wall may be of importance and so may also
need to be estimated.
Excavation 309

3. Bottom heave can take place due to soil squeezing into the excavation under the action
of the weight of soil outside. As well, if the water table is kept down within the excava-
tion, the flow of water up through the base of the excavation can wash away the soil if
the hydraulic gradient is high enough, and this can lead to failure.

Design of retaining walls can be carried out using computer programs. Although stability
of a wall can be calculated by hand, soil layering and soil deflections may require the use of
a program such as WALLAP (2013) (that uses finite element methods for calculating deflec-
tions and limit equilibrium methods for calculating stability) or a finite element program
such as PLAXIS or Phase2.
The modelling of excavation using finite element methods is discussed in Section 8.8, but it
is necessary to decide whether the excavation is to be modelled treating the soil as a one-phase
material (i.e. not treating the groundwater separately) or as a two-phase material (where pore
water and the soil skeleton are considered in the model). If the soil is treated as a two-phase
material, then the drawdown of the water table as excavation proceeds becomes an issue.

8.4  BASE HEAVE FOR CUTS IN CLAY

One way in which an excavation can fail is through heave of the base. This is most likely to
occur in soft clays, but can occur in other types of soils.
Analysis of the base heave problem may be carried out for simple two-dimensional cases
by using the method proposed by Terzaghi (1943).

8.4.1  Shallow excavations (H/B < 1)


In this case, we can assume that the failure surface reaches the ground surface. The assumed
failure mechanism depends on whether the excavation has a firmer stratum below the base
of the excavation or not. The two cases are shown in Figure 8.8; case (A) is where the exca-
vation is quite narrow and there is no layer of stiff material close to the base of the excava-
tion. Case (B) is where the excavation is wide and has a stiffer stratum close to the base.

Case A
In this case, the load per metre run due to the soil block (on a–b) is 0.7BHγ − Hsu and the
bearing capacity per metre run will be suNc0.7B, therefore the factor of safety F is given by

s u N c 0 .7 B
F = (8.1)
0.7 BH γ − Hsu

(a) B (b)

0.7B D
H su H
45°
a b
45° D
0.7B

1 su Nc
Figure 8.8 Failure modes for long shallow excavations in clay. (a) F = ⋅ narrow excavation with
H γ − su / 0.7B
1 sN
no stiff stratum close to base. (b) F = ⋅ u c wide excavation with stiff stratum close to base.
H γ − su /D
310  Geomechanics in soil, rock, and environmental engineering

where B is the width of the excavation, H is the depth, Nc is the bearing capacity factor, and
su is the undrained shear strength of the clay that is assumed to be uniform.

Case B
For the wide excavation, the load due to the soil block is DHγ − Hsu and the bearing capac-
ity is suNcD, therefore the factor of safety F is given by
su N c
F = (8.2)
H(γ − su /D)

8.4.2  Deep excavations (H/B > 1)


For a deep excavation, the failure surface will not reach the surface as shown in Figure 8.9.
For a long excavation, the factor of safety F against a base heave failure is given by Equation
8.3 (Bjerrum and Eide 1956).
su N c
F = (8.3)
γH

The bearing capacity factor Nc in the formula needs to be found for the appropriate value
of H/B. For a long excavation, Figure 8.10 can be used where L/B = ∞.

8.4.3  Excavations of rectangular shape in plans


For cuts that are of width B and length L in plan, Bjerrum and Eide’s formula (Equation 8.3)
may still be used, although the bearing capacity factor in the formula needs to be found for
the appropriate value of L/B. The bearing capacity factors are given in Figure 8.10.

c, γ H

c, γ

Figure 8.9 Failure mechanism for a deep excavation.


Excavation 311

9
L/B = 1
8
2
3
7

Nc

4
0 1 2 3 4 5
H/B

Figure 8.10 Bearing capacity factors for rectangular excavations.

8.4.4  Base failure in sands


With braced excavations in sands, the danger of base failure usually occurs when the water
level inside the excavation is lowered so that an upward flow of water can take place. Piping
of the sand can occur if the hydraulic gradient approaches a value of unity.
The hydraulic gradient at the base of the excavation can be estimated from a flow net such
as the one shown in Figure 8.11. The exit hydraulic gradient iexit is given by

∆H
iexit = (8.4)
a

Water level

h Water
level

1
a 8
7 2
6
5 3
4

Impervious layer

Figure 8.11 Flow net used for estimating the factor of safety against piping failure.
312  Geomechanics in soil, rock, and environmental engineering

where a is the smallest distance between the equipotential lines (in this case near the wall),
and ΔH is the drop in total head between the equipotential lines. In Figure 8.11, the head
drop between equipotential lines is given by

h h
∆H = = (8.5)
Nd 8

where h is the difference in the water levels inside and outside the excavation, and Nd is the
number of head drops (here there are 8). A factor of safety against a piping failure may be
defined as shown in Equation 8.6.

icr 1
Fpiping = ≈ (8.6)
iexit iexit

A factor of safety of 1.5 or more would generally be required to guard against a piping
failure.

8.5  GROUND SETTLEMENT CAUSED BY EXCAVATION

Lateral movement of walls used to brace excavations results in vertical movement of the
ground surface that is called ground loss. There is always some movement of retaining walls
before the bracing can be applied and therefore some ground loss will inevitably occur.
Peck (1969) provided some information on expected ground movements, and these are
shown in Figure 8.12. The amount of movement expected depends on the type of soil
encountered, but is largest in very soft to soft clays as may be expected. The magnitude of
the vertical deformation δv to depth of the excavation H depends on the distance from the
edge of the excavation as can be seen from Figure 8.12. The settlements can be divided up
into three regions called I, II, and III that give the maximum settlement envelopes for differ-
ent soil types as are shown in Figure 8.12.
Tomlinson (1995) has presented some data collected for different excavations in soils of
different types. He comments that the amount of horizontal movement that occurs is not
dependent on the type of wall and bracing system, and that there is little difference in the
movement of diaphragm walls and sheet pile walls. Plots of measured maximum horizontal
movements divided by excavation depth are shown in Figure 8.13 for (a) soft-to-firm nor-
mally consolidated clays; (b) stiff-to-hard overconsolidated clays; and (c) sands and gravels.
Average ratios of maximum inward movement to depth are (a) 0.30% for soft-to-firm NC
clays; (b) 0.16% for stiff-to-hard OC clays; and (c) 0.18% for sands and gravels.
Plots of vertical surface movements versus distance from the edge of the excavation have
also been presented by Tomlinson and are shown in Figure 8.14. The curves are for differ-
ent case studies in the same soil type, and show the range of movement that may occur. Also
on the figure is Peck’s curve for very soft to soft clay, where it may be noted that unlike the
other vertical settlement profiles, the curve reaches a maximum at the edge of the excavation.
The other measured profiles reach a maximum away from the edge of the excavation. This
is because of the upward movement of the soil as it is unloaded as soil is excavated, and is
observed in both field cases and in finite element simulations of excavation. The older curves
from Peck do not reflect this smaller vertical deflection near the edge of the excavation.
Clough and O’Rourke (1990) have also presented data for the vertical surface settle-
ment δv with distance d from a braced excavation. Figure 8.15 shows the vertical surface
Excavation 313

3
I – Sand and soft clay and average
workmanship

II – Very soft to soft clay. Limited in


depth below base of excavation

III – Very soft to soft clay. Great depth


2 below excavation

(%)
δv
H

1
III

II

0
0 1 2 3 4
Distance from the braced wall
H

Figure 8.12 Variation of vertical surface movement with distance from edge of excavation. (Adapted from
Peck, R. 1969. Proceedings of the 7th International Conference on Soil Mechanics and Foundation
Engineering, Mexico City, State-of-the-Art Volume, pp. 225–290.)

settlement for excavations in sand. The data measured falls within a triangular region if the
ratio of settlement to maximum settlement δv/δvm is plotted against the ratio of distance from
the excavation to maximum excavation depth d/H as shown in the inset to Figure 8.15.
For soft to medium clays, the scatter of data for surface settlement versus distance from
the excavation is shown in Figure 8.16. Also shown are Peck’s regions I, II, and III (see Figure
8.12). Again, if the settlement to maximum settlement is plotted, a plot can be made that
encloses the range of data, although unlike the plot for sand this time the plot has a trapezoi-
dal shape.
The settlement to maximum settlement δv /δvm has a limiting value of about 1.0 over a
range of d/H of about 0.7 as can be seen from the inset to Figure 8.16.
Mana and Clough (1981) have also examined the relationship between the maximum
horizontal movement δH(max) of the walls of a braced excavation and the maximum surface
settlements δV(max). They have found that the vertical movement is about 0.5–1.0 times the
horizontal movement as shown in Figure 8.17, that is,

δV (max) = 0.5 to 1.0 δ H (max) (8.7)

8.5.1  Effect of shape of excavation


More recent data has shown that the shape of the excavation and the size of the excavation
can have an influence on the deflections and forces that are observed (Tan and Wang 2013a,b).
For instance, if the excavation is cylindrical, the walls retaining the excavation can behave
like an arch. However, when the diameter of the cylinder becomes large, the effect is less
pronounced as may be expected (Tan and Wang 2013a).
314  Geomechanics in soil, rock, and environmental engineering

(a) (b)
30 30

Average 0.30%
Average
Excavation depth (m)

20 20 0.16%

10 10
#
Soft-firm NC clays Stiff-hard OC clays

0 0
0 0.2 0.4 0.6 0.8 0 0.2 0.4
Max. inward movement/depth (%)
(c)
20 Key
Diaphragm wall, anchored
Diaphragm wall, or secant pile
Excavation depth (m)

wall, strutted
Average 0.18%
Sheet pile, soldier pile with
10 concrete infill, strutted
Sheet pile, soldier pile with
concrete infill, timber, anchored
Sand and gravels
# Excluded from average
0
0 0.2 0.4 0.6 0.8
Max. inward movement/depth (%)

Figure 8.13 Observed maximum inward movement of braced excavations. (After Tomlinson, M.J. 1995.
Foundation Design and Construction, 6th ed. Longman Scientific and Technical, Harlow, UK.)

Large excavations may result in more extensive vertical ground settlements that extend
further from the excavation than for smaller (in plan) excavations. This is shown in Figure
8.18 for the soft Shanghai clays (Tan and Wang 2013b) in a plot similar to that of Clough
and O’Rourke (see Figure 8.16).
In Figure 8.18, it can be seen that the data for the Shanghai World Finance Centre (SWFC)
Annex excavation (that is a large excavation of about 30,000 m 2 in plan) extend further
than excavations for the Metro excavations that are for narrow rail corridors. Building
basement excavations that are smaller and the Metro data can be seen to extend to about a
d/H value of 2 as predicted by Clough and O’Rourke, but the larger SWFC excavation the
deflections extend to a d/H value of about 3.7.

8.6  FORCES ON BRACED EXCAVATIONS

Unlike ordinary retaining walls that can move away from the backfill and develop active
pressures in the soil, braced walls cannot move easily and therefore different pressure distri-
butions are developed on such walls.
Excavation 315

Distance from face of excavation


Depth of excavation
0 1 2 3 4 5
0
10
0.2
11
0.4
× 100

0.6

0.8
Excavation depth

1.0
Settlement

Peck’s (1969) curve for


1.2 8
2-level basements
1.4 7

1.6
Soft-to-firm normally
1.8 consolidated clay
7 = Diaphragm wall, soft-to-
2.0 firm N/C clay, strutted
Distance from face of excavation 9 = Sheet pile, soft-to-firm N/C
Depth of excavation clay, strutted
0 1 2 3 4 5 11 = Diaphragm wall, soft-to-
0 firm N/C clay, strutted
× 100

14
0.2 16 14 = Diaphragm wall, stiff-to-
hard O/C clay, strutted
Excavation depth

0.4 16 = Diaphragm wall, stiff-to-


Settlement

hard O/C clay, anchored


0.6
Stiff-to-hard 28 = Bored piles, sands and
0.8 overconsolidated clay gravels, strutted
32 = Diaphragm wall, sandy
1.0 decomposed rock, strutted
Distance from face of excavation
Depth of excavation
0 1 2 3 4 5
0
× 100

28
28
Excavation depth

0.10
Settlement

32

Sands and gravels

0.20

Figure 8.14 Maximum vertical movement of surface versus distance from edge of excavation. (After Tomlinson,
M.J. 1995. Foundation Design and Construction, 6th ed. Longman Scientific and Technical, Harlow, UK.)

Peck et al. (1974) have presented some empirically developed pressure envelopes that are
based on load measurements taken in struts. The apparent pressure envelope can then be
used to compute the forces in a strut at any given level. This pressure envelope is not the true
pressure distribution, but a device to allow computation of the forces in the bracing.
The forces in the bracing can be computed by finding the area under the pressure envelope
for each strut. This is done by assuming that each strut carries the pressure applied to the
316  Geomechanics in soil, rock, and environmental engineering

Distance from excavation d


=
Max. excavation depth H

0 0.5 1.0 1.5 2.0 2.5 3.0


0
(%)

0.1
δv
H

Legend
=

0.2 Hatfield
Max. excavation depth

Bershamra
Settlement

0.3
7th and G Sts.
G St. Test Site
Distance from excavation d 8th and G St.
=
Max. excavation depth H OCC Bldg.
δvm
δv

0 1 2 Charter station
0
=
Max. settlement

0.5
Settlement

Triangular bounds
1.0 on distribution of
settlement

Figure 8.15 Measured settlements adjacent to excavations in sand. (After Clough, G.W. and O’Rourke, T.D.
1990. Design and Performance of Earth Retaining Walls, ASCE, Geotechnical Special Publication
No. 25, pp. 439–470.)

wall over the region going half way to the next strut as shown in Figure 8.19. If no strut is
placed at the base of the excavation, it is assumed that part of the load is taken by the soil
at the base of the excavation.
For cuts made in sand, the apparent pressure envelope may be considered to be of constant
magnitude with depth and have a value of 0.65γHtan 2(45° − ϕ). It should be noticed that this
value only applies to dry or moist sands (see Figure 8.20).
For excavations made in clay, the pressure envelope depends on the parameter γH/su where
H is the depth of the cut. If this value is less than 4, the envelope shown in Figure 8.17c
should be used. In this case, the average magnitude of the apparent pressure envelope is
about 0.3γH. If the ratio exceeds 4, the pressure envelope of Figure 8.20d should be used
provided the envelope is greater than that in Figure 8.17c, otherwise the value of pressure
from (c) is used. The value of the undrained shear strength su is taken as the average value
over the depth of the cut.

8.7  STABILITY OF SLURRY-FILLED TRENCHES

As mentioned previously (Section 8.2.4), diaphragm walls can be constructed by excavating


under a bentonite slurry that is thick (dense) enough to prevent the sides of the excavation
from caving in. The density of the slurry needs to be high enough to prevent the sides of the
trench from failing, but not so dense as to prevent excavation of soil by the clamshell grab.
Simple analytic techniques can be applied to estimate if a trench is likely to fail. The
analysis is based on the assumption that the slurry supplies a hydrostatic force to the face
of the excavation. This is a reasonable assumption as in most cases the slurry tends to cake
against the sides of the excavation forming a water resistant seal and reducing water flow
into the surrounding soil.
Excavation 317

Distance from excavation d


=
Max. excavation depth H

0 0.5 1.0 1.5 2.0 2.5 3.0


0

I
(%)
δv
H

1 II
=
Max. excavation depth

Legend Chicago Transit


Vaterland 1 Grant Park
Settlement

Vaterland 2 Social Security


III
2 Vaterland 3 Harris Bank
Gronland 1 Northern Trust
Gronland 2 Water Tower
Zones I, II, III after Olav Kyrres Embarcadero III
Peck (1969)
3

0 0.5 1.0 1.5 2.0 2.5 3.0


0
δvm
δv
=
Max. settlement

0.5
Settlement

1.0

Max. settlement Transition zone

Figure 8.16 Measured settlements adjacent to excavations in soft-to-medium clay. (After Clough, G.W.
and O’Rourke, T.D. 1990. Design and Performance of Earth Retaining Walls, ASCE, Geotechnical
Special Publication No. 25, pp. 439–470.)

8.7.1  Wedge analysis


Shown in Figure 8.21a is the face of a slurry-filled trench, with a presumed wedge type
failure occurring. If such a failure were to occur, then the force polygon can be drawn for
the wedge assuming that it is in equilibrium, and the factor of safety (FoS) against a sliding
failure can be computed for the wedge.

8.7.2  Purely cohesive soil


For a purely cohesive soil such as an undrained clay, the force diagram shown in Figure
8.21b is applicable. In this case, we can compute the magnitude of the mobilised cohesive
force C m since the vector b − c is equal to Pf and therefore vector a − b is given by (W − Pf).

(W − Pf )
Cm = (8.8)
2
318  Geomechanics in soil, rock, and environmental engineering

San Francisco
Oslo
Chicago

δv (max) = δH (max)

2
(%)
δv (max)
H

δv (max) = 0.5δH (max)

0
0 1 2 3
δH (max)
(%)
H

Figure 8.17 Relationship between maximum lateral and maximum vertical ground movement. (After Mana,
A.I. and Clough, G.W. 1981. Journal of the Geotechnical Engineering Division, ASCE, Vol. 107, No.
GT6, pp. 759–777.)

Normalised distance behind retaining wall, d/He


0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0
0

0.2

0.4
δv/δvm

0.6

0.8

1.0
Clough and O’Rourke (1990)
Max. settlement Transition zone
Tan and Wang (2013b)
Max settlement Transition zone
SWFC-Annex Clough and O’Rourke (1990)
Building basement Hashash et al. (2008)
Metro excavations Hsieh and Ou (1998)
Tan and Wang (2013b)

Figure 8.18 Vertical settlement adjacent to excavations of different sizes. (After Tan, Y. and Wang, D. 2013b.
Journal of the Geotechnical and Geoenvironmental Engineering Division, ASCE, Vol. 139, No. 11, pp.
1894–1910.)
Excavation 319

d1 Strut
d2/2
d2
Strut d2/2 Apparent
pressure
d3/2 envelope
H d3
d3/2
Strut
d4/2
d4
d4/2

Soil reaction

Figure 8.19 Method of calculating strut loads from apparent pressure diagram. (Adapted from Peck, R.B.,
Hanson, W.E., and Thornburn, T.H. 1974. Foundation Engineering, 2nd ed. Wiley, New York.)

(a) (b) (c) (d)

γH Clay γH
Sand su ≤ 4 su > 4

0.25H 0.25H

H 0.50H
0.75H

0.25H

0.2γH
to γH – 4su
0.65 γH tan2(45 − φ ⁄ 2) 0.4γ

Figure 8.20 Apparent pressure diagram for calculating loads in struts of braced cuts. (a) Wall of height
H. (b) Dry or moist sand. (c) Clays if γH/su ≤4. (d) Clays if γH/su 4 provided that γH/sub does
not exceed about 4 (where sub is the undrained strength of the clay below excavation level).
(Adapted from Peck, R.B., Hanson, W.E., and Thornburn, T.H. 1974. Foundation Engineering, 2nd
ed. Wiley, New York.)

The mobilised cohesion is therefore

Cm (γH 2 − γ f H 2 )/ 2 2
cm = = (8.9)
2H 2H
where γ is the unit weight of the soil, γf is the unit weight of the slurry, and H is the height of
the trench. The factor of safety F against failure of the wedge may be defined as
su 4s u
F = = (8.10)
cm (γ − γ f )H

where su is the undrained shear strength of the soil. It may be seen from Equation 8.10, that
the closer the unit weight of the slurry is to the unit weight of the soil, the higher the factor
of safety.
320  Geomechanics in soil, rock, and environmental engineering

(a) (b) (c)

Slurry Ground
a
level level C
A C
45°

b
R
C
H R
Pf
R W
α 90° – (θ−α )
H/3 θ 45°
c
B Pf Pf

Figure 8.21 Assumed failure wedge (a) and force vector diagrams for (b) cohesive soil, and (c) cohesionless
soil.

8.7.3  Cohesionless soil


Cohesionless soils such as sands can also be supported by slurry such as bentonite. With
reference to the vector diagram of Figure 8.21a, we can see that the resultant force on the
plane of slip is inclined at the mobilised angle of friction α. From the vector diagram of
Figure 8.21c, we have

W π 
= tan  − θ + α  (8.11)
Pf  2 

The weight of the wedge is given by

π 
W = 0.5γH 2 tan  − θ (8.12)
2 

and the fluid force is given by

Pf = 0.5γ f H 2 (8.13)

therefore Equation 8.11 becomes on substitution for these forces

γ tan(π / 2 − θ) π 
= tan  − θ + α  (8.14)
γf 2 

It can be shown that the most critical plane occurs when θ = π/2 + α/2, so under these
conditions Equation 8.14 becomes

γ tan(π / 4 + α / 2) π 
= = tan2  − θ + α  (8.15)
γf tan(π / 4 + α / 2)  2 

Excavation 321

Therefore,

γ − γf
tan(α) = (8.16)
2 γ − γf

The factor of safety F against collapse can be defined as

tan φ′ 2 γ .γ f tan φ′
F = = (8.17)
tan α γ − γf

If the free water level is near the ground surface, it may be difficult to maintain trench
stability. It may be necessary to

1. Lower the water table in the vicinity of the excavation


2. Raise the level of slurry in the trench, or
3. Use a denser slurry

The factor of safety against a failure of the slurry-filled trench in this case is given by

2 γ ′.γ ′f tan φ′
F = (8.18)
γ ′ − γ ′f

where the submerged unit weights of soil and slurry γ ′, γ ′f are now used in the equation.

8.8  NUMERICAL ANALYSIS

Many different means of analysing excavations supported by anchors or struts are available,
and are today routinely used to design support systems. One such program WALLAP (2013)
uses a combination of limit equilibrium methods to assess stability and finite element or
subgrade reaction methods to estimate wall deflections (see Figure 8.22).
The program allows for different layers of soil (horizontally layered), surcharges at the
back of the wall, struts or anchors, analysis of soldier or king pile walls, water table differ-
ences on either side of the wall, and earthquake forces. Excavation can be applied in steps to
allow calculation of wall deflections.

8.8.1  Finite element analysis


The method of computing the forces that are needed to simulate excavation using finite ele-
ment methods was first published by Brown and Booker (1985). In an elastic medium, the
forces due to excavation of some of the finite elements in the mesh may be calculated by
noting that the finite element equations say that the stiffness matrix multiplied by the deflec-
tions are equal to the body forces and the applied tractions. The definitions of terms in the
following equations are explained in Chapter 3.

K δ = fb + ft (8.19)

322  Geomechanics in soil, rock, and environmental engineering

Stage No. 1 Excav. to elev. 0.00 on passive side


1.70

Fill

0.00
Fill Fill
–0.40
Post Post –0.50
Concrete Concrete
–0.90 –0.90

Fill Fill

–2.00 –2.00
10 5 0 5 10
Water pressure (kN/m2)

Figure 8.22 Cross section of retaining structure analysed by computer program WALLAP.

Expanding the above equation leads to

∫B DBδdV = ∫N
T T

γdV + N T tdS (8.20)
V V S

or in terms of the stresses

∫B σdV = ∫N
T T

γdV + N T tdS (8.21)
V V S

Hence, for equilibrium of the finite element mesh, we must have

∫B σdV − ∫N
T T

γdV − N T tdS = fnodes (8.22)
V V S

Generally, the integrations of Equation 8.22 will yield zero forces fnodes as the mesh is
in equilibrium, but if some of the elements are removed (as in excavation) the forces will
not be in equilibrium at the nodes around the sides of an excavation, therefore these forces
are the ones that need to be applied to the finite element mesh to simulate the excavation.
Excavation 323

Forces will also be calculated at the location of fixed nodes (where boundary conditions are
applied) and these forces must be reset to zero in the calculation.
If the material is a linear elastic material, then Brown and Booker (1985) show that the
sequence of excavation does not change the final result, that is, if elements are removed from
the whole excavation in one step or they are removed from the excavation in stages.
At the beginning of an excavation problem, it is necessary to set up the initial stresses in
the ground due to the weight of overlying soil layers (and perhaps other causes). The lateral
stresses σh can be set up through a knowledge of the stress regime in the ground. The lateral
stresses may be due to the coefficient of earth pressure at rest K0 and sometimes locked in
geostatic stresses σl due to earth movements and compression of the ground laterally. In
general, the formula

σ h = K0σ v + σ l (8.23)

may be used, although more complicated lateral stress variations can be used if it is known
that they exist. Different lateral stresses may exist in the two lateral axis directions for
example, or surface loads may exist.
It is very important to get the stress regime correct as the forces that drive the excavation
modelling are due to the initial stresses in the ground. If these are incorrect, then the whole
excavation analysis will yield incorrect results.

8.8.1.1  Non-linear analysis


If the soil is modelled as an elasto-plastic material, then it can be shown (Brown and Booker
1985) that the same method of computing excavation forces can be applied as shown in
Equation 8.22. The excavation process has to be modelled in incremental steps with layers
of elements removed in sequence and equilibrium of the finite element mesh obtained before
proceeding to the next step. Unlike excavation in an elastic material, the process depends on
the excavation sequence.
The process involves removing some of the elements and calculating the forces to be
applied to the sides and base of the excavation by integrating the stresses and self-weight over
the remaining elements in the finite element mesh and applying any surface tractions. This
ensures that the remaining elements in the mesh are in equilibrium as excavation proceeds.
If the mesh includes beam elements or other types of elements, then the stiffness of these
elements must also be included as they contribute to the overall force balance in the mesh.
Shown in Figure 8.23, is a simple finite element mesh consisting of eight-node isopara-
metric elements. Elements were removed from this mesh to simulate excavation of a tunnel
(half of mesh is used because of symmetry). The material was treated as being elasto-plastic
in this simple problem, and there are three layers of material.
Once excavation has taken place, the stress trajectories plot as shown in Figure 8.24 (a
plot of the directions of the principal stresses), shows the major principal stresses parallel to
the face of the excavation, and the minor principal stress almost zero perpendicular to the
tunnel face as may be expected.

8.8.2  Finite difference approach


The finite difference code FLAC (ITASCA Consulting Group 2014) can also be used for
excavation problems. Anchors and props may be added as excavation proceeds in the same
way as may be done in finite element codes.
324  Geomechanics in soil, rock, and environmental engineering

Figure 8.23 Finite element mesh for the tunnel excavation problem.

An example of an excavation problem involving a sheet pile wall with two sets of
anchors is shown in Figure 8.25. The excavation and placing of the anchors was mod-
elled, and after completion of the excavation, a c − ϕ reduction analysis was carried out.
This involves reducing the magnitude of the strength parameters c (cohesion) and tanϕ
(frictional shearing resistance) until failure occurs. The amount of reduction before fail-
ure allows a factor of safety to be calculated (shown on the Figure as 1.52). In this way,
the Factor of Safety against a collapse of the shoring system at the end of excavation can
be estimated.
It may be noted that the solution shows the failure surfaces going behind the anchors that
hold the wall back, with a wedge of soil moving down above the end of the anchors. In front
of the wall, a passive wedge of soil is shown being pushed forward and upward.

8.9  EXCAVATION INCLUDING GROUNDWATER

Excavation of soil containing water in the pores may involve removing just the soil as would
occur in an excavation under water, or removing both the soil and the water (included in the
pores of the soil) as may occur when a saturated clay is excavated.
Where the soil is reasonably permeable, the groundwater around the excavation will flow
into the excavation. This may occur through the sides and base of the excavation, but if
the sides are retained by an impervious wall (say a diaphragm wall), flow may only occur
through the base of the cut.
Excavation 325

Figure 8.24 Principal stress trajectories around a tunnel.

Job title : . (*101)


2.250
FLAC (Ver. 5.0)

Legend
1.750
14-Sep-11 12:35
step 53781
1.049E+01 <x< 4.443E+01
–1.045E+01 <y< 2.350E+01
1.250
Factor of safety 1.52
Max. shear strain-rate
0.00E+00
5.00E–07 0.750
1.00E–06
1.50E–06
2.00E–06
2.50E–06
3.00E–06 0.250
3.50E–06
4.00E–06

Contour interval = 5.00E–07 –0.250


Boundary plot

0 1E 1

Cable plot –0.750


Net applied forces
Anchored sheet pile wall
Excavation example
1.250 1.750 2.250 2.750 3.250 3.750 4.250
(*101)

Figure 8.25 Failure mode for excavation using c − ϕ reduction (FLAC2D).


326  Geomechanics in soil, rock, and environmental engineering

Finite element analysis of excavation in a two-phase soil (i.e. containing soil and water)
can be performed through use of the equations of consolidation (see Equation 3.70). In
Equation 8.24, we need to be able to compute the increment in force Δf due to removal of
the soil (and pore water).

 K − γ w LT   ∆δ   ∆f 
   =   (8.24)
− γ w L −(1 − α)∆tγ w Φ   ∆h   ∆t γ w Φht 

At any stage, the stress equilibrium is made up of the effective stresses and total ground-
water pressure being in equilibrium with the body forces (due to self-weight) and any applied
forces. We can therefore substitute for the total stress in Equation 8.22 as the total stress σ
is equal to the effective stress σ′ plus the pore water pressure q. In flow problems, the total
head hT is made up of the elevation hEL head plus the water pressure head hw, therefore we
can write the water pressure as

hw = hT − hEL (8.25)

Hence, the water pressure may be obtained by multiplying the water pressure head by the
unit weight of water. It may be noted that the water pressure can consist of any static water
pressure plus excess pore pressures (say, due to loading).

q = γ w hw i = γ w (hT − hEL )i (8.26)


The vector i = (1,1,1,0,0,0)T is introduced as the water pressure is only added to the direct
stress components and not the shear stress components.
The total stress is therefore given by

σ = σ ′ + q = σ ′ + γ w (hT − hEL )i (8.27)


Substitution of the total stress into Equation 8.22 gives the forces to be removed from the
sides of the excavation

∫B σ′dV − γ ∫d a(h
T
w
T
T
∫ ∫
− hEL ) − N T γdV − N T tdS = ∆fnodes (8.28)
V V V S

Therefore, as excavation proceeds the increments in force to be applied to the sides of the
excavation can be found from Equation 8.28 and used in Equation 8.24.
If the excavation is made underwater, and only soil is removed (as in say, a very perme-
able sand), then only the effective stresses are used in computing the force increment, that is,

∫B σ′dV − ∫N
T T

γdV − N T tdS = ∆fnodes (8.29)
V V S

Once again, if there are structural elements in the finite element mesh, these need to be
included in the calculation of the forces to maintain equilibrium of the finite element mesh.
Excavation 327

The process involves setting up the initial effective stresses in the mesh (generally at the
Gauss points in the elements) and the static water pressures (generally at the nodes of the
elements). Then any loads can be applied, and the excess pore water pressures that these
cause is added to the static water pressures. Then as elements are excavated, the equilibrium
equation can be used to compute the excavation forces to be applied to the mesh.
Because the equations of consolidation are written in incremental form in Equation 8.24,
they can be used for materials that are non-linear. The soil may be an elasto-plastic material
and the permeability of the soil may change with stress level for example.
Further information on simulation of excavation in a two-phase material is given in the
paper by Hsi and Small (1993).

8.9.1  Example excavation problem (no drawdown)


Shown in Figure 8.26 is a tunnel that is excavated in a poro-elastic layered material. In this
problem, the water table is assumed to be level with the ground surface (which is assumed
to be permeable), and is held there and not allowed to drop. The sides of the tunnel are
assumed to be permeable so that flow can occur into the tunnel, while the base of the finite
element mesh is assumed to be impermeable. Because of symmetry, only half of the tunnel
is analysed making the line of symmetry an impermeable boundary.
The solution shows the equipotential lines (lines of constant total head), and it may be
seen that the equipotential lines are equally spaced with respect to height around the tunnel

Total head

12.0 Contour legend


3.000E+00
3.500E+00
10.0 4.000E+00
4.500E+00
5.000E+00
8.0 5.500E+00
6.000E+00
6.500E+00
6.0
7.000E+00
7.500E+00
8.000E+00
4.0
8.500E+00
9.000E+00
9.500E+00
2.0

Isotropic permeability
3 soil layers
–2.0
–2.0 0 2.0 4.0 6.0 8.0 10.0 Elastic materials

Isotropic soil – excavation in a poro-elastic material

Figure 8.26 Equipotential contours and flow vectors for tunnel excavation in a layered poro-elastic material.
328  Geomechanics in soil, rock, and environmental engineering

boundary. This is because if the water pressure is zero (as specified), the total head is equal
to the elevation head. If the total head increments are the same in the plot, then the elevation
increments will be equal.
It may also be noticed from Figure 8.26 that the flow vectors (vectors showing the direc-
tion and magnitude of the groundwater flow) are perpendicular to the equipotentials as they
should be because the permeability was assumed to be isotropic in this example.

8.9.2  Excavation involving drawdown of the water surface


When an excavation takes place in a material that is reasonably permeable, such as a sand
or a silt, the groundwater will flow into the excavation if the sides or base of the excavation
are permeable. Even in low permeability soils, the groundwater will eventually flow into the
excavation with time if the groundwater recharge is low enough so that the groundwater
levels cannot be topped up, and the phreatic surface of the groundwater table will drop.
The drop of the phreatic (or free) surface involves the release of water from the elements
that are intersected by the free surface as it drops from one location to another as shown in
Figure 8.27.
Along the free surface, we have two boundary conditions: (1) the total head is equal to the
elevation head and (2) the amount of water released from the pores of the soil as the water
falls is equal to the flow across the free surface.
This may be mathematically written as shown in Equation 8.30.

hT = hEL = f (x, y, t)
∂h (8.30)
qw = Sy T cosβd Γ
∂t

where
kn is the permeability normal to the free surface
Sy is the specific yield of the soil (has no units). It is the volume of water given up per
unit volume of soil
β is the angle of the free surface to the x-direction (Figure 8.28)
Γ is the distance along the free surface
qw is the flow rate of yielded water

This introduces an extra term G into the set of equations (see Equation 8.24).

 K − γ w LT   ∆δ   ∆f 
   =   (8.31)
− γ w L −(1 − α)∆tγ w Φ − γ wG   ∆h   ∆t γ w Φht 

Water released as
surface drops

Figure 8.27 Drop of water table next to excavation.


Excavation 329

∆h

β
Flow
Γ

Figure 8.28 Water released from a region above the free surface.

ks

ks : Permeability of saturated soil

klimit = ks/1000

Plimit – + P

Figure 8.29 Pore pressure–permeability relationship.

where the matrix G in Equation 8.31 is defined as


G = aaT Sy cosβd Γ (8.32)
Γ

The integration of Equation 8.32 can be performed numerically by evaluating the shape
functions for the pore pressure in vector a along the free surface. The location of the free
surface is found from the value of total head at the nodes of the element that is cut by the free
surface. The values of total head can be used to compute the head within the element, and
where the total head is equal to the elevation head is the location of the free surface. More
details on this process are provided by Hsi and Small (1992).
In the region above the free surface, the permeability of the soil ks can be reduced gradu-
ally by linking the permeability of the soil to the pore pressure in the soil. When the pore
pressure becomes negative (as it will above the free surface), the permeability is reduced
linearly (as shown down to 1/1000 of its value when saturated) at a limit pressure plimit. For
higher suctions, the permeability remains at this limiting value (see Figure 8.29).

8.10  SOIL MODELS

When modelling excavation, the appropriate soil model needs to be used. The Mohr–
Coulomb model can overestimate bottom heave and often can predict heave of soil behind
330  Geomechanics in soil, rock, and environmental engineering

0.10%

0.07% 0.04% 0.01%


θ
CL Displacement vectors

Figure 8.30 Shear strain contours for a maximum wall displacement of 0.2% of wall height. (After Simpson,
B., O’Riordan, N.J., and Croft, D.D. 1979. Géotechnique, Vol. 29, No. 2, pp. 149–175.)

the wall. Experience has shown that small-strain models give better prediction of the
behaviour of retaining walls because the strain that occurs behind walls is generally fairly
small.
This was pointed out by Simpson et al. (1979) for the stiff London clay during excavation
for the New Palace Yard car park next to the Houses of Parliament in London.
The strains predicted from an elastic analysis are shown in Figure 8.30 where it can be
seen that the strains are between about 10 −4 and 10−3. Their finite element analysis using a
small strain model as well as a plasticity model at larger strains gave good agreement with
measured performance of the wall and of the surface settlement behind the wall.
Shown in Figure 8.31 is the strain that may be associated with various geotechnical
works, and it can be seen that retaining walls undergo the smallest strains. The small-strain
soil model is therefore best used with finite element analysis. In the commercial program
PLAXIS, the HS-small model is recommended as it gives good bottom heave prediction
independent of model depth and a more realistic settlement trough behind the wall (nar-
rower and deeper).
The variation of shear modulus with shear strain in the PLAXIS HS-small model is
based on the model proposed by Santos and Correia (2001) and is shown in Figure 8.32.

1 Retaining walls
Shear modulus ratio G/G0

Foundations

Tunnels
Very
small
strains Conventional soil testing
Small strains

Larger strains
0 Shear strains γs
10–6 10–5 10–4 10–3 10–2 10–1

Dynamic methods
Local gauges

Figure 8.31 Range of shear strain associated with various geotechnical works.


Excavation 331

1.0

0.8

G ⁄ G0 0.6

0.4

0.2

0
10–3 10–2 10–1 1 10 100 1000
γ/γ 0.7

Figure 8.32 HS-small model used in PLAXIS showing reduction in shear modulus with shear strain. (After
Santos, J.A.D. and Correia, A.G. 2001. Proceedings of the 15th International Conference on Soil
Mechanics and Geotechnical Engineering, Istanbul, Vol. 1, pp. 267–270.)

The equation for degredation of the shear modulus G is based upon the shear strain γ0.7 at
which the shear modulus drops to 70% of its initial value G 0 (at γ ≈ 10 −6). The relationship
is then given by

G 1
= (8.33)
G0 1 + 0.385(γ /γ 0.7 )

The constant 0.385 comes from fitting the degradation curve to experimental data. The
small-strain shear modulus may be found from seismic field tests as discussed in Sections
4.25 and 4.26.
This page intentionally left blank
Chapter 9

Retaining structures

9.1 INTRODUCTION

The excavation of any kind of trenches or cuts such as basements for buildings or cuts for
road works or railways may require some kind of support if the soil or rock is not of suf-
ficient strength to stand safely on its own. The process of excavation is closely related to
excavation support, therefore this chapter should be read in conjunction with Chapter 8.
There are a number of different systems that have been designed to support the sides of
excavations and which type is used depends on the proximity of existing structures, the
space available and the type of soil being retained.

9.2  EARTH PRESSURE CALCULATION

Pressures on the back of a retaining structure depend on the movement of the structure. If it
is moving away from the backfill, then the pressures that act are called the active pressures.
If the wall is being pushed into the soil, then the passive soil pressures apply. If the wall is
rigid and not moving, then an intermediate case of at rest pressures are applicable.
Active pressures apply behind retaining structures that can move forward such that the
soil begins to shear and the shear forces support some of the weight of the backfill. Passive
forces occur in front of a wall moving forward and pushing into the soil. The forces are
therefore being resisted by shear forces in the soil and are higher than the active forces.
For walls that are braced or are very rigid, and are not able to move forward, the at
rest condition applies. The forces involved are intermediate between the active and passive
forces.

9.2.1 ​Rankine’s theory
If a wall is smooth, then these forces can be calculated simply from Rankine’s theory
(Rankine 1857). This is shown in Figure 9.1, where it is assumed that the soil fails according
to the Mohr–Coulomb failure criterion. If the soil is at rest, then at any depth in the soil, the
vertical stress σv0 is due to the overburden pressure and the horizontal stress σh0 is related to
the vertical stress by the coefficient of earth pressure at rest K0, that is,

σ v0 = γz
(9.1)
σ h0 = K0σ v0

333
334  Geomechanics in soil, rock, and environmental engineering

Mohr–Coulomb failure surface

σ
σha σh0 σv0 σhp

Figure 9.1 Mohr’s circle at failure for active and passive cases.

Equation 9.1 is written in terms of total stress, but if water is present, the effective stresses
need to be used. The coefficient of earth pressure at depth may be found from the empirical
relationship

K0 = (1 − sin φ) (9.2)

If the soil is behind a wall and the wall moves forward, the lateral pressure σh0 will reduce
until it reaches the value σha and the soil will fail since Mohr’s circle has reached the failure
surface. If the lateral pressure is increased as it would as a wall pushes into a soil, then the
lateral pressure will increase until it reaches the passive pressure σhp at which time the soil
will fail as Mohr’s circle has reached the failure surface. It can be seen from Figure 9.1 that
the passive pressure is much larger than the active pressure.
For a smooth wall, the direction of the failure planes can be found from Mohr’s circle
at failure. For the passive case, the planes of failure (broken lines) pass through points C
and D, while for the active case, the failure planes pass through A and B. The directions
of the active planes of failure are therefore at 45° + ϕ/2 from the horizontal and the passive
planes of failure are at 45° − ϕ/2 from the horizontal. The planes of shearing are shown in
Figure 9.2.
From Mohr’s circle, we can find the magnitudes of the stresses σha and σhp using the fact
that Mohr’s circle touches the failure surface. For a vertical wall with a backfill having a
horizontal surface, we have

σ ha = Kaσ v − 2c Ka (9.3)

σ hp = Kpσ v + 2c Kp (9.4)

where c is the intercept of the Mohr–Coulomb failure surface on the vertical axis (the cohe-
sion) and
Retaining structures  335

Failure Failure
See Figure 9.1 for plane at A plane at B
points A, B, C, D

Failure Failure Active


plane at C plane at D failure

Wall
Passive movement
45°+ φ/2
failure 45°– φ/2
Front and back
of smooth wall

Figure 9.2 ​Directions of failure planes in active and passive cases (smooth wall).

1 − sin φ
Ka = (9.5)
1 + sin φ

1 + sin φ
Kp = (9.6)
1 − sin φ

The angle of shearing resistance ϕ is the slope of the Mohr–Coulomb failure surface.
Hence, at any depth down at the back of the wall, the vertical stress can be calculated, and
this can be converted to the horizontal stress through Equation 9.3. If there is a surcharge
q over the surface of the backfill, this is added to the vertical stress, that is, σv = γz + q. In
front of the wall, the vertical stress at any depth can be converted to the passive horizontal
stress using Equation 9.4. Because the vertical stress increases linearly (and is equal to q at
the surface), then the increase in lateral stress is linear with depth. Therefore, for a uniform
soil, with unit weight γ and for a wall of height H, we have

1
Pa = Ka γH 2 − 2c Ka H + qKa H (9.7)
2

If the wall is not smooth, then there will be shear stresses acting at the back (or front) of
the wall, and the stress state is not a point on the horizontal axis of Mohr’s circle diagram.
The shear planes in the soil are then curved as near the back of the wall or at the front of
the wall, the planes of failure will be at different angles than for the smooth wall case. The
planes of failure are shown in Figure 9.3 for the active case.
Use of Rankine’s theory is shown in Figure 9.4 where a soil possessing friction and cohe-
sion is saturated with water. First, the vertical effective stress is calculated and then the
vertical effective stress is converted to the horizontal effective stress using Equation 9.3.
This calculation results in negative horizontal stresses near the top of the wall (Figure 9.4c).
If the soil cannot sustain tension, then the soil will form a crack and provide no tensile
force on the wall (Figure 9.4d). If the water fills the cracks, then the water pressure must be
added to the horizontal effective stress as shown in Figure 9.4f to find the final total stress
on the wall.
336  Geomechanics in soil, rock, and environmental engineering

Rough
Slip

F′

Figure 9.3 ​Slip lines for active pressure on rough wall from plasticity theory.

(a) (b) (c) (d)


2c′/ γsub √ Ka
φ′
c′
H

σh′ = γH − γ w H σh′ = Ka σv′ − 2c′√ K σh′


a
Retaining = γsub H
wall Effective horizontal
Vertical Horizontal effective stress for no
effective stress effective stress tension

(e) γwH (f) σh

Water pressure if no Total stress =


weep holes (d) + (e)

Figure 9.4 ​Calculation of active pressure on a vertical wall with horizontal backfill using Rankine’s theory.
Retaining structures  337

9.2.1.1 ​Inclined backfill
If the backfill has an inclined surface (but the wall is vertical) then the solution of Mazindrani
and Ganji (1997) may be used to find the active and passive pressures at a depth z below the
top of the wall (Equation 9.8).

cos β
pp , pa = {2 γzcos2β + 2c cos φ sin φ ± √[4cos2β(cos2β − cos2φ)γ 2 z 2
cos2 φ (9.8)
2 2 2
+4c cos φ + 8c γzcos β sin φ cos φ]} − γz cos β

where β is the angle of inclination of the backfill to the horizontal.

9.2.2 ​Coulomb’s theory
Coulomb’s theory (Coulomb 1776) is based on the assumption that the failure surface is
planar, and while this is true for a smooth wall, it is not true for a rough wall where there is
friction causing a shear force against the wall as was discussed in the previous section. If the
backfill is sloping at an angle β to the horizontal, and the back of the wall is at an angle α to
the horizontal, and the angle of friction between the wall and the soil is δ, then it is possible
to draw the vector of forces for the soil wedge. In Figure 9.5, this is done for the active case
and the wedge of soil ABC. For the passive case, the forces Pp and F are inclined at angles
to the opposite side of the normal as the wedge is moving up and not down as for the active
case (see Figure 9.8).

9.2.2.1  Active case


For the case where there is no water behind the wall, trial slip planes can be selected and the
vector of forces drawn. This is shown in Figure 9.6, and the point at which the maximum
active force on the wall is found corresponds to the critical failure plane. The maximum
value of the active pressure is the value required.
If there is water behind the wall, then the water forces must be included in the force dia-
gram as shown in Figure 9.7. It can be seen that the weight of water in the wedge is balanced

(a) (b)

β B
A Pa c
W ψ
b

W
δ R

Pa ψ φ θ–φ
α θ
R a
C

Figure 9.5 Coulomb’s planar failure surface.


338  Geomechanics in soil, rock, and environmental engineering

c
Pa′

b
W

R′

a
Critical
wedge 4
3 c4
2
1 b4
c3

b3
Pa′ max c2
δ R′ φ′ c1
b2
Pa′
b1

Figure 9.6 C
​ oulomb wedge analysis (active case with no water).

by the reactive forces pw and Rw, therefore it would be possible to carry out the analysis by
omitting the lower part of the diagram aed and using the total weight of the soil W minus
the weight of the water Ww. Trial wedges will give a maximum force Pa′ that is only the effec-
tive force on the wall. The water pressure force must be added to get the total active force
on the wall (see Figure 9.7).

9.2.2.2  Passive case


For the passive case, the force diagram can be drawn as shown in Figure 9.8. The vector dia-
gram of forces can be drawn for trial wedges and this time, the minimum value is obtained.
This gives the value of the passive force on the wall.

Pa′ c

Pw b
Pa′

δ φ′ R′
Pa′ W
Pw Rw R′ Pw a
d
Rw Ww

Figure 9.7 ​Coulomb analysis for case where there is water behind a wall (active case).
Retaining structures  339

Pp′

Cw′ W
c2
C′
d2
R′
φ′ Pp′
c Pp′
d Minimum

c1
W d1
R′
a
e C′
w b1
a
b C′

Figure 9.8 ​Coulomb analysis for passive case in cohesive soil.

It may be noted that in this case cohesive forces have been included in the analysis.
Cohesive forces acting on the back of the wall and along the trial slip plane can be included
as extra forces when drawing the force diagram.
As for the active case, water pressures can be included if water is lying in front of a wall.
Again, the effective passive force is found and added on to the water force on the wall to get
the final passive force.

9.2.2.3 ​Surface loads
If there are surface loads such as distributed loads or line or point loads, these forces can
be added into the weight of the soil when drawing the vector diagram. For some of the trial
wedges of soil, part of a uniform load may not need to be included as the load is outside of
the top of the wedge.

9.2.2.4 ​Uniform materials
For uniform soils that do not have cohesion and do not involve surface forces, the maximum
(active) or minimum (passive) force can be calculated through the use of calculus. The soil
has to be totally uniform, and so must be totally drained or totally saturated. By looking at
force equilibrium on a wedge of soil, it is possible to calculate the critical wedge by writing
the force on the wall in terms of the angle of the wedge θ to the horizontal, and then dif-
ferentiating to find the minimum value of the force, that is, for the active case finding where
∂Pa /∂θ = 0. This leads to the result for the active case

1
Pa = Ka γH 2
2
sin2 (α + φ)
Ka = 2 (9.9)
2
 sin(φ + δ)sin(φ − β) 
sin α sin(α − δ)  1 +
 sin(α − δ)sin(α + β) 

340  Geomechanics in soil, rock, and environmental engineering

and for the passive case

1
Pp = Kp γH 2
2
sin2 (α − φ) (9.10)
Kp = 2
 sin(φ + δ)sin(φ + β) 
sin2 α sin(α + δ)  1 −
 sin(α + δ)sin(α + β) 

The angle of shearing resistance used in the Equations 9.9 and 9.10 is the drained angle of
friction as the equations do not apply to cohesive soils, that is, ϕ = ϕ′.
The values given by the formulae of Equation 9.10 for the passive case can be largely in
error, especially for high values of the angle of shearing resistance ϕ because of the assump-
tion of a planar failure surface. For this reason, it is prudent to use the values given by the
log spiral method or from the bound theorems given in Sections 9.2.3 and 9.2.4. The values
of the active earth pressure coefficient are not so sensitive to error for higher values of ϕ and
are sufficiently accurate for use in engineering applications.

9.2.2.5 ​Active earthquake forces


If there are earthquake forces acting on the fill behind the retaining wall, a pseudo-static
analysis can be undertaken using the Coulomb method. The horizontal force Fh acting on
any potential active failure wedge is assumed equal to the weight of the wedge times the
seismic coefficient kh. Similarly, the vertical force F v on the wedge is equal to the weight of
the wedge times the vertical seismic coefficient kv.

Fh = khW
(9.11)
Fv = kvW

The vector sum of the weight of the wedge and the two seismic forces gives a resultant
force acting at an angle ψ to the vertical (see Figure 9.9) where

 kh 
ψ = tan−1  (9.12)
 1 − kv 

β Wkv

Wkh
Failure
W plane
Wall
Cohesionless
soil
δ
φ′
α

Figure 9.9 ​Seismic forces acting on Coulomb wedge.


Retaining structures  341

The resultant force W′ may be seen to be equal to

 1 − kv 
W′ = W  (9.13)
 cos ψ 

This leads to the Mononobe–Okabe solution (Mononobe and Matsuo 1929) and Okabe
(1924) for the active earth force

1  1 − kv 
Pa = Ka γH 2 
2  cos ψ 
sin2 (α − ψ + φ) (9.14)
Ka = 2
 sin(φ + δ)sin(φ − β − ψ) 
sin2 α sin(α − ψ − δ)  1 +
 sin(α − ψ − δ)sin(α + β) 

Solutions for active wall pressures have also been presented by Iskander et al. (2013) based
on Rankine’s theory, and it is shown that these are not greatly different to the Mononobe–
Okabe solutions. For the passive pressure case, see Section 9.2.3.1.

9.2.3  Log spirals


A log spiral shaped failure surface can be used to find values of the passive earth pressure
coefficient, and this was employed by Caquot and Kerisel (1948) to calculate values for walls
with a rough surface. The values of Kp are shown in Figure 9.11 for a range of values of the
angle of shearing resistance ϕ and the angle of the soil surface to the horizontal β as shown
in Figure 9.10. The plot of Figure 9.11 is for a vertical wall, that is, α = 90° and for a purely
frictional material where the angle of shearing for the wall–soil interface is δ = ϕ. If the
angle δ is less than ϕ, correction of the value of Kp can be made through use of the values of
λ in Table 9.1 where Kpδ = λKp.

9.2.3.1  Passive earthquake forces


The effect of earthquakes on passive pressures can be somewhat in error if the Mononobe–
Okabe method is used because of the assumption of planar failure surfaces. Solutions using
a logarithmic spiral have been obtained by Subba Rao and Choudhury (2005) for this case.

β
α φ′, γ
Drained
cohesionless Potential slip line
δ

Figure 9.10 Passive pressure on retaining wall.


342  Geomechanics in soil, rock, and environmental engineering

1.0
80
70 0.8
60
50 0.6

40
0.4
30
0.2

20
0

–0.2
10 H δ
Kp

9
8 H/3 Log spiral –0.4
7
6
σP = KpγH
5
–0.6
4

3
–0.8
2

β
= −0.9
φ
1
0 10 20 30 40
φ (in degrees)

Figure 9.11 ​Passive pressure coefficient for vertical wall. (After Caquot, A.I. and Kerisel, J. 1948. Libraire
du Bureau des Longitudes, de L’école Polytechnique, Paris Gauthier-Villars, Imprimeur-Editeur,
p. 120.)

Table 9.1  ​Factor λ to be used with values from Figure 9.11

Values of δ/ϕ
ϕ (deg) 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0.0
10 0.978 0.962 0.946 0.929 0.912 0.898 0.880 0.864
15 0.961 0.934 0.907 0.881 0.854 0.830 0.803 0.775
20 0.939 0.901 0.862 0.824 0.787 0.752 0.716 0.678
25 0.912 0.860 0.808 0.759 0.711 0.666 0.620 0.574
30 0.878 0.811 0.746 0.686 0.627 0.574 0.520 0.467
35 0.836 0.752 0.674 0.603 0.536 0.475 0.417 0.362
40 0.783 0.682 0.592 0.512 0.439 0.375 0.316 0.262
45 0.718 0.600 0.500 0.414 0.339 0.276 0.221 0.174
Retaining structures  343

(a) (b)
10 15
φ = 40° kv = 0.0kh φ = 40°
kv = 1.0kh

8 12
kv = 0.0kh
kv = 1.0kh

6 9
Kpγd

Kpγd
φ = 30°
φ = 30°
4 6

φ = 20°
φ = 20°
2 3
α = 0° , β = 0° , α = 0° , β = 0° ,
φ = 10° φ = 10°
δ/φ = 0.5 δ/φ = 1.0
0 0
0 0.1 0.2 0.3 0.4 0.5 0 0.1 0.2 0.3 0.4 0.5
kh kh

Figure 9.12 ​Passive earth pressure coefficient Kpγd for self-weight of fill for different horizontal earthquake
acceleration factors k h. (a) δ/ϕ = 0.5; (b) δ/ϕ = 1.0.

They compute the passive pressure coefficient for the self-weight of the soil, any surface
surcharge and for cohesion separately and if all three are acting, then the effects are added
together.
Values of the passive pressure coefficient Kpγd are shown in Figure 9.12 for a vertical wall
with a horizontal backfill. More results are given in the paper by Subba Rao and Choudhury
(2005). The force on the wall is given (for self-weight only) by

Pp = 0.5γH 2Kpγd (9.15)


9.2.4 ​Upper- and lower-bound solutions


The upper- and lower-bound solutions of plasticity theory as discussed in Section 9.7.6 can
be used to find bounds on the passive and active pressures. This has been done by Shiau
et  al. (2008) for the wall shown in Figure 9.10 (the passive case). Bounds for the passive
earth pressure coefficients are shown in Figure 9.13 for the case where the wall is vertical
and the surface of the soil is horizontal. Results are presented for different angles of friction
δ acting on the front of the wall. To be conservative, the lower bound could be taken in any
calculation without too much error. However, the upper bound gives an indication of what
the error may be, as the solution must lie between the two values.
If the front of the wall is at an angle α to the horizontal (see Figure 9.10), then results can
be found for the passive coefficient of earth pressure Kp from Figure 9.14 for a backfill with
a horizontal surface. Variation of Kp with the slope of the backfill surface β (for one value of
wall friction angle δ = 19.2°) is shown in Figure 9.15 for a wall with a vertical back.
Table 9.2 (Reddy et al. 2013) shows values of Kp from different sources and it may be seen
that there are large differences especially when then angle of shearing resistance ϕ is large
344  Geomechanics in soil, rock, and environmental engineering

50
α = 90°
φ′ = 45°
β = 0°
40

30
Kp Upper bound
Lower bound
20 φ′ = 40°

φ′ = 35°
10 φ′ = 30°
φ′ = 25°
φ′ = 20°
0
0.0 0.2 0.4 0.6 0.8 1.0
δ/φ′

Figure 9.13 ​Passive earth coefficients from upper- and lower-bound theorems.

along with the angle of friction δ on the wall. The values of Kp from the bound theorems are
therefore the most desirable for use, as the error in the value is known.

9.3 ​EFFECT OF WATER

The formulae for calculating the pressures on retaining structures can apply to the drained
or undrained case. If the retaining wall has a saturated clay backfill or the soil in front of a
wall is a saturated clay, then any rapid loading will take placed under undrained conditions.
In this case

φ = φu
(9.16)
c = cu

16
α = 90°, δ/φ′ = 1
14 β = 0°

12 Upper bound
Lower bound α = 75°, δ/φ′ = 1
10
Kp,h

8
α = 60°, δ/φ′ = 1
6
α = 90°, δ/φ′ = 0
4
α = 75°, δ/φ′ = 0
2 α = 60°, δ/φ′ = 0

0
20 25 30 35 40
φ′

Figure 9.14 Values of passive coefficient for different angles of wall friction δ and slope of back of wall α.
Retaining structures  345

20

Upper bound
Fang et al. (failure)
15 Lower bound
φ′ = 30.9° Fang et al. (S/H = 0.30)
δ = 19.2° Fang et al. (S/H = 0.20)
α = 0°
Kp,h

10

Fang et al. (S/H = 0.10)


5 Fang et al. (S/H = 0.05)

0
–15 –10 –5 0 5 10 15 20
β°

Figure 9.15 ​Variation of Kp with slope of backfill β.

Table 9.2  V
​ alues of passive earth coefficient from different sources
Caquot Kumar and Soubra and
Wall angle of Coulomb and Kerisel Subba Rao Macuh Lancellotta Reddy et al.
Soil ϕ friction δ (1776) (1948) (1997) (2002) (2002) (2013)
20 2.04 2.04 2.04 2.04 2.04 2.04
25 2.46 2.46 2.46 2.46 2.46 2.46
30 0 3.00 3.03 3.00 3.00 3.00 3.00
35 3.69 3.69 3.69 3.69 3.69 3.69
40 4.60 4.59 4.60 4.60 4.60 4.60
20 2.41 2.35 2.38 2.39 2.35 2.40
25 3.12 3.03 3.06 3.07 3.07 3.08
30 1/3 ϕ 4.14 4.00 4.02 4.03 4.03 4.05
35 5.68 5.28 5.42 5.44 5.44 5.46
40 8.15 7.25 7.58 7.62 7.62 7.62
20 2.64 2.60 2.50 2.57 2.48 2.61
25 3.55 3.40 3.40 3.41 3.22 3.46
30 1/2 ϕ 4.98 4.50 4.60 4.65 4.29 4.73
35 7.36 6.00 6.60 6.59 5.88 6.71
40 11.8 9.00 9.80 9.81 8.38 10.00
20 2.89 2.65 2.73 2.75 2.58 2.85
25 4.08 3.56 3.72 3.76 3.41 3.91
30 2/3 ϕ 6.11 5.00 5.26 5.34 4.63 5.57
35 9.96 7.10 7.78 7.95 6.51 8.32
40 18.7 10.7 12.24 12.6 9.57 13.27
20 3.53 3.01 3.07 3.13 2.7 3.4
25 5.60 4.29 4.42 4.54 3.63 4.95
30 ϕ 10.1 6.42 6.68 6.93 5.03 7.58
35 22.9 10.2 10.76 11.3 7.25 12.33
40 92.6 17.5 18.86 20.1 11.03 21.64
346  Geomechanics in soil, rock, and environmental engineering

and these values are used with the formulae for earth pressure along with the total stresses
in the ground.
If the soil in front of or behind a retaining structure is very pervious such as a sand, or
has been in place long enough to drain, then the drained strength parameters and the effec-
tive stresses in the ground are used to compute the forces on the wall. If the soil is beneath
the water table, these forces are the effective lateral stresses, so the water pressures must be
added to them to obtain the total pressure on the wall. In this case

φ = φ′
(9.17)
c = c′

The water pressures can be very large, therefore in the design of walls it is always essential
to supply drains behind the wall to eliminate water pressures. The drains must be main-
tained to ensure that blocking does not occur, as this would result in water pressures acting
on the wall and possible failure.

9.4 ​SURFACE LOADS

The force exerted on retaining walls can be calculated using hand methods, and most of
these are based on solutions that treat the soil as a uniform elastic body. For point and line
loads, a distance x = mH from the wall, pressure distributions can be calculated from Figure
9.16. The magnitude Ph and location yp of the resultant force can be calculated for use in
stability (overturning) calculations. The pressure distribution for a horizontal line load Qh
is also given in the figure.

9.4.1 ​Compaction stresses
When soil is compacted behind a wall, the compaction can induce higher stresses than apply
for a wall that has not had the soil specifically compacted. According to Ingold (1979), the
locked in stresses due to compaction of each layer result in a stress diagram like that shown
in Figure 9.17c.
In the figure, the compaction stresses are developed to a depth dc where

1  2Q/w 
dc = (9.18)
Ka  πγ 

The roller load is Q which includes dead and dynamic loads from the roller, the roller width
is w, γ is the unit weight of the backfill, and K a is the active earth pressure coefficient.
From the surface to a depth zc , the stress increase is linear. The depth zc is given by

 2Q/w 
zc = Ka  (9.19)
 πγ 

Horizontal stresses σ ′ha in the range zc ≤ z ≤ dc are given by the formula

2Qγ
σ ′ha = (9.20)
πw
Retaining structures  347

(a) (b) (c)


x
Line load, Qℓ Point load, Qp
z = nH x = mH x = mH Line load, Qh

z = nH
σh

z
Ph
Ph

He
σh Ph σh σh
A A Qp

Wall
H
H

θ
σhθ
yp

yp
x = mH

For m ≤ 0 .4 For m ≤ 0.4 Section A–A

σh H =
Qℓ ( )
0.20n
(0.16 + n2)2
σh ( )
H
Qp
2
=
0.28n2
(0.16 + n2)3
(
He = x tan 45° +
φ′
2 )
Ph = 0.55 Qℓ Ph = 0.69Qp
σhθ
= σh cos2 (1.1 θ)
σh ( ) (
He
Qp
=2 1−
z
He )
For m > 0.4 For m > 0.4 Resultant Ph = Qh

σh H =
Qℓ ( ) 1.28m2n
(m2 + n2)2
σh ( )
H2
Qp
=
1.77m2n2
(m2 + n2)3

( )
0.64Qℓ m (1 – m2) 1
Ph = Ph = 0.48Qp + tan−1
(1+ m2) (1 + m2)2 m
Lateral pressure on
Lateral pressure on wall Lateral pressure on wall wall due to horizontal
due to vertical line load, Qℓ due to vertical point load, Qp line load, Qh
(d) (e)
0

0.2
H
z

0.4
Value of n =

0.6 Ph yp
Symbol m Ph yp
Qℓ H Symbol m
Qp H
0.4 0.55 0.61
0.8 0.5 0.51 0.56 0.4 0.69 0.58
0.6 0.47 0.52 0.5 0.65 0.52
0.7 0.43 0.49 0.6 0.59 0.47
1.0
0 0.2 0.4 0.6 0.8 1.0 0 0.5 1.0 1.5 2.0
Value of σ h H
Qℓ ( ) Value of σ h
( )
H2
Qp
Pressure distribution due Pressure distribution due
to vertical line load, Qℓ to vertical point load, Qp

Figure 9.16 ​Pressure distributions on retaining walls due to point and line loads. (After Geotechnical
Engineering Office, Geoguide 1 2000. Guide to Retaining Wall Design, Civil Engineering
Department, The Government of Hong Kong Special Administrative Region.)
348  Geomechanics in soil, rock, and environmental engineering

(a) ′
Phm (b) ′
Phm

zc

zc
Depth below surface, z

Depth below surface, z


Resultant pressure
distribution due to ′
Locus of Phm
compacting surface for successive
layer compacted
layers

Ph′
Horizontal earth pressure distribution Horizontal earth pressure distribution
in uncompacted fill resulting from resulting from successively compacted
compaction of surface layer only layers of fill

(c) ′
Phm

2Q1γ
zc

P ′hm =
π
Depth below surface, z

hc

Ph′ = Kγz

2Q1
zc = K
πγ

1 2Q1
hc =
K πγ

Ph′
Design diagram for horizontal earth
pressure induced by compaction

Figure 9.17 Compaction stresses. (After Ingold, T.S. 1979. Géotechnique, Vol. 29, No. 3, pp. 265–283.)

For walls that are unyielding, the active pressure coefficient should be replaced by the at
rest coefficient K0.
Other methods include those of Duncan and Seed (1986) that involves a simple hand
calculation to find the stress distribution. Chen and Fang (2008) have measured compaction
stresses in sands. At the top of the compacted zone, they found that the lateral pressures
were close to the passive earth pressure, but below this, they approached the at rest (Jacky)
K0 pressure.

9.5 ​SHEET PILE WALLS

Often, temporary walls are made from sheet piling driven into the soil and then the soil
is excavated in front of the sheet piling. For unsupported sheet piles, the wall can flex and
rotate forward as shown in Figure 9.18. The soil behind the wall will mobilise the active
Retaining structures  349

Active

Passive

X
Passive
Point of rotation

Figure 9.18 Sheet pile wall.

earth pressure and the soil in front of the wall will mobilise part of the passive pressure.
Below the point of rotation, the wall will move back into the soil mobilising a passive force
on the back of the wall.
One simple method that can be used to analyse the wall is to assume that the passive pres-
sure at the front of the wall is only partially mobilised by applying a factor of safety of F to
the full passive force.
The following assumptions are then made:

1. Full active pressure acts on the back of the wall.


2. Only part of the full passive pressure is mobilised; F can be taken as 1.5–2.0 depend-
ing on the risk involved.
3. The passive pressure behind the wall can be considered to be a force P 2 acting on
the base of the wall. By taking moments about the base of the wall, the depth of
embedment of the wall can be found without having to know what the value of the
force P 2 is.

With reference to Figure 9.19, the active Pa and passive Pp forces acting on the wall are

H
Pa = (Ka γH) (9.21)
2

H
Pa
d Pp/F

P2
X

Figure 9.19 Sheet pile wall showing assumed force P 2 at base of wall X.


350  Geomechanics in soil, rock, and environmental engineering

d
Pp = (Kp γd) (9.22)
2

where d is the depth of embedment and H is the overall length of the sheet pile.
Each of these forces acts at the centroid of the pressure diagram, therefore by taking
moments about the base of the wall, we have

 Kp γd 2  d  Ka γH 2  H
 2F  3 =  2  3 (9.23)

If a value of H = d + h is substituted (where h is the depth of retained soil), we have

 d + h  Kp 
 d  = 3
 K F  (9.24)
a

This equation can be solved for the depth of embedment d required. However, because
of the approximations made in the analysis, the sheet piling is usually driven an extra 20%
further than the calculated value.
If there is likely to be a difference in water level from one side of the sheet pile wall to the
other (say, because of tides), then the water forces should be included in the analysis, as the
water forces can be considerable.

9.6 ​ANCHORED WALLS

Walls of various types can be held back by anchors whether they are sheet pile walls or
diaphragm walls such as those discussed in Section 8.2.4 in Chapter 8. The anchor cable is
attached at one end to the wall and may be attached at the other end to an anchor block or
may be grouted into the soil or angled downward so that it reaches a firm stratum or rock
where it may also be grouted.
Hand calculations can be carried out by taking moments about the point of application of
the anchor (assuming one row of anchors is used). This is the ‘free earth’ method of design
for the wall. An anchored sheet pile wall is shown in Figure 9.20 where it has been assumed
that there is a difference in water levels between the front and back of the wall due to a time
lag in the water draining from behind the wall (see Table 9.3).
By choosing values of the factor of safety F and the depth of the wall d, the negative and
positive moments can be calculated and the required value of d is the one that produces
equilibrium.

9.6.1 ​Anchors
The force that an anchor can supply can be found from the upper- and lower-bound theo-
rems. For strip anchors with a vertical face, solutions have been found by Merifield and
Sloan (2006) for sands. The pull-out force depends on how deep the anchor is buried as
shown in Figure 9.21. This is assuming that the anchor has been placed far enough back
from the wall so that it does not lie within the active wedge of soil that would arise during
failure.
Retaining structures  351

dw1 da
A A
dw2 Pa1

Pa3 h1 h2

d Pp Pa2 Pw1
Pw2
F
γsubd γdw1 γwh2 γwh1
γsubh1

Figure 9.20 A
​ nchored sheet pile wall.

In the charts, H is the depth to the base of the strip anchor and B is the height of the
anchor. The pull-out force per unit length Qu is given in terms of the break-out factor N γ
from the following equation:

Qu = γHBN γ (9.25)

Values of the break-out factor N γ are given in Figure 9.21a for the lower-bound solution
and in Figure 9.21b for the upper-bound solution. The true solution lies between the two
extremes; however, the lower-bound solution gives a conservative estimate of the pull-out
force.

Table 9.3  ​Forces and lever arms for taking moments about the anchor (point A)
Force number Magnitude of force Lever arm about A Moment
1 2 Pp
1 Pp = K p γ subd 2 lp = H + d − da × lp
2 3 F
1 2
2 Pa1 = Ka γ d w21 l a1 = d w1 − d a − Pa1 × la1
2 3
1 2
3 Pa 2 = Ka γ sub h12 la2 = h1 + ( d w1 − d a ) − Pa2 × la2
2 3
1
4 Pa3 = Kaγdw1h1 l a3 = h1 + ( d w1 − d a ) − Pa3 × la3
2
1 2
5 Pw1 = γ w h12 l w1 = h1 + ( d w1 − d a ) − Pw1 × lw1
2 3
1 2
6 Pw 2 = γ w h22 lw 2 = h2 + ( d w 2 − d a ) Pw2 × lw2
2 3

Note: h1 = H + d − dw1.
352  Geomechanics in soil, rock, and environmental engineering

(a) Lower bound

100

H Rough
B H/B = 10

Nγ 10

H/B = 1

qu = γHNγ
1
20 25 30 35 40
φ′
(b) Upper bound

100

H Rough H/B = 10
B

Nγ10
H/B = 1

qu = γHNγ
1
20 25 30 35 40
φ′

Figure 9.21 ​Upper- and lower-bound solutions to the vertical anchor break-out factor N γ for different
angles of shearing resistance (sands).

For line anchors in clay, Merifield et al. (2001) have provided the solutions given in Figure
9.22. The pull-out force per unit length Qu of the anchor is given by

Qu = B(su Nc0 + γH a ) (9.26)


where the bearing coefficient Nc0 lies between the upper and lower bounds given in the fig-
ure. The undrained shear strength su of the soil is assumed to be constant with depth in the
plot of Figure 9.22. Ha is the depth to the centre of the anchor.

9.7 ​REINFORCED EARTH

Reinforced walls consist of facing panels connected to reinforcing strips or geofabrics that
are incorporated into the backfill (Figure 9.23). The facing panels can be made of steel or
Retaining structures  353

(a)
Finite element upper bound
10
Finite element lower bound
Upper bound (five-variable)
Finite element (Rowe 1978)
8

6
Nc0

2 Rough
H
B

(b)

10 Finite element upper bound


Finite element lower bound
Meyerhof (1973)
8 Das et al. (1980)
Ranjan and Arora (1980)

6
Nc0

2 Rough
H
B

0
1 2 3 4 5 6 7 8 9 10
H/B

Figure 9.22 ​Bearing coefficients for strip anchor in clay: (a) comparison with numerical results; (b) compari-
son with experimental results.

concrete, and the reinforcing may be made from steel or polypropylene or geosynthetic
meshes. The original method (Terre Armée) that was invented by Henri Vidal in the 1950s
used steel strips, as these were less extensible than geosynthetics. Steel strips can be galvan-
ised to increase their service life or can be ribbed to give higher pull-out resistance.
Design of reinforced earth walls requires considering the failure modes that may occur
and these are shown in Figure 9.24. These modes are

a. Sliding of the block either at the base or part way up the wall
b. Failing in bearing
c. Rupture of the reinforcement
d. Pull-out of the reinforcement
e. Overall slip failure or failure through the reinforced block
f. Excessive settlement or tilt
354  Geomechanics in soil, rock, and environmental engineering

Wall facing panels

Reinforcing strips

Compacted soil

Figure 9.23 ​Reinforced earth wall showing reinforcing strips.

There are several design approaches available, but as the British (BS 8006-1:2010) and
Australian (AS 4678, 2002) Standards and the Roads and Maritime Services of NSW
Australia (RMS Specification R57 2012) use partial factor methods this approach will be
used here. In this approach, the soil strengths are factored down by ‘material factors’ and
the loads are factored up.
If the forces causing failure are still less than the resisting forces, then the design is
acceptable. This approach is more complicated than the single factor of safety method,
but has the advantage that different factors may be applied to different kinds of load and
to frictional and cohesive forces. The various failure modes are addressed in the following
sections.

(a) L (b) (c)

H Wv
P
Rupture

(d) (e) (f)



Settlement

Rotation

Translation

Figure 9.24 M
​ odes of failure for reinforced earth walls.
Retaining structures  355

9.7.1 ​Sliding
The reinforced block may slide as shown in Figure 9.24a where sliding may be on soil-to-soil
contact or on soil-to-reinforcement contact. The force causing the sliding are assumed to be
the active earth pressure behind the wall, and the resisting forces are those due to shearing
along the base of the sliding block.
There are two cases to consider: (1) long-term stability where the drained strength param-
eters are used and (2) short-term stability where undrained strength parameters are used.
For long-term stability, where there is soil-to-soil contact, we have

tan φ′ c′
fsRh ≤ Rv + L (9.27)
fms fms

where
Rv is the resultant of all factored vertical load components
R h is the factored horizontal disturbing force
fms is the partial materials factor applied to the soil strength
fs is the partial factor against base sliding
L is the effective base width for sliding
c′ and tan ϕ′ are the drained strength parameters

For long-term stability where there is sliding on a reinforcement-to-soil contact, then


Equation 9.28 is used.

α ′ tan φ′ α ′bcc′
fsRh ≤ Rv + L (9.28)
fms fms

where
α′ is the interaction coefficient relating soil/reinforcement bond angle with tan ϕ′,
α′bc is the adhesion coefficient relating soil cohesion to soil/reinforcement bond

For short-term stability where there is soil-to-soil contact at the base of the structure

su
fsRh ≤ L (9.29)
fms

and su is the undrained shear strength of the soil.


For short-term stability where there is reinforcement-to-soil contact

α ′bc su
fsRh ≤ L (9.30)
fms

where there are reinforcing strips, the sliding will partly occur on the strips and partly on
soil, so the modifying factor α′ needs to reflect this.

9.7.2  Bearing failure


The weight of the reinforced earth block plus the vertical force due to surface loads plus the
effect of any horizontal loads and the active earth pressure behind the block will create a
356  Geomechanics in soil, rock, and environmental engineering

SL

FL w2
w1

1
2
hi Self-weight
3

Ti
i
Hℓ
Li
Resultant loading
acting on rear of wall

Figure 9.25 L​ oads acting on reinforced earth block.

non-uniform pressure distribution across the base of the block. The resultant force can be
treated as a uniform pressure block centred at an eccentricity e from the centreline of the
block. The length of the pressure block is then L − 2e as shown in Figure 9.25.
The pressure acting qr is then (BS 8006)

Rv
qr = (9.31)
L − 2e

where Rv is the resultant of all factored vertical load components.


The imposed bearing pressure should be less than the ultimate bearing pressure as shown
in Equation 9.32.

qult
qr ≤ + γDm (9.32)
fms

where
qult is the net ultimate bearing pressure of the soil beneath the reinforced earth block
(see Section 5.3.1)
Dm is the wall embedment depth
γ is the unit weight of the foundation soil

9.7.3 ​Rupture of the reinforcement


The load carried by the reinforcement Tj is assumed to be due to the horizontal force acting
on a section of the facing at the level hj of the reinforcing that has a vertical height of svj as
shown in Figure 9.25. Generally, svj is the vertical spacing of the reinforcing.
Retaining structures  357

The horizontal force per unit length on a strip j is therefore given by

 c′ 
Tj =  Kaσ ′vj − 2 Ka  svj (9.33)
 fms 

where σ ′vj is the factored vertical stress acting on the reinforcing at depth hj

Rvj
σ ′vj = (9.34)
Lj − 2 e j

Rvj is the factored vertical load acting on the jth layer of reinforcement. Forces due to other
loads such as surface loads can be included in the force Tj, and the methods for calculating
these are given in BS 8006.
The tensile strength of the reinforcing needs to be high enough to withstand the tensile
force applied and so we can write

TD
Tj ≤ (9.35)
fn

where
T D is the design strength of the reinforcement
fn is the partial factor for the economic ramifications of failure

9.7.4 ​Pull-out of the reinforcement


In the active wedge method of design, the reinforcement is checked for pull-out failure for
the length of reinforcing that is outside of the active failure wedge as shown in Figure 9.26.
This length for a reinforcing strip j is L ej.

1
Active Resistant
2
zone zone
3

i
H La Lei

45° + φ /2

Figure 9.26 Pull-out of reinforcing strips of length Le behind an active wedge.


358  Geomechanics in soil, rock, and environmental engineering

If the perimeter of the reinforcing strip is Pj, then the bond strength around the outside of
the strip consists of frictional and cohesive components, and the force required to pull the
strip out Fj is given by

Pj Lej  α ′bcc′Lej 
Fj = (ffs γhj + ff ws )α ′ tan φ′ +  (9.36)
f p fn  fms 

where
ffs is the partial factor applied to soil self-weight
f f is the partial factor applied to dead loads ws
f p is the partial factor for reinforcement pull-out resistance
fn is the partial factor for the ramifications of failure

9.7.5 ​Overall slip failure


Slip failure through or beneath the reinforced soil block is a possible mode of failure, so limit
equilibrium methods can be used to check the stability of the reinforced soil block. This is
discussed in Section 9.8.1, where the effect of reinforcing is included in an analysis using the
method of slices. Circular and non-circular failure surfaces can be trialled cutting through
either the reinforced soil block or ground beneath it as shown in Figure 9.24e.

9.7.6 ​Excessive deformation
The settlement of the block of reinforced soil can be calculated by hand using the simple
methods outlined in Section 5.5.2 or computed using finite element methods. Guidance is
provided in BS 8006 for the acceptable deformations of retaining structures and these are
reproduced in Table 9.4.

9.8 ​COMPUTER METHODS

Today, hand calculations based upon simple methods such as those listed in the previous
sections are used as checks on values obtained from computer programs or for simple cases.
For more complex cases where excavation is being carried out in stages and anchors are
being placed as excavation proceeds, computer based methods are used.
One such commercial code WALLAP is especially designed for performing calculations of
pressures on retaining structures and has the added advantage that it will carry out moment
and shear force calculations for sheet pile walls.

Table 9.4  ​Usually accepted tolerances for faces of retaining walls and
abutments (BS 8006)
Location of plane of structure Tolerance ±  50 ​mm
Verticality ± 5 ​mm per metre height
(i.e. ±40 mm per 8 ​m)
Bulging (vertical) and bowing ± 20 mm in 4.5 ​m template
(horizontal)
Steps at joints ± 10 mm
Alignment along top (horizontal) ± 15 mm from reference alignment
Retaining structures  359

13.20
10.00
Rockfill 7.75
Fill/alluvium Fill/alluvium 4.90
Upper alluvium Upper alluvium 1.90
Monoman sand Monoman sand
–3.00
–4.00
Sand Sand
–10.00
200 100 0 100 200
Water pressure (kN/m2)
Rough Smooth

Figure 9.27 Schematic diagram of sheet pile wall in a river.

Stage No. 2 Excav. to elev. 10.00 on passive side

Bending moment (kN m/m run) Displacement (m)


1000 0 –1000 –0.4000 0 0.4000
12 12
Passive GL
Active GL

6 6

Elev. Elev.

0 0

–6 –6

–200.0 0 200.0
Shear force (kN/m run)

Figure 9.28 Moments and shears in a sheet pile wall.

Figure 9.27 shows a sheet pile wall that was used to hold back water in a river while con-
struction was carried out on the dewatered side of the wall. A small berm of rockfill was
provided on the dewatered side to assist with stability of the wall.
Figure 9.28 shows the bending moments and shear forces (per m run) in the sheet pile wall
and the predicted deflection of the wall. The large deflection predicted at the top of the wall
of 350 ​mm was confirmed by monitoring of the wall deflection.

9.8.1 ​Limit equilibrium methods


Both circular and non-circular slip surface methods can be used to calculate the Factor of
Safety that retaining structures have against an overall failure. Such an analysis is shown in
Figure 9.29 where it can be seen that soil consisting of several layers and under the action
of a surface surcharge can be analysed. Anchors can be included in the analysis as shown
in the figure.
360  Geomechanics in soil, rock, and environmental engineering

Track surcharge :
108 kPa

General surcharge : 12 kPa


172 Elv. 171.2
Silty clay fill
170 Sandy silt/silty sand fill
168 Sand and silt (I)
Sand and silt (II) 10.2 m
166
164 Clayey silt (II)
162 Elv. 161.0
160
Elevation (m)

158 Clayey silt (I)


156
154 Silty clay
152
150
148
146 Silty sand/sandy silt
144
142
140
138
0 10 20 30 40 50 60 70

Figure 9.29 N
​ on-circular slip surface analysis of an anchored wall (SLOPE/W).

(a) A A
A A

Phase 4

(b) A A A
A

Phase 5

Figure 9.30 Design of anchored sheet pile walls using finite element methods: (a) first set of anchors placed;
(b) second set of anchors placed.
Retaining structures  361

The forces that are available in the reinforcing are applied as forces to the slice through
which the anchor or reinforcement passes at its base. This involves calculating the length
of reinforcement that lies outside the slip surface and calculating the pull-out force that is
acting on this length.
Another way to apply anchor loads is to distribute the anchor forces to all of the slices
that the anchor intersects. This gives a more stable solution and does not create large normal
forces on the base of the slice through which the reinforcement passes.
Limit equilibrium methods can be applied to problems involving

1. Anchors
2. Soil nails
3. Steel reinforcing (reinforced earth)
4. Fabric reinforced soil

The commercial stability analysis codes SLOPE/W (2014) or STARES (2014) can be used
to estimate the stability of walls held back or reinforced by structural elements such as those
mentioned above.

9.8.2 ​Finite element methods


Retaining wall design can be performed using finite element methods and the analysis of
excavation was examined in Section 8.8.1 in Chapter 8. The advantages of the finite element
method is that the excavation sequence can be modelled with sheet pile walls being added
and anchors being added as the excavation is deepened. Anchors can have a free length
(shown as black in Figure 9.30) or be grouted (grey in Figure 9.30). Loads can be applied to
the surface of the soil, and lowering of the water table can be simulated. The moment and
shear force can be calculated in the wall and the axial force in the anchors computed.
This page intentionally left blank
Chapter 10

Soil improvement

10.1 INTRODUCTION

If the ground beneath a site is not suitable for the purpose for which the site is intended, the
site may either have to be abandoned and a more suitable one found elsewhere, or the soil
may be improved by some means. The cost of improvement and the benefits achieved will be
critical in the decision to go ahead with soil improvement.
Generally, it is soft clays that pose most problems, but loose sands and dispersive soils
are other examples of soils that can be modified. Sites that are contaminated or landfill
sites are not considered in this chapter, but are examples of unsuitable sites that can be
modified. Some examples of rehabilitation of contaminated sites are given in Section 11.9
of Chapter 11.

10.2  SOFT SOILS

Soft clays are generally those that have been recently deposited and are typically found on
the flood plains of rivers, in estuaries, and in mangrove swamps.
If the soil has never been loaded to a greater stress than the current overburden stress (i.e.
due to the self-weight of the soil), then the soil is said to be normally consolidated. If the
soil has been subjected in the past to higher stresses than the current overburden stress, then
the soil is said to be overconsolidated. The maximum past pressure that has been applied
to the soil is called the pre-consolidation pressure pc′ . The London clay and Frankfurt clay
are examples of overconsolidated soils, as they were once loaded by stresses from thick ice
sheets that have since receded hence unloading the clay layer. As a result of being loaded
to greater pressures in the past, overconsolidated clays are stiffer and have greater shear
strength than normally consolidated clays.
Normally consolidated clays can have an overconsolidated crust and this is commonly
found in soft clay deposits. The overconsolidation is caused by wetting and drying of the
surface layer of the soil. The swelling and shrinking cause compression and then unload-
ing of the soil, and this causes it to be overconsolidated. The strength profile for a soft
clay deposit in Malaysia is shown in Figure 10.1. It may be seen from Figure 10.1, that
the ­undrained shear strength su is higher at the surface, reduces to a minimum, and then
increases linearly with depth, which is due to the crust being overconsolidated.
Soft normally consolidated clays are generally too soft for civil works and some kind of
improvement needs to be applied. Some of these techniques are discussed in the following
sections.

363
364  Geomechanics in soil, rock, and environmental engineering

Vane shear strength su (kPa)


0 20 40 60 80
0
+ ×
2 + ×
×
4 ×
Legend
×
+ Field tests 1–9
6
Lab test BH1
×
8 × Lab test BH2
+
Depth (m)

×
10 +×
×
12

14 × +
×
16 ×

18

20

Figure 10.1 Undrained shear strength versus depth for a normally consolidated soil with an overconsoli-
dated crust.

10.3  SURCHARGING AND WICK DRAINS

A common way to improve soft soil is to remove water from the soil by causing the soil to
consolidate under the weight of a surcharge. During consolidation, the effective stress in the
clay increases and therefore the shear strength and stiffness of the soil increases.
The water in the soil may be left to drain to the surface of the clay layer or to any pervious
layer of underlying soil (such as a sand layer), or the consolidation process may be speeded
up by providing wick or sand drains in the clay layer. This greatly reduces the distance that
any water in the soil has to move to reach a drainage boundary, and this speeds up the con-
solidation process. As the time for consolidation depends on a square power of the drainage
distance, a reduction of 10 in the drainage path will speed up consolidation by 100 times.

10.3.1 Surcharging
Often an embankment on soft soil is constructed to a height greater than the final height
so as to apply a surcharge to the soil and thereby speed up consolidation. The surcharge is
often removed before complete consolidation of the soil has taken place and monitoring of
the settlement or of pore pressures can be used as a guide as to when the surcharge can be
removed. A method commonly used to estimate when primary consolidation is complete is
Asaoka’s method (Asaoka 1978). In this approach, the settlement at a time t (St) is plotted
against the settlement at a time t − Δt (St−1) as shown in Figure 10.2. The time increments Δt
are taken to be equal and this may require some interpolation between recorded settlement
points. Where the line through the field data intersects the 1:1 line gives the point at which
primary consolidation is complete.
Soil improvement  365

320

270

220
St (mm)

170

120

70 Field data

1:1 line

20
20 70 120 170 220 270 320
St–1 (mm)

Figure 10.2 An Asaoka plot used to estimate completion of primary consolidation.

For wide embankments, the settlement can be calculated using Terzaghi’s one-dimensional
theory and this may be done by testing a sample of the soil in an oedometer. An oedometer
contains a ring holding the soil, and the soil is loaded vertically between two porous plates.
Because the sample is confined, it is compressed under one-dimensional conditions (see AS
1289.6.6.1 1998 or ASTM D2435-04 2004).
From the oedometer test, we can make a plot of the void ratio e versus the vertical effec-
tive pressure applied p′ (usually plotted to a log scale). The pressure applied in the oedometer
test once all pore pressures have dissipated is the effective pressure p′ and this is what causes
the soil to deform. The change in height of the sample (or its settlement S) is given by

∆e
S = δz (10.1)
1 + e0

where δz is the sample height. This equation can be used to calculate the settlement of a layer
of soil where the stress goes from A to B as shown in Figure 10.3.
The soil will compress at slope Cr on a plot of void ratio e versus log p′ (the slope of line
A–P) until it reaches the pre-consolidation pressure pc′ after which the soil will compress
more readily at a slope Cc (the slope of line P–B). This information can be used to predict the
amount of settlement with embankment height assuming that all pore pressures are allowed
to dissipate. The definitions of the compression index Cc and the recompression index Cr are
(see Figure 10.3) given by Equation 10.2.

e0 − ei  p′ 
Cr = = ∆e log  c 
log pc′ − log p1′  p1′  (10.2)
ei − e2  p′ 
Cc = = ∆e log  2 
log p2′ − log pc′  pc′ 
366  Geomechanics in soil, rock, and environmental engineering

A Pre-consolidation
e0 P
pressure
ei

Void ratio

Unload–reload path
B
e2

p′1 p′c p′2


log p′

Figure 10.3 Plot of void ratio versus effective stress.

We can find the change in void ratio Δe from Equation 10.2 depending whether the
changes in void ratio occur before or after the pre-consolidation pressure and calculate
the settlement by substituting for the void ratio change in Equation 10.1. If the original
overburden pressure in the ground at any depth is σ ′v0 and the increase in stress due to an
embankment is ∆σ ′v so that the final stress is σ ′f , the vertical settlement of a layer of soil δz
thick is given by

Cr  p′  Cc  σ ′f 
S = δz log  c  + δz log   (10.3)
1 + e0  σ ′v0  1 + ei  pc′ 

The above equation is based on the assumption that the pre-consolidation pressure lies
between the initial stress and the final stress in the ground, that is, we are going from A to
B in Figure 10.3.
If the pre-consolidation pressure is greater than the final stress (the soil remains overcon-
solidated), we would only use the first part of the equation, that is,

Cr  σ ′f 
S = δz log  (10.4)
1 + e0  σ ′v0 

If the soil is normally consolidated so that there is no pre-consolidation, we would only


use the second part of the equation, that is,

Cc  σ ′f 
S = δz log  (10.5)
1 + e0  σ ′v0 

The process in applying the one-dimensional consolidation theory to embankments is


to divide the soil up into layers δz thick and calculate the initial and final effective stresses
at the centre of each layer. Then the settlement of each layer can be found from either
Soil improvement  367

Equation 10.1 or Equation 10.3, and the settlement of all layers added up to obtain the
overall settlement.
The rebound on unloading (surcharge removal) can be calculated using C r as the soil will
unload along the unload path shown in Figure 10.3.
The rate of primary consolidation can be found for wide embankments by using one-
dimensional theory. The degree of consolidation can be found by calculating the time factor
Tv and using a plot such as that of Figure 5.26 to find the degree of consolidation Uv. The
settlement at any time can be estimated from St = UvS ∞ where the final settlement S ∞ is cal-
culated from either Equation 10.1 or Equation 10.3. More advanced numerical methods are
discussed in Section 10.13.
Additional settlement can occur through creep of the soil, especially in soft clays.
Allowance can be made for this by calculating the creep settlement St from (Mesri 1973)

Cα t
St = δz log   (10.6)
1 + e0  t0 

where C α is the coefficient of secondary compression or creep, t is the time, and t 0 is the
s­ tarting time of the creep (often taken as 90% of primary consolidation). From the for-
mula, the creep settlement is linear on a logarithm of time scale, as this is often observed
for clays.
The coefficient of secondary compression is related to compression index C c and so esti-
mates of C α can be made from the compression index as indicated by Mesri and Godlewski
(1977) and shown in Table 10.1 (see also Figure 1.10 for direct measurement).
Once a surcharge has been removed, the clay beneath an embankment is partially
­overconsolidated, and so the creep rate reduces. The formula of Equation 10.7 may be
used (Wong 2007, 2010) to estimate the value of the creep coefficient for the overconsoli-
dated soil C αε(OC) from the value for a normally consolidated soil C αε(NC) where OCR is the
­overconsolidation ratio.

Cαε (OC) 1−m


= (OCR − 1)n + m (10.7)
Cαε (NC) e

and Cαε(NC) = Cα(NC) /1 + e0 .


The quantities m and n need to be found from experiment, but typical values are m = 0.1
and n = 6. If the OCR is large, then the ratio of Equation 10.7 tends to m, therefore m is
equal to the ratio Cr/Cc. A plot of experimental data is shown in Figure 10.4, along with
values from the expression of Equation 10.7 where m = 0.05 and n = 6.

Table 10.1  Relationship of C α and Cc


Material Cα/Cc
Granular soils including rockfill 0.02 ± 0.01
Shale and mudstone 0.03 ± 0.01
Inorganic clays and silts 0.04 ± 0.01
Organic clays and silts 0.05 ± 0.01
Peat and muskeg 0.06 ± 0.01
368  Geomechanics in soil, rock, and environmental engineering

1.0
Ng (1998)
Ladd (1989)
0.8 Ballina Bypass (2007)
Gateway upgrade (2007)
Exponential fit (m = 0.05 and n = 6)

(OC)/Cα (NC)
0.6

0.4
Cα ∋

0.2

0.0
1.0 1.2 1.4 1.6 1.8 2.0
OCRf

Figure 10.4 Reduction of coefficient of secondary consolidation with OCR. (After Wong, P.K. 2010. Australian
Geomechanics Society Seminar on Ground Improvement, Perth, WA.)

The time at which creep begins ts after removal of a surcharge to the time that the sur-
charge is removed tR is found to be given by

t 
{
log  s  = 0.0208 (100Rs′ ) ± 0.156
 tR 
} (10.8)

and Rs′ = OCR −1. The rebound due to unloading is shown in Figure 10.5.

10.3.2  Field observations


Generally, embankments on soft soils are instrumented when they are constructed so that
the progress of consolidation can be monitored. Settlement plates are often placed at the cen-
tre of the embankment, and piezometers placed along the centreline. Other instrumentation

tP – End of primary consolidation


∋ tR – Remove surcharge
υ
tS – Start of secondary consolidation

C
Ps = Σ(HiCα′ log t/ts) 1 C′
1

0.01 0.1 1 10 102 103 104


log (t/tp)

Figure 10.5 Pre-loading to reduce post-construction settlement.


Soil improvement  369

such as slope indicators is placed at the toe of the embankment to monitor lateral movement
with depth.
For embankments with no drains, Tavenas and Leroueil (1980) have shown that the pore
pressure response in a soft clay foundation is typically as shown in Figure 10.6. Initially,
the clay is in an overconsolidated state, and the pore pressure increases along the line shown
as B1 . The pore pressure coefficient is defined as the increase in pore water pressure Δu
divided by the increase in vertical principal stress Δσ1, that is,

∆u
B = (10.9)
∆σ1

Once the soil reaches its pre-consolidation pressure, then it yields and the pore pressure
rise is about equal to the increase in applied pressure from the embankment. This is shown
as B2 = 1.0 in Figure 10.6. At failure, the pore pressures begin to increase more rapidly than
the increase in embankment pressure, and the value of Bf >1.
This illustrates the need to construct trial embankments to a height greater than the pre-
consolidation pressure in the soil (usually >2 m high), or there will be no yield of the soil
and the embankment response will not be the same as for the full height ­embankment. In
addition, when modelling a clay foundation, it is necessary to use a numerical model that
allows yield and failure such as the Cam Clay type models (see Section 3.10).
Tavenas and Leroueil (1980) give several formulae for observed behaviour of embank-
ments on soft clay. The first is the threshold height of an embankment Hnc at which the clay
beneath the embankment first begins to yield

(σ ′p − σ ′v0 )
γH nc = (10.10)
I(1 − B1)
. 0

.0
=1

>1
Excess pore pressure Δu

Δσ u
= Δ



f
B
B

F
Δue
0
1.


=
2
B

Δσ′v threshold
P

B1

Δσv threshold Δσve = I γH


Δσ ′p − σ′v0 Applied vertical stress Δσv = IγH

Figure 10.6 Pore pressure response beneath an embankment.


370  Geomechanics in soil, rock, and environmental engineering

In the above equation, I is an influence factor that gives the vertical stress at any depth
from elasticity theory, γ is the unit weight of the embankment fill, σ ′p is the pre-consolidation
pressure of the clay and σ ′v0 , is the initial effective stress in the ground.
During the initial construction of an embankment, when the clay is still in an overcon-
solidated state, Tavenas and Leroueil (1980) suggest the settlement can be calculated from

 D   σ ′p − σ ′v0 
S =  nc   (10.11)
 m  σ ′p 

where Dnc is the depth of clay in a normally consolidated state and m is the modulus number
of the soil (which is of the order of 25–80 in the intact Champlain clays that Tavenas and
Leroueil investigated).
Beyond the threshold height, the change in settlement ΔS of an embankment is
approximately

∆S
= 0.07 ± 0.03 (10.12)
H − H nc

For changes in lateral movements Δym of embankments with side slopes in the range of
1.5–2.5 horizontal to 1 vertical, the following formulae are suggested:

a. Below the threshold height: ∆ym = (0.18 ± 0.09)∆s


(10.13)
b. Above the threshold height: ∆ym = (0.91 ± 0.02)∆s

10.3.3  Sand or prefabricated vertical (wick) drains


If consolidation beneath an embankment or other structure is likely to be too slow, then the
consolidation can be speeded up by the use of sand or wick drains.
Wick drains (often called prefabricated vertical drains PVDs) consist of a plastic core (to
allow drainage) covered by a geotextile filter fabric (see Figure 10.7). The wicks are pushed
into the ground with a steel mandrel by a special crane, and are wound from a reel as they
are pushed into the ground.
Sand drains are formed by drilling a hole in the clay and filling the hole with free draining
sand. Water squeezed from the clay by any surcharge can flow into the sand drain and then

Corrugated plastic
core for drainage

Geotextile filter

Figure 10.7 Section through a wick drain (PVD).


Soil improvement  371

to the surface. Stone columns, as discussed in Section 10.5 also act as drains and help speed
up consolidation as well as strengthening the soil. A typical surcharge and subsurface drain
layout is shown in Figure 10.8.
Where settlements are large (2 m or more is common in soft clays), the PVDs may become
kinked or bent. Care must be taken that this is avoided as the flow of water from the soil
will be cut off and the drains become ineffective.
Consolidation theory may be applied to the prediction of consolidation rates of soils
drained by wick or sand drains. Each individual drain can be considered to collect water
from a cylindrical volume of soil around itself, therefore the problem reduces to that of
analysing a circular cylinder of soil with an impermeable outer surface but for which flow
can occur into the central drain or to the top and bottom surfaces. Such a drain is shown in
Figure 10.9 where a soil layer of depth 2H is drained by a well of radius rw. If the drain is
of the shape shown in Figure 10.7, the equivalent diameter can be calculated as the average
dimension, that is, dw = 2rw = (L + t)/2. Depending on whether the drains are laid out in a
triangular pattern or in a square pattern, the cylindrical region which is drained by each
well can be determined and its diameter is found. The diameter of the cylindrical region of
soil is given by de, where

de = 1.05 × well spacing for a triangular pattern


(10.14)
de = 1.14 × well spacing for a square pattern

This is shown for an isolated single drain in Figure 10.9.


The equation governing the radial and vertical flow may be written as

∂u  ∂ 2 u 1 ∂u   ∂2u 
= cr  2 + + c (10.15)
r ∂r 
v
∂t  ∂r  ∂z 2 

where u is the excess pore pressure, r is the radial distance, z is the vertical distance, and cr
and cr are the coefficients of consolidation in the radial and vertical directions, respectively.

Stability berms Settlement plate


(if required) Temporary surcharge fill
Piezometer gauges Permanent fill
Sand drainage blanket

Water Piezometers
drainage
pattern

Wick or sand drain


Firm soil or rock Clay soil

Figure 10.8 Typical wick drain (PVD) layout beneath an embankment.


372  Geomechanics in soil, rock, and environmental engineering

Well
Definitions Example: To determine equivalent
Spacing radius of drain without smear whose
For triangular pattern effect is equal to the actual drain
de = 1.05 (well spacing) with smear
For square pattern
a de a de = 1.14 (well spacing)

Drain well
de = Effective diameter
of sand drain

rw = 1.5 ′
Drain wells in rs = 1.8 ′
trianglar pattern re = ~7.5 ′

Actual sand drain


de
n = 5 kh/ks = 7
kh = Horizontal permeability Estimated s = 1.2
Drain Well Determine neq from Figure 10.11
ks = Shear zone permeability
ks 7.5
w de r neq = 15 rw = = 0.5 ′
No flow across Floaths kv h n= = e 15
outer surface p dw rw

Drain well
kh Effective radius
2H =
Radius of drain
rs
s=
rw rw = 0.5 ′
Smear zone
rw Radius of shear zone re = 7.5 ′
=
rs Radius of drain
re Equivalent sand drain no smear
Section a-a

Figure 10.9 Allowance for smear effect in a sand drain (or PVD) design.

Equation 10.15 may be solved in two parts; firstly for the radial flow (with c v = 0), and
secondly for the vertical flow (with cr = 0). The vertical flow solution is merely the Terzaghi
one-dimensional consolidation solution.
If the degree of consolidation for the radial consolidation is Ur and for the purely verti-
cal flow is Uv, then the degree of consolidation when both vertical and horizontal flow are
operative Urv is given by Carillo (1942) as

(1 − Urv ) = (1 − Ur )(1 − Uv ) (10.16)

Plots that can be used to compute the degrees of consolidation for both radial and vertical
flow are shown in Figure 10.10. The broken curve is Terzaghi’s one-dimensional solution for
vertical flow only and is used with time factor Tv. The solid curves are used for the degree
of radial consolidation Ur and are used with time factor Tr. The time factors are defined as

cv t
Tv =
H2
(10.17)
ct
Tr = r 2
4re

For the vertical time factor, the full soil depth is taken as 2H (two-way drainage) or H
(one-way drainage) and t is time. For the radial time factor, re = de /2 is the effective radius.
Soil improvement  373

Several curves are provided for the radial flow solution, and these have been obtained for
different values of the parameter n where

de r
n = = e (10.18)
dw rw

and rw is the radius of the well. By knowing the radial and vertical time factors, the degrees
of radial and vertical consolidation may be found from Figure 10.10, and then the degree of
consolidation for both radial and vertical flow are found from Equation 10.16.
The above theory applied to problems involving wick or sand drains has sometimes been
found to give inaccurate predictions of consolidation rates. This is due to the fact that when
a hole is drilled or a wick drain is pushed into the soil, the sides of the hole are remoulded or
‘smeared’. The smeared zone has a lower permeability than the surrounding soil and slows
down the flow of water into the drain.
The procedure for designing a drain layout considering smear can be summarised as
follows:

1. The time factor Tv for vertical drainage after a time t is calculated, and the degree of
consolidation for vertical drainage Uv found from the broken line on Figure 10.10.
2. If the degree of consolidation Uv without drains is greater than that needed, then
drains are not needed. However, if Uv is less than the required degree of consolidation
Ureq, then drains will be needed as vertical drainage to the upper (and perhaps lower)
drainage boundaries will not provide adequate pore pressure dissipation.
3. The required degree of consolidation from radial drainage Ur may be calculated from
Equation 10.16 as

 (1 − U req ) 
Ur = 1 − (10.19)
 (1 − Uv ) 

0 100

10 90

20 80

Values of n
– (%)
Consolidation, Uv, Ur (%)

30 70
100
Excess pressure, u–v, ur

40 60 40
10
50 50
5
60 40 (a) Vertical flow
70 30 (b) Radial flow

80 20

90 10

100 0
0.004 0.01 0.04 0.10 0.40 1
Time factor Tv, Tr

Figure 10.10 Rate of consolidation for vertical and radial drainage by a sand drain (or PVD).
374  Geomechanics in soil, rock, and environmental engineering

100
Relation of actual and equivalent n values (kh/ks)
70
2 24 2 48 12 4 8
50
40 Legend
30 s = 1.2
s = 1.5
20
s = 2.0 s=1
Actual n

10
7
5 Example
4
2 24 2 48 4 12 8 (kh/ks) 8
3
2

1
1 2 3 4 5 7 10 20 30 4050 70 100 200 400 700
Equivalent neq = n for drain with no smear

Figure 10.11 Allowance for smear effects in a sand drain (or PVD) design.

4. A drain diameter dw (and hence r w) is chosen and a drain spacing is selected. From
Equation 10.14, the radius of influence may be found from knowledge of the drain
layout pattern (see also Figure 10.9). Hence, the ratio n = re /rw may be calculated.
5. From the estimated values of s = rs /rw and kh /ks, an equivalent value of n called neq can
be found by using Figure 10.11.
6. For this value of neq and Tr, the value of Ur can be found from Figure 10.10. The effect
of using an equivalent n value instead of the actual n value with Figure 10.10 is to slow
down the consolidation rate.
7. The procedures in Steps 4–6 can be repeated for different drain spacings (and drain
diameters if desired) and a plot made of Ur versus drain spacing can be made.
8. From this plot (see Figure 10.12), the spacing necessary to achieve the required degree
of consolidation (as calculated in Step 3) can be obtained.

100
Value required to
achieve overall
consolidation after time t

degree of
consolidation of
Radial degree of

Usp

Theoretical curve for


Required chosen drain diameter
spacing
0
Drain spacing s

Figure 10.12 Finding the required drain spacing.


Soil improvement  375

The permeability ratio can be found by performing oedometer tests on undisturbed soil
(kh) and on remoulded soil (ks), and calculating the permeabilities from the tests. Vertical cv
and horizontal ch coefficients of consolidation can be found by performing oedometer tests
with the clay sample oriented in the appropriate direction.

10.3.4  Vacuum consolidation


Consolidation using PVD systems can be accelerated by applying a vacuum to the drains
so as to create a greater pressure difference between the water in the drains and the clay.
A typical vacuum setup is shown in Figure 10.13.
This system was trialled near Ballina in NSW (Kelly et  al. 2008) where a vacuum of
80 kPa was applied to the PVD system in soft clay. The top 0.5 m of clay had an undrained
shear strength of 5 kPa and below 0.5 m it increased at 1 kPa/m depth. An impervious mem-
brane was used to seal the system of drains as shown in Figure 10.13.
As a result of the trial, it was found that embankments constructed using a vacuum con-
solidation system were able to be built more rapidly and to greater heights than embank-
ments constructed using standard surcharge filling methods because the vacuum pressure
enhances the stability of the embankment.

10.4 VIBROFLOTATION

Loose sands are very suitable for densification through vibration or shock, since the sand
particles will rearrange and form a more compact structure if vibrational energy is applied.
Sands that have a low relative density will liquefy on being vibrated by earthquake, and
therefore loose sand deposits are not suitable for foundations or as fill for wharf structures
or tailings dam embankments wherever there is any risk of an earthquake.
Vibroflotation is a process that involves improving granular soils through the application
of vibration, and was first established in Europe in the 1930s. The method involves push-
ing the vibroflot into the ground with the aid of water jets and vibration. Loose granular
materials such as sands and gravels can be densified with the vibration, and so their bearing
capacity can be improved and they become less compressible.

Vacuum gas
Atmospheric pressure phase booster
Impervious Vacuum air
membrane water pump
Surcharge
fill
Airflow
Draining
layer
Horizontal Water
drains treatment
station
Peripheral trenches filled
with bentonite and
polyacrylate

Vertical vacuum
transmission
pipes

Figure 10.13 Schematic diagram of vacuum consolidation.


376  Geomechanics in soil, rock, and environmental engineering

(a) 1 2 3

Feed
of
soil

Water Water Compacted


feed feed zone

Vibro-compaction method

(b) 1 2 3

Feed
of soil
material

Water
feed Compacted
column

Vibro-replacement method

Figure 10.14 The vibroflotation process: (a) vibro-compaction; (b) vibro-replacement.

The vibro-compaction process is shown in Figure 10.14a, where the vibroflot is advanced
into the soil using water jets that emerge from the end of the device. When the vibroflot is raised
vibration and the injection of water help compact the granular material around the vibroflot
probe. In addition, extra soil is added to the hole to compensate for the compacted material.
The process of vibroflotation has also been adapted to clayey or cohesive soils that can-
not be compacted by vibration. It involves replacing the cohesive material with a column
of granular material that can be compacted in place by vibration. This process is generally
referred to as vibro-replacement.

10.5 VIBRO-REPLACEMENT

Replacement of soft clays or silts with stronger, free draining materials can be used where it
is necessary to reduce settlements under structures or embankments. The process can also
Soil improvement  377

be used where it is necessary to speed up consolidation as well as strengthen the soil as the
granular columns of material act as drains. A further use is for reducing the liquefaction
potential of soils subject to seismic vibration. In this case, the granular column of material
also acts as a drain and dissipates the pore pressures in the soil that can lead to liquefaction.
This section is mainly concerned with vibro-replacement processes which can be used for
constructing stone or sand columns in soft clays and silts.
Figure 10.14b shows the vibro-replacement technique. The probe of the vibroflot is again
advanced into the soft soil by vibration and water jets, however as it is withdrawn, granular
material is introduced into the hole. The granular material can be coarse gravel, or crushed
rock or slag generally graded from 20 to 80 mm. The diameter of the column formed depends
on the properties of the clay. Larger diameter columns are formed in softer clays and smaller
diameter columns in stiffer clays. Typical column sizes are from 0.5 to 1.0 m.
Another alternative to the crane hung vibrator is the Bullivant method that is used in the
United Kingdom. The probe of the vibrator is tapered and the vibration is applied vertically.
As well, a pull-down force is applied to the probe from winches on the probe mast. This
method can achieve better densities in the material around the probe and can avoid the need
to use pre-boring in stiffer soils.
Design of stone or sand columns requires both settlement and bearing capacity calcula-
tions to be carried out in a similar fashion to a pile or surface foundation. These two aspects
are examined in the following sections.

10.5.1  Bearing capacity analysis


The bearing capacity of stone or sand columns can be carried out individually or for the
group. It may be necessary to compute the individual carrying capacity if only a few col-
umns are used under a foundation or the single column capacity is used in formulae for the
group capacity.

10.5.1.1  Bearing capacity of single columns


The load carrying capacity of stone columns is different to that of conventional steel or con-
crete piles. This is because they compress and bulge under applied loads much more than a
conventional pile. The column may undergo relatively large lateral distortion to a depth of
four diameters under the ultimate load, and therefore an important component of the load
carrying capacity is the maximum resistance that the soil can provide in the radial direction.
Greenwood (1970) assumed that the lateral resistance from the soil would create a triaxial
stress system within the column, and that the lateral pressure would be the passive pres-
sure of the soil (at failure). A more sophisticated design method is to use cavity expansion
theory to obtain the lateral pressures on the column. If the lateral expansion of the column
is idealised as the expansion of a cylinder in an elastic–perfectly plastic material, then the
limiting radial stress σRL can be obtained from the theory of Gibson and Anderson (1961)
for clay soils.

  G
σ RL = σ R0 + su 1 + log e    (10.20)
  su  

where
σR0 = total in situ lateral stress
G = shear modulus of the clay =E′/2(1 + ν′) = Eu /2(1 + νu)
378  Geomechanics in soil, rock, and environmental engineering

E′, Eu = drained, undrained modulus of the clay


su = undrained shear strength of the clay
ν′, νu = drained, undrained Poisson’s ratio for the clay

The results of quick pressuremeter tests (undrained tests) show that Equation 10.20 can
be approximated by

σ RL = σ ′R0 + u + 4su (10.21)

where
σ ′R0  = effective in situ lateral stress
u = pore water pressure

If the gravel in the bulged zone has yielded, then from the Mohr–Coulomb failure criterion

 1 + sin φ′ 
σ ′v =  σ ′RL (10.22)
 1 − sin φ′ 

where
σ ′v  = vertical effective stress in the column
ϕ′ = angle of internal friction of the column
σ ′RL  = lateral effective stress

The ultimate load that the column can carry is then given by

 1 + sin φ′ 
σ ′v =  (σ R0 + 4su − u) (10.23)
 1 − sin φ′ 

Equation 10.23 relies on the assumption that the load is applied to the column only, whereas
it may be applied to both the column and the surface of the soil. There is also the possibility
that the column may fail like a pile, and in this case, the usual pile bearing capacity formula
can be used. For clay, the ultimate base pressure may be taken as 9su and the ultimate shaft
load computed using the full undrained shear strength of the soil.
This may be used to design a column so that it will fail in bearing and bulging at about the
same load. This is not so straightforward if the shear strength of the clay varies with depth
and a choice for the value of su for use in Equation 10.23 must be made.
Dunbavan and Carter (1994) noted that the radial effective stress ultimately increases
by approximately four times the original undrained shear strength of the clay only under
fully drained conditions. Research on driven piles (Carter et al. 1979) has shown that in the
short term, the limiting value of the radial effective stress after cavity expansion is increased
by approximately 2.5–3 times the undrained shear strength of the soil. Francescon (1983)
has validated these predictions for normally and lightly overconsolidated inorganic clay. As
some consolidation will take place during loading, the value is probably somewhere between
2.5su and 4su , therefore for the undrained case, Equation 10.21 with a value of 2.5su (replac-
ing the value 4su) should be used.

 1 + sin φ′ 
σ ′v =  (σ ′R0 + 2.5su ) (10.24)
 1 − sin φ′ 

Soil improvement  379

30
σv

ϕ′ su
25
Vesic (1972) Gibson and
Anderson
(1961)

20

Hughes and Withers


(1974)
σv /su

15
Hughes et al.
(1975)
Greenwood
(1970)
10
Bell (1915)

5
Field data

0
35° 40° 45°
Internal friction angle of granular material (ϕ′)

Figure 10.15 Ultimate load carried by a single stone column. (From Bergado, D.T. and Lam, F.L.L. 1987. Soils
and Foundations, Vol. 27, No. 1, pp. 86–93.)

Measured data by Bergado and Lam (1987) is shown in Figure 10.15 along with predic-
tions of other investigators. The measured values are reasonably close to the values predicted
by Equation 10.24 if the lateral soil effective stress σ ′R0 is close to zero (as it would be near
the surface where bulging would occur).
Thornburn (1975) has presented an empirical design method initially recommended by
Thornburn and MacVicar in which the total building loads are supported entirely by the stone
columns. Thornburn considers such an approach ensures adequate factors of safety against a
bearing capacity failure and provides the ground with considerable stiffness. In Figure 10.16,
the relationship between the allowable working load recommended for preliminary design, and
the undrained shear strength of the cohesive soil is reproduced. This relationship was obtained
from the Rankine theory of passive earth pressure modified for radial deformation and from
correlating field measurements of the average diameters of stone columns with the undrained
shear strengths of the soils in which they were constructed. This correlation was established
using performance data from stone columns formed by Cementation and Keller vibrators.
Allowable stress that can be placed on stone columns is shown in Figure 10.17 together
with the values derived from the theory of Hughes and Withers (1974). The ultimate load
values as computed from Hughes and Wither’s theory have been divided by 3 to compute the
allowable stress levels shown in the plot.
380  Geomechanics in soil, rock, and environmental engineering

150 0.6

Effective diameter of stone column (m)


Allowable load on stone column (kN)
125 0.7

100 0.8

75 0.9
Allowable load
Effective diameter

50 1.0
10 20 30 40 50 60
Undrained shear strength su (kN/m)2

Figure 10.16 Allowable load on a stone column. (After Thornburn, S. 1975. Géotechnique, Vol. 25, No. 1,
pp. 83–94.)

The methods described above for determining the load carrying capacity of a single isolated
stone column should be sufficiently accurate for a wide variety of sites (Hughes et al. 1975).
However, the predictions made using these methods will become less reliable if the cohesive
soil deposit is not uniform, or the surface loading is actually applied to the surface of the soil.

10.5.1.2  Bearing capacity of column groups


The bearing capacity of a group of stone columns will be different to that of a single column
just as the bearing capacity of a group of piles is different to that of a single pile.

600

Stone column
Allowable stress on stone column (kN/m)2

500 Thorburn (1968)


Hughes and Withers
(1974)
400

300
Circular footing
5.14 × 1.3su
qA =
200 3

100

0
0 25 50 75
Undrained shear strength su (kN/m)2

Figure 10.17 Allowable stress on stone columns.


Soil improvement  381

Four modes of failure are possible for a group of stone columns:

1. A general bearing capacity failure similar to that which occurs for a conventional foot-
ing placed on the surface of a deep clay layer.
2. Squeezing of the clay between the stone columns, that is, between the surface of the
soil and the base of the layer.
3. A sliding failure through the material applying the load (e.g. an embankment or stock-
pile) and through the stabilised clay.
4. An end bearing failure with the columns pushing into the clay and behaving as stiff
piles.

10.5.1.2.1  Failure mode (1)


Bearing capacity factors are needed for a loading on a soil layer of finite depth, and such
factors have been produced by Mandel and Salençon (1969). Generally, the layer thickness
has to be taken into account as the loading on the stone columns is of large extent compared
to the thickness of the soil layer.
The bearing capacity factors are a function of B/H where

B = footing width
H = thickness of soil layer

The bearing capacity coefficients are presented in Tables 5.3 through 5.5 in Chapter 5 on
‘Shallow Foundations’.

10.5.1.2.1.1  DRAINED LOADING

For long-term or slow-loading cases, the drained friction angle of the clay needs to be used.
As the proportion of the area taken up by the stone columns is small and because the angle
of friction of the column material is usually close to that of the soil (typically friction angles
for the clay are 30° and that of the column material is 40°), the bearing capacity can be
calculated using the angle of friction of the clay, and treating the column-reinforced soil as
a uniform material.

10.5.1.2.1.2  UNDRAINED LOADING

A simple method of calculating the bearing capacity in undrained clay is to use the formula

σ ult = (1 − α s )σ c + α sσ s (10.25)

where
αs = cross-sectional area of stone/gross cross-sectional area.
σc = maximum vertical stress that the clay can support. This corresponds to the value
calculated using bearing capacity theory for a uniform layer of clay without stone
columns.
σs = maximum stress that the stone column can support. This corresponds to a bulging
failure of the column.

It is recommended that Equation 10.24 be used when calculating the value of σs.
382  Geomechanics in soil, rock, and environmental engineering

10.5.1.2.2  Failure mode (2)


Where the width of the load to the depth of the clay layer is large, the clay can squeeze out
sideways past the stone columns. In this case, the ultimate load that can be applied can be
estimated from the same formula as given in Equation 10.25 but where in this case, the
value of σc can be computed from the solution for a layer of cohesive material being squeezed
between two rough rigid plates for which

 w
σ c ≈ su  4 + π +  (10.26)
 h

where
su = undrained shear strength of the clay
w = the width of the loaded area
h = the thickness of the clay layer

The maximum vertical stress that the columns can support is calculated using Equation
10.24 that uses a value of 2.5su for the maximum lateral restraint provided by the clay.

10.5.1.2.3  Failure mode (3)


The stability of the stone columns can be assessed by carrying out a slip circle analysis;
however, care needs to be taken in doing so. This is because the stone columns are stiffer
than the surrounding clay and so they carry a greater normal stress. This makes the columns
stronger in resisting shearing, as the shear strength is proportional to the normal stress in a
frictional material.
An equivalent cohesion and angle of friction may be used in the slip circle analysis (Priebe
1995) where the cohesion and angle of friction of the composite material c eq and ϕeq are
given by Equations 10.27 and 10.28.

ceq = (1 − α s )cc (10.27)


tan φeq = m tan φ s + (1 − m)tan φc (10.28)


and cc is the cohesion of the clay, ϕs and ϕc are the angles of shearing resistance of the stone
column and the clay, respectively, and

As A
αs = = s (10.29)
As + Ac A

where A s is the cross-sectional area of the stone column, A is the total area of a unit cell, and
αs is the area ratio (area of stone columns to total area).
The stress carried by the stone columns is higher than just the stress from an embankment
as the stress is concentrated into the stiffer stone columns. Priebe (1995) has suggested using
Equation 10.28 in any slip circle analysis as it takes account of the stress concentration.
Values of m are provided in the plot of Figure 10.18. Priebe (1995) recommends using the
full lines in this plot rather than the broken lines as they allow for bulging of the columns.
Soil improvement  383

1.0
Dashed lines:
m = (n–1+Ac/A)/n
0.8
vs = 1/3
Proportional load m

ϕc = 45°
0.6
ϕc = 42.5°
ϕc = 40°
0.4
ϕc = 37.5°
ϕc = 35°
0.2 Solid lines:
m = (n–1)/n

0.0
1 2 3 4 5 6 7 8 9 10
Area ratio A/Ac

Figure 10.18 Proportional load carried by stone columns. (Priebe, H.J. 1995. Ground Engineering, Vol. 28,
No. 10, pp. 31–37. Reprinted with permission from Ground Engineering.)

He also suggests using a reduced equivalent cohesion (Equation 10.30) to allow for damage
to the soil structure where m is taken from the full lines in Figure 10.18.

ceq = (1 − m)cc (10.30)


10.5.1.2.4  Failure mode (4)


This failure mode considers the column to behave as a stiff pile. Failure will occur if the
load on the column is equal to or greater than the ultimate load that the column can carry.
The ultimate load that the column can carry can be computed as for a pile, that is, the sum
of the side friction force on the column (computed using the full shear strength su) and the
base resistance = 9subase Abase.

10.5.2  Settlement analysis of column groups


When the load is applied to the stabilised soil deposit, the clay will deform under undrained
conditions, while the stone columns remain drained. The deformation of the column group
depends upon whether the load can be considered to be rigid or flexible.

10.5.2.1  Flexible foundation


Finite element analyses of stone columns using representative properties for the clay and the
stone column (Balaam and Poulos 1978) have shown that elastic solutions for the settlement
of the column is sufficiently accurate if the applied load q satisfies

γh
q ≤ (10.31)
2

where
γ = the unit weight of the clay
h = the depth of the clay layer
384  Geomechanics in soil, rock, and environmental engineering

100

de/d = 2 2 5 5 10 10

10
h/d

de

Ep/Es′ = 20
L
h L/h = 1.0
d L/h = 0.5

1
0 0.2 0.4 0.6 0.8 1.0
Settlement ratio

Figure 10.19 Settlement ratio from finite element analysis.

Results for elastic solutions are shown in Figure 10.19. In this plot, the settlement ratio
is defined as the ratio of the maximum settlement occurring if a column is present to the
settlement that would occur if there were no pile.

10.5.2.2  Rigid foundation


A column-clay unit such as that shown in Figure 10.20 can be used to compute the settle-
ment of the stone column group. A solution can be obtained by assuming that the unit
remains in an elastic state (Balaam and Booker 1981).
Shown in Figure 10.21 is the reduction in settlement with respect to the spacing of the
stone columns a/b. The reduction in settlement is given by εz /qAmvs which is equal to the
strain in the column-clay unit divided by the strain in the layer of clay without any column.
It can be seen that there is a rapid reduction in settlement up to an a/b ratio of about 0.4
for all ratios of the column to clay stiffness ratios Ep /Es′.
If the stone column yields (with the clay remaining plastic), then the settlement of the col-
umns can be computed by treating the stone column as an elastic–perfectly plastic material.
Correction factors that can be applied to the elastic solutions are presented in Figures 10.22
and 10.23. These correction factors are only for certain cases where de /d ≤ 3 and the angle
of dilation of the granular material is zero (de is the equivalent diameter; see Figure 10.9).

10.6  COLUMN-SUPPORTED EMBANKMENTS

Stiff columns can be used to support embankments that are to be constructed in areas of
soft clay. The columns can be installed to a depth where there is soil of reasonable strength,
and then a layer of geotextile reinforced granular material placed over the heads of the col-
umns to allow them to support the embankment without punching through the fill. Several
layers of geosynthetic reinforcement can be used and such a foundation system is shown in
Figure 10.24.
Soil improvement  385

r
Uniform pressure qA

Smooth rigid

Stone column Clay Outer boundary


Ep Es smooth rigid
νp νs

Smooth rigid
a
b

Figure 10.20 A column-clay unit.

There are various ways in which such a column-supported embankment can fail and may
include:

a. A slope failure at the edge of the embankment


b. Lateral sliding
c. Failure of the supporting columns
d. Excess strain in the geosynthetics
e. Excess settlement

1.0
νp = 0.3
νp′ = 0.3
0.8
(1 + νs′)(1 – 2 νs′)
mvs =
Es′ (1 – νs′)

0.6
qAmvs

Ep /Es′
z

0.4 10

20
0.2 30

40
0
0 0.2 0.4 0.6 0.8 1.0
a/b

Figure 10.21 Vertical strain of column-soil unit for various spacings.


386  Geomechanics in soil, rock, and environmental engineering

1.0

0.8
10

0.6
δelas
δ 20

0.4 30
de
qA Ep/Es′= 40
0.2
h

d
0
0 1 2 3 4 5
qA
γh

Figure 10.22 Correction factors for elastic settlement (de /d) = 2, ϕ = 40°, ψ = 0, νs = 0.3.

The British code BS 8006 and the FHWA (Federal Highway Administration in the United
States) require that the column support goes within a distance Lp of the toe of the embank-
ment as shown in Figure 10.25. This is to ensure that the edge of the embankment is stable,
and this distance can be calculated from Equation 10.32.

Lp = H[n − tan(θ p )] (10.32)


1.0

0.8
10

0.6
δelas

20
δ

0.4 30
de
qA Ep/Es′ = 40
0.2
h

d
0
0 1 2 3 4 5
qA
γh

Figure 10.23 Correction factors for elastic settlement (de /d) = 3, ϕ = 30°, ψ = 0, νs = 0.3.
Soil improvement  387

Embankment

Geosynthetics
Geosynthetic
load transfer
platform

Column caps

Vertical columns

Firm soil or bedrock

Figure 10.24 Column-supported embankment with geosynthetic reinforcement.

where
n is the side slope of the embankment (n horizontal to 1 vertical)
φ′emb is the angle of shearing resistance of the embankment material
 φ′ 
θ p =  45 − emb  is the angle shown in Figure 10.25
 2 
Lateral sliding or spreading is another way that the embankment could fail as shown in
Figure 10.26. Both the British Standard BS 8006 and the FHWA give simple calculation
methods for this case. The required tensile force that is needed in the reinforcement to pre-
vent lateral spreading Tds is given by

 γH 
Tds = Ka H  + ws  (10.33)
 2 

Embankment
Ls

Lp

n
H 1
θp
Fill: ϕ′cv

Pile cap

Pile

Figure 10.25 Edge failure of a column-supported embankment.


388  Geomechanics in soil, rock, and environmental engineering

Surcharge ws

Embankment
Fill: γ, ϕ′cv
Outward shear stress
Pfill H
Le
Reinforcement
Lb

Lp Tds
Pile caps
Soft foundation
Piles

Figure 10.26 Sliding of an embankment on the geotextile.

where
K a is the active earth pressure coefficient = tan2 (45° − φ′emb / 2)
ws is any surcharge on top of the embankment
H is the height of the embankment

It should be pointed out, however, that the geotextiles do not necessarily need to be
designed to take this force as it may lead to a conservative design.
The minimum length of reinforcement required to stop the embankment sliding sideways
across the geotextile L e (see Figure 10.26) is given by Equation 10.34.

Tds
Le = (10.34)
0.5γH µ tan φemb

where μ is the coefficient of friction for sliding of the embankment material against the
geotextile.
The design of the load transfer platform can be done several ways, including the Collin
method (Collin 2004, Collin et al. 2005), the tension membrane theory (BS 8006-1:2010)
and the enhanced arching method (Guido et al. 1987).

10.6.1  Collin beam method


In the Collin method, the assumption is made that the thickness h of the load transfer
platform is greater than one-half of the clear span between the columns s − d where s is
the centre-to-centre spacing of the columns and d is the column diameter, and the distance
between the layers of reinforcement is a minimum of 150 mm (6 inches). The method also
assumes that there are at least three layers of geotextile used, and that the angle of the region
of arching is 45° (see Figure 10.27).
The purpose of the reinforcing fabric is to provide lateral confinement in the select fill
layer at the base of the embankment and to facilitate arching. This means that the vertical
load from the embankment will arch onto the columns. The secondary function of the geo-
textile is to support the weight of fill below the arch.
The weight of fill that each layer of reinforcing is required to carry is the weight of fill
between that reinforcing layer and the layer above. Hence, the ‘uniform pressure’ W Tn on
layer n is given by
Soil improvement  389

WTn = [ An + An + 1 ]hn γ / 2 An (10.35)


where the height of each layer is hn and the unit weight of the select fill is γ. An is the area
of reinforcement layer n, and this will be different for different column layouts (square or
triangular) as shown in Figure 10.27.
The tension in the reinforcement Trpn at layer n is calculated from the applied pressure
W Tn , and the design span for the membrane D.

Trpn = WTnΩD/2 (10.36)


The design span for a tensioned membrane D is the diagonal length of a square (=1.41*side
length) or the mid-side to corner distance for a triangle (=0.866*side length). The term Ω is
a factor that comes from membrane theory and is a function of the strain in the membrane
as given in Table 10.2.
Because the strain in the membrane is not known, the tension can be found from Equation
10.36 and plotted against a tension–strain diagram from the manufacturer. Where the two
curves cross gives the strain to use in the calculation.

Embankment fill
Well graded granular fill
h4
Geogrid 3
L4 h3
Geogrid 2 h
L3 h2
Geogrid 1
L2 h1
Angle of arching
45° L1 = (s – d) 45°

n=1
Ln = (s − d) − 2 ∑ hi/tan 45°
i=1

For n = 2, 3, 4, etc.

(s – d) (s – d)
(s – d)

(s – d) = Length between pile caps, L1 (s – d) = Length between pile caps, L1


Square column spacing Triangular column spacing

Figure 10.27 Collin beam method of design for membranes. (Adapted from Collin, J.G. 2004. Proceedings
52nd Annual Geotechnical Conference, University of Minnesota, Mineapolis, MN, February 27,
2004.)
390  Geomechanics in soil, rock, and environmental engineering

Table 10.2  Values of Ω


Ω Reinforcement strain ε%
2.07 1
1.47 2
1.23 3
1.08 4
0.97 5

10.6.2  BS 8006 method


The method suggested for calculating the tension in the reinforcing in BS 8006-1:2010 is
based on the tension membrane theory. First, the pile cap stress pc′ to embankment stress σ ′v
is calculated from

2
pc′ C a 
=  c  (10.37)
σ ′v  H 

where a is the size of the cap, H is the embankment height and C c is an arching coefficient
that can be found from Table 10.3.
The distributed load W T carried by the reinforcement can be found from the following
equations where s is the centre-to-centre spacing of the piles.
For H > 1.4(s − a)

1.4sffs γ (s − a)  2 2  pc′  
WT = 2 2  s − a    (10.38)
s −a  σ ′v 

For 0.7(s − a) ≤ H ≤ 1.4(s − a)

s(ffs γH − fqws )  2 2  pc′  


WT = 2 2  s − a    (10.39)
s −a  σ ′v 

s2 p′
But if ≤ c
a2 σ ′v

WT = 0 (10.40)

Table 10.3  Values for the arching coefficient, Cc


Pile arrangement Arching coefficient
End bearing piles (unyielding) H
C c = 1.95 − 0.18
a
Friction and other piles H
(normal) C c = 1.5 − 0.07
a
Soil improvement  391

100

Percentage of ultimate tensile strength (%)


80

60

40

20

0
0 2 4 6 8 10 12 14
Strain (%)

Figure 10.28 Short-term load–strain curve for a geotextile.

The partial load factors used in BS 8006, f fs and fq can both be taken as 1 for the service-
ability limit state and 1.3 for the ultimate limit state.
The tensile force Trp generated in the reinforcement per metre run is then given by

WT (s − a) 1
Trp = 1+ (10.41)
2a 6ε

In Equation 10.41, ε is the strain in the geotextile. Again, the strain is not known and a plot
of the strain versus the tensile force can be compared with the stress–strain curve for the
geotextile being used. A typical plot is shown in Figure 10.28.

10.7  CONTROLLED MODULUS COLUMNS

Stiffening of soft soil can be carried out through the use of controlled modulus columns
(CMCs) in a similar manner to stone columns. The columns are made of a weak concrete
so that they will reduce settlement and provide extra strength to the soil. The modulus of
the column is designed so that the load from an embankment or structure is applied to the
original soil as well as the columns, unlike stiff piles that carry most of the load.
The columns are created by using a tapered auger (Figure 10.29) that is pushed down into
the soil using a high downward thrust. The auger is then withdrawn, and grout is pumped
from the base of the auger into the cavity thus formed. This method has the advantage that
there is very little spoil produced, is vibration free, and compacts the soil laterally.
An example of the use of CMCs was the construction of 875 CMCs with a 450 mm
diameter to depths of between 7 and 11 m under an approach embankment located on the
left bank of the Macleay River near Frederickton as part of the Kempsey Bypass project.
392  Geomechanics in soil, rock, and environmental engineering

Figure 10.29 Auger used for installing controlled modulus columns.

The design included six rows of ‘stepped’ columns, stopped 2 m short of the embedment
level to act as a transition zone between pile reinforced soil and unreinforced soil. The mix
design included a majority of 10 MPa concrete columns with the two rows at the periphery
of the treated area made of 40 MPa concrete.

10.8  DYNAMIC COMPACTION

Dynamic compaction (DC) involves dropping a large weight from a crane onto the ground
causing it to densify. The process was originally developed for sands or granular materials,
but it has since been used for clayey soils as well.
The weights that are dropped can be made from steel or concrete typically weighing
4.5–18 tonnes (5–20 tonnes) from heights of up to 30 m (100 ft). The soil can be compacted
up to 15 m with heavy weights and high drop heights (Mayne et al. 1984). In the initial ‘high
energy’ phase, the weight is dropped on a square pattern and the area then levelled pushing
soil into the craters. Then a second pass is performed over the area, dropping the weight at
Soil improvement  393

the centre of the initial square pattern. Finally, an ‘ironing’ phase is carried out using lower
energy drops to compact the soil missed in previous phases.
For silts and clays, the impact creates pore pressures in the ground that eventually dissi-
pate producing a reduction of volume in the ground and strengthening of the soil.
The potential energy available from the weight when raised is WH where W is the weight
of the block that is dropped and H is the drop height. Field data indicates that the depth
of compaction dmax that can be achieved is proportional to the square root of the potential
energy, that is,

1 WH
dmax = (10.42)
2 n

The quantity n in Equation 10.42 is a unit factor which is equal to 1 tonne/m or 672 lb/ft.
The data that the above relationship is based upon can be seen in Figure 10.30. The depth
of influence of the compaction is plotted versus the energy of the drop to a log–log scale
from which it may be seen that the average of the data can be well represented by the square
root relationship of Equation 10.42 although there is a good deal of scatter. The data for
this plot has come from sites that were underlain by sands (50% of sites) but also silt, silty,
clay, or clay as well as rubble fills.
The effects of the compaction process and the depth to which the improvement reaches
can be assessed in the field by carrying out field tests before and after compaction. Tests
such as SPT (standard penetration tests – Section 4.15) and CPT (cone penetration tests –
Section 4.18) as discussed in Chapter 4, can be used.
An example of DC is the Penrith Lakes site in Sydney where a 20-tonne pounder was
dropped from a height of 23 m. The spacing between drop locations was 4.5 m and the
pounder was dropped 16 times in each location. This was successful in compacting the

50
Maximum observed depth of influence
Dmax = Maximum depth of influence (metres)

Influence greater than depth tested


20

10

2
Dmax = 0.8(WH)0.5
1 Dmax = 0.5(WH)0.5

0.5 Dmax = 0.3(WH)0.5

0.2

0.1
0.1 1 10 100 1000 10,000
Energy per blow = WH (tonne m)

Figure 10.30 Maximum depth of influence versus energy per blow.


394  Geomechanics in soil, rock, and environmental engineering

uncontrolled fill (silty sand [SM], sandy clay [CL], and sandy silt [ML]) that had been placed
at the site to depths of between 10 and 12 m (Moyle 2013).

10.8.1  Impact rollers


Impact rollers apply a dynamic force to the ground like a falling weight but they do so in a
slightly different way. The roller can have three, four, or five sides, and is towed behind a
tractor. Because the roller is not circular, it rises up on the high point of the roller and then
falls down impacting the soil below. A three-side roller is shown in Figure 10.31.
Field data reported by Berry et al. (2004) shows that the peak densification occurs at a
depth of 0.67–1.0 B where B is the width of the roller (typically 0.9 m) and the depth of
influence of the roller is 2–3 B. In impact trials in South Africa, compaction settlements of
over 500 mm and compaction down to depths of more than 2 m were achieved with a three-
side roller.
Avalle (2004) gives the example of compaction carried out with a four-side impact roller
for a new building at Adelaide Airport. The impact roller used was a 1.3 m wide, 1.5 m high
square steel concrete-filled module that had a mass of approximately 8t. It was drawn in its
6t frame by a 200 kW four-wheel drive towing unit at a speed of 10–12 km/h.
The ground conditions at the site comprised 1.3–1.5 m of existing fill overlying firm to
stiff clay and loose to medium dense sand with the water table at 2–1.5 m. The fill also con-
tained bricks, concrete, and rock fragments.
The settlement of the site was measured for different numbers of passes of the roller, and
these are plotted in Figure 10.32. Settlements of up to 85 mm were achieved with 40 passes
of the roller as may be seen from the figure.

Figure 10.31 Three-side impact roller.


Soil improvement  395

80
Average settlement (mm) Polynomial trend

60

40
Average settlement

20

0
0 5 10 15 20 25 30 35 40
No. of impact roller passes

Figure 10.32 Settlement versus number of passes. (Adapted from Avalle, D.L. 2004. Proceedings of the 23rd
Southern African Transport Conference [SATC 2004], pp. 44–54.)

The vibration caused by impact rolling may be of concern in some cases, especially if
the compaction is being carried out near residential areas or historic structures. Data on
the peak particle velocity (PPV) versus distance from the compactor has been presented by
Bouazza and Avalle (2006) for an 8-tonne, four-side impact roller (Figure 10.33). The area
had been filled with refuse, observed to be about 3–4 m thick, and capped with 2–3 m of
quarry overburden. Further information on the criterion for damage to structures in terms
of PPV is given in Section 12.9 in Chapter 12.

100
Peak particle velocity (mm/s)

10
C
B

1
A

Closest house at
21 m

0.1
1 10 100
Distance (m)

Figure 10.33 Peak particle velocity versus distance for an impact roller.


396  Geomechanics in soil, rock, and environmental engineering

10.9  DEEP SOIL MIXING

In the deep soil mixing (DSM) process, an additive is mixed with the soil in either a dry state
or a wet state. A mixing tool such as the one shown in Figure 10.34 mixes the grout that
is ejected from the base of the mixer, with the surrounding soil. The mixer may be moved
up and down to achieve a more complete and uniform mix. Often there are three mixers
used side by side so as to create a wall of grouted soil. The blades of the mixers overlap (see
Figure 10.34) so that there is no unmixed material between each of the mixing blades.
The grout mixture used depends upon the soil type and trials can be done to establish
the best mix. Cement or lime or combinations of both have been used as admixtures. DSM
carried out in the Shanghai clay (Chen et al. 2013) used water/cement ratios in the range
1.2–1.8 and amounts of cement in the range 360–450 kg/m3. Penetration speeds varied from
0.25 to 0.4 m/min with withdrawal speeds in the range 0.4–0.6 m/min.
Madhyannapu et al. (2010) report using DSM columns to stabilise expansive soils hav-
ing a PI of 30% (Site 1) and 50% (Site 2). For a lime (3%)–cement (9%) binder used at
the rate of 200 kg/m3, the field mixed samples had an unconfined compressive strength qu
of 1140–1176 kPa and a stiffness G max of 108–114 MPa. The untreated soil had strengths
of 105–300 kPa and stiffnesses of between 35 and 67 MPa. Laboratory mixed specimens of
the clay had higher stiffnesses and strengths indicating that field-mixing trials give a better
indication of actual performance (see Table 10.4).
Dry mixing can be used in soils with a high enough moisture content to allow hardening
of the admixture used that can be cement, lime–cement, or blast furnace slag. The admix-
ture is forced from the end of the mixer by compressed air. An advantage of the dry method
is that it is cheaper to implement.
Cutter soil mixing (CSM) is a more recent process that involves using the double-headed
rotor shown in Figure 10.35. The cutters do not need a guide wall (as for the construction of

Figure 10.34 Deep soil mixing process showing slurry injection at base of mixers.
Soil improvement  397

Table 10.4  S trength and stiffness ratios of field


and laboratory treatments
Site Gmax,field/Gmax,lab qu,field/qu,lab
1 0.43–0.67 0.67–0.70
2 0.56–0.65 0.83–0.86

Figure 10.35 Cutters used for cutter soil mixing (CSM).

diaphragm walls with a clamshell). On insertion, water and compressed air can be used to
aid in the penetration of the cutting head. On reaching the design depth, the cutter is raised
and cement slurry is injected between the cutting heads. Finally, steel reinforcement can be
added to the grout–soil mix if desired either by penetrating under its own weight or with the
assistance of vibration.

10.10  JET GROUTING

Jet grouting involves a process where grout is sprayed under high pressure (that can be up
to 60 MPa) from the sides of a shaft (called a monitor) that is drilled into the soil. The jet
of grout issues horizontally at high velocity (100 m/sec) and cuts into the soil churning it up
and mixing the soil with the grout (Figure 10.36). The shaft is raised and rotated so that
398  Geomechanics in soil, rock, and environmental engineering

Grout Grout
backflow

Grout jet

Figure 10.36 Jet grouting system.

a cylinder of soil mixed with the grout is formed. There are various types of jet grouting
systems, some having two and three jets instead of one, and some having air shrouded grout
jets or air shrouded water jets. The process is different to conventional grouting where the
grout has to be injected into the pores of the soil, because the grout jet cuts into the soil and
mixes with it.
There are many applications for the jet grout process, and it has been used for retaining
systems, to seal the base of excavations in pervious material prior to excavation, to form
cut-off barriers, and in underpinning for foundations. Care is needed when using the system
for sealing the base of excavations as lack of overlap of the jet grout plugs can lead to water
ingress and flooding of the excavation.
The system was used to repair the new Sydney Airport runway. The runway was con-
structed of sand pumped from Botany Bay and was supported around the perimeter by a
reinforced earth wall. Sand was being eroded between the gaps in the facing panels of the
wall, therefore over 5000 jet grout columns were constructed behind the wall with diam-
eters of between 1 and 2.7 m and to depths of 9 m to seal the sand backfill.

10.11 GROUTING

Grouting of soils can be performed by pumping a grout into the soil under pressure and
allowing it to flow into the pores of the soil. Grouting of rock is different to soil grouting,
as the grouts used are generally cement–water mixtures that are forced into fractures and
fissures in the rock. These grouts are generally unsuitable for soils as the cement particles
are too large to penetrate the voids in between the soil particles that act like a filter to
stop the cement particles penetrating. However, they can be used for gravels and coarse
sands.
Soil improvement  399

Many different types of grouts can be used and selection may be based on cost, the degree
of penetrability and permanence of the grout. Penetrability is a primary factor and for
grouts which consist of solid particles (such as cement and clay) the relative sizes of the grout
particle and the void size are important. For grouts such as acrylates and phenols which do
not contain solid particles, the viscosity of the grout determines ease of penetration.
The types of grouts and the soil types that they can penetrate are shown in Figure 10.37,
however, further information is available in Baker (1985).
At present, the most satisfactory method which can be used for grouting deposits of allu-
vium is the French tube à manchettes TAM (sleeved tube) process. The grouting pipe used
in this process is shown in Figure 10.38. The PVC grouting pipe is placed into a borehole
and grouted in with a weak clay cement grout. At regular intervals along the PVC pipe are
holes surrounded with a rubber sleeve much like a large rubber band. This sleeve acts like
a valve, stopping grout from coming back into the tube but allowing the grout to flow out.
The grouting pipe is lowered to the desired depth and then packers are inflated to seal off
a section of the PVC tube. The grout is then pumped out from the inner grout pipe and
expands the rubber sleeve, thus flowing out into the soil.
The groutability of a soil using a particulate grout depends on its grain size distribution.
A guide can be obtained from Mitchell and Katti (1981) who define an N value as given in
Equation 10.43.

D15(soil)
N = (10.43)
D65(grout)

Particulate grouting is considered feasible if N > 24 and not feasible if N < 11.


Although the process is expensive, it has the advantage that the same pipe can be re-
centred as often as desired to perform additional grouting or to use different grouts at dif-
ferent levels. Alluvium such as sands, gravels, and cobbles has been grouted to depths of over
100 m by this process (e.g. Aswan, Terzaghi dams).

Gravel Sand Silt Clay

Cement, soil

Clay

Silicate chemicals

Chrome and lignins

Polymers (resins, phenoplasts)

10 1.0 0.1 0.01 0.001


Grain size (mm)

Figure 10.37 Soil types that can be penetrated by various grouts.


400  Geomechanics in soil, rock, and environmental engineering

63.5 mm pipe with


circumferential rows of
holes each spaced at
300 to 1000 mm
Brittle clay-cement in

330 mm
annular space
Rubber Packers inside
sleeve 63.5 mm pipe
Grout enters soil
through cracks
330 mm

Holes
Grout pipe
in pipe
Packers inside
63.5 mm pipe

140 mm

Figure 10.38 The tube à manchettes grouting pipe.

10.12  OTHER METHODS

The methods of ground improvement that have been mentioned so far are the most com-
monly used methods in the author’s experience. There are a number of other methods that
can be used and some of these are listed below.

10.12.1  Ground freezing


The ground may be artificially frozen by inserting pipes and circulating a coolant through
the pipes until the water in the ground freezes, thus forming a temporary hardened region.
Calcium chloride (brine) is the most common cooling agent and is cooled down to between
−15°C and −25°C before circulation. In some smaller operations, liquid nitrogen has
been used because it can rapidly solidify the soil and it can produce high strength in the
frozen material.
Ground freezing can be used on just about all soil types unlike other soil improvement
methods. It has been used to construct vertical shafts, as crown support for tunnels, and
horizontal tunnels.
Ou et al. (2009) describe the freezing of the ground at the face of a shaft where a tunnel-
boring machine was to enter. Finite element analysis was undertaken based on the governing
equation

∂  ∂T  ∂  ∂T  ∂T
 kx  +  ky  +Q = λ (10.44)
∂x  ∂x  ∂y  ∂y  ∂t

This heat flow equation is similar to the equation for steady state flow of groundwater,
and can be solved numerically in a similar fashion. In the equation, k x and ky are the thermal
conductivities in the x and y directions, Q is an applied flux, λ is the heat storage capacity,
T is temperature, and t is time. The term λ is defined as
Soil improvement  401

∂w u
λ =C+L (10.45)
∂T

and C is the volumetric heat capacity (kJ m−3 °C −1), L is the latent heat of water (=334 kJ kg−1),
and w u is the unfrozen volumetric water content.

10.12.2  Electro-osmotic or electro-kinematic stabilisation


Electro-osmosis involves applying a direct current to the soil through the use of electrodes
implanted into the ground. It is most applicable to clays and silts where dewatering is
required. As the soil contains charged particles, the particles are attracted to the anode (+ve)
or the cathode (−ve). Generally, there is a net flow towards the cathode and water can be
removed thus draining the soil and causing strengthening.
The velocity of flow vx created in the pore water (in one dimension) is given by Wan and
Mitchell (1976)

kx ∂u ∂V
vx = − − ke
γ w ∂x ∂x (10.46)

where ke is the electro-osmotic permeability and V is voltage in the soil. The first term in
the equation is Darcy’s law for flow under a pore water pressure gradient in a soil with per-
meability kx. At equilibrium, it can be shown that the pore pressure u(x) generated by the
gradient in voltage is

k 
u(x) =  e  γ wV (x) (10.47)
 kx 

One system uses anodes and cathodes that are placed into a slope. The application of a
current removes water from the soil and finally the anodes and cathodes are used as soil
nails to further reinforce the slope.
Electro-kinematics involves using electro-osmosis to introduce stabilising chemicals into
the ground, although the term seems to be used for electro-osmosis in many cases. One
form of electro-kinematic stabilisation introduces calcium chloride at the anode and sodium
silicate at the cathode.
In recent years, conductive polymer materials have been developed that do not suffer cor-
rosion like steel or copper anodes.

10.13  NUMERICAL ANALYSIS

Analysis of improved soil systems can be performed through the use of numerical means,
such as finite element and finite difference analyses. Stability of embankments on soft soils
and improved soils may be carried out by these means. As well, slip circle or non-circular
analyses (Sections 7.2 and 7.3 in Chapter 7) may be employed, and where stone columns or
CMCs are involved, uniform soils with equivalent properties can be used (Section 10.5.1.2).
However, it is difficult to analyse problems involving surcharging and consolidation with
slip circle analyses because as the soil consolidates, the effective stresses in the soil increase,
and this means that the undrained strength also increases. It is necessary to monitor pore
402  Geomechanics in soil, rock, and environmental engineering

pressures so that the effective stress can be estimated and then the undrained strength can
be found from a relationship of the kind (see Section 3.10 in Chapter 3).

su(OC) = Kσ ′v (OCR)m (10.48)


The undrained shear strength su(OC) is related to the vertical effective stress σ ′v and the over-
consolidation ratio OCR (see Ladd 1991).
For soft clays modelled using numerical (finite element) analysis, it is therefore essential to
use the proper soil model, and the Cam Clay type models are most suitable since they allow
yield of the soil to take place and can model normally consolidated and overconsolidated
behaviour. Therefore, an embankment that is built up to an initial height and then allowed
to consolidate before being raised again (with this cycle repeated if required) can be anal-
ysed because the pore pressure build up and dissipation can be modelled in the analysis. This
is addressed in detail in Section 3.12. A suitable model that can be used in the commercial
code PLAXIS is the Soft Soil Model.

10.13.1  Three-dimensional analysis


For modelling embankments on soft soils, two-dimensional analysis is sufficient. Modern
finite element programs allow the embankment fill to be added through the addition of
elements. For analyses involving PVD or wick drains, a slice through the drains may be
modelled using three-dimensional elements (see Figure 10.39). This allows edge movements
and the three-dimensional nature of the drains to be considered (Borges 2004). Vacuum
consolidation can also be modelled by specifying the negative pressure at nodes of the mesh
in the location of the drains.

10.13.2  Equivalent two-dimensional analysis


To make the numerical analysis simpler, wick drains or stone columns can be treated
approximately by smearing the individual columns into continuous two-dimensional rows.

5.3 m
3.0
y

2.0

5.0

Figure 10.39 Three-dimensional mesh including vertical drains. (After Borges, J.L. 2004. Computers and
Geotechnics, Vol. 31, No. 8, pp. 665–676.)
Soil improvement  403

(a) (b)

Oblique view

Plan view: Arrows


indicate direction of
flow

Figure 10.40 Conversion of (a) a cylindrical drain to (b) an equivalent two-dimensional drain.

This is shown in Figure 10.40 where B is the half-width of the two-dimensional unit cell
and R is the radius of the axi-symmetric unit cell. To calculate R for triangular and square
patterns of well layout, the formulae of Equation 10.14 may be used.
Hird et al. (1995) suggested the use of an equivalent permeability for the two-dimensional
drain where there are three alternatives. Firstly, the width of the equivalent plane strain
unit cell can be calculated by keeping the permeability the same as for the axi-symmetric
case; secondly, the permeability of the two-dimensional drain can be calculated keeping the
half-width of the unit cell the same as the radius of the axi-symmetric cell; and finally, the
half-width of the two-dimensional cell can be selected and a permeability calculated for the
2D case. Keeping the unit cell width equal to the unit cell radius B = R is probably the most
useful in finite element work, therefore the formula for calculating the equivalent perme-
ability kpl is given in Equation 10. 49.

kpl 2
= (10.49)
kax 3 {ln(R/rs ) + (kax /ks )ln(rs /rw ) − (3/ 4)}

In the above equation, kax is the lateral permeability of the soil, subscripts (pl) and (ax) stand
for plane strain and axi-symmetric, B is the half-width of the two-dimensional unit cell, and
R is the radius of the axi-symmetric cell. As before, rs is the radius of the smeared zone, ks is
the permeability of the smeared zone, and r w is the radius of the well.
If other matching formulae are required, these may be found in Hird et al. (1995).
When used in a two-dimensional finite element analysis, the soil can be given a vertical
permeability that is the actual permeability of the soil, and an equivalent horizontal perme-
ability (from Equation 10.49). There is no need to try to model the actual drain, only to use
the equivalent parameters. Half of the unit cell is subdivided into a mesh, and pore pressure
boundary conditions are applied at the top (and if needed at the bottom) of the mesh. At the
location of the drain, the excess pore pressure is set to zero. Lateral movement is restricted
at the sides and the base of the mesh (if rough). Alternatively, the unit cells can be used under
a full two-dimensional representation of an embankment.
This page intentionally left blank
Chapter 11

Environmental geomechanics

11.1 INTRODUCTION

Today, the environment is playing an ever increasing role in geotechnical engineering, as


waste has to be disposed of in belowground or aboveground repositories, and there is the
likelihood of pollutants from the waste seeping into the groundwater, being washed into
rivers, or being blown into the air by wind.
Waste from previous eras when regulations were not as strict as they are today, may exist
in the ground and have to be cleaned up. Today, environmental investigations are often
performed at the same time as routine site investigations, with the same borehole used for
sampling pollutants as well as soil.
Waste from domestic sources and industry is generally placed into landfills, while waste
from mining operations and large-scale industrial operations such as power station waste is
placed into large dams. Both of these types of waste and the geotechnical aspects of dealing
with the waste are examined in the following sections.

11.2 LANDFILLS

In most developed countries today it is required that the waste placed into landfills be con-
tained so that pollution from the waste does not contaminate the surrounding areas. This
generally requires a liner be constructed across the base and sides of the waste repository. If
the waste is placed in a very low permeability layer of clay, this may be unnecessary, but in
many cases a liner is required.
Once the landfill reaches its capacity, the waste needs to be sealed with a cover to keep the
waste confined for many years into the future.
Designs of liners and covers are examined in the following sections.

11.2.1 Liners
Liners may be constructed of geosynthetic materials or from compacted clay that has a low
permeability. The main purpose of these materials is to reduce the possibility of leachate
(pollutants) passing through the liner and into the surrounding soil and groundwater. The
various types of liner materials used are ideally resistant to any form of chemical attack or
degradation so that they remain as a barrier to the leachate for a long period of time.
Leachate from the leachate collection layers is collected in slotted pipes (Figure 11.1) and
can be placed back into the repository or treated and released elsewhere. Some different
liner designs are shown in Figure 11.2 where a double barrier is used with leachate collection
between the two barriers.

405
406  Geomechanics in soil, rock, and environmental engineering

Gas vent
Cover soil
Drainage layer
Geomembrane
Compacted clay
Gas collection layer
Waste
Primary leachate collection layer
Geomembrane
Compacted clay layer
Leak detection layer
Geomembrane
Compacted clay layer
Original soil
Drainage pipe

Figure 11.1 Liner and cover for landfill.

Liners can be constructed of compacted clay in combination with geomembranes such


as high-density polyethylene (HDPE) sheet or geomembrane clay liners (GCLs). Permeable
layers of sand or geonets can be used to collect the leachate. Types of liners are examined in
more detail in Sections 11.3 through 11.5.
A clay liner being compacted is shown in Figure 11.3 and leachate collection pipes being
laid in a trench surrounded by gravel is shown in Figure 11.4. A geotextile membrane being
laid above the compacted clay layer is shown in Figure 11.5.
Once in operation, waste is compacted into the landfill with a heavy compactor
(Figure  11.6) and generally covered with soil at each lift to prevent rubbish being blown
around by the wind and to discourage birds and other vermin from being attracted by the
waste.

Geonet leachate Geonet leachate


Sand leachate collection layer collection layer detection layer
Sand leachate detection layer Protective soil layer

Geomembrane
top liner

Composite
bottom liner Clay
Clay

Composite
top liner

Composite
bottom
liner

Figure 11.2 Double liner designs.


Environmental geomechanics  407

Figure 11.3 Impervious liner being constructed over the base of landfill.

11.2.2 Covers for landfills


Once a landfill has been filled to capacity the waste needs to be covered to prevent ingress of
water and to stop gases produced by the waste from venting into the atmosphere. The cover
may need to be designed to intercept and collected gases such as methane that are produced
by the large volume of decaying vegetable matter in the landfill. In many cases, the gas is
collected and burnt off, but in very large landfills, the methane can be used for commercial
purposes such as generating power.

Figure 11.4 Drainage pipe being laid in a leachate collection ditch.


408  Geomechanics in soil, rock, and environmental engineering

Figure 11.5 Membrane being laid on the base of a landfill.

An example of a cover is shown in Figure 11.7 where different types of protective layers
are shown. Not all of these layers need to be used in every situation but potentially could
be used in a cover design. In areas where wind erosion is likely to be a problem, cobbles
can be used for layer 1 as they can resist strong wind. Otherwise, layer 1 may be a grassed
topsoil layer to prevent erosion by runoff from rainfall. Runoff can be channelled down
lined gutters and the water allowed to sediment into a pond before discharge from the site
(see Figure 11.8).
A vent for burning off methane gas from a landfill is shown in Figure 11.9.

Figure 11.6 Waste being compacted into a landfill.


Environmental geomechanics  409

(a) Surface layer of topsoil or cobbles

Cover soil
(b) Protection layer of locally available soil

(c) Drainage layer of sand or geosynthetic

(d) Hydraulic/gas barrier layer (compacted clay


liner, geosynthetic clay liner, or geomembrane)

(e) Gas collection layer (geosynthetic or soil)

(f) Soil foundation

(g) Waste

Figure 11.7 Typical cover design.

The sides of the landfill may also have to be covered by an impervious seal and this can
be done by extending the clay liner or by using a geomembrane in conjunction with a clay
liner on the sides of the landfill. The geomembrane needs to be anchored at the top of the
slope and this is often done by digging a trench along the top of the slope and placing the
membrane into it. Backfilling the trench with soil will anchor the membrane (Figure 11.10).
Because the angle of shearing resistance between the soil and the geomembrane on the
slope can be low, there is potential for a sliding failure to take place on the slope. This is
discussed in Section 11.6.

11.3 COMPACTED CLAY LINERS

Compacted clay liners (known as CCLs) have been commonly used for creating barriers
for waste containment in the mining industry and for landfills. Because the permeability of
clays is low (10−6 to 10−9 cm/sec) any fluid takes a long time to move through the clay, and
hence is contained.
Where there are naturally occurring clay deposits, like the old brick pits around Sydney,
then the natural clay acts as the barrier, and waste can be placed into the pit directly. Where
there is no natural clay barrier, then clay needs to be imported and compacted to form a
lining for the waste. As mentioned before, often clay liners are used in conjunction with
geomembranes to form the barrier.

11.3.1 Compaction of clay
Naturally occurring clays that can be used as a CCL are those that can be classified under
the unified soil classification system as CH (high plasticity clay), CL (low plasticity clay), or
SC (sand with clay content).
410  Geomechanics in soil, rock, and environmental engineering

Figure 11.8 Channel for draining water from a landfill cover with a water treatment pond.

Figure 11.9 Vent for burning off methane produced by a landfill.


Environmental geomechanics  411

0.61 m 0.91 m
(24″) (36″) min

Geotextile
Primary HDPE geomembrane
Geonet or geocomposite
Secondary HDPE
0.61 m (24″) geomembrane

HDPE = high-density polyethylene

0.61 m 0.61 m 0.61 m 0.91 m


(24″) (24″) (24″) (36″) min

Primary HDPE liner

0.61 m (24″)

Secondary HDPE liner

Figure 11.10 Single and double anchor trenches for geomembranes at the top of a side slope.

Clay can be mixed with existing soils to lower their permeability. For example, bentonite
can be mixed with natural soil to create a composite material more resistant to seepage
losses.

11.3.2 Compaction method
The compaction of clay liners is carried out in similar fashion to the compaction of soils for
other purposes, in that the maximum dry density to be achieved is specified. For example,
the specification may state that the clay is to be compacted to 98% of the maximum dry
density found in the standard compaction test.
The standard (or modified) compaction tests are performed in the laboratory, and a plot
of the water content of the soil versus its dry density is obtained. The maximum dry density
is achieved at the optimum water content as shown in Figure 11.11.
It is then up to the contractor to meet the specifications by

1. Selecting the moisture content. If the clay is too wet it may need to be scarified and left
to dry; if it is too dry, water can be added by a water truck.
2. Selecting the thickness of each lift of clay to be compacted.
3. Selecting the type of compaction equipment. Heavier rollers and rollers with studded
drums may be used to improve compaction. The number of passes of the roller also has
to be determined.

Once the clay is compacted, it must not be allowed to dry out and crack as this will lead
to paths through which leachate can leave the landfill.
412  Geomechanics in soil, rock, and environmental engineering

Dry unit weight (γ d)


Maximum dry
unit weight

γ d max Zero air voids curve

Optimum water
content

mopt
Water content m

Figure 11.11 Dry unit weight versus moisture plot.

11.3.3 Compaction control
Once the clay is compacted, testing needs to be performed to determine if the specifications
have been met. This is done by using field density testing, and some of the techniques are
listed below.

1.
Nuclear density testing: A nuclear source of fast neutrons is used to pass neutrons
through the soil. Water slows these down and so the number of fast neutrons detected
at the gauge is a measure of the water content m.
  A gamma ray source is then used to determine the density γ of the soil either by
placing a probe containing the source down into the soil (direct transmission) or by
emitting the rays from the surface (backscattering). The two measurements allow the
dry unit weight γdry to be calculated.

γ
γ dry = (11.1)
  1 + m%

2.
Sand replacement: A hole is dug in the compacted fill and the soil carefully collected
so that its weight can be determined. Then the volume of the hole is found by filling it
with sand of known density.
  The weight of the soil and the volume of the hole are used to compute the bulk unit
weight of the soil and from water content samples taken from the excavation, the dry
unit weight can be found.
3.
Balloon densometer: A water-filled balloon is used to find the volume of the hole exca-
vated in the fill. The balloon is inflated by pumping water into it from a piston.
4.
Other methods: There are many other methods in use, such as cutters that are driven
into the clay, and the soil is extracted inside the cutting ring. The weight and volume
(the volume of the ring after the soil is trimmed) give the bulk unit weight of the soil.
Most of the methods employ different ways of determining the volume of the hole left
by the excavated clay.
Environmental geomechanics  413

11.3.4 Permeability of clay
Although clay that is compacted at optimum moisture content will have the greatest dry
unit weight and therefore the greatest shear strength, the permeability is not necessarily the
lowest at that point.
It can be seen from Figure 11.12 that the permeability is less on the wet side of optimum.
The curves on the plot are for low compactive effort A, medium compactive effort B, and
high compactive effort C. The data shows that heavier compaction equipment and therefore
greater compactive effort gives lower permeabilities.
Therefore, it is desirable to compact slightly wet of optimum to get a lower permeability,
but this may lead to the clay shrinking more when it dries out, and hence cracks forming
in the CCL. If the clay can be kept moist with a sand cover or a geotextile cover before the
waste is placed, then this can be an effective solution.
In addition, any clods in the clay should be broken up by more passes of the roller or
heavier compaction equipment, as these can lead to paths along which water can flow (i.e.
next to the clods).

(a)

10–7 A
Hydraulic conductivity, k (m/s)

B
10–8
C

10–9

10–10

(b)
18
Dry unit weight, γd (kN/m3)

C
S = 100%

17
B

16

A Silty clay
15

11 13 15 17 19 21 23 25
Moulding water content, w (%)

Figure 11.12 Permeability versus water content for a silty clay. (a) Permeability versus water content; (b) dry
unit weight versus water content for a silty clay. (Adapted from Mitchell, J.K., Hooper, D.R.
and Campanella, R.G. 1965. Journal of the Soil Mechanics and Foundations Division, ASCE, Vol. 91,
No. SM4, pp. 41–65.)
414  Geomechanics in soil, rock, and environmental engineering

11.3.5 Measuring permeability of CCLs


The permeability of the clay liner is best determined in the field as laboratory testing can
return incorrect values due to the limited size of specimen tested and the fact that in the field
the clay may have defects that are missing in the laboratory specimens. Field methods for
measuring permeability are examined in the following sections.

11.3.5.1 Ring infiltrometer
The ring infiltrometer is a large ring 0.5–2 m in diameter that is filled with water. The ring is
sealed by embedding it into the clay so that water cannot leak out but must infiltrate down-
wards into the liner. The rings may be single or double rings and sealed or open as shown
in Figure 11.13.
The test takes a long time to perform, as infiltration of water into the liner is slow.
However, the permeability can be calculated from (Daniel 1989)
Q
k = (11.2)
Ati

where
k is the permeability of the clay liner
Q is the quantity of water lost
A is the area of the infiltrometer ring
i is the hydraulic gradient

However, the problem is to determine the hydraulic gradient i as the difference in head
driving the flow is uncertain, as the wetting front of the water as it infiltrates is not easy to
determine. The suction in the soil at the wetting front that is used in calculating the head
difference is also not easily found.
Tensiometers can be used to locate where the wetting front is, and the suction is often
assumed to be zero. Alternatively, post-test excavation can be used to locate the wetting
front. Then the hydraulic gradient can be calculated from

H L + Lf
i = (11.3)
Lf

Clay liner

Single ring; open Double ring; open

Water
HL
Lf
Clay liner H

Single ring; sealed Double ring; sealed

Figure 11.13 Infiltrometer rings.


Environmental geomechanics  415

where
HL is the depth of water in the infiltrometer
Lf is the depth to the wetting front (see Figure 11.13)

11.3.5.2 Borehole test
Another way of determining the permeability is to note the drop in water level in a standpipe
connected to a casing set into the clay liner (Boutwell and Tsai 1992).
The water is first allowed to drop in the casing that is flush with the bottom of the bore-
hole. The vertical permeability is determined from this test. Then the Stage II test is per-
formed allowing water to infiltrate laterally as well as vertically into the soil. From this, both
the lateral and vertical permeabilities can be deduced.
The process is as follows:

1. Falling head tests are performed and the hydraulic conductivity k1 from Stage I is com-
puted using Hvorslev’s formula

πd 2 H 
k1 = ln  1  (11.4)
11D(t2 − t1)  H 2 

In the formula, the heads H1 and H 2 are measured at two times t1, t 2 as the water falls
down the tube, and D and d are the diameters of the casing and standpipe, respec-
tively, as shown in Figure 11.14.
The test is performed several times (which may take up to 2 weeks) until a steady
value of k1 is obtained.

2. The falling head test is performed in a hole deepened beyond the casing. In Stage II of
the test, k 2 is calculated from

A  H1 
k2 = ln (11.5)
B  H 2 

d Casing
Grout
D D
H H
z

L
Clay
Clay

Stage I Stage II

Figure 11.14 Two-stage borehole test.


416  Geomechanics in soil, rock, and environmental engineering

4
L/D = 2.0
1.5
3
1.0

k2
k1
2

1
1 3 5 7 9 11 13
m

Figure 11.15 Graph for obtaining m for different values of L/D.

where

  2 
 L  L  
A = d 2 ln  + 1 +  D   (11.6)
  D 


L    L   
B = 8D (t2 − t1) 1 − 0.562 exp  −1.57     (11.7)
D    D   

Again, Stage II is repeated until a steady state result is obtained. The final stage is then to
use Figure 11.15 to obtain a value of m from the ratio k 2 /k1. Finally, the permeabilities in the
horizontal kh and vertical kv directions can be found from

kh = mk1
1 (11.8)
kv = k1
m

11.3.5.3 Lysimeters
Large-scale seepage collectors (called lysimeters) can be installed below clay liners to collect
seepage through the liner. They are basically a sand drain under the liner that is lined on
the bottom and sides with an HDPE membrane as shown in Figure 11.16. The sand is about
15 m × 15 m in plan and about 20–30 cm thick.
The water is collected with drainage tubes and the permeability calculated from Q = kiA
where the area of the lysimeter is A and Q is the quantity of water collected in time t. The
hydraulic gradient i across the liner is the change in total head divided by the thickness of
clay above the sand.

11.3.5.4 Porous probes
Porous probes, which are really standpipes, can be sealed into the clay liner and the perme-
abilities can be calculated according to the formulae in Figure 11.17. The clay is assumed to
be isotropic and saturated in deriving these formulae, and the test needs to be run for long
enough to saturate the soil around the tip of the probe.
Environmental geomechanics  417

Landfill waste
Leachate level
Flow of leachate

HDPE
≈1.5 m membrane
Lysimeter sand ≈20–30 cm

Clay liner Subsoil


Discharge Q

Figure 11.16 Lysimeter used to determine the permeability of a clay liner.

(a) (b)

d d

H H
Seal Seal

L L

Figure 11.17 Porous probe tests: (a) Case A – Probe with permeable base; (b) Case B – Probe with imper-
meable base.

q
Constant head: k =
FH
πd 2 /4  H1 
Falling head: k = ln
F(t2 − t1)  H 2 
2πL
Case A: F =
(
ln (L /D) + 1 + (L /D)2 )
2πL
Case B: F = − 2.8D

(
ln (L /D) + 1 + (L /D)2 )
11.4 FLEXIBLE MEMBRANE LINERS

Flexible membrane liners (FMLs) or geomembranes are used to line landfills and prevent the
escape of leachate. The various geomembrane materials are made from parent resins with
other additives.
The membranes need to be as watertight as possible, thus the joining of individual sheets
and the problem of puncturing of the membranes on site (due to trafficking) is of interest
to engineers.
418  Geomechanics in soil, rock, and environmental engineering

The various types of geomembranes most commonly used in landfill applications are
listed in the following sections.

11.4.1 Types of geomembranes
Several types of polymers are used in the manufacture of geomembranes, and the ones more
commonly used in landfill applications are listed in the following.

11.4.1.1 High-density polyethylene
High-density polyethylene liners are known by the acronym HDPE liners. They are made
from polyethylene resin, carbon black, and other additives.
The carbon black is added as a stabiliser against ultra-violet light and is generally added
as a pellet concentrate to the resin or can be added in powder form. It is usually added in the
proportion of 2%–3% by weight so as to give sufficient long-term UV protection. Additives
are included to prevent oxidisation, to increase durability, and to act as a lubricant. It is called
‘high density’ because the density of the final product is in the range 0.941–0.954 g/cc.
The components of the mixture (the resin, carbon black, and additives) are fed into a
continuous screw feed, where they are heated and then forced through a die. The HDPE
sheet that is produced is typically 0.75–3.0 mm thick and about 4.5 m wide. Generally, the
edges of the sheet are trimmed during manufacture to give an exact width of sheet before it
is wound onto a spool ready for shipment.
Wider sheets can be produced by heat welding sheets together at the factory in a similar
way to the way in which sheets are connected in the field.
Textured sheet can be produced that has a rough surface. This is an advantage for liners
placed on slopes as the rough surface gives better sliding resistance between the membrane
and any waste materials or soil. The roughened surface can be produced by (1) spraying hot
HDPE particles onto the HDPE sheet, (2) laminating with a hot HDPE foam, (3) passing the
hot sheet through a patterned roller, or (4) use of a blowing agent that expands and creates
a textured surface (see Figure 11.18 for two of the methods).

11.4.1.2 Very low density polyethylene


Very low-density polyethylene (VLDPE) is so called because the density is lower (0.89–
0.912 g/cc) than HDPE. It is made from a polymer resin of ethylene and other alpha-olefins.
Carbon black is added at about 2%–3% by weight as for HDPE to give ultra-violet light
protection.

(a) Hot polyethylene


foam
Spreader bar

Roll of smooth Roll of textured


sheet sheet
(b)

Roll of textured
Smooth
Extruder sheet
sheet

Figure 11.18 Adding texture to HDPE sheet: (a) lamination with polyethylene foam; (b) patterned calendering.
Environmental geomechanics  419

The sheet is produced in a similar way to HDPE sheet, by passing it through an extrusion
die. The finished sheet should be free from any pinholes, surface blemishes, or carbon black
agglomerates.
Texture can be added to the sheet in the same way as for HDPE sheet.

11.4.1.3 Polyvinyl chloride
Polyvinyl chloride (PVC) is made by mixing a PVC resin as a dry powder with a plasticiser
together with a filler and mixing in a blender. The mixture is then heated to 180°C, which
melts the mixture. The viscous mixture is then fed into rollers and rolled into sheet (called
calendering).
The thickness of sheet produced is determined by the separation of the rollers. The thick-
ness of the final sheet should not be less than 95% of the nominal thickness. The sheets of
PVC are generally 1–2 m wide and are transported in rolls weighing up to 700 kg.
PVC sheet can be factory seamed, with one sheet being heat welded or chemically sealed
to the next sheet.

11.4.1.4 Chlorosulfonated polyethylene
Chlorosulfonated polyethylene (CSPE) is made by mixing the CSPE resin with carbon black,
fillers, and lubricants. The polymer is heated and made into a viscous material that can be
calendered into a sheet.
In some cases, a woven fabric, called a reinforcing scrim, is introduced between two
sheets of the CSPE to make it stronger. It is then called a CPSE-R sheet (R = reinforced). The
reinforcing is commonly a woven polyester yarn.

11.4.2 Placing geomembranes
The geomembranes are rolled off the reels and placed onto the subgrade, which should be
compacted to specification, and should not contain any ruts due to construction traffic.
The presence of sharp stone fragments will puncture the membrane, so stones over 12 mm,
especially if angular, should not be present in the subgrade.
The sheets are placed or ‘spotted’ in their correct location, and are tack welded to keep
them in position. This is done with hot air guns generally. At this stage, the geomembrane
may be vulnerable to wind damage as wind can get under the membrane and lift it.

11.4.3 Seaming
The geomembranes can be joined together by (1) extrusion welding, (2) thermal fusion,
(3) chemical fusion, or (4) adhesive seaming.
In extrusion welding, a ribbon of molten polymer is placed along the edge of the sheets to
be joined and melts and fuses with the sheets. However, with fusion seams, a hot air gun is
used to melt the polymers and fuse them together.

11.5 GEOSYNTHETIC CLAY LINERS

Geosynthetic clay liners (GCLs) consist of a clay core supported on its upper and lower faces
by geotextiles. They are sometimes called ‘clay blankets’ or ‘bentonite mats’ because the
predominant clay type used is bentonite, which is a clay from the smectite group.
420  Geomechanics in soil, rock, and environmental engineering

11.5.1 Types of GCLs
There are several variations on the way in which the GCLs are manufactured.

1. The geotextile is attached to the clay with an adhesive.


2. The geotextile can be stitched to the clay that is provided in powdered form.
3. Needle punching can be used to push fibres into the geotextile and bond it to the lower
geotextile.
4. A geotextile can be attached by adhesive to just one side of the clay blanket.

The sheets are generally about 5–7 mm thick and weigh about 3.2–6.0 kg/m 2 . The width
of rolls is generally 2–5 m.
Figure 11.19 shows some of the different types of geosynthetic clay liners.

11.5.2 Manufacturing
The GCLs are manufactured in different ways depending on whether they are glued, stitch
bonded, or needle punched, and some of the methods used are shown in Figure 11.20.
The clay is dried and pulverised, and fed on top of a geotextile along with an adhesive
in the first type of manufacturing process. The adhesive and clay are added in a number of
layers. It then has the upper goetextile added and goes through rollers to produce a sheet of
the desired thickness.
If the product is needle punched, the dry powder is fed onto the lower geotextile, then the
upper geotextile added and needles punch the fibres through into the clay and lower layer.
A line may be drawn on each side of the GCL sheet to indicate the overlap that is needed
in the field as the sheets cannot be joined as geomembrane liners can (i.e. by heat welding).
The GCL is then wrapped in a plastic sheet to prevent it from picking up moisture as this
will make the clay swell as it is supposed to do when wetted in the field.

11.5.3 Placement
When the GCL is taken to the site, it needs to be stored off the ground so that it will not be
affected by moisture. The rolls should also be treated carefully so as not to damage them,
and not stacked high as this can thin the sheet.

(a)
Clay plus adhesive
~ 4.5 mm
Geomembrane

(b)
Clay core Upper geotextile

~ 5 mm Stitch bonding
Lower geotextile

(c)
Clay core Upper geotextile
~ 4–6 mm Needle punched fibres
Lower geotextile

Figure 11.19 Different means of manufacturing geosynthetic clay liners: (a) adhesive bound clay to geotex-
tile; (b) stitch bonded geotextiles; (c) needle punched geotextiles.
Environmental geomechanics  421

(a) Bentonite hoppers

Needle
punching or
stitching

To winding reel

Lower geotextile Lower geotextile


or geomembrane or geomembrane

(b) Upper geotextile or


geomembrane (if
Adhesive used)
Bentonite

To winding
reel

Lower geotextile
or geomembrane

Figure 11.20 Manufacturing processes used for the production of GCLs.

The GCLs are then laid on the subgrade that should be free of ruts from construction
traffic. This is done by rolling the sheet from the spool, and overlapping the sheet onto the
neighbouring sheet. Sharp gravel pieces are to be avoided as well, as these can puncture
the sheet.
Overlap is generally 150–300 mm, and in some cases, bentonite is placed over the join to
seal it. Any punctures or tears in the GCL can be repaired in the same way, that is, by plac-
ing a patch over the hole, and sealing it with dry bentonite or bentonite paste.
The GCL is then covered with a geomembrane or other soil liner. This should ideally be
done before any rain wets the liner as this will cause the GCL to hydrate and swell, making
covering more difficult.
When the liner is placed on side slopes, it is better to cover the liner with soil working
from the bottom up as this puts less tensile stress in the liner.

11.5.4 Examples of use
Some examples of where GCL liners have been used in Australia are:

1.
Swanbank (Ipswich)
Because the clay underlying the site of a proposed landfill was too permeable
(k = 1 × 10−7 m/sec), a GCL was used to cover the compacted clay, and then a geomem-
brane was laid on top of the GCL.
  The GCL used was a needle punched Bentofix X1000 liner (dry bentonite = 4000 g/
m2 and k = 3 × 10−11 m/s) and the covering geomembrane was a 1.5 mm thick HDPE
sheet. A Bidim geotextile was then placed on top of the HDPE liner to protect it from
the leachate collection stone and waste that would tend to puncture the membrane.
422  Geomechanics in soil, rock, and environmental engineering

2.
Gordonvale (near Cairns)
A Bentofix X2000 GCL was used to cap a landfill rather than using a 400 mm thick
compacted clay cover. The GCL sheets were overlapped by 300 mm (no joining was
necessary), but transverse joints were covered with bentonite paste. In all, 23,890 m 2
of GCL was used in the project.
  The GCL was finally covered by 400 mm of a mixture of green waste (50%) and
sandy soil (50%), and monitoring has indicated that leachate production due to water
infiltration has been greatly reduced from that of the uncapped landfill.

11.6 STABILITY OF LINERS

The use of liners at the side of a landfill can lead to a slip of the fill on the liner, and so some
calculation of the stability of the slope needs to be performed before the slope is designed.
Figure 11.21 depicts is a slope that has a membrane liner and a soil liner on top at the side
of a landfill.
In the figure, γc is the unit weight of the clay liner, ϕc is the angle of shearing resistance of
the clay, and cc is the cohesion of the clay liner. The clay liner runs across the base and up the
side of the landfill and has a thickness tc and the side slope is at an angle β to the horizontal.

11.6.1 Tension in the membrane


A geosynthetic or geotextile or combinations of the two lie under the clay liner, and the
minimum angle of shearing resistance of the geomembranes is ϕi. This is because if there are
several types of geotextiles used in combination, slip will occur on the weakest interface.
Two cases are shown in the figure: the first is where the fill is not at the top of the liner
(Figure 11.21a) and the second where the fill is above the top of the liner (Figure 11.21b).
We can draw the vector of forces for a likely mode of failure, which is a two block failure
as shown in Figure 11.22. The vector of forces (Figure 11.23) is drawn assuming that the
inter-block force P is acting parallel to the slope, and the resultant force on the base of block
2 F 2 is acting at the angle ϕi to the normal. An additional force α, which is the tensile force
in the membrane, is shown acting parallel to the slope.

(a)
tc
D
0 A
=
, cc
γ c, ϕ c
H
C ϕi
B
ϕc
β

(b) D
A
tc
H = 0 A′
, cc
γ c, ϕ c
C
B ϕi
ϕc
β

Figure 11.21 Liner on a side slope of a landfill. (a) Fill below top of liner; (b) fill above top of liner.
Environmental geomechanics  423

D A

2
B′ W2

P
1 P
C W1

B
F1 φi
φc

Figure 11.22 Two block failure mechanism.

F1 W1
1

α P

W2

F2

β – φi

Figure 11.23 Vector of forces acting on sliding blocks (α is the tension in the membrane).

We can write the expressions for the weight of each block as shown in Equations 11.9
and 11.10.

1 tc2 γ t2
W1* = γc = c c (11.9)
2 sin β cos β sin 2β

γ ct c  tc  γ ctc2  2H cos β 
W2* =  H −  =  − 1 (11.10)
sin β  2 cos β  sin 2β  tc 

The magnitudes of F 1 and F 2 are not known but their directions are. They can be eliminated
by taking a projection perpendicular to their direction.
424  Geomechanics in soil, rock, and environmental engineering

W1* sin φc = P cos(β + φc ) (11.11)


W2* sin(β − φi ) = (P + α)cos φi (11.12)



Eliminating P from the previous two equations gives
sin(β − φi ) sin φc
α = W2* − W1* (11.13)
cos φi cos(β + φc )

Substituting for the values of W1* and W2* gives the value for the tension in the liner α.

γ ctc2  2H cos β  sin(β − φi ) sin φc 


α =  − 1 −  (11.14)
sin 2β  tc  cos φi cos(β + φc ) 

The analysis shows that there is no need for reinforcement (in the way of a geogrid) if

H < H max (11.15)

where

H max 1  sin φc cos φi 


= 1+ (11.16)
tc 2 cos β  cos(β + φc )sin(β − φi ) 

The fill can be placed in layers in the landfill so that Hmax is not exceeded, and this may
require several cycles of filling and raising of the side liner.

11.6.2 Factor of safety
We can draw the vector diagram for forces on each of the sliding blocks (Figure 11.24) if
we assume that the angles of shearing resistance are only partially mobilised on the lower
triangular block and on the membrane on the upper block but assume that the same factor
of safety applies to each block, that is,

tan φc tan φi
tan φcm = ; tan φim = (11.17)
F F

P2

F1 W1
F2 W2

P1

Figure 11.24 Vector diagrams for forces on each block.


Environmental geomechanics  425

P1

P2

Actual F of S

Figure 11.25 Forces acting on the block interface.

This enables the inter-block forces P1 and P2 to be found. Then a plot can be made of P1 and
P2 versus the factor of safety (Figure 11.25). The required factor of safety is found at the
point where the two P forces are equal.

11.7 PROCESSES CONTROLLING POLLUTANT TRANSFER

Pollutants can be carried through the soil by several processes. The first is by advection,
where the pollutants move along with the pore water as it flows through the soil. The pol-
lutants are therefore moving at the speed of the liquid.
The second is by diffusion. Chemicals dissolved in the groundwater can move from regions
of high concentration to regions of low concentration even if the water is not flowing at all.
If the velocity of flow of the groundwater is low, then the movement of pollutants may be of
the same order of magnitude, and the effects of both need to be considered. For example, if
water was flowing into a landfill due to advective flow and pollutants were moving out due
to diffusion, then there could be a net inward movement of pollutants.
The third is due to dispersion which can be caused by turbulent flow and mixing of the
pollutants in the groundwater.

11.7.1 Advective transport
The flow of polluted water through soil (leachate) from a landfill or other waste repository
is governed by Darcy’s law as we saw previously in Section 3.3 in Chapter 3. The equation
of flow is given for the x direction by

∂h
vdx = kxix = kx (11.18)
∂x

where
vdx is the Darcy velocity in the x-direction
kx is the permeability or hydraulic conductivity of the soil in the x-direction
ix is the hydraulic gradient, that is, the change in total head with distance in the
x-direction

Similar expressions can be written for the other two axis directions.
426  Geomechanics in soil, rock, and environmental engineering

The Darcy velocity is merely an average velocity as it is measured in the laboratory by


collecting a volume of water Q passing through an area A in a given time t, that is k = Q/At.
The real groundwater velocity v depends on the actual area through which the water is
passing, and this depends on the porosity n of the soil as the water is flowing through the
pores in the soil only. Hence
vdx ki
vx = = xx (11.19)
n n

The mass of pollutant transported by advection per unit area per unit time is called the
mass flux F and can be computed from

F = nc(vx + vy + vz ) (11.20)

where c is the concentration of pollutants (having units of mass per unit volume).

11.7.2 Diffusive transport
The mass flux F of pollutants due to diffusion is given by Fick’s First Law (1855) which states
that the mass flux is proportional to the gradient of the concentration and may be written as

 ∂c ∂c ∂c 
F = −De  + +  (11.21)
 ∂x ∂y ∂z 

where De is the effective (molecular) diffusion coefficient, generally taken to be the same in
the three axis directions.
Because diffusion occurs mainly in the pores of a soil, Fick’s Law needs to be written:

 ∂c ∂c ∂c 
F = −nDe  + +  (11.22)
 ∂x ∂y ∂z 

11.7.3 Dispersive transport
Dispersion can occur when there is non-homogeneity in the flow of the water that can cause
mixing and spreading of pollutants. Dispersion can be caused at molecular scale through
Brownian movement or at larger scale by flow turbulence.
Dispersion is often dealt with by lumping together the coefficients of dispersion and diffu-
sion through the use of a ‘coefficient of hydrodynamic dispersion’, D where

D = De + Dm (11.23)

In the above equation, De is the coefficient of diffusion for the chemical of interest. The
mechanical dispersion Dm depends on the velocity of flow v, and in aquifers the coefficient
of dispersion can be described as

Dm = αv (11.24)

where α is the dispersivity. This is a scale-dependent property and not a material constant.
In aquifers, the flow rate tends to be relatively high and dispersion dominates diffusion.
Studies have suggested that the coefficient of hydrodynamic dispersion can be estimated from
D = De + 1.75 dv (11.25)

Environmental geomechanics  427

where
d is the mean grain diameter of the soil in metres
v is velocity in metres per year giving D in m 2 /year.

For clay liners, the velocity of flow is low, therefore we can assume that D = De.

11.7.4 Sorption
Due to reactions between the chemicals in the pore water and the clay particles in the soil,
some chemicals may be removed from solution and sorbed onto the clay particle.
There are several chemical mechanisms responsible for adsorption. These include

• Ion exchange: For example, if the waste has high concentrations of sodium ions Na+
these can exchange with calcium cations Ca 2+ from the clay causing a reduction in Na+
concentration but an increase in Ca2+ concentration in the pore water.
• Precipitation: An example is precipitation of heavy metal ions due to a change in pH.
• Removal of radio nuclei.

There are several relationships that have been proposed for the mass of contaminant
removed from solution S (the mass of solute removed from solution per unit mass of solid).
The first is a simple linear relationship where

S = Kd c (11.26)

The sorption is therefore proportional to the concentration of pollutant c and the constant
of proportionality Kd called the distribution coefficient (having units of L3 M−1).
Another relationship is given by the Langmuir isopleth

Smbc
S = (11.27)
1 + bc

It can be seen that as the concentration becomes large, the sorption tends to S m and so
this term represents the maximum amount of mass that can be sorbed. The parameter b
represents the rate of sorption of a pollutant species.

11.7.5 One-dimensional transport
The above equations can be simplified if the problem is one-dimensional in nature such as at
the base of a wide repository. The total combined mass flux is then given by

∂c
Ftot = nvc − nDe (11.28)
∂z

Considering the conservation of mass within a small volume, we can write the following
equation (if radioactive decay is not included):

∂c ∂F ∂S
n = − tot − ρ (11.29)
∂t ∂z ∂t
428  Geomechanics in soil, rock, and environmental engineering

This equation states that the rate of increase in the concentration of pollutants in a small
volume is equal to the increase in mass due to advection, diffusion, and sorption onto the soil
particles. Substituting for the flux from Equation 11.28 and for the sorption using Equation
11.26 into Equation 11.29 gives

∂c ∂ 2c ∂c ∂c
n = nD 2 − nv − ρKd (11.30)
∂t ∂z ∂z ∂t

or in more compact form

∂c ∂ 2c ∂c
= D* 2 − v* (11.31)
∂t ∂z ∂z

where
c is the concentration of pollutants
t is the time
z is the depth

and

D * v
D* = , v = and R = (1 + ρKd /n)
R R

11.7.5.1 Ogata–Banks solution
The Ogata–Banks solution (1961) to Equation 11.30 is for an infinitely deep deposit of soil
with a constant velocity of flow v and a constant surface concentration, c 0. The boundary
conditions that apply are

c(z, 0) = 0 z ≥0
c(0, t) = c0 t ≥0 (11.32)
c(∞, t) = 0 t ≥ 0

The solution for the above boundary conditions is

c0   z − vt  (vz / D)  z + t 
c = erfc   + e erfc   (11.33)
2   2 Dt  2 Dt  

where erfc is the complementary error function.


The solution of Equation 11.33 is presented graphically in Figure 11.26 so there is no need
to evaluate the complimentary error function erfc. For pollutants that are retarded due to
sorption, this expression may be used by scaling the time of interest T so that t = T/R where
R is the retardation coefficient (R = 1 + ρK/n).
Environmental geomechanics  429

0.999

0.995

0.98
0.95 D = 100
νz
0.90
50
0.80 20
10
0.70 5
c0

0.50 2
c

c = c0
0.30 1
0.5
z
0.10 0.2
0.05
0.1
0.01
0.005 0.05

0.001
0.1 0.5 1.0 5 10
νt
z

Figure 11.26 The Ogata–Banks solution to the dispersion–advection equation (constant surface


concentration).

11.7.5.2 Booker–Rowe solution
If the concentration of the contaminant is not constant, which it would not be if the con-
taminant is leaving a landfill and not being replaced, then the Booker–Rowe solution is
applicable (Booker and Rowe 1987).
If the finite mass of leachate in a landfill is represented by a height of leachate Hf, then the
solution to the concentration of contaminants with depth z and time t is given by

2  bf (b, t) − df (d , t) 
c(z, t) = c0e(ab − b t )   (11.34)
 (b − d) 

where

2  a 
f (b, t) = e(ab + b t )erfc  + b t
2 t 
2  a 
f (d, t) = e(ad + d t )erfc  + d t
2 t 
 n + ρKd 
a = z  (11.35)
 nD 
 n 
b = v 
 4D(n + ρKd ) 
nD  n + ρKd 
d = −b
H f  nD 

430  Geomechanics in soil, rock, and environmental engineering

In order to evaluate the function f(p,t) where p can be either a or d, it may be noted that

2
f (p, t) = e −(a / 4t )
φ(x) (11.36)

and where

 a 
x =  + p t (11.37)
2 t 

The function ϕ(x) can be found from a plot once x has been calculated, therefore the solu-
tion can be found by using a calculator. The plot of ϕ(x) versus x is shown in Figure 11.27.

EXAMPLE 11.1
Use the Booker–Rowe solution to find the concentration of contaminants in a layer of
clay given the following data:

Downward Darcy velocity va = 0.002 m/annum


Diffusion coefficient D = 0.01 m2/annum
Porosity n = 0.4
Sorption ρKd = 1.2
Initial source concentration co = 2000 mg/L
Time of interest t = 150 years
Depth of interest z = 2 m
Height of leachate in landfill Hf = 1 m

0.002
v = = 0.005 m /a
0 .4

1.0

For x > 5
0.8
5 ϕ (5) 0.5535
ϕ (x) ≅ =
x x
Note:
0.6 2
ϕ (–x) = 2ex − ϕ(x)
ϕ(x) 2
f (p, t) = e–a /4t ϕ(x)
0.4 α
where x = p √ t +
2 √t

0.2

0
0 1 2 3 4 5
x

Figure 11.27 Function ϕ(x) used in the Booker–Rowe solution.


Environmental geomechanics  431

½ ½
 n + ρKd   0 .4 + 1 .2 
a = z = 2 = 40
 nD   0.4 × 0.01 

½ ½
 n   0.4 
b =v   = 0.005  = 0.0125
 4D (n + ρKd )   4 × 0.01(0.4 + 1.2) 

½ ½
nD  n + ρKd  0.4 × 0.01  0.4 + 1.2 
d = −b = 125 = 0.0675
− 0.01
H f  nD  1  0.4 × 0.01 

 a2   0.5a 
f (b, t) = exp  −  φ  bt ½ + ½ 
 4t   t 

= 0.0695φ(1.786)

From the graphical solution of Figure 11.27, ϕ(1.786) ≈ 0.3

∴ f (b, t) = 0.0183 × 0.3 = 0.0208


 a2   0.5a 
f (d, t) = exp  −  φ  dt ½ + ½ 
 4t   t 

= 0.0695φ(2.459)

From the graphical solution of Figure 11.27, ϕ(2.459) ≈ 0.21

∴ f (d, t) = 0.0183 × 0.21 = 0.0146


c = coexp (ab − b2t) [bf (b, t) − df (d, t)] / (b − d)


= 2000 × 0.0212
= 42.4 mg/l

Using the Ogata–Banks solution (with the same data), we have

ρKd 1 .2
R = 1+ = 1+ = 4
n 0 .4

D 0.01
= =1
vz 0.005 × 2

In this case, we must use the transformed time because there is sorption, and real time
T = 150 years.

T 150
t = = = 37.5
R 4
432  Geomechanics in soil, rock, and environmental engineering

vt 0.005 × 37.5
= = 0.0938
z 2

From graph in Figure 11.26, c/co ≈ 0.05,


Hence, the concentration for a constant surface concentration is

c = 0.05 × 2000 = 100 mg/L


This is more than double the concentration than for the case where the concentration of
leachate reduces as leachate flows into the soil (i.e. the Booker–Rowe solution).

11.8 FINITE LAYER SOLUTIONS

The finite layer method that was discussed in Chapter 2 can also be applied to the spread
of contaminants. The process involves applying integral transforms to the governing equa-
tions along with a Laplace transform. Once the equations are solved in transform space,
then numerical inversion is employed to obtain the solutions in real time. The method is
explained in detail in the book by Rowe et al. (1997).

11.8.1 Three-dimensional solutions
If we have a source of contamination as shown in Figure 11.28, pollutants can potentially
flow downwards through the soil and into any more permeable layer lying beneath.
The equation governing the three-dimensional diffusion–advection of contaminants in
soil is given by the following equation:

∂ 2c ∂ 2c ∂ 2c ∂c ∂c ∂c  ρKd  ∂ c
Dxx + Dyy 2 + Dzz 2 − vx − vy − vz = 1 + (11.38)
∂x 2
∂y ∂z ∂x ∂y ∂z  n  ∂t

Landfill

Hf
x
1
2
1
3

j k

Permeable layer
z

Figure 11.28 Three-dimensional contamination problem for layered soil.


Environmental geomechanics  433

The finite layer method can now be used to solve these governing equations by firstly
applying a Laplace transform to simplify the time element of the equations and then a double
Fourier transform to simplify the three-dimensional aspect of the equations. For example,
the transform of the concentration c is given by
∞ ∞

C =
∫ ∫c e − i ( ηx + γy)
d ηd γ (11.39)
−∞ −∞

C =
∫C e − st
dt (11.40)
−∞

If the flow of groundwater is only in the z-direction (i.e. downward), then we have on
transforming Equation 11.38

∂ 2C ∂C ρKd 
(
− η2Dxx + γ 2Dyy C + Dzz ) ∂z 2
− vz
∂z

= 1 +
 n 
sC (11.41)

It may be noted that this is an equation similar to the two-dimensional case if we make
the substitution ξ2 = (η2 + γ2(Dyy /Dxx)).

∂ 2C ∂C  ρKd 
Dzz − vz = ξ 2DxxC +  1 + sC (11.42)
∂z 2
∂z  n 

This equation has the simple solution

C = Aeαz + Beβz (11.43)

If we make a substitution for the right-hand side of Equation 11.42 of X, that is,

 ρKd 
X = ξ 2DxxC +  1 + sC (11.44)

 n 

then the solutions for the quantities α and β become

vz  vz2 X 
α = +  4D2 + D  (11.45)
Dzz zz zz

vz  v2 X 
β = −  z2 + (11.46)
Dzz  4Dzz Dzz 

In a layered soil system, we want to ensure continuity of the flux Fz between one layer of
soil and the next at the interface of these layers. The flux is given by

∂c
Fz = nvzc − nDzz (11.47)
∂z
434  Geomechanics in soil, rock, and environmental engineering

therefore the transformed flux equation becomes

∂C
Fz = nvzC − nDzz (11.48)
∂z

If the fluxes are matched at the top and bottom of the layers we can solve for the constants
A and B and can write the concentration anywhere within the layer in terms of the concen-
trations at the faces j and k of the layer, C j and Ck

 eα(z − zk ) − eβ(z − zk )   eα(z − zj ) − eβ(z − zj ) 


C = C j  α(zj − zk )  + Ck  α(zk − zj )  (11.49)
e − eβ(zj − zk )   e − eβ(zk − zj ) 

and so we can write the relationship between the transformed flux at the top and the bottom
of the layer

 Fzj  Qk Rk  C j 
  =    (11.50)
 − Fzk   Sk Tk  Ck 

In the above matrix,

nDzz (βe µβ − αe µα )
Qk =
e µβ − e µα
nD (β − α)
Rk = − µβzz
e − e µα
(11.51)
nDzz (β − α)e µ(β + α)
Sk = −
e µβ − e µα
nDzz (βe µβ − αe µα )
Tk =
e µβ − e µα

where

µ = zk − zk − 1

This relationship is analogous to the finite layer matrices for stress analysis where the
transformed displacements are related to the transformed forces at the top and bottom of
each layer. Hence, we can assemble the layer matrices in the same way, here noting that the
fluxes will cancel at the layer interfaces because of continuity requirements.

11.8.2 Boundary conditions
We need to apply some boundary conditions to the transformed equations and these can be
of two kinds. First, at the top of the layered soil profile, we have the landfill and the concen-
tration of contaminants can be specified there. These can be diminishing with time if there
is only a finite amount of contaminant available in the landfill.
At the base of the layered soil profile, there can be an aquifer that carries the pollutants
away through horizontal flow.
Environmental geomechanics  435

11.8.2.1 Boundary condition at the base


The condition for the concentration at the base is

t
 fb vb ∂cb ∂ 2cb ∂ 2cb 
cb =
∫  n h − n ∂x + Dxxb ∂x2 + Dyyb ∂y 2  d τ
b b
(11.52)
0

The subscript b denotes that the quantities are for the permeable base layer, and the thick-
ness of the base layer is h. It is assumed that the concentration is constant across the depth
of this base layer.
Applying the double Fourier transform and the Laplace transform to Equation 11.52 gives

1 Fb v
Cb = − b i ηCb − Dxxb η2Cb − γ 2DyybCb (11.53)
s nbh nb

Hence, we can find a relationship between the flux and concentration at the base

Fb = YbCb (11.54)

where

 i η vb  Dyyb 
Yb = nbh  s + + Dxxb  η2 + γ 2
 nb  Dxxb 

11.8.2.2 Boundary condition at the surface


At the surface, we have the condition that the concentration in the landfill c LF is the initial
concentration c 0 minus a reduction in concentration due to flow of contaminant into the
landfill. This can be expressed as (Rowe and Booker 1986)

t  
1
cLF = c0 −
∫ ∫
ALF H LF 
 fLF (x, y, τ)dxdy d τ
0  ALF

(11.55)

where A LF is the area in plan of the landfill, H LF is the height of the contaminant source, c 0
is the initial concentration, and f LF is the flux of the contaminant into the surface of the soil.
Taking the Laplace transform of Equation 11.55 gives

c0 1
cLF =
s

sALF H LF ∫f LF (x, y) dxdy (11.56)
ALF

Suppose we can write the surface flux as a function of the surface concentration

FLF = ψCLF (11.57)


436  Geomechanics in soil, rock, and environmental engineering

where ψ is the transformed flux at the surface for a unit value of the transformed concentra-
tion. The value of ψ can therefore be found by solving the Equations 11.66 for a unit value
of the transformed concentration at the surface, and calculating the flux at the surface from
Equation 11.50.
If the transform of the concentration can be written

CLF = T cLF (11.58)


we can calculate the transform parameter T, by taking the double Fourier transform of the
concentration. For a rectangular landfill as shown in Figure 11.28 that has a total width B
and total length L this can be written

+ B / 2+ L / 2

CLF =
∫ ∫c LF ei (ηx + γy) dxdy (11.59)
−B / 2−L / 2

so for this case, the value of T is

sin(ηL / 2)sin(γB/ 2)
T = 4 (11.60)
ηγ

The flux f LF can be represented by an inverse transform of F LF, hence

+∞ +∞
1
fLF =
4π 2 ∫ ∫F LF e − i(ηx + γy)d ηd γ (11.61)
−∞ −∞

therefore if we substitute the value of f LF from the equation above into Equation 11.56, we
have

c 1  +∞ +∞ 
cLF = 0 −
s 2
s4π ALF H LF ∫ ∫∫
ALF 

 −∞ −∞
ψTcLFe − i(ηx + γy)d ηd γ  dxdy

(11.62)

which becomes after performing the integral over the area of the landfill

c 1  +∞ +∞ 
cLF = 0 −
s 2
s4π ALF H LF
∫∫
 −∞ −∞
T .T ψcLFd ηd γ 

(11.63)

Solving for the transformed concentration at the surface gives

c0
cLF = (11.64)
1+ Λ
Environmental geomechanics  437

where

1  +∞ +∞ 
Λ = 2
s4π ALF H f
∫∫
 −∞ −∞
T .T ψd ηd γ 

(11.65)

Hence, we can find the solution of the finite layer equations for a unit value of the trans-
formed concentration at the surface. The final solution of the finite layer equations is then
multiplied by the value of c LF from Equation 11.64 to get the actual solution.

11.8.3 Assembly of finite layer matrices


The global matrix for all of the layers of soil may now be assembled by adding the individual
layer matrices together (Equation 11.50) and noting that the flows will cancel at the inter-
faces of the layers. The resulting matrix is a diagonal matrix with a full bandwidth of three,
so that solution of the set of equations is very rapid.

Q1 R1   C1   FT 
S    
 1 T1 + Q2 R2   C2   0 
 S2 T2 + Q3 R3   C3   0 
    
     =   
       
    
 Sn − 2 Tn − 2 + Qn − 1 Rn − 1   0
 Sn − 1 Tn − 1 + Qn Rn     
C 0
   n −1   
 Sn Tn + Yb   Cn   0 
(11.66)

11.8.4 Inversion of transforms
Once the finite layer equations are solved, the transformed variables need to be inverted to
obtain the concentrations and fluxes in real time and at Cartesian coordinates. The inverse
integrals are not easily evaluated algebraically, so numerical integration is used.
For the inversion of the double Fourier transform, Gaussian integration is used (see
Section 3.7 in Chapter 3) and for the inverse Laplace transform the numerical algorithm due
to Talbot (1979) is used.

11.8.5 Solutions for a three-dimensional problem


The solution was programmed by the author for a three-dimensional landfill that is rectan-
gular in plan. The soil beneath the landfill was assumed to consist of a single layer of clay
underlain by a more permeable sand layer. The depth of the upper clay layer was assumed
to be 2 m thick and the underlying aquifer 1 m thick. The height of the leachate in the fill
initially was assumed to be Hf = 1 m and it is assumed that there is a finite mass of contami-
nant in the landfill so that the concentration diminishes with time. The properties shown in
Table 11.1 have been assigned to the soil layers.
The results of the analysis are shown in Figure 11.29. The full line shows the concentra-
tion to initial concentration ratio across the centreline of the fill (at y = 0) for a strip landfill
438  Geomechanics in soil, rock, and environmental engineering

Table 11.1  P
 roperties of soil used in analysis
Layer Quantity Symbol Units Value
Clay Vertical Darcy velocity va m/a 0.0
Porosity n 0.4
Sorption potential ρKd 0.0
Coefficient of hydrodynamic dispersion D m2/a 0.01
Sand Horizontal Darcy velocity vb m/a 1.0
Porosity nb 0.3
Sorption potential ρKd 0.0
Coefficient of hydrodynamic dispersion
Horizontal DH m2/a 1.0
Vertical Dv m2/a 0.2

(i.e. very long in the y-direction). It can be seen from the plot (which is made at time = 300
years) that the concentration is largest just near the edge of the landfill (x = 100 m).
If we now consider a landfill that has all of the same soil properties and geometry, except
that the landfill is 200 × 200 m in plan, then the concentration is given by the broken line.
It may be seen that if the landfill is rectangular, the concentration of pollutants is less than
for the strip case.

11.9 REMEDIATION

In situ treatment methods depend on the soil conditions, the extent of the contamination,
and the type of pollution present. The treatment may be physical, chemical, biological,

200 m
0.24
Hf =1 m
x Clay
2m z
0.20 1m vb Sand Strip 200 m wide
cb1

0.16

ρKd = 0
c0
cb

0.12 cb1

0.08
Square 200 m × 200 m
0.04
CL t = 300 years

–100 –50 0 50 100 150 200 250 300 350 400


Lateral distance (m)

Figure 11.29 Concentration in aquifer across base of landfill at 300 years for two- and three-dimensional
analyses.
Environmental geomechanics  439

thermal, or combinations of these methods (Vidic and Pohland 1995). The key feature of an
in situ method is that the contaminated soil is treated where it is found, and is not moved
elsewhere for treatment (CIRIA 1995).
The major difficulty in treating soils in place is ensuring effective contact between the
contaminant and the treatment agents. This is affected by the permeability of the soil, and
contaminants that can clog the pores. Methods such as hydraulic fracturing, pneumatic
fracturing, electrokinetics, and ultrasonic methods plus a variety of other techniques can be
used to try to improve the penetration of treatment agents.

11.9.1  I n situ leaching and washing/flushing


The term ‘leaching’ applies to processes where the contaminants are dissolved in the leach-
ing agent and are removed from the soil in this manner. Washing or flushing refers to a
process where a fluid is used to flush the contaminants out (e.g. oil flushed with water).
Figure 11.30 shows some of the techniques that can be used for flushing contaminants.
Some techniques make use of a hydraulic barrier, where water is pumped into the ground
so as to raise the water table, and then water is extracted so as flow of water is always
through the contaminated region and does not go outside (Figure 11.30b). Other tech-
niques make use of a barrier wall that is constructed to contain the contaminated material
(Figure 11.30a).
Acidic solutions are used for metal recovery (e.g. cadmium) or for basic organic materials
(amines, ethers, and anilines). Surfactants aid the desorption of oily materials and form an
emulsion that can be flushed from the ground.
Once extracted, the water or extraction fluid can be decontaminated by using materi-
als that adsorb the contaminants, chemically reacts with them, or achieves microbial
degradation.

(a)
Monitoring Extraction Containment
well well wall
Water infiltration

Contaminated
zone

Water level

(b) Temporary
Monitoring Extraction Infiltration
well sheet piling well
well
Ponded area

Contam-
inated zone

Figure 11.30 Soil flushing systems.


440  Geomechanics in soil, rock, and environmental engineering

11.9.2 
I n situ chemical treatment
Chemical treatment may involve oxidation, reduction, polymerisation, or precipitation.
Whatever process is used, the treatment is aimed at destroying or detoxifying the pollutant
in the soil.
The chemicals used can be sprayed or ponded on the surface so that they seep into the soil,
or can be injected in wells.

11.9.2.1 Oxidation
An oxidising agent may be added to the soil such as ozone, hydrogen peroxide, or hypochlo-
rates. One drawback is that oxidising agents will react with vegetable matter in the soil (if
present) and be consumed by this non-target material.
Chemicals such as alcohols and glycols, cyanides, metals and metal compounds, phe-
nols and cresols, chlorinated organics, and sulphides can be converted to more mobile
products through the use of oxidants. For example, a formaldehyde spill was treated by
using hydrogen peroxide and this reduced the concentration from 30,000–50,000 mg/kg to
500–1000 mg/kg.

11.9.2.2 Chemical reduction
Chemical reduction is the process in which a reducing agent (that is an electron donor) is
added to the soil. An example is iron powder, which can be used to treat organic compounds
such as chlorobenzene and cyclohexanol.
Another example is the use of ferrous sulphate to reduce hexavalent chromium to trivalent
chromium for a site that had been used for a chromate smelter.

11.9.2.3 Polymerisation
Smaller molecules can be linked to form larger molecular chains by polymerisation. The
larger molecules are less mobile and generally less toxic. Chemical polymerisation can be
used for oxygenated monomers such as styrene, vinyl chloride, and isoprene. Agents such as
sulphates can be used to initiate the polymerisation.

11.9.3 
I n situ biological treatment
Biological treatments mainly involve the use of microbial processes, but may include the use
of enzymes or the use of plants that will uptake pollutants through their root systems.

11.9.3.1 Microbial treatment
One process is to use water conditioned with nutrients such as nitrogen and phosphates,
oxygen, and (in some cases) biological agents such as bacteria. To do this, the soil must
be permeable enough to allow the nutrients to infiltrate and the permeability should be
>10−2 m/s. A simple system for achieving this is shown in Figure 11.31.
Sources of oxygen can be air, pure oxygen, hydrogen peroxide, or ozone. Of these, hydro-
gen peroxide is the most favoured, but it is toxic to micro-organisms at high concentrations,
so this must be controlled.
An example of bioremediation is the treatment of groundwater contaminated with
methylene chloride, n-butyl alcohol, and dimethylene. Water extracted from the ground
Environmental geomechanics  441

Biological agents
Discharge and nutrients
Monitoring wells
Infiltration Gas Water
trench
Water
extraction Contaminants

Figure 11.31 Simple in situ microbial treatment system.

can be treated with air, nutrients, and micro-organisms, and then pumped to a sec-
ond tank where the micro-organisms settle out and are re-used. Water from the settling
tank is then infiltrated back into the soil. This can reduce the contaminant levels by
around 90%.
Biological treatments can be used on contaminants such as petrol, asphalt, jet fuel,
and oil.

11.9.4 Soil venting
Soil venting or soil vapour extraction (SVE) is a process whereby injected air vapourises a
contaminant that is then removed through extraction wells. The extracted vapour is either
incinerated, condensed by cooling, adsorbed onto activated carbon, or allowed to vent into
the atmosphere.
A PVC pipe is installed down to the zone of volatile pollutants and then (see Figure 11.32)
is extracted by a vacuum pump and pumped to the treatment plant. The pipes need to have
a screen (a slotted or perforated section) at the base that is installed in permeable material
(i.e. sand) and then sealed in with bentonite along the upper part of the tube. This stops air
leakage along the side of the pipe.
This method is applicable to volatiles such as petrol, diesel, kerosene, and gas oil, but
there are a number of other liquids such as vinyl chloride, carbon tetrachloride, and toluene
that can be vaporised by this method.

Vapour–liquid
separator
Groundwater
Pressure Vacuum Vapour
well – to
gauge pump/blower treatment
treatment
Monitoring
installation

Vapour
extraction

Contaminated Groundwater
zone level

Figure 11.32 Vapour extraction system.


442  Geomechanics in soil, rock, and environmental engineering

11.9.5 Thermal desorption
The process in general involves the injection of steam or hot air into the ground to strip vola-
tiles from the soil and vaporise them. This can be done in two ways: by injecting the steam
or hot air through a cutting/mixing head or injecting it straight into the ground.
Shown in Figure 11.33 is a typical arrangement whereby the steam is injected into the
ground through the cutter bits and comes back to the surface where it is collected in the
collector shroud. Hot air is injected down the outer pipe to assist in the recovery and the
combination of hot air and steam raise the temperature of the soil to 80°C.
This system is applicable to organic compounds with high volatility such as petroleum
wastes and halogenated solvents.

11.9.6 
I n situ stabilisation/solidification
Stabilisation involves mixing stabilising materials such as cement (with other additives), or
lime or fly ash with the soil so that the permeability of the soil is reduced and the contami-
nants are effectively bound up in the soil.
One such process is the deep mixing process where the grout is injected into the ground
down the auger and is mixed into the soil by the mixing blades. This is shown schematically
in Figure 11.34.
The group of augers (that may be three abreast) and mixing arms are attached to the
boom of a crane. The augers are used to bore into the soil, and as they are lifted up, the
stabilising agent is added and mixed with the soil by the mixing blades.
Another process is jet grouting. In this process, the grout is sprayed under high pressure
from the central grout tube, and cuts into the soil and is mixed with it. Cement or cement/
bentonite mixes can be mixed into the soil in this manner. Permeabilities can be lowered to
between 10−6 or 10−9 m/s reducing advective transport of contaminants.
The process is suitable for containing polychlorinated biphenyls (PCBs), asbestos, inor-
ganic cyanides, radioactive materials, and volatile and non-volatile metals.

Steam
generator

Air
Cooling compressor
tower

Water Air Shroud

Condensed Process
volatiles train

Mixing blade
Cutter blade
Cutter head

Figure 11.33 Steam injection system.


Environmental geomechanics  443

Generator
Water supply

Power supply

Cement

Cement storage Grout Grout Agitator Mixer


mixer pump blade

Figure 11.34 Slurry mixing unit for in situ stabilisation.

11.9.7 Electro-remediation
Low level DC currents are passed through the soil to remove contaminants through electro-
kinetic and electro-chemical means. Electro-osmosis causes a movement of the groundwater
towards the cathode, and this carries the contaminants with it. Positively charged species such
as Cd, Pd, Cu, and As, Zn, will also move toward the cathode whereas negatively charged
species will move toward the anode (e.g. CN−, CrO24 −, and AsO3− 4 ) as shown in Figure 11.35.
Material migrating to the cathode and anode needs to be flushed to remove the contami-
nants and permit adjustment of the pH there.

Generator or
mains supply
+ –

Purification Purification

Conditioning Conditioning

Pumps

2+
OH− Cu
2− H3O+
SO4
2+
CN− Pb
Zn 2+
NO3 −

F− H2O
2+
PO3− Cd
4
Cl−
Contaminated
soil

Figure 11.35 Electro-remediation process.


444  Geomechanics in soil, rock, and environmental engineering

Starter
material
Electrodes
Backfill

Contaminated Molten
zone soil

Electrodes
Solidified soil and
lowered
contaminants

Figure 11.36 In situ vitrification.

The process works best in saturated soils, and can be used in fine-grained soils. As the
groundwater migrates, it can carry pollutants with it and can be effective in removing phe-
nol, benzene, and toluene.

11.9.8 In situ vitrification


In situ vitrification involves inserting electrodes into the ground and passing a current
between them so as to melt the contaminated solids at temperatures of between 1600°C and
2000°C. Any organic compounds are evaporated or destroyed and inorganic pollutants are
immobilised in the liquid mass when it solidifies.
A starter mixture is placed on the surface of the soil between the electrodes as shown
in Figure 11.36. The current initially passes along the starter path and melts the soil and
once this occurs, the molten soil will act as the conductor. The electrodes can then be low-
ered melting the soil and contaminants. Once complete, the vitrified soil is left to cool and
harden.
A hood is placed over the electrodes and molten soil and gas is extracted from within the
hood and taken to a treatment plant, but in most cases the hot gases oxidise into less harm-
ful compounds. The primary requirement is for a large amount of power (say, 13.8 kV at
4.25 MW) to pass through the soil. Silica or alumina need to be present as these are the best
for creating a melt at the temperatures achievable.
The method has been used for immobilising mercury, pesticides, dioxins, and PCBs.

11.10 MINING WASTE

In many mining operations, the hard rock ore is mined and crushed to a fine sand consis-
tency and the minerals are removed by floatation or chemical processes. The remaining
fine material left over at the tail end of the process is called the tailings. This material,
which often contains toxic or environmentally undesirable processing chemicals, has to be
disposed of economically and safely. This is done by pumping the tailings away from the
processing plant and storing them in a tailings impoundment.
The embankments which retain the tailings must be constructed as cheaply as possible,
and as a result they are often constructed from the tailings themselves. Only the coarser (and
therefore drier) fraction of the tailings is used to construct the embankments; the fines and
water are confined in the lagoon.
In Australia, this type of construction, although used in some cases is not common, and
tailings embankments are mostly constructed to higher standards from borrow materials or
Environmental geomechanics  445

from mine overburden rockfill much like conventional water storage dams. These dams are
however staged (i.e. increased in height during the mining process) so as to spread the cost
of construction throughout the life of the mining project and the design must be such as to
allow the height of the embankment to be increased.
Embankments constructed from the tailings themselves are, by necessity, constructed in
stages, with a small embankment or starter dike being constructed at the outset of a proj-
ect. As more tailings are produced, the embankment is increased in height. This leads to
economy of operation as not all of the cost of the tailings dam has to be outlaid at the start
of the mining process.
With huge volumes of tailings being produced worldwide, embankment heights and
impoundment areas have increased. The largest dams are the Syncrude Dam (Mildred Lake
Settling Basin) in Alberta, Canada, holding 540 million m3 of tailings (in 2014) and are up
to 88 m high and the New Cornelia tailings dam in Arizona, holding 209 million m3.

11.10.1 Properties of tailings
The tailings produced from different mining operations have quite different compositions
and characteristics. Some examples are given below:

1.
Gold: The ores are often treated with sodium cyanide and so the tailings are toxic.
2.
Aluminium: ‘Red mud’ tailings are typically 35% sand-sized particles.
3.
Coal: Washing of coal results in waste consisting of coarse shale particles, fine coal
particles, and clay. Water draining from tailings can contain high concentrations of
sulphates.
4.
Uranium: Waste may have a very low pH (i.e. are acidic) and high metal content.
Seepage must be monitored and collected.

Tailings produced from different ores and different mills have quite a wide variation in the
grain size distribution. Some typical grading curves are shown in Figure 11.37.

11.10.2 Tailings dam construction


Many different methods of construction are employed although they may be broadly
grouped as upstream, centreline, or downstream methods.

11.10.2.1 Upstream method
With this method, the dam is built up with successive retaining dikes being constructed on
top of previously deposited tailings. This is shown in Figure 11.38.
The dikes are formed from the coarser fraction of the tailings. Two methods may be used
in construction.

11.10.2.2 Spigotting
The tailings are pumped onto a tailings ‘beach’ by a series of pipes running from a main
feed pipe. As the tailings run down the ‘beach’, the heavier and coarser sand particles settle
out first and remain on the downstream face of the dam, while the finer fraction (the slimes)
run into the lagoon. To provide the required freeboard on the crest of the dam, drag lines,
or bulldozers may be used to reclaim the material close to the crest.
446  Geomechanics in soil, rock, and environmental engineering

Sand sizes
Gravel sizes Coarse Med Fine Silt sizes Clay sizes

6" 3" 1½" ¾" ¼" 4 10 20 40 60 100 200 U.S. sieve sizes
100
90

80

70
Percentage finer

60 Typical range of
material sizes
50

40

30

20

10

0
100 10 1 0.1 0.01 0.001
Grain size (mm)

Figure 11.37 Gradation of typical mine tailings.

Spigots discharging tailings into the tailings impoundment are shown in Figure 11.39.

Advantages
1. The advantage of the upstream method over other methods is simplicity of
construction

Disadvantages
1. Since the dike is built over the slimes as the embankment is raised, this leads to
reduced stability since
a. Water in these layers will tend to seep towards the face of the embankment
b. Slimes have less strength than the coarser fractions

Spigot

Sand “beach”
Sand dikes

Downstream face

Slimes

Free draining
starter dike

Figure 11.38 Upstream method of construction using spigotting.


Environmental geomechanics  447

Figure 11.39 Spigots discharging tailings.

c. There is little control over construction and therefore no real basis for engi-
neering design
d. Earthquake or nearby blasting can shock and liquefy the sand in the dike

This method is rarely used in countries where earthquakes are likely and is no longer con-
sidered satisfactory for major tailings dams in such areas.

11.10.2.3 Cycloning
A hydrocyclone is used to separate the coarse fraction (sands) from the fines (slimes). A sche-
matic diagram of a hydrocyclone is shown in Figure 11.40. This shows the coarse material
(called the underflow) and the fine material (called the overflow) being separated out from
the slurry feed from the mill.
As the slurry spins inside the cyclone, the coarse material is centrifuged to the outside and
exits from one end of the device, while the fines are forced to the centre and exit from the oppo-
site end. The resulting sand fraction would typically contain from 10% to 20% of material finer
than the 75 μm sieve. An embankment constructed in this manner is shown in Figure 11.41.

Advantages
An embankment constructed using cycloned coarse fraction has advantages over one
constructed using spigotting since
1. It is possible to form and control the width of the sand zone forming the down-
stream shell
2. The slimes do not come as close to the downstream face

Disadvantages
1. The downstream sand shell still has to be constructed over previously deposited slimes
448  Geomechanics in soil, rock, and environmental engineering

Overflow (fine fraction)

Slurry feed

Underflow
(coarse or sand fraction)

Figure 11.40 A hydrocyclone.

Hydrocyclone

Downstream face

Starter dike
(free draining) Slimes

Figure 11.41 Upstream method of construction using a hydrocyclone.

This leads to reduced stability of the embankment compared to an embankment con-


structed using the centreline or downstream methods.

11.10.3 Centreline method of construction


With this method, a cyclone is used to deposit sand above the centreline of a starter dike
as shown in Figure 11.42. The cyclone is shifted vertically upward as the embankment is
raised.

Advantages
1. Requires significantly less sand than the downstream method but the factor of
safety is only slightly less than for the downstream method
2. The embankment is much more stable than one built using the upstream method
as it is not built over the slimes
Environmental geomechanics  449

Hydrocyclone

Downstream face
Sand
Slimes

Starter dike

Figure 11.42 Centreline method of construction.

Disadvantages
1. Usually limited to where cyclones can separate sand and slimes
2. Large amounts of sands are needed in the tailings and the volume needed increases
with the height of the embankment. Generally, it is not practical to construct a
dam by this method if the tailings contain more than 75% slimes
3. As the downstream slope is continually changing, it is difficult to apply slope pro-
tection to prevent erosion

11.10.4 Downstream method
For this method, a hydrocyclone is used to produce sands which are placed on the downstream
side of the starter embankment or dike (see Figure 11.43). The cyclones are moved downstream
as the embankment rises and material is added to the downstream face of the embankment.

Advantages
1. The embankment is not built over the weaker slimes

Hydrocyclone

Ultimate
downstream
face Slimes
Sand

Starter dike
Free draining
rock toe

Figure 11.43 Downstream method of construction.


450  Geomechanics in soil, rock, and environmental engineering

2. As the embankment is built entirely of competent material, it can be designed and


constructed to the required factor of safety. Allowance can be made for the effects
of seismic shock
3. Seals on the upstream face of the embankment or drainage layers beneath it can be
added at various stages of construction

Disadvantages
1. As for centreline method of construction

11.10.5 Embankments built entirely of borrowed materials


Embankments can be constructed of waste rockfill or imported materials. Where rockfill is
used, an impervious clay core with filter zones needs to be used as for water storage dams.
Impervious upstream seals (e.g. bitumen, HDPE) may be used on homogeneous dams made
from borrow materials (see Figure 11.44a,b).

11.10.6 Tailings storages
Storage lagoons may be constructed on flat ground, on hillsides, or in valleys.

1.
Flat ground storages: Such storages are not subject to runoff inflows, only from direct
precipitation. However, they require containing embankments which surround the
entire impoundment area (see Figure 11.45a).
2.
Hillside storages: Tailings storages constructed on hillsides are often tiered as shown
in Figure 11.45b. Runoff from the slope needs to be intercepted and channelled around
the tailings impoundment.
3.
Valley storages: Storages built in valleys can consist of single embankments or multiple
embankments however such storages are subject to large inflows of water from the sur-
rounding countryside (see Figure 11.45c) and this can cause problems. It is best that

(a) Upstream seal


and filter

Compacted borrow
Tailings material

Drainage layer with filter

(b) Sloping clay core Filters

Tailings
Rock fill

Figure 11.44 Tailings dams constructed from borrow or waste material: (a) homogeneous dam; (b) waste
rock dam with clay core.
Environmental geomechanics  451

(a)

(b)
Stage II

Stage I
Stage I Slimes

(c)

Tailings

Retaining embankment

Figure 11.45 Types of tailings storages: (a) flat ground; (b) hillside; (c) valley.

the impoundment is constructed at the head of the valley to minimise inflows; how-
ever, diversion ditches can be placed on the valley sides to channel water away. Water
from any stream in the valley can be channelled beneath the tailings in a conduit.

11.10.7 Control of water
Unlike a water storage dam, the tailings embankment is usually not designed to store water.
Water in the tailings occupies valuable space which could be occupied by solid waste mate-
rial and in many cases schemes (such as underdrains) are used to remove water from the
solid waste. However, in some cases it is desirable to retain water in the tailings lagoon. For
example, in the uranium mining industry, the tailings may be kept beneath a surface layer
of water in order to prevent dangerous radon gas from escaping into the atmosphere (sub-
aqueous deposition). This however can lead to the tailings having very low densities which
makes rehabilitation of the waste more difficult. Sub-aerial deposition is most favoured for
achieving high densities.
In other cases, water may be contaminated with toxic or undesirable chemicals and it
needs to be prevented from overtopping the embankments, entering the groundwater system
and streams or overflowing through a spillway system.
452  Geomechanics in soil, rock, and environmental engineering

(a) (1) Evaporation (4) Emergency spillway


(3) Decanting

(2) Seepage

(b) (2) Direct precipitation


(1) Runoff (3) With tailings

Figure 11.46 (a) Water outflows from tailings pond; (b) water inflows to tailings pond.

It is therefore important to be able to control the water entering or leaving the tailings
lagoon. There are several ways in which water can enter and leave; these are shown sche-
matically in Figure 11.46a,b.

11.10.7.1 Inflows
Inflow of water may come from several different sources:

1.
Precipitation: It is not possible to control direct water influx by precipitation, however
this is not usually a problem as in theory, 50 mm of rainfall will only cause a 50 mm
rise in water level.
2.
Water entering with tailings: The amount of water entering with the tailings can only
be regulated by a small amount, as certain minimum levels are required for processing
or pumping the slurry in pipes.
3.
Runoff: The amount of runoff entering a lagoon depends on where the dam is con-
structed. Large amounts of runoff would be expected if the dam is built in a valley.
Three possibilities exist for control
a. Retain water in tailings lagoon and remove it through a decant system.
b. Pass water around tailings lagoon by constructing a diversion system.
c. Allow water to flow through an emergency system, that is, a spillway.

11.10.7.2 Outflows
Water can leave the tailings impoundment by several different mechanisms:

1.
Evaporation: The rate of evaporation is governed by the pond area. Evaporative losses
can be minimised by using a small area for the free pond water (i.e. deep pond with a
small surface area is better than a shallow pond with a large surface area, if losses are
to be minimised). Surface chemicals may also be used.
2.
Decanting: Water may be removed from the surface by syphoning water or by using
a floating pump. Alternatively, a buried decant line may be used. Water enters a pipe
which is buried at the toe of the tailings embankment.
Environmental geomechanics  453

3.
Spillway: A spillway may be provided to deal with the maximum expected flood. For
some types of tailings, which are potential pollutants, overtopping is not allowed.
Sufficient freeboard must be allowed to accommodate the expected influx of water.
4.
Seepage: Depending on the type of operation, it may be necessary to stop all seepage
losses (i.e. toxic chemicals in tailings). In other cases, it is only necessary to con-
trol seepage which can cause failure of the embankment due to piping. Water in
the embankment lowers the factor of safety against collapse. Drains such as those
described in Section 11.10.5 may be used to control seepage (i.e. toe drain or blanket
drains).

Remedial action may be taken if seepage is observed discharging from the face of the tail-
ings, by dumping filter material over the zone of emerging water (see Figure 11.47).

11.10.8 Stability of embankments
Stability of embankments can be assessed using any of the methods discussed in Chapter 7
on slope stability. Slip circle methods may be coupled with seepage analysis so that pore
pressures in the embankment are more accurately modelled.

11.10.9 Piping
If a tailings embankment is constructed from borrow materials, then the soil should be
tested to see if it is dispersive. Dispersive soils are those that easily go into suspension and
therefore will wash away with any water seeping through the embankment. The soil is
eroded through a small cylindrical hole or ‘pipe’ initially but eventually the hole will widen
until collapse of the embankment occurs (see Figure 11.48).
There are many tests that can be performed on soils to discover if they are likely to be
dispersive. A very simple test is the Emerson crumb test that involves dropping a crumb of
soil into a beaker of water. If the soil is highly dispersive, it will begin to go into suspension
and make the water cloudy. Depending on the reaction of the soil in water, the soil is placed
into one of eight Emerson class numbers; Class 1 is the most dispersive and Class 8 is the
least dispersive.
Details of how to perform the test are given in Australian Standard AS 1289.3.8.1 (2006)
and ASTM D6572-13e1 (2013).
Another more direct test of a soil’s resistance to dispersion and erosion is the pinhole
test. In this test, developed by Sherard et al. (1976), water is passed through a small hole in
a sample of the soil. The rate of erosion determines the classification (see ASTM D4647).

Coarse filter
Fine filter

Seepage from toe of


embankment and from
foundation

Figure 11.47 Remedial measure to prevent piping.


454  Geomechanics in soil, rock, and environmental engineering

Figure 11.48 Piping occurring in an earth dam embankment.

11.10.9.1 Filters
Filters can be used to resist erosion of the soil in a tailings embankment, even if the soil is
dispersive although it is not good practice to use dispersive soil. A filter consists of a soil
which has a grain size distribution that is fine enough to stop the particles of the protected
soil from washing through it. Filters may be constructed in stages with a sand filter to pro-
tect a clay or silt, and then a gravel filter to protect the sand filter.
Figure 11.44b shows an embankment dam constructed from waste rock with a sloping
clay core. Downstream of the core is the filter zone that protects the core from erosion. A
sloping clay core such as the one shown has the advantage that it is less prone to cracking as
any settlement or compression in the core will result in the core and rockfill above moving
down without cracks opening across the width of the core.
Filters can be designed by using the design approach of Sherrard and Dunnigan (1989).
Filter requirements are given for the protection of four different soil groups.

Group 1: Fine silts and clays with more than 85% passing the No. 200 (75 μm) sieve.

For these soils, the filter requirement is

D15f ≤ 9 × D85b ; but not smaller than 0.2 mm (11.67)



D15f is the diameter of soil particles for which 15% of the sample is finer, and the subscript
f denotes the filter. This diameter can be found from a grading curve for the filter material.
Similarly D 85b is the grain size for which 85% of the protected (or base) soil is finer.

Group 2: Silty and clayey sands and sandy silts and clays with 40%–85% passing the No.
200 (75 μm) sieve.

D15f = 0.7 mm (11.68)



Environmental geomechanics  455

Group 3: Silty and clayey sands and gravelly sands with 15% or less passing the No. 200
(75 μm) sieve.

D15f ≤ 4 × D85b (11.69)


Group 4: Soils intermediate between Groups 2 and 3.

For these soils, interpolation is used between Groups 2 and 3 requirements giving

(40 − A)
D15f ≤ (4 × D85b − 0.7 mm) + 0.7 mm (11.70)
(40 − 15)

where A is the percentage of the base soil passing the No. 200 (75 μm) seive after regrading
of the soil so that 100% passes the No. 4 (4.75 mm) seive.
Generally, for Group 1 and 2 soils, the condition that the filter contains less than 5% fines
is stipulated so that the filter is permeable enough to drain water away from the filter, that
is, <5% of sample is finer than the No. 200 (75 μm) seive.
Although the rules of Sherard and Dunnigan (1989) do not specify additional criteria
other than those listed above, often the criterion

D15f
≥ 4 or 5 (11.71)
D15b

is used to ensure that the filters are more permeable than the soils they protect and to pre-
vent build-up of seepage forces and hydrostatic pressures. By applying this rule to the sand
and gravel filters as well, the grading range for the filters can be established.
The U.S. Army Corps of Engineers (2004) basically use the above rules for design of filters
with some small modifications. For Group 3 soils, they specify

D15f ≤ (4 to 5) × D85b (11.72)


while their permeability criterion allows a wider range than Equation 11.71, that is,

D15f
≥ 3 to 5 (11.73)
D15b

An example of a filter design is shown in Figure 11.49. Filter ‘A’ is the sand filter to protect
the clay core, and filter B is the gravel filter to protect the sand filter. The sand filter is placed
downstream of the clay core, then the gravel filter placed downstream of the sand filter.
Rockfill can then be placed downstream of the gravel filter.

11.10.10 Cut-offs and barriers


Pollutants from tailings storages can enter groundwater and rivers and may therefore harm
the environment. Various schemes have been used to reduce seepage losses and these may be
different to those used for municipal waste repositories in some cases. Some of the methods
are listed in the following.
456  Geomechanics in soil, rock, and environmental engineering

U.S. standard sieve sizes in inches U.S. standard sieve numbers Hydrometer
18 12 8 6 4 3 2 1½ 1 ¾ ½ 3 4 6 10 14 20 30 40 50 70 100 200
100

90
e d85

Filte
80

r
B (gr

Filt
70 Im
pe

e
avel

rA
rv
Percentage finer

60 io
us

(san
fi
lter) b ≤ 9 d85 cl
ay

d fi
50 co
re

lter
40

)
30 b ≥ 3 to 5c
f ≥ 3 to 5a
20
d ≤ 4 to 5e d f a b c
10

0
100 10 1 0.1 0.01 0.001
< 5%
Grain size in millimetres

Figure 11.49 Design of filters to protect a clay soil. (Adapted from U.S. Army Corps of Engineers. 2004.
General Design and Construction for Earth and Rockfill Dams, EM1110-2-2300, July 30.)

11.10.10.1 Controlled placement of tailings


This approach involves using the fine fraction of the tailings (the ‘slimes’) for lining the base
of the storage thus providing resistance to seepage. If this is done, the whole of the base of
the impoundment should be lined with slimes as water can flow around any partially con-
structed liner.
For a slimes liner to be effective, it would be necessary to have at least 40% passing the
No. 200 (75 μm) sieve. If spigotting is used, the area next to the spigot discharge will con-
tain the sand fraction and so it is necessary to move the spigot around or to put an imperme-
able membrane liner under the area which contains the sand fraction.
Slimes liners have the advantage that they are cheap to construct and are not susceptible
to cracking or rupture as clay or synthetic liners may be. However, they are not able to
reduce pollutant levels due to sorption as natural clay liners are able to do for some con-
taminant species.

11.10.10.2 Clay liners
Clay liners for tailings impoundments are similar to those constructed for municipal waste
repositories, and such liners have been discussed in Section 11.3. The liner may be exposed
to the sun for a period of time before it is covered with tailings, and this can cause cracking
and increased permeability. Covering the liner with 150 mm of clean or silty sand can pre-
vent drying of the clay. Thicknesses of clay liners can vary from 150 to 900 mm thick, but
only need to be 150–600 mm thick if properly constructed.
Underdrains can be used in conjunction with clay liners to reduce the required liner thick-
ness as shown in Figure 11.50. The total head of water acting on the clay liner is greatly
reduced from the height h (without drains) down to the height hmax (with drains).
Environmental geomechanics  457

(a) (b)

Tailings h Tailings
Underdrains

Liner hmax hmax

Figure 11.50 Clay liner (a) with and (b) without underdrains.

The pipes used for the underdrainage must be protected from clogging by filters. Suitable fil-
ters can be designed by using the theory given in Section 11.10.9.1. Geotextiles are also used as
filters, however they should not be exposed directly to the tailings or they may clog. This can be
overcome by placing a sand layer between the tailings and the geotextile (Scheurenberg 1982).

11.10.10.3 Cut-off trenches
When the depth of permeable soil is not great, a barrier to seepage may be constructed by
excavating out the permeable material and replacing it with material of low permeability
(Figure 11.51). Such barriers are usually taken completely through the pervious foundation
down to bedrock, but can be terminated at impervious material above bedrock. Common
depths of cut-off are from 6 to 10 m as beyond this depth difficulty may arise in dewatering
the trench and there may be problems with the stability of the sides of the trench. At greater
depths, slurry trenches and other types of cut-off are likely to be more economic.
In order to prevent any loss of material from the rolled earth cut-off, a filter zone may be
constructed on the downstream side.

11.10.10.4 Slurry cut-off walls


An excavation from 1.5 to 3 m wide is made with a backhoe, excavator, clamshell bucket, or
dragline bucket. The trench is kept open with a bentonite slurry which prevents the trench

326
Possible future raising
322
313 Planned future raisings
314.5

2 2.75
All gravels 1 Rock facing
1
stripped Zone 1 Zone 2
Original surface

Upstream Strip foundation


interceptor drain Rolled earth Downstream
cut-off interceptor drain

Figure 11.51 Rolled earth cut-off.


458  Geomechanics in soil, rock, and environmental engineering

walls from caving in. The depth to which the excavation can be made depends on the equip-
ment used, but approximate depths of excavation are

1. Excavator – 10 m
2. Dragline – 25 m
3. Clamshell – 25 m

When complete, the trench is backfilled from one end, displacing the bentonite. The back-
fill should be well graded but impermeable, the coarse fraction being desirable to limit post-
construction settlement.
A typical composition for the backfill may be a bentonite slurry with a Marsh funnel
viscosity of greater than 40 s and with water losses not greater than 15 cm3 in 30 min. The
Marsh viscosity is obtained by placing 1500 cm3 of slurry in a funnel and recording the
time for 946 cm3 (1 U.S. quart) to flow through the nozzle. The percentage of bentonite may
range from 5% to 15%. Aggregate added to the mixture should have a continuous grading
with particle sizes ranging from 0.02 to 30 mm and if necessary aggregate may have to be
imported from elsewhere to adjust the grading (ICOLD 1985).
The backfill mixture can be made with a bulldozer beside the trench or in a concrete type
batching plant and the aggregate and bentonite mixture blended until a homogeneous mate-
rial with a standard concrete cone slump of 100–200 mm is obtained.
Khancoban Dam, which is part of the Snowy Mountains Scheme in NSW, was constructed
using a slurry cut-off trench and the grading of the material used is shown in Figure 11.52.
Care must be taken that the mixture does not segregate as it is placed, so to avoid this, the
trench can first be filled at one end using a clamshell bucket and then the backfill material
placed at a shallow slope of between 6H:1V and 8H:1V.
Post-construction settlement of material in a trench constructed in this manner would be
of the order of 25–150 mm for trenches of 15–20 m deep.
Owaidat and Day (1988) describe a slurry trench that was constructed for a copper mine
in Arizona. The soil–bentonite mix had a 100 mil (1/10 inches) thick HDPE membrane
placed vertically into the trench to create a sealing wall. The HDPE panels were each 8 feet
1.18 mm

75 mm
75 μm

2.36

4.75

37.5
150

300

600

9.5

19

100

80
Percent passing

60

40

20 Khancoban Dam
Backfill material

0
3/8"

3/4"

1½"
30

4
50

16

3"
200

100

U.S. Standard sieve size

Figure 11.52 Grading of material used for a slurry trench – Khancoban Dam.


Environmental geomechanics  459

wide and joined together with interlocking joints that contained a cord that would swell on
contact with water thus sealing the joint. The soil that was mixed with the bentonite con-
sisted of 5.6% gravel, 39.4% sand, 55% silt, and 3% bentonite was added to this to form
the wall.

11.10.10.5 Grout diaphragm walls


These cast-in-place walls are usually of thickness between 0.5 and 1.5 m. A trench is exca-
vated under a cementitious bentonite slurry referred to as grout, and when complete the
grout in the trench is allowed to harden. The construction sequence is shown in Figure 11.53.
Panels are excavated in the sequence A–B–C–D–E–F–G with the secondary panels being
excavated before the primary panels have fully hardened. This has the advantage that there
are no end stop joints (as are used for concrete walls).
Grouts are produced in batching plants which consist of the cement and bentonite storage
bins, and the bentonite slurry storage bins for storing the grout mixture. These plants are
highly automated and can produce 20–50 m3 per hour.
A grout may typically contain

1. 80–350 kg cement per m3


2. 30–50 kg of bentonite per m3
3. Cement/water ratios in the range 0.2–0.3 if Portland cement is used or 0.1–0.25 if
granulated blast furnace slag is used.
4. If there is a possibility of a groundwater attack on the grout, fly ash is added in propor-
tions varying between 10% and 100% of cement by weight. Pure water can dissolve
free lime in the cement and selenic water can destroy the grout by forming a salt (tri-
calcium sulphoaluminate) which expands in the pores of the hardening mixture.
5. A set retarder can be added. The set retarder allows the grout to remain as a slurry for
a longer period of time (about 15 h) before hardening to allow more time to construct
the adjacent panel.

Order of constructing
panels
A C B E D F

A C B E D G F

Figure 11.53 Construction sequence for A grout diaphragm wall.


460  Geomechanics in soil, rock, and environmental engineering

This type of wall needs to be flexible so that it can endure the deformations which are
caused by the self-weight of the dam and the water loads without cracking or joints open-
ing. The modulus of the wall needs to be only 4–5 times that of the surrounding soil. On
average, the unconfined compressive strength of the grout is only 100 kPa at 28 days and
150 kPa at 90 days (ICOLD 1985).
An example of a grout diaphragm wall is the 27 m deep cut-off wall that was constructed
beneath the Harris Dam, which is 10 km north of Collie in Western Australia (see Bradbury
1990, Potulski 1990). The wall, that was constructed using an 8 tonne clamshell grab, was
800 mm wide and was constructed in panels 2.7 m long that had an overlap of 350 mm at
the surface.
The grout consisted of

1. Cementing agent that was 65% blast furnace slag and 35% ordinary Portland cement
(225 kg/m3)
2. Bentonite (30 kg/m3)
3. Water (913 L/m3)
4. Retarder 0%–2% of cement weight (Daratard combined plasticiser/retarder)

The properties obtained for the grout were

1. Laboratory permeability (average) 2.5 × 10−8 m/s


2. Elastic modulus (average) 67 MPa
3. Unconfined compressive strength 500–800 kPa
4. Plastic behaviour occurred up to 5% strain (often up to 10% strain)
5. ND1 classification in the pinhole test (see Section 11.10.9)

The mix design had to achieve a modulus in the grout which was equal to that of the sur-
rounding soil, so that the wall and soil would settle by the same amount. This was because
it was expected that the embankment construction would cause 500 mm of settlement
of the soil and if the wall were more compressible than the surrounding soil it would not
carry load and shed it to the surrounding soil, thus making the wall vulnerable to hydraulic
fracture. If less compressible, the slurry wall would attract load and it could be overstressed
and cracked.

11.10.10.6 Grouting
Grouting of soil has been discussed in Section 10.11 in Chapter 19 (“Soil Improvement”)
and the same techniques can be used for reducing the permeability of soils to reduce losses
of contaminated water from tailings impoundments.
An example of the use of chemical grouts was the sealing of the foundation for the refinery
catchment lake dam, which was part of the Worsley Aluminium project in Australia. The
foundation consisting of silts, clays, and sand lenses was grouted using a resin forming grout
(Geoseal MQ4). This grout seal was used in conjunction with a conventional rolled earth
cut-off, which was taken through the more porous surface soils, and downstream monitor-
ing bores that could accept a pump if it was necessary to return any seepage to the storage.
If it is necessary to grout the rock below cut-off level because the rock is fractured, holes
can be drilled into the rock and cement or chemical grout injected under pressure so that it
is forced into the fissures in the rock (Houlsby 1990).
Curtain grouting usually involves drilling a row (or several rows: 3–5) of holes which are
drilled and grouted in sequence. This is shown in Figure 11.54. First, the primary holes are
Environmental geomechanics  461

P T S T P

(P) Primary grout holes


(S) Secondary grout holes
(T) Tertiary grout holes

Figure 11.54 Sequence of grouting.

grouted, then the secondary holes which are at intermediate positions and then the tertiary
holes and so on, with the distance between holes being halved each time. This allows packer
testing of the foundation to determine whether the grouting has been successful before
further grouting is carried out. Water/cement ratios (by weight) used for the grout typically
vary between 0.5 and 5 depending on the ease of grouting.
One of the problems with cement grouts is that they tend to ‘bleed’. This occurs if the
grout is being injected slowly or is motionless so that the cement particles can sediment out
from the water. For this reason, the grout must be continuously agitated in a grout mixer
before injection. For slow grout takes that may occur in low permeability rocks or toward
the end of a grouting programme, the grout hole must be bled off to allow the thin grout and
water which rises to the surface to be replaced by thicker well mixed grout.
An example of a grouting operation has been described by Houlsby (in Baker 1985) for
the foundation of Copeton dam. The dam foundation consisted of coarse-grained massive
biotite granite with an inclusion of a small mass of fine-grained granite. The main joints in
the rock were relatively widely spaced and of considerable extent with wide-open surface
joints.
The primary grout holes were drilled at 24.4 m (80 ft) apart, then intermediate holes were
drilled at spacings of 12.2 m (40 ft), 6.1 m (10 ft), and finally 1.5 m (5 ft). Water cement
ratios used varied depending on the water takes indicated in the holes but varied from 0.6:1
up to 3:1 with the thinner mixes being used to start grouting and the thicker mixes to finish
(e.g. start at 2:1 and finish at 0.6:1). The permeability of the foundation was reduced from
21 lugeons down to about 5 lugeons at the final hole opening of 1.5 m. Average grout takes
dropped from 0.8 L/cm in the primary holes down to 0.1 L/cm in the quinery holes.
A lugeon is defined as a water loss of 1 litre per minute per metre length of a test section
at an effective pressure of 1 MPa. The test is performed in a borehole section sealed by
inflatable packers. The borehole should be between 46 and 74 mm in diameter. Rock masses
with lugeon values which are greater than about 20 may have their permeabilities effectively
lowered by grouting. If the lugeon values are less than this, then a great deal of expensive
grouting may not make much improvement.

11.10.11 Synthetic liners
Synthetic liners which have been used for tailings impoundments to control seepage include
concrete, shotcrete, asphalt, synthetic rubbers such as butyl rubber, and EDPM (ethylene
propylene diene monomer), although such materials are not commonly used. More common
462  Geomechanics in soil, rock, and environmental engineering

are the thermoplastic membranes. These materials nearly all have good resistance to acids,
alkalis, and salts produced by the ore processing. The following thermoplastics can be used
in tailings applications, and are similar to plastics used for landfill liners.

1.
Polyvinyl chloride: PVC is among the least expensive of liners; however, it is subject
to degradation by sunlight (ultra-violet light). It must therefore be kept under water
or under cover by soil or tailings. PVC used in tailings impoundment applications is
generally from 20 to 30 mil thickness (a mil is 1/1000 inches).
2.
High-density polyethylene: HDPE has good chemical and sunlight resistance and is
available in thicknesses ranging from 20 to 140 mil. It is easy to heat ‘weld’ the sheets
together to form strong joints that do not leak.
3.
‘Hypalon’ – chlorosulphonated polyethylene: This is among the most commonly used
liner in tailings applications. It has good ageing characteristics and is resistant to sun-
light (UV radiation). It is commonly used in 30–36 mil thicknesses but is about twice
the cost of the equivalent PVC sheet. It is not as easy to join sheets together as it is for
materials such as HDPE.
4.
Chlorinated polyethylene (CPE): This material is similar to Hypalon in most charac-
teristics, but has the advantage that it is cheaper.
5.
Elasticised polyolefin (EP): This material is resistant to sunlight and also has a high
resistance to petroleum derivatives. It costs about the same as Hypalon.

Tests carried out by Abbott and Corless (1993) indicated that PVC, HDPE, and CPE were
satisfactory for containing acid wastes. However, there was only limited experience for
highly alkaline wastes stored for long periods of time.
Tests performed on PVC liners, which had been in service for 8 years and had been per-
manently inundated and seasonally inundated, showed that the PVC performed satisfacto-
rily with only a slight loss in elongation properties.

11.10.12 Seepage return systems


Some seepage, which may be small, will always get past artificial barriers of the kinds
mentioned previously. In some cases, it may be necessary to collect seepage water because
of its toxic nature or it may be cheaper to collect and return seepage water than to invest in
expensive cut-off systems. There are several kinds of collection systems that can be used,
and these are examined in the following sections.

11.10.12.1 Collector ditches
Ditches can be a useful way of intercepting seepage if the seepage emerges in the region of
the ditch. Deeper flows will pass beneath the ditch unless it is made deep enough to go down
to an impermeable stratum.
Such drains will also collect water that runs off the face of the dam and groundwater
from downstream of the dam, and such flows may exacerbate water management problems
if water is returned to the tailings storage.

11.10.12.2 Collector wells
Collector wells can be constructed in a line in order to intercept seepage, which can then
be pumped back into the storage. Like collector ditches, wells are most effective if they are
Environmental geomechanics  463

(a)

Pervious layer

Impervious layer

(b)

Well

Drawndown
water surface

Figure 11.55 Seepage return systems: (a) collector ditch or sump; (b) collector wells.

taken down to a depth where there is a less permeable stratum. There are several disadvan-
tages of using collector wells:

1. The pumps lower the phreatic surface downstream of the storage and therefore tend to
increase seepage rates.
2. The wells have to be pumped (perhaps continuously) and this is costly.
3. The wells draw in water from downstream (of the wells) and this can add to water
management problems.
4. After close down of the storage, the wells will usually not be operated and so other
collection schemes are necessary.
5. The well screens and pumps are liable to corrode or to block, and therefore require
maintenance or replacement.

Hydraulic barriers may be set up in some circumstances. An injection well and an extrac-
tion well are used in conjunction in this case. Water is pumped into the outer line of wells,
and because the hydraulic head is higher there than at the line of extraction wells, water will
always flow towards the extraction wells.
To do this, it is necessary to have a supply of fresh water to inject into the injection wells.
This is because water can flow downstream from the injector wells.
A schematic diagram of a collector ditch and collector wells is shown in Figure 11.55
(Vick 1983).

11.10.12.3 Collection and dilution dams


Seepage dams can be used where seepage can be collected in a natural channel. Usually, the
water is collected and pumped back into the storage, so the collector dam must be close to
the storage so that excess water from runoff is not collected.
Alternatively, the collector dam can be located far enough downstream so that it will col-
lect sufficient water to dilute any seepage to acceptable standards.

11.10.13 Acid rock drainage


Acid generation is caused by the exposure of sulphide minerals to oxygen and water. Minerals
such as galena (PbS) and sphalerite (ZnS) can be oxidised to form acid, however by far the
464  Geomechanics in soil, rock, and environmental engineering

most reactive mineral is pyrite (FeS2). Pyrite can be found in the tailings from lead–zinc,
nickel, gold, copper, coal, and uranium processing. The oxidation of pyrite is represented
by the following equation:

FeS2 + (15/ 4)O2 + (7 / 2)H 2O → Fe(OH)3+2SO24 − +4H + (11.74)

Normally, the sulphide bearing minerals lie at depth in rock beneath the water table, so
that their contact with oxygen is minimised and very little acid is formed. When these min-
erals are processed and exposed to the air, they will oxidise and acid will form much more
readily.

11.10.13.1 Factors affecting acid generation


The mining of the rock, or the crushing of the ore, greatly increases the surface area of the
sulphate minerals and this will accelerate oxidation. Waste rock dumps are permeable to
both air and water, so that the sulphide minerals can be more easily oxidised. Most tailings
deposits have low permeability and remain saturated during the life of the storage so it is
less likely that oxidation will take place. Figure 11.56 shows how acid generation can occur
in waste rock dumps and tailings impoundments.

Potential
sources of
water Rainfall infiltration
Surface water Potential airflow
runoff pathway

Groundwater
inflow

Potential acid
rock drainage
(ARD)

Rainfall infiltration
Zone of oxidation
Surface water Oxidation front
runoff Air

Partially saturated
zone

Groundwater Saturated zone


seepage
Mixture of residual
process fluids and
ARD

Figure 11.56 Acid rock drainage (ARD) formation in a waste rock dump and in a tailings impoundment.
Environmental geomechanics  465

Factors which assist in the chemical oxidation of pyrite are

1.
The pH of the media: Chemical oxidation of pyrite is sensitive to the pH of the reaction
media. At low pH levels, ferric iron acts as an oxidant of pyrite.
2.
Oxygen concentrations: Oxygen is the major factor contributing to the chemical oxi-
dation of pyrite as mentioned previously.
3.
Temperature: Cooler conditions will reduce the rate of reaction and slow down the
rate of acid formation. Temperatures as high as 60–80°C can be generated in waste
dumps and this rise in temperature causes a greater flow of air into the dump as the
hot air rises, sucking fresh air into the base of the dump.
4.
Moisture content: Small amounts of moisture are necessary to assist in the oxidation
process. However, if the sulphides are totally submerged in water, the process will be
slowed considerably due to lack of oxygen.
5.
The presence of micro-organisms: Certain bacteria such as thiobacillus ferro-oxidans
can speed up the oxidation process. These organisms are most active at pH 2.0–5.0.
The rate of ferrous iron oxidation can increase 10–50 times through the addition of
these bacteria to the pyrite system. The optimum conditions for the bacteria are a pH
of about 3.2, a moist environment, and the presence of carbon dioxide and oxygen.
6.
The presence of acid neutralising chemicals in the tailings: Acid produced from sul-
phide oxidation can be totally or partially neutralised by minerals contained in the
ore or from chemicals added during processing. Minerals such as calcite CaCO3 and
dolomite CaMg(CO3)2 are the most common minerals responsible for neutralisation.

11.10.13.2 Control of acid generation


Several measures can be used to combat acid generation. Many of these methods are ways
of counteracting the conditions mentioned above which are conducive to acid generation.
These measures are

1.
Pre-treatment of tailings: Flotation can be used prior to disposal to remove sulphide
minerals, or limestone, cement or bentonite can be added to the tailings.
2.
Covers and seals: Clay seals can be placed over the top of the tailings which will reduce
the ingress of water. Multi-layer covers can cut the infiltration to 1%–3% of rainfall.
3.
Covering tailings with water: If the tailings are covered with water, the oxygen avail-
able is greatly reduced. Tailings can be dumped into existing water bodies such as lakes
or oceans, although today this would not be common. Alternatively, they can be kept
underwater in an old open pit or underground workings or in a constructed impound-
ment where water levels would have to be kept high enough to cover the tailings.
4.
Addition of chemicals to control pH: Alkaline additives such as lime, limestone, or
sodium hydroxide NaOH can be added to neutralise any acid formed.
5.
Bactericides: As bacteria can increase the rate at which oxidation of sulphides occurs,
killing the bacteria will slow the process. Bactericides such as sodium lauryl sulphate
can be used for this purpose.

11.10.13.3 Control of acid migration


Not only can measures be taken to stop the production of acid, but measures can be taken
to try to reduce the migration of any acid which may be produced. Measures which can be
taken are
466  Geomechanics in soil, rock, and environmental engineering

1.
Covers and seals: As discussed above, clay covers with underlying sand drainage lay-
ers can be used to intercept water infiltrating the tailings and transporting the acid
produced into the groundwater.
2.
Interception and diversion of surface water: Surface water can be intercepted and
channelled away from the tailings by surface drains.
3.
Collection and treatment of acid drainage: The leachate from the waste rock dump or
tailings impoundment can be collected downstream from the waste and monitored to
see if acid levels are too high. If so, the leachate can be contained or treated to bring
down the acid level, although the second option is more costly and would have to be
continued for a long period of time after mining had ceased.

11.10.14 Remediation
After mining operations cease, it is generally required by government bodies that the tailings
repositories be made secure for the future. Drainage and leakage from tailings may contain
unacceptable levels of toxicity for many years into the future, and this has to be taken into
consideration. It is therefore necessary to make sure that

1. The impoundment has long-term mass stability.


2. The waste has long-term erosion stability.
3. Contamination of the environment will not occur in the long term.
4. The area may be returned to productive use.

11.10.14.1 Hydrological stability
The best way to guard against damage due to water inflows is to prevent water entering the
waste by providing a cap. The cap needs to be provided with a slope to assist drainage and
prevent ponding of water that could infiltrate the tailings.
Placing the capping may not be easy in some cases. The additional weight of the cap can
cause soft tailings to consolidate, requiring large quantities of capping material to infill
depressions which would cause water to pond if left unfilled.
An example of this is the residue areas in Kwinana, Western Australia, where historically
tailings were placed into the impoundments in a wet state (Coreless and Glenister 1990).
The tailings sands from perimeter beaches have been used to cover the slimes. This was a
time consuming process, because if it was carried out too fast, displacement of the mud at
depths of up to 8 m would occur. One technique which was used was to produce ditches
using an amphibious screw tractor. The ditches would drain the upper layers of the tailings
allowing them to form a crust to a depth of about 500 mm. This was then strong enough to
support the cover of sand.
One of the problems with providing a cover is that often large settlements can occur over
the soft slimes zones with less settlement over firmer regions. This means that large amounts
of fill can be required to infill depressions which can be of the order of 3 m deep. Providing a
slope for drainage on the cap can also be a problem as a slope of only 0.5%–1.0% can mean
significant thicknesses of cap material are required at the centre of a large area.
An example of the design of a cover for a gold mine in Washington State, has been given
by Frechette (1994). Conventional oedometer tests were carried out on the tailings, and
Terzaghi’s one-dimensional consolidation theory was used to predict the amount of consoli-
dation that would be caused by the cover. The compression index C c , of the tailings ranged
from 0.18 to 0.33.
Environmental geomechanics  467

Ramps were constructed on the tailings initially so that access could be obtained for plac-
ing the cover. These ramps consisted of a single layer of geotextile fabric 5.3 m wide, or a
double layer of fabric 7.6 m wide in softer areas. Fill of between 0.9 and 2.4 m thickness and
4 m wide at the crest was placed above the fabric.
Fill placed over the tailings from the ramps ranged in thickness, but was up to 6 m thick
with an average of 3 m.
The design of the cap for another gold mine in northern California is also described by
Frechette (1994). First, a geofabric was laid on the tailings, and an initial layer of rock fill
was placed on the geofabric. The fabric and rockfill (which was up to 6 m thick in some
areas) were placed to allow initial drainage and consolidation of the tailings. The rockfill
was placed in lifts (initial lift 60 cm) and kept 15–30 m back from the edge of the fabric.
The final cover (see Figure 11.57) consisted of (from the bottom up):

• A geofabric
• An initial rock layer
• An impermeable soil layer (90 cm min)
• A soil/bentonite mix (15 cm min)
• A topsoil layer (30 cm min)
• A gravel cover

Settlements of up to 3 m were predicted for the cover in this case.


Experience has shown that for gold tailings reclaimed immediately after deposition,
the tailings will consolidate a further 10% during the initial 2–3 years following closure.
Placement of fill over the tailings to re-contour the impoundment so as to shed surface
Soil with bentonite mix

Foundation rock
Top layer of soil
First layer of soil

Gravel cover

Finished grade
Impoundment limits

Initial foundation
rock layer

Geofabric

0 50 100

Scale in meters

Figure 11.57 Construction sequence for cover of gold tailings.


468  Geomechanics in soil, rock, and environmental engineering

water, may require quantities of up to 10% of the tailings thickness. Additional settlement
caused by cover fill placement may be of the order of 25% of the cover thickness.

11.10.14.2 Long-term erosion stability


On flat, unbroken surfaces, wind erosion is primarily a factor but on embankment slopes
erosion by water is most often a problem. Vick (1983) states that slopes flatter than about
3:1 are usually required for ‘reasonable erosion resistance and establishment of vegetation’.
Eurenius (1990) suggests slopes should be flatter than 2:1 or 3:1 for dams with heights of
between 15 and 30 m.

11.10.14.3 Vegetation
Resistance to erosion (both wind and water) is best provided by vegetation, and it is neces-
sary to provide the correct type of grasses and plants for the climatic conditions in the region.
Native plants are an obvious choice, and should be used wherever possible (Pickersgill 1994).
Where tailings have a low pH (acid) or a high pH (alkaline) or high concentrations of
heavy metals or salts, establishment of vegetation may be difficult. Adjustment of soil pH
can be carried out (e.g. with lime), or drainage layers can be added beneath the final cover
to prevent migration upwards of pollutants by capillary action.
It is often necessary therefore, to strip the natural topsoil before constructing the tailings
impoundment and storing it for later use for establishing vegetation. Topsoil stored too long
in this manner can lose much of its organic matter, nutrients, and bacteria and therefore can
be much less suitable for plant growth than it was originally.
Forrest et al. (1990) describe the rehabilitation of a copper–silver tailings impoundment.
Soil which was stockpiled was placed to a depth of 30 cm over the top of the tailings and
the face of the embankment. Grass was then planted by hydro-mulching and conventional
means. Riprap was placed on the crest and upper part of the embankment (which had a
slope of 3:1) to prevent erosion gullies forming. The estimated cost of this work was esti-
mated at about 19% of the construction cost of the impoundment.
Brett (1990) reports that coal washery discard at the Duncan colliery in Tasmania was
rehabilitated by establishing native flora. The site was firstly contoured and then deep ripped
to a depth of 500 mm and fertiliser added at the rate of 300 kg per hectare. Both native
seedlings and native seeds were planted and these consisted of acacias, casuarinas, banksias,
and eucalypts.

11.10.14.4 Riprap
Wind erosion may also be prevented by placing a layer of riprap over the surface of the tail-
ings. This can consist of either gravel- or rock-sized particles and smelter slag has also been
used successfully.
The disadvantages of riprap are, however, that it can be costly to place and that it prevents
natural plant growth.

11.10.14.5 Prevention of environmental contamination


Long-term prevention of environmental contamination is also of great importance. As has
been discussed before, in many cases the loss of water from pyrite rich tailings will result in
exposure of the tailings to oxygen. One long-term measure against acid drainage is to seal the
tailings with a clay cap to prevent the entry of water which can carry acid from the tailings.
Environmental geomechanics  469

Top soil Top soil

0.3 m

Moraine 1.0 m Moraine

0.2 m
Limestone Bentonite
moraine

Tailings
≈ 100 m

Figure 11.58 Cover design for uranium tailings.

Uranium waste is another example. Upon desaturation, the tailings will be able to emit
radon gas to the atmosphere (nominally saturated tailings will not emit the gas). Generally,
a cover of about 3 m is required to prevent long-term radon gas emission.
An example of a cover for uranium mine tailings is presented by Euranius (1990). A test
pile of tailings was covered by two types of covers as shown in Figure 11.58. One consisted
of limestone, moraine, and topsoil, while the other consisted of a bentonite–moraine layer,
overlain by a moraine layer and a topsoil layer. Overall, each cover was 1.5 m in total thick-
ness. Tests on both of these covers have shown that the infiltration is about 1%–4% of pre-
cipitation, that the weathering is reduced 98%–99% as compared to uncovered tailings and
that radon attenuation has about the same value as the natural background.
This page intentionally left blank
Chapter 12

Basic rock mechanics

12.1  INTRODUCTION

The behaviour of rock and rock masses is different to that of soils mainly due to the fact
that the intact rock generally has a higher strength and deforms less than soil. In addition,
rock mass behaviour is often dictated by any discontinuities or planes of weakness such as
fractures and fissures in the rock, especially if the fissures are infilled with soft material.
Special methods of testing rock and for analysis of rock slopes and structures constructed
on rock (as opposed to soil) have been developed, and some of these methods are included
in this chapter.
The study of rock mechanics is a large field, so only basic rock mechanics are covered
here. The reader is referred to the many specialist texts and papers that have been r­ eferenced
if further information on some of the topics covered is required.

12.2  ENGINEERING PROPERTIES OF ROCKS

The properties of rocks may be found from a number of specialist laboratory or field tests.
Many of the test procedures are outlined in national or international codes or standards
that provide details of how the test should be performed, and where possible these stan-
dards are referred to in the following sections.

12.2.1  Point load strength index


This is a simple test that can be performed in the field or in the laboratory, using a hand
operated hydraulic jack. Generally, rock core is placed between two cone shaped platens
(of included angle 60° and with a 5 mm radius spherical tip, see Figure 12.1) and loaded
until the core undergoes a tensile failure, although the test can be performed on blocks or
lumps of rock. For rock core, the test may be performed by loading across the diameter or
by ­loading in the direction of the axis of the core.
The uncorrected point load index Is is given by Equation 12.1.

P ×1000
Is = (12.1)
D2

where
P is the load at failure in kilonewtons
D is the platen separation in millimetres

471
472  Geomechanics in soil, rock, and environmental engineering

Conical platens

Hydraulic pump

Figure 12.1 Point load testing machine.

The value is then corrected to a value that would have been measured for core of 50 mm
diameter (or dimension) using the following equation:

0.45
 D (12.2)
I s(50) = I s  
 50 

Details of the test procedure are given in the standards AS 4133.4.1 (2007) or ASTM
D5731-08 (2008).
For axial, block, or irregular lump tests, an equivalent distance De is used instead of D in
Equations 12.1 and 12.2 where

4×A
De 2 = (12.3)
π

and A is the minimum cross-sectional area of the plane through the two platen points.
Typical values of the point load index are given in Table 12.1.

Table 12.1  Point load index values for various rock types
Type of rock Is(50) (MPa)
Tertiary sandstone 0.05–1
Coal 0.2–2.0
Limestone 0.25–8
Mudstone 0.2–8
Volcanic lava 3–15
Dolomite 6–11
Hawkesbury sandstone (Sydney) 0.3–1.6
Basic rock mechanics  473

12.2.2  Unconfined compression test


The unconfined compressive strength (UCS) of an intact rock sample can be found by com-
pressing a piece of core axially until the core fails in compression. Generally, the core needs
to have a length to diameter ratio of between 2.5 and 3 so that the end effects are m
­ inimised.
No lateral pressure is applied during the test, and so the sample is ‘unconfined’.
The UCS qu(max) is then simply calculated from

P
qu(max) = (12.4)
A0

where
P is the applied load at failure and
A 0 is the original cross-sectional area of the rock sample.

Details of the test are given in standards AS 4133.4.3.2 (2013), ASTM D7012-14 (2014), or
BS EN 1926 (2006).
The relationship between the UCS and the point load index can be established by test-
ing any particular rock type (in unconfined compression and point load) but for hard rock
Bieniawski (1975) indicates a relationship of

UCS = kI s(50) = 24 I s(50) (12.5)


For Sydney sandstones, the relationship of UCS to axial point load index value has been
found to have the best fit to the data if UCS = 20 Is(50) for axial point load tests (on saturated
specimens) but with a range of k from 15 to 29, and UCS = 24 Is(50) for diametral tests on
saturated specimens with a range of k from 14 to 35 (Pells 2004).

12.2.3  Modulus of rock from unconfined compression test


Rock core can be instrumented during the unconfined compression test so as to obtain the
deformation characteristics. Changes in length and diameter can be measured with LVDTs,
optical devices, or strain gauges. If strain gauges are used the circumferential strain can be
measured, and this is equal to the diametral strain.
The elastic modulus is calculated from

σ P l
E = = (12.6)
εa A0 ∆l

where P is the axial load at any stage of loading, A0 is the original cross-sectional area, and
l is the length of the specimen. The Poisson’s ratio ν of the core is the ratio of the diam-
etral strain to the axial strain. Details of the test are provided in AS 4133.4.3.1 (2009), AS
4133.4.3.2 (2013), and ASTM D7012-14 (2014).
Typical values of UCS and elastic modulus are presented in Table 12.2 for various rock
types.

12.2.4  Confined compressive strength


Rock core may be tested under triaxial conditions much like soil, where the core is placed
in a cell and all round hydraulic pressure applied to the sample. The pressures used are gen-
erally higher than those required for soil, and so the cell is made from steel with specially
474  Geomechanics in soil, rock, and environmental engineering

Table 12.2  Typical rock strengths and moduli


Rock type UCS (MPa) Elastic modulus E (GPa)
Granite 150–200 50–70
Limestone 100–150 50–100
Sandstone 20–50 2–10
Hawkesbury sandstone (Sydney) ~20 ~6

Spherical seat
for axial ram

Steel cell
body

Rock sample

Inlet for cell oil


Axial strain
gauges

Diametral
strain gauges

Rubber sealing sheath

Figure 12.2 Steel triaxial cell for testing of rock samples.

designed membranes that self-seal against the cell (Hoek and Franklin 1968). This cell can
apply about 70 MPa pressure to the specimen. The lateral pressure applied to the sample as
well as the axial pressure allow the strength envelope of the rock to be obtained, and this
is examined later (Section 12.3), as the strength envelopes of rock are found to be different
to those of soils.
Design of the cell is such that strain gauges can be attached to the rock sample and the
wires passed out of the cell through the base (see Figure 12.2), hence avoiding the need to
pass the wires through the membrane. The strain gauges that are attached axially and radi-
ally to the rock sample, allow the moduli of the rock and its Poisson’s ratio to be calculated.
The procedures for the test may be found in ASTM D7012-14 (2014).

12.2.5  Sonic velocity test


Elastic constants can be found for intact rock by propagation of sonic waves through a
rock sample (usually core). Both p and s waves can be passed through the rock and the
Basic rock mechanics  475

compressive vp and shear vs wave velocities measured. Assuming that the rock is homoge-
neous, isotropic, and linear elastic, the shear modulus G, modulus of elasticity E, and the
Poisson’s ratio ν of the rock can be calculated from Equations 12.7 through 12.9.
Since

G = E / 2 /(1 + ν) = ρvs2 (12.7)


 3(vp /vs )2 − 4 
E = ρvs2  2  (12.8)
 (vp /vs ) − 1 

1  (vp /vs )2 − 2 
ν =   (12.9)
2  (vp /vs )2 − 1 

where ρ is the density of the rock.


ASTM D2845-08 (2008) gives the details for performing the test.

12.3  FAILURE CRITERION FOR ROCK

It has been found through testing of rock under triaxial conditions, that the failure surface
when plotted in σ1–σ3 space (σ1 and σ3 are the maximum and minimum principal stresses) is
not a straight line like the Mohr–Coulomb failure surface for soils, but is curved. A typical
curved failure surface is shown in Figure 12.3.
σ1 (MPa)

σci = 100 MPa 190


σtB = –10 MPa
σtU = –9.9 MPa
mi = 10 160
s=1
a = 0.5
130
Triaxial compressive strength

100

Brazilian tensile 70
strength Uniaxial compressive strength

40

Uniaxial tensile strength


10

–15 –5 5 15

–20 σ3 (MPa)

Biaxial tensile strength

Figure 12.3 Curved failure surface for rock.


476  Geomechanics in soil, rock, and environmental engineering

12.3.1  Hoek–Brown criterion for intact rock


The curved failure surface of Figure 12.3 can be approximated by an empirical equation
that was developed by Hoek and Brown (1980a) and is therefore known as the Hoek–Brown
failure surface. The equation for the failure envelope is given in Equation 12.10 where the
constant m is used to fit the equation to test data.

0.5
 σ 
σ1 = σ 3 + qu  m 3 + 1.0 (12.10)
 qu 

Values of m will range from about 5.4 for limestone or marble up to 28 for hard rocks like
granite. More values are given in Hoek and Brown (1980b).

12.3.2  Hoek–Brown criterion for rock mass


The failure criterion of Equation 12.10 was developed for intact rock. However, rock masses
contain fractures or jointing that means that the equation for intact rock will not apply to
rock masses. Hoek and Brown (1980b) proposed the Equation 12.11 for rock masses where
they introduced the parameter s. For intact rock, s is equal to 1.0, but for a completely
granulated rock mass s = 0.0.

0.5
 σ′ 
σ1′ = σ ′3 + qui  mi 3 + s (12.11)
 qui 

Later, Hoek and Brown (1997) introduced a more general equation that had the form

a
 σ′ 
σ1′ = σ ′3 + qui  mb 3 + s (12.12)
 qui 

where mb, s, and a are all material constants for the rock mass (qui is the unconfined com-
pressive strength of the intact rock) and mb is found from the intact rock value of mi.

 GSI − 100 
mb = mi exp 
 28 − 14D 
 GSI − 100  (12.13)
s = exp 
 9 − 3D 

a = +
2 6
e(
1 1 −(GSI /15)
+ e −(20 / 3) )

where GSI is the geological strength index which can be found from Table 12.3, and D is
the damage index which is a measure of blast and stress relaxation around an excavation.
D can range from 0 where there is no damage to 0.7–1.0 where the rock is badly damaged.
Hoek et al. (2002) give examples of obtaining values of D, and their descriptions are given
in Table 12.4.
Basic rock mechanics  477

Table 12.3  Values of GSI

Source: Adapted from Hoek, E. and Brown, E.T. 1997. International Journal of Rock Mechanics and Mining Sciences &
Geomechanics Abstracts,Vol. 34, No. 8, pp. 1165–1186.
478  Geomechanics in soil, rock, and environmental engineering

Table 12.4  Suggested values of the parameter D


Description of rock mass Suggested value of D
Excellent quality controlled blasting or excavation by tunnel boring machine results D = 0
in minimal disturbance to the confined rock mass surrounding a tunnel.
Mechanical or hand excavation in poor quality rock masses (no blasting) results in D = 0 or D = 0.5
minimal disturbance to the surrounding rock mass. Where squeezing problems No invert
result in significant floor heave, disturbance can be severe unless a temporary
invert is placed.
Very poor quality blasting in a hard rock tunnel results in severe local damage, D = 0.8
extending 2 or 3 m, in the surrounding rock mass.
Small scale blasting in civil engineering slopes results in modest rock mass damage, D = 0.7
particularly if controlled blasting is used. However, stress relief results in some Good blasting
disturbance. D = 1.0
Poor blasting
Very large open pit mine slopes suffer significant disturbance due to heavy D = 1.0
production blasting and also due to stress relief from overburden removal. Production blasting
In some softer rocks excavation can be carried out by ripping and dozing and the D = 0.7
degree of damage to the slopes is less. Mechanical excavation
Source: Adapted from Hoek, E., Carranza-Torres, C., and Corkum, B. 2002. Proceedings 5th North American Rock Mechanics
Symposium and 17th Tunnelling Association of Canada Conference, University of Toronto, Toronto,Vol. 1, pp. 267–273.

12.4  CLASSIFICATION OF ROCKS AND ROCK MASSES

The purpose of rock classification systems is to group together rocks having similar
­characteristics. This is useful in designing structures in rock where engineers need informa-
tion about the intact rock and the rock mass.
Rock mass classifications can be used with empirical schemes based on experience and
judgement to carry out engineering design.

12.4.1  Classification on strength


The UCS of the intact rock can be used as a means of classifying rocks. This is not indicative
of the rock mass strength or behaviour if the rock mass is fractured, but is a means of clas-
sifying the parent rock by its strength. Table 12.5 shows such a classification (after Deere
and Miller 1966).

12.4.2  Classification by jointing


As mentioned in the previous section, the rock mass behaviour is generally affected by
the jointing in the rock. Hence, rock classification can be carried out for the rock mass on
the basis of the joint spacing. Such a classification is shown in Table 12.6.

Table 12.5  Classification of rocks by intact rock strength


Uniaxial compressive
Description strength (MPa) Examples of rock types
Very low strength 1–25 Chalk, rock salt
Low strength 25–50 Coal, siltstone, schist
Medium strength 50–100 Sandstone, slate, shale
High strength 100–200 Marble, granite, gneiss
Very high strength >200 Quartzite, dolerite, gabbro, basalt
Basic rock mechanics  479

Table 12.6  Classification of rock mass due to joint spacing


Rock description Spacing of joints (m) Rock mass grading
Very wide >3 Solid
Wide 1–3 Massive
Moderately close 0.3–1 Blocky/seamy
Close 0.05–0.3 Fractured
Very close <0.05 Crushed and shattered

12.4.3  Rock quality designation


One system for indicating the quality of rock is the rock quality designation (RQD) that
was originally introduced by Deere et  al. (1967). The RQD is defined as the percent-
age of intact core pieces longer than 10 cm (4 inches) in the total length of core. The
test was developed for NX size core (54.7 mm or 2.16 inches in diameter) but NQ core
(47.5 mm or 1.87 inches) is also commonly used. The core should be drilled with a double
or triple-tube core barrel so that good recovery is achieved. A section of rock core is
shown in Figure 12.4 showing how the RQD should be calculated. Deere and Deere (1988)
­recommend that the length of core used be about 1.5 m as the result is influenced by the
length of core selected.
The method is outlined in detail in the Standard ASTM D6032 (2008) or British Standard
BS 5930. The description of the rock can then be made as shown in Table 12.7.

Length of
core pieces >10 cm (4″)
RQD = × 100%
Total core run length
L = 38 cm
38 + 17 + 20 + 43
RQD = × 100%
200
L = 17 cm
Core run total length = 200 cm

RQD = 59% (FAIR)


L = 0 cm as no
centreline L = 0 cm
pieces longer
than 10 cm
L = 20 cm RQD
(rock quality Description of
designation) rock quality
0%–25% Very poor
L = 43 cm 25%–50% Poor
50%–75% Fair
75%–90% Good
90%–100% Excellent
Mechanical break
L = 0 cm
caused by drilling
No recovery
process

Figure 12.4 Method of assessing RQD value from rock core. (After Deere, D.U. and Deere, D.W. 1988. Rock
Classification Systems for Engineering Purposes, Ed. Kirkaldie, L., ASTM Special Publication No.
984, pp. 91–101. Philadelphia: American Society for Testing of Materials. Reprinted with per-
mission from ASTM STP 984 Rock Classification Systems for Engineering Purposes, Copyright
ASTM International, 100 Barr Harbor Drive, West Conshohocken, PA 19428.)
480  Geomechanics in soil, rock, and environmental engineering

Table 12.7  Description of rock mass based on RQD


Description of rock mass RQD Ratio Efield/Elab
Very poor 0–25 0.2
Poor 25–50 0.2
Fair 50–75 0.2–0.5
Good 75–90 0.5–0.8
Excellent 90–100 0.8–1.0

12.4.4 
C lassification of individual parameters used
in the NGI tunnelling quality index
The Norwegian Geotechnical Institute (NGI) has developed a system of classifying rock
based on several factors that include the RQD, jointing set numbers, joint roughness,
joint alteration, joint water, and stress reduction. The effect of these factors is quantified
by assigning numerical values. The factors are listed in Appendix 12A and are defined
below:

RQD is the rock quality designation


Jn is the joint set number
Jr is the joint roughness number
Ja is the joint alteration number
Jw is the joint water reduction number
SRF is the stress reduction factor

Three parameters that are crude measurements of the block size RQD/Jn , the inter-block
shear strength Jr/Jn , and the active stress Jw /SRF can then be calculated and multiplied
together to give the Q index for the rock (see Equation 12.24).
The Q index can be used for design of tunnels and this is discussed more fully in Section
12.6.5.1.

12.4.5  Rock mass rating method


The Council for Scientific and Industrial Research CSIR (South Africa) method proposed
originally by Bieniawski (1973) provides a general rock mass rating (RMR) with a ‘score’
between 1 and 100 given to the rock mass.
The RMR is based on five rock parameters and one parameter depending on use. These
parameters are

1. The strength of the intact rock from UCS or point load test Is
2. The quality of the drill core RQD
3. The groundwater conditions
4. The spacing of joints or fractures
5. The condition of joints
6. Joint orientations relative to the tunnel or foundation or slope

The ratings are given in Appendix 12B from an updated scheme by Bieniawski (1989).
Use of the method for tunnel design is given in Section 12.6.5.2.
Basic rock mechanics  481

12.5  PLANES OF WEAKNESS

Planes of weakness include fissures, joints, shear zones formed by interlayer slip during
f­ olding, and faults which are major dislocations caused by tectonic forces. Such planes cause
the rock mass to be weaker than the intact rock and it may be more deformable and aniso-
tropic because of the jointing. In addition, the rock may be highly permeable parallel to the
planes of weakness.
It is rare to find planar weaknesses distributed in a truly random pattern, and often planes
of weakness are oriented in one or more preferred directions. The orientation of joints can
be measured by the dip and the dip direction as shown in Figure 12.5.

12.5.1  Stereographic projections


In order to be able to plot the directions of all planes of weakness in a rock mass, ­stereographic
projection is often used. The idea is to take a sphere and allow the plane of weakness to cut
through the sphere (also passing through the centre of the sphere).
There are different types of projections that are then used to obtain a projection of the
great circle where the plane intersects the sphere. The trace of the great circle can be plotted
or, to make the plot simpler, the pole of the plane can be plotted. The pole is a point that
shows where the normal to the plane of weakness would cut the sphere. Hence, the pole is
just a single point rather than an arc such as the trace of the great circle.
Shown in Figure 12.6 is a plane of weakness cutting through a sphere. The very top point
of the sphere is the zenith, and if lines are drawn from the zenith to the great circle (where
the plane cuts the sphere) then the trace of this on the horizontal plane through the centre
of the sphere gives the stereographic projection of the great circle. This particular type of
projection is called an equal angle projection.
The normal to the plane of weakness through the centre of the sphere will cut the sphere
at the pole. The pole can be projected onto the horizontal plane as well by joining the zenith
to the pole and plotting where it cuts the horizontal plane.
A stereonet can be produced (as shown in Figure 12.7) and the poles or great circles plot-
ted by hand on the net. For example, for the net shown in Figure 12.7, if the dip direction
was 120° and the dip was 50°, then the great circle and pole can be plotted as shown in the
figure. The dip direction is found by rotating clockwise from the north position as shown in
Figure 12.7a. Then the dip angle is found by moving in from the circumference of the circle

N
Strike

Dip direction

Dip

Plane of interest

Figure 12.5 Dip and dip direction of a plane of weakness.


482  Geomechanics in soil, rock, and environmental engineering

Zenith

Stereographic
Stereographic projection of
projection of pole great circle

Pole

Plane of Great circle


weakness

Figure 12.6 Plane of weakness (shaded) cutting sphere and equal angle projection onto a horizontal plane.

to the given angle of dip, that is, 50° as shown in Figure 12.7b. By using tracing paper, the
great circle can be copied and then rotated to the 120° position as shown in Figure 12.7c.
Because the pole is found from a perpendicular to the plane of weakness, the dip direction
of the pole will plot at 180° away from the dip direction of the plane of weakness. The dip
of the pole will be (90° – dip angle) of the plane of weakness.
Computer programs that can be downloaded from the Internet such as GEOrient (2014)
or DIPS (2014) are available to perform the plotting and it is not necessary to plot the
­jointing data by hand on a stereonet any longer. Figure 12.8 shows the equal angle plots of
two sets of joint data. Because the joints are in a similar direction, the traces of the planes
can be seen to fall into two groups. The plots of the poles in Figure 12.8 also shows that the
poles fall into two regions, and such groupings indicate that there are probably two joint
sets in the rock.
Stereonets are very useful for finding intersections between planes. For instance, they can
be used to find the dip and dip direction of the line of intersection of two planes of weakness.
Such information is useful when examining the stability of rock wedges (see Section 12.7.2).

(a) (b) β = 50° β = 50° (c)


N α = 120°

W E Pole
Pole 90°

Great
circle
S Great circle

Figure 12.7 Drawing a great circle (and pole) manually on a stereonet where the dip direction is 120° and
the dip is 50°. (a) Dip direction; (b) dip angle; (c) rotation of great circle to dip direction.
Basic rock mechanics  483

(a) (b)

Figure 12.8 Plots showing (a) projections of planes; (b) plots of the associated poles for two joint sets.
(Adapted from GEOrient. 2014. V9.x http://www.holcombe.net.au.)

This can be done manually by using a stereonet and tracing paper, but this is an outdated
method, and the intersection data can easily be found using computer programs such as those
listed above.

12.5.2  Roughness of joints


The roughness of joints has an effect on the shearing resistance of the joint. Smooth, slick-
ensided joints will slip easily, while joints that are rough with a lot of interlocking will be
most resistant to shearing.
Barton (1973) suggested a method of finding the shear resistance τ of a joint based on
the joint roughness coefficient (JRC). This was expressed in terms of an empirical Equation
12.14.

τ   JCS  
= tan  JRC log10   + φ′r  (12.14)
σ ′n   σ ′
n 

In this equation, JCS is the joint wall compressive strength, φ′r is the residual angle of friction
of the joint, and σ ′n is the normal stress acting on the joint. The equation is therefore similar
to the Mohr–Coulomb criterion where the angle of friction is the residual angle increased by
the joint roughness term (the first term in the square brackets of Equation 12.14).
The joint roughness coefficient ranges from 0 for very smooth joints to 20 for very rough
joints. A means of estimating what the joint roughness is can be seen in Figure 12.9 where a
section of joint 10 cm long is depicted along with the JRC to be assigned to that roughness
(Barton and Choubey 1977).
The JCS is equal to the unconfined compressive strength of the rock if the joint is
­unweathered, but less if it is weathered. The value for weathered joints can be found by
using a Schmidt hammer on the joint surface (see Barton and Choubey 1977).
484  Geomechanics in soil, rock, and environmental engineering

JRC = 0–2

JRC = 2–4

JRC = 4–6

JRC = 6–8

JRC = 8–10

JRC = 10–12

JRC = 12–14

JRC = 14–16

JRC = 16–18

JRC = 18–20

0 5 cm 10

Figure 12.9 Joint roughness profiles. (After Barton, N.R. and Choubey, V. 1977. Rock Mechanics, Vol. 10,
No. 1–2, pp. 1–54.)

12.6  UNDERGROUND EXCAVATION

Underground excavations may be carried out for highway, railway, water supply, and sew-
age tunnels as well as tunnels for mine access and hydro-power. Chambers may be exca-
vated for hydro machine halls, mine plant rooms, drawpoints in mines, or for radioactive
waste disposal.
Rock mechanics can be used in planning the dimensions, shapes, and orientations of the
openings. It can also be used in the design of the support systems, the planning of blasting
and excavation operations, and in monitoring the excavation process.

12.6.1  Support systems


If the stresses in the rock are too large, then there may be failure in the rock surrounding
the excavation. This can manifest itself as rock bursts and spalling of the walls and floor.
Even if there is no failure in the rock, movements may be too large and may increase with
time (creep).
The rock may therefore need to be supported by bolting or by linings. If the rock is
allowed to move and stabilise before the support systems are put in place, then this may
lead to more economic design as the support systems do not have to deform as much and
therefore carry as much load.
Design of support systems is examined in Section 12.6.6.
Basic rock mechanics  485

12.6.2  Design process


The design process for excavations can involve performing two- or three-dimensional stress
analysis, and this commonly involves finite element or finite difference (FLAC) analysis. This
can be used as a preliminary analysis to refine the shapes and spacings of tunnels. Joints can
be included either as major discontinuities or as ‘ubiquitous joints’ in numerical analyses.
The major driving force for the analysis is the initial stress state used, as it is the gravity
stresses and stresses locked into the rock that cause the rock to move when it is unloaded
by excavation.
If only loosening of the rock is expected from the initial analyses, then the primary
­support  system can be based on experience, that is, empirical methods based on rock
­classification schemes. However, if the analysis indicates the strength of the rock being
exceeded then the support systems need to be placed rapidly after excavation. Design can
still be performed using empirical means in this case, but more sophisticated numerical
methods may need to be used.

12.6.3  In situ stresses


The behaviour of excavations and tunnels is largely driven by the stresses being released as
the rock is removed. If there are very high lateral stresses, a tunnel will be squeezed inwards
(along the horizontal diameter) when the rock is excavated.
In some cases, any high lateral stresses can be of benefit to the engineer when a tunnel is
being excavated near the surface. In this case, the lateral stresses tend to compress the rock
in the roof of the tunnel, thus holding any blocks of rock in place. For example, the roof of
the Opera House car park in Sydney is comprised of 8–9 m of rock spanning over 17.5–19 m
(Pells et al. 1991). This is possible because the stress field is such that the lateral stress is 2–5
times as large as the vertical stress, that is, σh = 2–5σv.
The stress in rocks can be measured by using overcoring methods (see Goodman 1989).
The process involves drilling a borehole into the rock and then installing a strain gauged
device into a smaller pilot hole at the base of the borehole. The smaller hole is then over-
cored, thus releasing the stresses in the rock as shown in Figure 12.10. Strain gauges in

Main hole

Pilot hole

Overcoring

Figure 12.10 Overcoring stress measurement.


486  Geomechanics in soil, rock, and environmental engineering

the device record movements that can be used to calculate the stress released. The CSIRO
HI cell is a device used in a pilot hole (Worotnicki and Walton 1976). The South African
CSIR doorstopper gauge system (Thompson et al. 1997) is slightly different as the gauge is
attached to the base of the drill hole before overcoring.
Another way to measure stresses in rocks is to use hydraulic fracturing. A section of
borehole is isolated by packers, and this section is subjected to hydraulic pressure until the
rock fractures at the fracture initiation pressure pf. The pressure in the borehole is then
reduced and reapplied over several cycles. The fracture re-opening pressure pr is the pressure
required to open the fracture on re-pressurising (see Figure 12.11). It is assumed that the
minimum horizontal principal stress σh is equal to the re-opening pressure, that is,

σ h = pr (12.15)

This is then used (Hubbert and Willis 1957) in calculating the maximum horizontal stress
in the ground σH through use of Equation 12.16 (assuming that the tensile stress in the rock
is almost zero).

pr − p0 = 3(σ h − p0 ) − (σ H − p0 ) (12.16)

where
pr is the fracture re-opening pressure
σh is the minor lateral pressure in the rock
σH is the major lateral stress in the rock
p0 is the groundwater pressure

Details of the test are given in ASTM standard D4645-08 (2008).


The direction of the fracture in the rock can be found by using an impression packer,
which is a partially cured thin rubber sleeve that is pressurised against the side of the
­borehole to obtain an imprint of the fracture.

To pump, flowmeter,
To pump
pressure transducers

To pump

Drill rod
Pf = Fracture initiation pressure
Pressure High Pr = Fracture re-opening pressure
transducer pressure Ps = Shut-in pressure
Shut-in
housing hoses Shut-in Ps = Shut-in pressure
Compass
Pressure

Formation pore
pressure
≈1.1 m
Cycle 1 Cycle 2 Cycle 3
Impression
Fracture packer ≈1.2 m
interval ≈0.9 m
Straddle
Flow rate

packer

Drillhole
Time
Drillhole

Figure 12.11 Hydraulic fracturing in borehole.


Basic rock mechanics  487

12.6.4  Stresses around underground openings


Analytic solutions to the stresses caused by tunnel excavation can be found in the book by
Poulos and Davis (1974) and in Hoek and Brown (1980a), and some of these are reproduced
here.

12.6.4.1  Circular tunnel


For a circular tunnel in an infinite mass under conditions of plane strain, the stresses can
be calculated for a stress field that involves a vertical stress pz and a lateral stress px = Npz
The  radial σr tangential σθ and shear τrθ stresses as shown in Figure 12.12 are given
­separately for each of the vertical and lateral stresses, but can be superimposed if both are
acting. The stresses due to each of the vertical and lateral stresses can be scaled and added
also because the solution is based on linear elasticity theory.
Stresses due to uniform vertical pressure pz

pz  a2  pz  3a4 4a2 
σr =  1 − + 1 + − 2  cos 2θ
2 r 
2
2  r4 r 
pz  a2  pz  3a 4 
σθ =  1 + − 1 + cos 2θ (12.17)
2 r 
2
2  r 4 

pz  3a4 2a2 
τ rθ = 1 − + 2  sin 2θ
2  r4 r 

Stresses due to uniform lateral pressure px

px  a2  px  3a4 4a2 
σr =  1 − − 1 + − 2  cos 2θ
2 r 
2
2  r4 r 
px  a2  px  3a4 
σθ =  1 + + 1 + cos 2θ (12.18)
2 r 2  2  r 4 
px  3a4 2a2 
τ rθ =  1 − + 2  sin 2θ
2 r4 r 

pz
σr

z τθr
τrθ σθ
a θ
r
px = Npz
x

Figure 12.12 Circular tunnel in an infinite mass.


488  Geomechanics in soil, rock, and environmental engineering

It may be noted that at the boundary (r = a) the stresses become

σ r = 0, τ rθ = 0
(12.19)
σ θ = pz {(1 + N) − 2(1 − N)cos 2θ}

At the floor and roof (θ = 0° and 180°)

σ θ = pz {3N − 1} (12.20)

and at the sidewalls θ = 90° and 270°

σ θ = pz {3 − N } (12.21)

Other observations are that there are zones of tensile stress if N < 1/3 or N > 3 and that
the stress concentrations reduce rapidly away from the opening as they vary with a 2 /r 2 , and
at values of r/a ≈ 3 the stress field is not very different to the original stress field.

12.6.4.2  Elliptical tunnel


For an elliptical tunnel under the action of the stress state shown in Figure 12.13, the stresses
in the tangential direction at the points A and C are given by

 a 
σ A = pz 1 + 2 − N  (12.22)
 c 

  c 
σC = pz N  1 + 2  − 1 (12.23)
  a 

12.6.5  Support design


Many types of direct support are available including props, arches, rock bolts, and ­precast
concrete segments. Different support systems are used in mining to civil engineering
­applications, as support does not need to be permanent.
Empirical systems have been developed from years of experience in tunnelling and
­mining, and the design methods are based on various rock classification schemes.

pz

z β=0 z
a a α2 β1 σα
β2
A c σβ
x α1 α0
0 C c x
β= π
2
px = Npz

Figure 12.13 Elliptical tunnel in an infinite mass.


Basic rock mechanics  489

Table 12.7a  Stresses on axes of elliptical tunnel (a/c = 0.5)


pz = 1.0 px = 0 pz = 0 px = 1.0 pz = 1.0 px = 0.25
z/a or x/a σv σh σv σh σv σh
Stresses along the z-axis
2.00 0 −1.000 0 5.000 0 0.250
2.03 −0.042 −0.790 0.254 4.423 0.021 0.316
2.06 −0.565 −0.636 0.421 3.974 0.049 0.357
2.12 −0.039 −0.413 0.630 3.281 0.118 0.408
2.35 0.155 −0.102 0.746 2.155 0.341 0.437
2.88 0.491 0.017 0.547 1.470 0.628 0.385
4.67 0.828 0.019 0.192 1.114 0.876 0.298
7.89 0.943 0.008 0.065 1.035 0.959 0.266
∞ 1.000 0 0 1.000 1.000 0.250
Stresses along the x-axis
1.00 2.000 0 −1.000 0 1.750 0
1.06 1.925 0.028 −0.900 −0.013 1.700 0.025
1.11 1.859 0.069 −0.814 −0.021 1.656 0.064
1.23 1.747 0.092 −0.657 −0.026 1.583 0.085
1.59 1.476 0.166 −0.311 0.023 1.398 0.172
2.30 1.219 0.188 −0.030 0.234 1.211 0.247
4.33 1.043 0.099 0.055 0.660 1.057 0.264
7.70 1.010 0.038 0.028 0.875 1.017 0.257
∞ 1.000 0 0 1.000 1.000 0.250

Table 12.7b  Stresses on axes of elliptical tunnel (a/c = 2.0)


pz = 1.0 px= 0 Pz = 0 px = 1.0 pz = 1.0 px= 0.25
z/a or x/a σv σh σv σh σv σh
Stresses along the z-axis
0.50 0 −1.000 0 2.000 0 −0.500
0.53 −0.013 −0.900 0.028 1.925 −0.006 −0.419
0.56 −0.021 −0.814 0.069 1.859 −0.003 −0.349
0.61 −0.026 −0.657 0.092 1.747 −0.003 −0.221
0.80 0.023 −0.311 0.166 1.476 0.065 0.058
1.15 0.234 −0.030 0.188 1.219 0.281 0.274
2.1 0.660 0.055 0.099 1.043 0.685 0.316
3.85 0.875 0.028 0.038 1.010 0.885 0.280
∞ 1.000 0 0 1.000 1.000 0.250
Stresses along the x-axis
1.00 5.000 0 −1.00 0 4.750 0
1.01 4.423 0.254 −0.790 −0.042 4.226 0.243
1.03 3.974 0.421 −0.636 −0.057 3.815 0.407
1.06 3.281 0.630 −0.413 −0.039 3.178 0.621
1.18 2.155 0.746 −0.102 0.155 2.129 0.785
1.44 1.470 0.547 0.017 0.491 1.474 0.670
2.33 1.114 0.192 0.019 0.828 1.119 0.399
3.94 1.035 0.065 0.008 0.943 1.036 0.301
∞ 1.000 0 0 1.000 1.000 0.250
490  Geomechanics in soil, rock, and environmental engineering

12.6.5.1  Q index method


Several rock properties, that is, the J factors and RQD discussed in Sections 12.4.3 and
12.4.4 need to be estimated to calculate the Q index. Once the factors have been estimated
from the tables of Appendix 10A, the rock mass quality Q can be calculated from

RQD Jr J
Q = × × w (12.24)
Jn Ja SRF

Another quantity called the excavation support ratio (ESR) can be introduced which is
related to the intended use and the importance of a tunnel. Barton et al. (1974) apply the
values suggested in Table 12.8.
These values of Q and ESR can be used with a chart such as that from Grimstad and
Barton (1993) (Figure 12.14) to make estimates of the bolt spacing and the thickness of
shotcrete for the rock support system of a tunnel.
Figure 12.14 also gives rock bolt lengths L for an ESR of 1. However, if the ESR is other
than 1, Barton et al. (1980) provide additional information on rock bolt length L as given
by Equation 12.25, where B is the excavation width.

0.15B
L = 2+ (12.25)
ESR

They also give the formula of Equation 12.26 for estimating the unsupported tunnel span

Unsupported span = 2 ⋅ ESR ⋅ Q0.4 (12.26)


The estimated permanent roof support pressure Proof can also be found in terms of the
NGI Q index factors as shown in Equation 12.27.

2 Jn Q−(1/ 3)
Proof = (12.27)
3 Jr

Although these systems provide a practical method of designing tunnel support, they
have their limitations and should be used with caution. Palmstrom and Brock (2006) have
discussed the limitations of the system and Pells and Bertuzzi (2007) have pointed out that
in the Sydney sandstone, the Q method suggests less support than has actually been used on
major projects (as designed by more sophisticated methods).

Table 12.8  The various excavation support ratio (ESR) categories


Class Description ESR
A Temporary mine openings 3–5
B Permanent mine openings, water tunnels for hydro-power (excluding high pressure penstocks), 1.6
pilot tunnels, drifts, and headings for large excavations
C Storage rooms, water treatment plants, minor road and railway tunnels, surge chambers, and 1.3
access tunnels
D Power stations, major road and railway tunnels, civil defence chambers, and portal intersections 1.0
E Underground nuclear power stations, railway stations, sports and public facilities, and factories 0.8
Source: Adapted from Barton N., Lien R., and Lunde J. 1974. Rock Mechanics,Vol. 6, No. 4, pp. 189–236.
Basic rock mechanics  491

Rock classes
G F E D C B A
Exceptionally Extremely Very Very Extremely Excep.
poor poor poor Poor Fair Good good good good
100 2.3 m 2.5 m
20
a
rete d are 1.7 m
2.1 m

Bolt length in m for ESR = 1


50 in shotc 1.5 m
11
cing 1.3 m
Span or height in metres

spa 1.2 m
Bolt 1m 7
20 5
9 8 7 6 5 4 3 2 1
ESR

10 m 4.0 m rea 3

m
0m m m da

m
25 m m m te

m
3.0 m
0m 90 cre

40
0

50
5 15 12 t
2.0 m n sho 2.4
3 1.6 m gi nu
in
2 s pac 1.5
1.3 m
Bolt
1.0 m
1
0.001 0.004 0.01 0.04 0.1 0.4
4 10 1
40 100 400 1000
RQD Jr J
Rock mass quality Q = × × w
Jn Ja SRF

Reinforcement categories
(1) Unsupported bolting (5) Fibre reinforced shotcrete, 50–90 mm, and bolting
(2) Spot bolting (6) Fibre reinforced shotcrete, 90–120 mm, and bolting
(3) Systematic bolting (7) Fibre reinforced shotcrete,120–150 mm, and bolting
(4) Systematic bolting with 40–100 mm (8) Fibre reinforced shotcrete, >150 mm with reinforced
unreinforced shotcrete ribs of shotcrete and bolting
(9) Cast concrete lining

Figure 12.14 Estimated support categories based on the tunnelling quality index Q. (After Grimstad, E. and
Barton, N. 1993. Proceedings of the International Symposium on Sprayed Concrete – Modern Use
of Wet Mix Sprayed Concrete for Underground Support, Fagernes, Eds. Kompen, Opsahl and Berg,
Norwegian Concrete Association, Oslo, pp. 46–66.)

12.6.5.2  RMR method


RMR (Section 12.4.5) can also be used to design tunnel openings in rock. Table 12.9 from
Bieniawski (1989) provides a guide to the support that may be needed for tunnels in rock
for which the RMR has been calculated. Steel sets and shotcrete and rock bolts that are sug-
gested as support systems are discussed in Section 12.6.7.

12.6.6  Support types


Where the rock is not of sufficient quality to support itself or the tunnel spans are wide,
linings are used to support the tunnel. Linings can be concrete, made of steel sets or be
shotcrete used in conjunction with rock bolts. Rock bolts alone can be used if no shot-
crete is deemed necessary to support the rock. Steel sets used as support are shown in
Figure 12.15.
In cases where the rock tends to squeeze into the opening, it is of advantage to delay the
placing of the lining as this will allow the rock to carry some of the stresses released on
excavation and results in less load on the lining. In Austria, innovative linings that contain
a gap that can close as the rock squeezes have been used. Metal cylinders that can collapse
are placed in the gap to control the gradual closing of the lining (Schubert 2007).
492  Geomechanics in soil, rock, and environmental engineering

Table 12.9  Guidelines for excavation and support of 10 m span rock tunnels in accordance with the RMR
system
Rock bolts
(20 mm diameter,
Rock mass class Excavation fully grouted) Shotcrete Steel sets
I – Very good rock Full face 3 m advance Generally no support required except spot bolting
RMR: 81–100
II – Good rock Full face Locally, bolts in 50 mm in crown None
RMR: 61–80 1–1.5 m advance crown 3 m long, where
Complete support 20 m spaced 2.5 m required
from face with occasional
wire mesh
III – Fair rock Top heading and bench Systematic bolts 50–100 mm in None
RMR: 41–60 1.5–3 m advance in top 4 m long, crown and
heading spaced 1.5– 30 mm in sides
Commence support after 2 m in crown
each blast and walls with
Complete support 10 m wire mesh in
from face crown
IV – Poor rock Top heading and bench Systematic bolts 100–150 mm in Light to medium
RMR: 21–40 1.0–1.5 m advance in top 4–5 m long, crown and ribs spaced
heading. Install support spaced 1–1.5 m 100 mm in 1.5 m where
concurrently with in crown and sides required
excavation, 10 m from walls with wire
face mesh
V – Very poor Multiple drifts 0.5–1.5 m Systematic bolts 150–200 mm in Medium to heavy
rock RMR: <20 advance in top heading 5–6 m long, crown, ribs spaced
Install support spaced 1–1.5 m 150 mm in 0.75 m with
concurrently with in crown and sides, and steel lagging and
excavation. Shotcrete as walls with wire 50 mm on face fore-poling if
soon as possible after mesh. Bolt required.
blasting invert Close invert
Source: Adapted from Bieniawski, Z.T. 1989. Engineering Rock Mass Classifications. Wiley, New York.

Figure 12.15 Steel sets used to support a tunnel roof.


Basic rock mechanics  493

Support can be primary or secondary support. Primary support is applied immediately


after excavation to make the tunnel safe for subsequent excavation, after which a more
­permanent secondary support system is constructed.

12.6.7  Rock bolts and shotcrete


The support for tunnels can be through rock bolts or by rock bolts and shotcrete lining.
For permanent bolts in civil structures such as railway or road tunnels, the bolts need to be
protected from corrosion (if steel) or they can be made from stainless steel. Carbon steel bolts
in a plastic sheath and grouted can last for more than 100 years without corrosion of the steel.
The bolts serve to tie any rock wedges back to the rock behind the tunnel walls and roof, and
are usually tensioned so as to compress the rock and increase the shear resistance of potential
sliding blocks. Various types of bolts are described in the book by Brady and Brown (2006).
Bolts may be made of fibreglass, stainless steel, or black steel that can be epoxy or resin
coated. The steel bolts are placed inside a plastic sheath and grout is pumped into the annu-
lus between the bolt and the plastic sheath and between the region outside the plastic sheath
and the rock. A section through a grouted rock bolt is shown in Figure 12.16, and a drawing
of a grouted bolt is shown in Figure 12.17.
Shotcrete is applied by spraying the cement, sand, and fine gravel mixture onto the ­surface
of the rock. A steel mesh can be placed first to act as reinforcement for the shotcrete, or
steel or propylene fibres can be incorporated in the shotcrete mixture. Often admixtures
are added, such as air entraining agents, set accelerators or retarders, and water reducers, to
obtain the required properties of the grout.
The shotcrete is sprayed using compressed air, and this may be done wet or dry. In the dry
process, the water is added at the shotcreting nozzle, whereas in the wet process, the grout
already has the water added and is pumped wet. The dry process has the advantage that the
dry mix can be conveyed over longer distances.

12.7  ROCK SLOPES

Slopes are cut in rock for civil works such as roads, railways, and canals or for spillways,
and foundations for dams. Open pit mines also require stable slopes that are cut to access

Figure 12.16 Rock bolt section showing plastic sheathing and surrounding grout.
494  Geomechanics in soil, rock, and environmental engineering

45 mm diameter hole
(nominally)

300 mm max mechanical or


resin grout anchorage
Mechanical anchor

Extra high strength


deformed bar of
24 mm diameter

Water seal

Cement grout
injection
Plastic sheath 36 mm
Embedded length varies

diameter and min


2 mm thick

150 mm diameter
stainless steel plate

Hemispherical dome and grout


injection hole

Steel nut

Isolation washer 50 mm min cover

Figure 12.17 Typical rock bolt used in tunnel support.

the ore body. For mining applications, the slopes need to be as steep as possible so that the
volume of rock removed is not large while still being safe. Often slopes are steeper than they
would be for civil applications, but monitoring allows warning of any impending failures.
The stability of slopes in rock is often dictated by the planes of weakness in the rock.
Modes of failure can include

1. Planar sliding
2. Wedge sliding
3. Block toppling
4. Circular slip, in highly fractured or highly weathered rocks

12.7.1  Planar sliding


If a rock slide can be taken as being a two-dimensional planar slide, then the forces acting
on the wedge of sliding rock are as shown in Figure 12.18. The weight of the wedge of rock
is W and the water forces acting on the wedge along the plane of weakness is U and from
any water in a crack behind the wedge is V.
Basic rock mechanics  495

Tension crack
Slope behind slope
face
V zw z
U
H

Failure surface β
W
α
l
Tension crack in slope face

Slope z
face
H V zw
β α U

Failure surface W
l

Figure 12.18 Planar failure of a rock wedge.

By resolving forces, we can calculate the force normal to the plane of sliding as

N = W cos α − U − V sin α (12.28)

and the force acting parallel to the plane of sliding

T = W sin α + V cos α (12.29)

The factor of safety F can then be calculated from the force acting parallel to the plane of
sliding divided by the force resisting sliding (Equation 12.30)

c′ + N tan φ′ c′ + (W cos α − U − V sin α) tan φ′


F = = (12.30)
T W sin α + V cos α

where the angle of shearing resistance on the plane of sliding is ϕ′. As well, the forces in the
above equation can be calculated from

 = (H − z) / sin α
U = 0.5γ w zw (H − z)/ sin α (12.31)
2
V = 0 .5γ zw w

If the vertical joint is behind the crest of the slope

   z  
2 
W = 0.5γH 2  1 −    cot α − cot β  (12.32)
   H  

496  Geomechanics in soil, rock, and environmental engineering

If the vertical crack is on the slope face

   z  
2 
W = 0.5 γH 2  1 −    cot α(cot α tan β − 1) (12.33)
   H  

where the plane of weakness is at the angle α to the horizontal and the slope face is at an
angle β to the horizontal.

12.7.2  Wedge failure


A wedge of rock may slide out of a slope and if the slide occurs on two planes as shown
in Figure 12.19, then the factor of safety may be calculated through the use of charts.
The angle of shearing resistance on each of the two planes of sliding can be different; in
Figure 12.19, these are called ϕA on plane A and ϕB on plane B. The factor of safety for the
wedge (considering friction only) can then be calculated from Equation 12.34.

F = A tan φ A + B tan φB (12.34)


where the coefficients A and B may be found from charts (Hoek and Bray 1981; Wyllie and
Mah 2004). The values of the dip direction and the dip of each of the planes need to be
known before the charts can be used to find A and B.
The charts are presented for various differences in the dip of the two planes of sliding, and
an example for a dip difference of 0° is given in Figure 12.20. Further charts for different
dip differences are provided in Appendix 12C.

12.8  FOUNDATIONS ON ROCK

Foundations may be constructed on the surface of rock or they may be socketed into the
rock. Drilled or socketed foundations may be used to resist downward or upward forces
or large lateral forces, so it is of interest to be able to make estimates of the ultimate load
that  the sockets can support as well as what the expected deflections of the foundation
may be.

Plane B
Note: The flatter of the two
planes is always called plane A

Factor of safety F
Plane A
F = A · tan ϕA + B · tan ϕB

Figure 12.19 Wedge failure in a rock slope.


Basic rock mechanics  497

A/B chart – dip difference 0°


5.0

4.5

4.0

3.5

3.0
Ratio A or B

2.5 Dip of plane (degrees)

2.0 20

30
1.5
40
50
1.0
60 70
0.5 80

0
0 20 40 60 80 100 120 140 160 180
360 340 320 300 280 260 240 220 200
Difference in dip direction (degrees)

Figure 12.20 Chart for rock wedge stability (dip difference 0°). See Appendix 12C for a complete set of
charts.

12.8.1  Surface foundations


For surface foundations on rock, the allowable vertical bearing pressure can be found from

qba = quKsp (12.35)



where qu is the UCS of the rock.
The coefficient K sp can be found from the plot shown in Figure 12.21 (Canadian
Geotechnical Society 2006), where it may be seen that the coefficient depends on the spac-
ing of discontinuities c and the aperture of the discontinuities δ. The factor K sp contains a
factor of safety of 3 against the lower bound bearing capacity.
Carter and Kulhawy (1988) have suggested a method for obtaining the ultimate capacity
through the use of the Hoek and Brown strength criterion

(
qbult =  s + m s + s ) q
0.5
(12.36)
  u

where m and s are the coefficients in the Hoek–Brown failure criterion for the rock mass
(see Section 12.3).

12.8.2  Shafts in rock


Shafts constructed in rock gain their bearing resistance from end bearing and from friction
on the shaft. In order to get good resistance, the base of the shaft should be clean and free of
debris, and the sides of the shaft should not be smooth or caked with drilling mud.
498  Geomechanics in soil, rock, and environmental engineering

0.5

=0
0.4 δ/c
1
0.00 Ksp =
3 + c/B
. 0 0 2
0
0.3 10 1 + 300 δ/c
Value of Ksp

0.005 c = Spacing of discontinuities


δ = Aperture of discontinuities
0.2 0.010
B = Footing width
0.020
Valid for 0.05 < c/B < 2.0
0 < δ/c < 0.02
0.1

0
0 0.4 0.8 1.2 1.6 2.0
Ratio c/B

Figure 12.21 Bearing pressure coefficient K sp. (Adapted from Canadian Geotechnical Society. 2006. Canadian
Foundation Engineering Manual, 4th ed. 488pp.)

12.8.2.1  Base resistance


The end resistance can be calculated from an empirical relationship such as that given by
Ladanyi and Roy (1971) in Equation 12.37, where the allowable base load qba is calculated
by modifying the surface foundation bearing pressure by a factor λ.

qba = quKspλ (12.37)


where λ = (1 + 0.4Ls /Bs ) ≤ 3.4 is a factor to allow for the depth of the foundation, and L s
is the depth and Bs is the diameter or width of the socket. The coefficient K sp is found from
the plot of Figure 12.21 (for footings), and includes a factor of safety of 3. qu is taken as the
UCS of concrete in the shaft or the rock, whichever is the lower.
Another formula often used is one such as that in Equation 12.38 that gives the ultimate
base load for a socketed pile (Zhang and Einstein 1998).

qbult = Kb (qu )0.5(MPa) (12.38)

The constant Kb is determined empirically for any given rock mass, but for the Sydney
sandstone it is taken as Kb = 4.8 if qu (the UCS) is in MPa.

12.8.2.2  Shaft resistance


The ultimate side friction of shafts (rock sockets) may be expressed in terms of a similar
power law to that for the base bearing capacity, that is,

0.5
qsf q 
= λ u  (12.39)
pa  pa 

Basic rock mechanics  499

where the expression has been normalised by using the atmospheric pressure pa which can
be taken as 100 kPa, and the UCS qu is the lower of the values of the shaft concrete or of
the rock.
Values of λ have been given by several authors, and the value depends on whether the
shaft has been grooved or whether it is smooth. Shafts can have a groove cut in them by a
rotary grooving tool to make the shaft rougher and therefore more resistant to slip.
For conventional drilled shafts with walls that have not been grooved, values of λ have
been given as 1.41 (Rowe and Armitage 1984), 0.63–0.94 (Horvath et al. 1983) while Carter
and Kulhawy indicate that λ = 0.63 is a conservative value that may be used in design.
For artificially grooved socket walls, Rowe and Armitage (1984) suggest using a value of
λ = 1.89 in Equation 12.39.
Williams and Pells (1981) provided a method that took into account the roughness of a
shaft and of the discontinuities in the rock. The ultimate shaft friction is found from

qsf = αβqu (12.40)


where
α is a factor for strength and includes a roughness effect
β is a factor expressing the effect if discontinuities

The factors α and β may be found from the plots of Figures 12.22 and 12.23.
Zhang and Einstein (1998) also give the formula of Equation 12.41 for the ultimate shaft
friction

qsf = Ks (qu )0.5(MPa) (12.41)


where they found for a smooth socket K s = 0.4 and for a rough socket K s = 0.8. For the
Sydney sandstone, the factor K s is often taken as being about 0.3 for an unroughened socket.
This equation is the same as Equation 12.39, without non-dimensionalisation so units must
be specified as MPa.

1.0

0.8 ×
Reduction factor α

0.6 ×

×
0.4
×
×
0.2
× ×
× ×× × ×× ×
0
0.1 1 10 100
Unconfined compressive strength (MPa)

Figure 12.22 Factors α for use in Equation 12.40. (After Williams, A.F. and Pells, P.J.N. 1981. Canadian
Geotechnical Journal, Vol. 18, No. 4, pp. 502–513.)
500  Geomechanics in soil, rock, and environmental engineering

1.0

0.8

0.6

Factor β
0.4

0.2

0
0 0.2 0.4 0.6 0.8 1.0
Emass/Eintact

Figure 12.23 Reduction factor β for rock mass stiffness.

12.8.2.3  Lateral capacity


The ultimate lateral resistance of rock sockets has not been as well researched as vertical
load capacity, but there are some results that may be used as guidance. Zhang et al. (2000)
have suggested that the lateral resistance is a combination of the direct pressure in front of
the pier plus a component at the sides of the pier due to shear. This leads to the formula in
(Equation 12.42) for the ultimate lateral pressure pult

pult = (pL + τ max ) (12.42)


The limit pressure in front of the pile or pier pL can be found from the Hoek–Brown
­criterion for a rock mass (Equation 12.11), where the minor principal stress is given by
the pressure due to the self-weight of the rock, that is, σ3 = γ ′z and z is the depth below the
surface. This is shown in Figure 12.24. Hence, we can write the limit pressure (in terms of
effective stress) as

a
 γ ′z 
pL = σ1′ = γ ′z + qu  m + s (12.43)
 qu 

The parameters m, s, and a can be found for the rock mass as outlined in Section 12.3.2.
The side shear can be calculated as for the vertical side shear (Equation 12.41) where
Zhang et al. suggest that the factor K s be taken as 0.2, that is,

τ max = 0.2(qu )0.5 (12.44)


Carter and Kulhawy (1988) use a similar formula to that of Equation 12.42, but they
calculate pL from cavity expansion theory (as the limit pressure of an expanding cylindrical
cavity). These authors make the assumption that the lateral resistance increases linearly to
a depth of 3 diameters and then remains constant (at greater depth). Hence, they give the
following two equations (Equation 12.45) for the maximum lateral force Hult that can be
applied to a rock-socketed pier.
Basic rock mechanics  501

M
H
Side shear resistance τmax

Enlarged Front normal


cross section resistance pL
D

M
H
γz = σ3′

z
Enlarged pL = σ1′

rock mass element

Figure 12.24 L aterally loaded pier in rock socket.

p L 
H ult =  L + τ maxD L for L ≤ 3D
 6 
(12.45)
p 
H ult =  L + τ max  3D2 + (pL + τ max )(L − 3D)D for L > 3D
 6 

In the above formulae, L is the length of the pier and D is the diameter.

12.8.2.4  Uplift capacity


Shafts in rock may be used as anchors for such things as power transmission cables and
structures subjected to wind load or earthquake. The capacity in uplift is different in some
aspects to the capacity in compression loading, and so different approaches are taken in
design.
In uplift, it is assumed that the base of the pier does not take any load and is able to break
away from the base of the hole. In tension sockets also, there is a negative Poisson’s ratio
effect, with the shaft of the pier contracting under load (i.e. the diameter becomes smaller).
However, this effect is only important for a flexible shaft, and not for rigid shafts. Carter
and Kulhawy (1988) have shown that if (Ec /Er)(D/L)2 > 4 (D and L are the diameter and
length of the pier and Ec and Er are the elastic moduli of the concrete in the pier and the
rock mass, respectively) then the pile can be considered rigid, and the uplift capacity can be
­calculated by reducing the side resistance by 30% from that of a pier in compression loading.
For uplift in a highly fractured rock mass, it is often assumed that the rock can provide
no resistance to the uplift force and that the only resistance is due to the weight of a cone of
the rock that is pulled out by the drilled pier foundation.
502  Geomechanics in soil, rock, and environmental engineering

12.8.2.5  Piles on sandstones and shales of the Sydney region


The design of piles or piers in rock for piles socketed into the shales and sandstone of the
Sydney region is based on experience and has been developed for a specific region. It may
not be applicable to other locations, but similar design charts have been developed for other
cities.
Sockets can be excavated by hand or by drilling and can be constructed dry or under
water. The sockets must have clean walls and bases so that there is good contact between
the concrete of the pier and the rock. Sockets drilled in moist weathered shale or sandstone
can have very smooth sides or crushed rock can be smeared on the walls. The base may also
contain debris, so sockets must be inspected and cleaned before concreting.
Inspection can be carried out by descending into the hole in a cage or using a special
TV camera. Safety rules apply if personnel descend into the hole in NSW, where breathing
­apparatus and safety harness must be used.
If sidewalls are not clean, they can be roughened with a roughening tool. The tool can be
attached to a drilling rig and can cut grooves in the walls of the socket. Appendix 12D con-
tains Tables 12.1 through 12.3 that give the rock class classification and the recommended
design values for end bearing pressure and of shaft adhesion for sockets in shale and in
sandstone (see Pells et al. 1998). It may be seen that the design pressures depend on defects
in the rock as well as rock strength.

12.8.3  Deformation of foundations on rock


The deformation of foundations on rock is often analysed using the theory of elasticity, and
for rock, the rock mass stiffness or modulus. The rock mass stiffness depends upon any
fractures or fissures in the rock as these will influence the deformation behaviour as they
will tend to close upon loading.
Hoek et al. (2002) suggest that the modulus of the rock mass can be estimated from the
formula in Equation 12.46.

 D  qui
Em (GPa) =  1 −  .10([GSI − 10]/ 40) (12.46)
 2  100

where the GSI and damage factor D have been given in Tables 12.3 and 12.4 previously and
qui is the unconfined compressive strength of the intact rock.

12.8.3.1  Vertical deformation


In compression, Carter and Kulhawy (1988) suggest that the compressive deformation yc
can be found (for a rigid shaft in contact with the base) from

Pc
yc = 2
(12.47)
(EbD /1 − ν ) + (π / ζ)(EmL /1 + νm )
b

In the above equation, Pc is the compressive load, L is the socket length, Eb is the modu-
lus of the rock below the base of the socketed pier, Em is the rock mass modulus, νb is the
Poisson’s ratio below the base of the pier and νm is the Poisson’s ratio of the rock mass.
ζ = ln[5(1 − νm)L/D] where D is the diameter or width of the foundation as Em ≈ Eb and
ν ≈ νb the equation simplifies to
Basic rock mechanics  503

Pc (1 − ν2m )/(EmD)
yc = (12.48)
1 + (π /ζ)(L /D)(1 − νm )

In uplift, the same authors suggest that the deflection y u of a rigid shaft in a shear socket
can be obtained by taking Eb = 0 in Equation 20.47 giving

 ζ  1 
yu =    Pu (12.49)
 2π   Gm L 

where Gm is the shear modulus of the rock mass and Pu is the uplift load.

12.8.3.2  Lateral deformation


Lateral deflection u and rotation θ of flexible piers in rock sockets may be estimated by sim-
ple formulae given by Randolph (1981) and used by Carter and Kulhawy (1988) (Equations
12.50 and 12.51).

−1 / 7 −3 / 7
 H  E   M  E 
u = 0.50  *   e*  + 1.08  * 2   e*  (12.50)
 G D  G  G D  G 

−3 / 7 −5 / 7
 H  E   M  E 
θ = 1.08  * 2   e*  + 6.40  * 3   e*  (12.51)
G D  G  G D  G 

In the above equations, H is the horizontal load applied, M is the moment applied to the
head of the pier, D is the diameter of the pier, and

 3ν r  (12.52)
G* = Gr  1 +

 4 

where G r and νr are the shear modulus and Poisson’s ratio of the rock (G r = E r/2(1 + νr))
for  an isotropic rock mass. The effective E e Young’s modulus of the concrete shaft is
given by

(EI)c
Ee = (12.53)
πD4 /64

where I is the second moment of inertia of the pier cross section if the pier is not circular.

12.9  VIBRATION THROUGH ROCK

A common problem encountered when excavating rock is the vibration caused by the equip-
ment used for breaking the rock. Measurements of ground vibrations are generally made in
terms of peak particle velocities (PPV).
504  Geomechanics in soil, rock, and environmental engineering

100

Peak particle velocity PPV (mm/s)


10

0.1
1 10 100
Distance (m)
× Hammer > 1500 kg

Figure 12.25 Peak particle velocities versus distance for >1500 kg hammers in class I/II sandstone. (Adapted
from Hackney G.A. 2002. Proceedings 5th Australian New Zealand Young Geotechnical Professional
Conference, New Zealand.)

The Australian code for blasting, AS 2187.2-1993, provides a recommended limit for PPV
of less than 10 mm/s for residential structures; while the German standard, DIN 4150-3
1999, recommends values of less than 5–20 mm/s depending on the frequency of the vibra-
tion (the higher the frequency, the greater the allowable peak particle velocity). Higher PPV
values are allowed for industrial buildings; AS2187.2-1993 allows up to 25 mm/s and DIN
4150-3 1999 allows between 20 and 50 mm/s at ground level, again depending on frequency.
Measurements for several different types of machinery have been measured by Wiss (1981)
and presented on a plot of PPV to distance using a log–log scale in which the data plotted as
a straight line. However, it is not clear which soil or rock types this data was collected for.
Measurements have been taken in Sydney sandstone by Hackney (2002) for different
kinds of rock breaking equipment. These were classified into 250–500 kg, 500–1000 kg,
100–1500 kg, and >1500 kg rock hammers (used for breaking rock) as well as rotary rock
grinders. The results were also plotted for different classes of sandstone, classes I/II and
classes II/III. Sandstone classes have been discussed previously in Section 12.8.2.5 and also
in Appendix 12D.
Generally the PPV versus distance from the source of vibration plots as a straight line on a
log–log plot as shown in Figure 12.25, but there is a good deal of scatter, so upper and lower
bounds to the data are shown as straight lines on the plot.
From the data collected, the plot of Figure 12.26 may be produced showing how far
from the source of vibration a PPV of 10 mm/s will be likely to occur for different rock
hammer weights. The data is plotted for classes I/II sandstone and classes II/III sandstone
(broken line). The data for rotary rock grinders lies at approximately the same location as
for 250–500 kg hammers.

12.10  NUMERICAL METHODS

Numerical methods can be used to model the rock mass when designing footings, piers, cut-
tings, and tunnels in rock. The models that are used depend on the type of jointing in the
rock. When the jointing is spaced widely or there are just a few well-defined joints, finite ele-
ment programs can be used where the joints are specifically modelled in a continuum that is
used to model the remaining rock mass. Such a model is shown in Figure 12.27 for a tunnel
Basic rock mechanics  505

15

Distance to maintain
<10 mm/s PPV (m)
10

0
250 to 500 500 to 1000 1000 to >1500
1500
Rock hammer size (kg)
Class II or III sandstone Class I or II sandstone

Figure 12.26 Distance to cause <10 mm/s peak particle velocity PPV for various rock hammer sizes in dif-
ferent classes of sandstone.

Total displacement and deformation vectors (scaled up by 100) at end of basement excavation (stage 8)
Total
60

displacement
m
0.0015
0.0014

0.0000
0.0015
0.0013

0.0002
0.0015

0.0004
0.0006
58

0.0008 0.0012
Defects
0.0010 0.0015
0.0011
0.0013
0.0015 0.0007
0.0014
0.0017
0.0019
56

0.0014 0.0006
0.0021
0.0023
0.0013 0.0006
0.0025

0.0013 0.0006
54

0.0004 0.0006
0.0005

0.0005

0.0005
0.0004

0.0005
0.00

End of basement excavation – defects, staged tunnel excavation


0.0006

–20 –18 –16 –14 –12 –10 –8 –6

Figure 12.27 Specific joints modelled for tunnel excavation by Phase2 finite element software.

in a jointed rock mass where the commercial finite element software Phase2 (Rocscience
2014) has been used.
Joint networks can be generated in finite element analyses to simulate jointing patterns,
and software such as Phase2 allows this to be done. Rock bolts and lining systems can also
be included in such numerical analyses.
If the rock has several joint sets and the jointing is closely spaced, discrete element m
­ odelling
can be carried out, where each rock block can be modelled individually. Commercial
­software such as UDEC (2014) may be used to model the rock behaviour.
When the rock is fairly competent, non-linear finite element analysis may be used, and the
rock treated as a continuum. An example of this is shown in Figure 12.28 (Liu et al. 2008)
where the finite element software ABACUS (2014) has been used to model the effects of a
new tunnel being constructed next to an existing tunnel.
506  Geomechanics in soil, rock, and environmental engineering

Figure 12.28 Finite element (ABACUS) analysis of two parallel tunnels in sandstone.

Figure 12.28 shows a cutaway view of the two tunnels and the shotcrete lining and rock
bolts in the roof and sides of the tunnels. Construction of the new tunnel is modelled by
removing elements to simulate excavation, then adding shell elements for the shotcrete and
then adding structural elements to model the rock bolts.

APPENDIX 12A
Table 12A  Classification of individual parameters used in the NGI tunnelling quality index
1 Rock quality designation RQD
A Very poor 0%–25%
B Poor 25%–50%
C Fair 50%–75%
D Good 75%–90%
E Excellent 90%–100%
Notes:
1. Where RQD is reported or measured as ≤10, (including 0)
a nominal value of 10 is used to evaluate Q.
2. RQD intervals of 5, i.e. 100, 95, 90, etc., are sufficiently accurate.

2 Joint set number Jn


A Massive, no or few joints 0.5–1.0
B One joint set 2
C One joint set plus random 3
D Two joint sets 4
E Two joint sets plus random 6
F Three joint sets 9
G Three joint sets plus random 12
H Four or more joint sets, random, 15
heavily jointed, ‘sugar-cube’, etc.
J Crushed rock, earthlike 20
Notes:
1. For intersections use (3.0 × Jn).
2. For portals use (2.0 × Jn).
(Continued )
Basic rock mechanics  507

Table 12A (Continued)  Classification of individual parameters used in the NGI tunnelling quality index
3 Joint roughness number Jr
a.  Rock wall contact and
b.  Rock wall contact before 10 cm shear
A Discontinuous joints 4
B Rough or irregular, undulating 3
C Smooth, undulating 2
D Slickensided, undulating 1.5
E Rough or irregular, planar 1.5
F Smooth, planar 1.0
G Slickensided, planar 0.5
c.  No rock wall contact when sheared
H Zone containing clay minerals thick enough to prevent rock wall contact 1.0
J Sandy, gravelly, or crushed zone thick enough to prevent rock wall contact 1.0
Notes:
1. Add 1.0 if the mean spacing of the relevant joint set is greater than 3 m.
2. Jr = 0.5 can be used for planar slickensided joints having lineations, provided the lineations are orientated for minimum
strength.

4 Joint alteration number Ja ϕr (approx.)


a. Rock wall contact
A Tightly healed, hard, non-softening, impermeable filling, i.e. quartz or epidote 0.75 –
B Unaltered joint walls, surface staining only 1.0 (25–35°)
C Slightly altered joint walls. Non-softening mineral coatings, sandy 2.0 (25–30°)
particles, clay-free disintegrated rock, etc.
D Silty-, or sandy-clay coatings, small clay fraction (non-soft) 3.0 (20–25°)
E Softening or low friction clay mineral coatings, i.e. kaolinite or mica. Also 4.0 (8–16°)
chlorite, talc, gypsum, graphite, etc., and small quantities of swelling clays
b. Rock wall contact before 10 cm shear
F Sandy particles, clay-free disintegrated 4.0 (25–30°)
rock, etc.
G Strongly overconsolidated non-softening clay mineral fillings (continuous, 6.0 (16–24°)
but <5 mm thick)
H Medium or low overconsolidation softening clay, mineral fillings 8.0 (12–16°)
(continuous but <5 mm thickness)
J Swelling-clay fillings, i.e. montmorillonite (continuous, but <5mm thick). 8.0–12.0 (6–12°)
Value of Ja depends on percent of swelling clay-size particles, and access
to water
c. No rock wall contact when sheared
K Zones or bands, of disintegrated or crushed rock and clay (see G, H, 6.0 (6–24°)
L J for clay conditions) 8.0
M 8.0–12.0
N Zones or bands of silty- or sandy clay, small clay fraction (non-softening) 5.0
O Thick, continuous zones or bands of clay (see G, H, J for clay conditions) 10.0–13.0 (6–24°)
P 13.0–20.0
R
Notes:
1. Values of ϕr, the residual friction angle, are intended as an approximate guide to the mineralogical properties of the altera-
tion products, if present.

(Continued )
508  Geomechanics in soil, rock, and environmental engineering

Table 12A (Continued)  Classification of individual parameters used in the NGI tunnelling quality index
Approx. water
5 Joint water reduction factor Jw press. (kgf/cm2)
A Dry excavations or minor inflow, i.e. <5 l/min locally 1.0 <1
B Medium inflow or pressure, occasional outwash of joint fillings 0.66 1–2.5
C Large inflow or high pressure in competent rock with unfilled joints 0.5 2.5–10
D Large inflow or high pressure, considerable outwash of joint fillings 0.33 2.5–10
E Exceptionally high inflow or water pressure at blasting, decaying with time 0.2–0.1 >10
F Exceptionally high inflow or water pressure continuing without decay 0.1–0.05 >10
Notes:
1. Factors C–F are crude estimates. Increase Jw if drainage measures are installed.
2. Special problems caused by ice formation are not considered.

6 Stress reduction factor SRF


a. Weakness zones intersecting excavation, which may cause loosening of rock mass when tunnel is excavated
A Multiple occurrences of weakness zones containing clay or chemically 10.0
disintegrated rock, very loose surrounding rock (any depth)
B Single weakness zones containing clay or chemically disintegrated rock 5.0
(depth of excavation <50 m)
C Single weakness zones containing clay or chemically disintegrated rock 2.5
(depth of excavation >50 m)
D Multiple shear zones in competent rock (clay-free), loose surrounding 7.5
rock (any depth)
E Single shear zones in competent rock (clay-free) (depth of 5.0
excavation < 50 m)
F Single shear zones in competent rock (clay-free) (depth of excavation 2.5
>50 m)
G Loose open joints, heavily jointed or ‘sugar-cube’ (any depth) 5.0
b. Competent rock, rock stress problems
H Low stress, near surface σc/σ1 σt/σ1 2.5
>200 >13
J Medium stress 200–10 13–0.66 1.0
K High stress, very tight structure (usually favourable to stability, may be 10–5 0.66–0.33 0.5–2.0
unfavourable for wall stability)
L Mild rock burst (massive rock) 5–2.5 0.33–0.16 5–10
M Heavy rock burst (massive rock) <2.5 <0.16 10–20
c. Squeezing rock plastic flow of incompetent rock under the influence of high rock pressure
N Mild squeezing rock pressure 5–10
O Heavy squeezing rock pressure 10–20
d. Swelling rock chemical swelling activity depending on the presence of water
P Mild swelling rock pressure 5–10
R Heavy swelling rock pressure 10–20
Notes:
1. Reduce these values of SRF by 25%–50% if the relevant shear zones only influence but do not intersect the excavation.
2. For strongly anisotropic virgin stress field (if measured): when 5 ≤ σ1/σ3 ≤ 10, reduce σc to 0.8σc and σt to 0.8σt.
When σ1/σ3 > 10, reduce σc and σt to 0.6σc and 0.6σt, where, σc = unconfined compression strength, and σt = tensile
strength (point load), and σ1 and σ3 are the major and minor principal stresses.
3. Few case records available where depth of crown below surface is less than span width. Suggest SRF increase from 2.5 to
5 for such cases (see H).
Basic rock mechanics  509

ADDITIONAL NOTES ON THE USE OF THESE TABLES


When a making estimates of the rock quality (Q) the following guidelines should be fol-
lowed, in addition to the notes listed in the tables.

1. When a borehole core is unavailable, RQD can be estimated from the number of joints
per unit volume, in which the number of joints per metre for each joint set are added.
A simple relation can be used to convert this number to RQD for the case of clay-free
rock masses:

RQD  = 115–3.3 Jv (approx.) where Jv = total number of joints per m3 (RQD = 100 for
Jv < 4.5).

2. The parameter Jn representing the number of joint sets will often be affected by folia-
tion, schistosity, slatey cleavage, or bedding. If strongly developed these parallel ‘joints’
should obviously be counted as a complete joint set. However, if there are few ‘joints’
visible, or only occasional breaks in bore core due to these features, then it will be
more appropriate to count them as ‘random joints’ when evaluating Jn.

3. The parameters Jr and Ja (representing shear strength) should be relevant to the weakest
significant joint set or clay filled discontinuity in the given zone. However, if the joint
set or discontinuity with the minimum value of (Jr/Ja) is favourably oriented for stabil-
ity, then a second, less favourably orientated joint set or discontinuity may ­sometimes
be of more significance, and its higher value of (Jr/Ja) should be used when evaluating
Q. The value of (Jr/Ja) should in fact relate to the surface most likely to allow failure to
initiate.

4. When a rock mass contains clay, the factor SRF appropriate to loosening loads
should be evaluated. In such cases, the strength of the intact rock is of little interest.
However, when jointing is minimal and clay is completely absent the strength of the
intact rock may become the weakest link, and the stability will then depend on the
ratio r­ ock-stress/rock strength. A strongly anisotropic stress field is unfavourable for
stability and is roughly accounted for as in Note 2 in the table for stress reduction
factor evaluation.

5. The compressive and tensile strengths (σc and σt) of the intact rock should be evaluated
in the saturated condition if this is appropriate to present or future in situ conditions.
A very conservative estimate of strength should be made for those rocks that deterio-
rate when exposed to moist or saturated conditions.
510  Geomechanics in soil, rock, and environmental engineering

APPENDIX 12B

Table 12B  Rock mass rating system RMR


A. Classification parameters and their ratings
Parameter Range of values
1 Strength of Point-load >10 MPa 4–10 MPa 2–4 MPa 1–2 MPa
For this low
intact rock strength range—Uniaxial
material index compressive test
is preferred
Uniaxial comp. >250 MPa 100– 50–100MPa 25–50 MPa 5–25 1–5 <1
strength 250 MPa MPa MPa MPa
Rating 15 12 7 4 2 1 0
2 Drill core quality RQD 90%–100% 75%–90% 50%–75% 25%–50% <25%
Rating 20 17 13 8 3
3 Spacing of discontinuities >2 m 0.6–2 m 200– 60–200 mm <60 mm
600 mm
Rating 20 15 10 8 5
4 Condition of Very rough Slightly Slightly Slickensided Soft gouge > 5 mm
discontinuities (see E) surfaces rough rough surfaces or thick or
Not surfaces surfaces Gouge < Separation > 5 mm
continuous Separation Separation 5 mm thick Continuous
No < 1 mm < 1 mm or
separation Slightly Highly Separation
Unweathered weathered weathered 1–5 mm
wall rock walls walls Continuous
Rating 30 25 20 10 0
5 Groundwater Inflow per None <10 10–25 25–125 >125
10 m tunnel
length (L/m)
(joint water 0 <0.1 0.1–0.2 0.2–0.5 >0.5
pressure)/
(major
principal σ)
General Completely Damp Wet Dripping Flowing
conditions dry
Rating 15 10 7 4 0

B. Rating adjustments for discontinuity orientations (see F)


Strike and dip orientations Very favourable Favourable Fair Unfavourable Very unfavourable
Ratings Tunnels and mines 0 −2 −5 −10 −12
Foundations 0 −2 −7 −15 −25
Slopes 0 −5 −25 −50

C. Rock mass classes determined from total ratings


Rating 100 ← 81 80 ← 61 60 ← 41 40 ← 21 <21
Class number I II III IV V
Description Very good rock Good rock Fair rock Poor rock Very poor rock

(Continued )
Basic rock mechanics  511

Table 12B (Continued)  Rock mass rating system RMR


D. Meaning of rock classes
Class number I II III IV V
Average stand-up 20 years for 1 year for 1 week for 10 h for 2.5 m 30 min for
time 15 m span 10 m span 5 m span span 1 m span
Cohesion of rock >400 300–400 200–300 100–200 <100
mass (kPa)
Friction angle of >45 35–45 25–35 15–25 <15
rock mass (deg)

E. Guidelines for classification of discontinuity conditions


Discontinuity < 1 m 1–3 m 3–10 m 10–20 m >20 m
length
(persistence)
Rating 6 4 2 1 0
Separation None <0.1 mm 0.1–1.0 mm 1–5 mm >5 mm
(aperture)
Rating 6 5 4 1 0
Roughness Very rough Rough Slightly rough Smooth Slickensided
Rating 6 5 3 1 0
Infilling (gouge) None Hard filling Hard filling Soft filling Soft filling
< 5 mm > 5 mm < 5 mm > 5 mm
Rating 6 4 2 2 0
Weathering Unweathered Slightly Moderately Highly Decomposed
weathered weathered weathered
Rating 6 5 3 1 0

F. Effect of discontinuity strike and dip orientation in tunnellinga


Strike perpendicular to tunnel axis Strike parallel to tunnel axis
Drive with dip – Dip Drive with dip – Dip Dip 45–90° Dip 20–45°
45–90° 20–45°
Very favourable Favourable Very unfavourable Fair
Drive against dip – Dip Drive against dip – Dip Dip 0–20° Irrespective of strikeb
45–90° 20–45°
Fair Unfavourable Fair
Source: Adapted from Bieniawski, Z.T. 1989. Engineering Rock Mass Classifications. Wiley, New York.

a Modified after Wickham et al. (1972).


b Some conditions are mutually exclusive. For example, if infilling is present, the roughness of the surface will
be overshadowed by the influence of the gouge. In such cases, use A4 directly.
512  Geomechanics in soil, rock, and environmental engineering

APPENDIX 12C

Table 12C  Rock wedge stability factors A and B

A chart – Dip difference 10° B chart – Dip difference 10°


5.0 5.0

4.5 4.5

4.0 4.0
Dip of plane A (degrees)
3.5 3.5

3.0 3.0
Dip of plane B (degrees)
Ratio A

Ratio B
2.5 20 2.5

40

20
2.0 2.0

20
30

30

50 40 0
3
1.5 40 1.5

0
6 0

70 6
5
70 0

1.0 1.0

80
90
80

0.5 0.5

0 0
0 20 40 60 80 100 120 140 160 180 0 20 40 60 80 100 120 140 160 180
360 340 320 300 280 260 240 220 200 360 340 320 300 280 260 240 220 200
Difference in dip direction (degrees) Difference in dip direction (degrees)

A chart – Dip difference 20° B chart – Dip difference 20°


5.0 5.0
4.5 4.5
4.0 4.0
3.5 Dip of plane A
3.5
(degrees)
3.0 3.0
Ratio A

20
Ratio B

2.5 2.5
Dip of plane B (degrees)
2.0 30 2.0
1.5 40 1.5 30 0
50 4 50
60 60 70 0
30

1.0
40 50

1.0
8 0
70 9
0.5 0.5
0 0
0 20 40 60 80 100 120 140 160 180 0 20 40 60 80 100 120 140 160 180
360 340 320 300 280 260 240 220 200 360 340 320 300 280 260 240 220 200
Difference in dip direction (degrees) Difference in dip direction (degrees)

(Continued)
Basic rock mechanics  513

Table 12C (Continued)  Rock wedge stability factors A and B

A chart – Dip difference 30° B chart – Dip difference 30°


5.0 5.0

4.5 4.5

4.0 4.0
Dip of plane A
3.5 3.5
(degrees)
3.0 3.0
20

Ratio B
Ratio A

2.5 2.5
30
2.0 2.0
5 0
4

Dip of plane B (degrees)


60 0

1.5 1.5

5040
80 70 60
400
1.0 1.0

5 0
6

90
0.5 0.5

0 0
0 20 40 60 80 100 120 140 160 180 0 20 40 60 80 100 120 140 160 180
360 340 320 300 280 260 240 220 200 360 340 320 300 280 260 240 220 200
Difference in dip direction (degrees) Difference in dip direction (degrees)

A chart – Dip difference 40° B chart – Dip difference 40°


5.0 5.0

4.5 4.5

4.0 Dip of plane A (degrees) 4.0

3.5 3.5

3.0 20 3.0
Ratio A

Ratio B

2.5 2.5
30
40
50

2.0 2.0

1.5 1.5 Dip of plane B (degrees)


90 0 0
50
8 70 6

1.0 1.0
500
6

0.5 0.5

0 0
0 20 40 60 80 100 120 140 160 180 0 20 40 60 80 100 120 140 160 180
360 340 320 300 280 260 240 220 200 360 340 320 300 280 260 240 220 200
Difference in dip direction (degrees) Difference in dip direction (degrees)

(Continued)
514  Geomechanics in soil, rock, and environmental engineering

Table 12C (Continued)  Rock wedge stability factors A and B

A chart – Dip difference 50° B chart – Dip difference 50°


5.0 5.0

4.5 4.5

4.0 4.0
Dip of plane A (degrees)
3.5 3.5

3.0 20 3.0
Ratio A

Ratio B
2.5 2.5
30
2.0
40
2.0 Dip of plane B (degrees)

1.5 1.5 70

60
1.0 1.0

90
0.5 0.5 80

0 0
0 20 40 60 80 100 120 140 160 180 0 20 40 60 80 100 120 140 160 180
360 340 320 300 280 260 240 220 200 360 340 320 300 280 260 240 220 200
Difference in dip direction (degrees) Difference in dip direction (degrees)

A chart – Dip difference 60° B chart – Dip difference 60°


5.0 5.0
4.5 4.5
4.0 Dip of plane A 4.0
(degrees)
3.5 3.5
3.0 20 3.0
Ratio A

Ratio B

2.5 2.5
30

2.0 2.0
1.5 1.5
Dip of plane B (degrees)
1.0 1.0
80

70 0
9
0.5 0.5
0 0
0 20 40 60 80 100 120 140 160 180 0 20 40 60 80 100 120 140 160 180
360 340 320 300 280 260 240 220 200 360 340 320 300 280 260 240 220 200
Difference in dip direction (degrees) Difference in dip direction (degrees)

(Continued)
Basic rock mechanics  515

Table 12C (Continued)  Rock wedge stability factors A and B

A chart – Dip difference 70° B chart – Dip difference 70°


5.0 5.0

4.5 4.5

4.0 4.0
Dip of plane A
3.5 3.5
(degrees)
3.0 20 3.0
Ratio B

Ratio B
2.5 2.5

2.0 2.0

1.5 1.5
Dip of plane B (degrees)
1.0 1.0
80
0.5 0.5 90
0 0
0 20 40 60 80 100 120 140 160 180 0 20 40 60 80 100 120 140 160 180
360 340 320 300 280 260 240 220 200 360 340 320 300 280 260 240 220 200
Difference in dip direction (degrees) Difference in dip direction (degrees)

APPENDIX 12D

Table 12D  Engineering classification of shales and sandstones in the Sydney region – A summary guide
(Adapted from Pells, Douglas, Rodway, Thorne, and McMahon, 1978, as revised by Pells et al. 1998.)
The classification system is based on rock strength, defect spacing and allowable seams as
set out below. All three factors must be satisfied.

TABLE I  Classification for sandstone


Unconfined compressive
Class strength qu (MPa) Defect spacing Allowable seams
I >24 >600 mm <1.5%
II >12 >600 mm <3%
III >7 >200 mm <5%
IV >2 >60 mm <10%
V >1 N.A. N.A.

Classification for shale


Unconfined compressive
Class strength qu (MPa) Defect spacing Allowable seams
I >16 >600 mm <2%
II >7 >200 mm <4%
III >2 >60 mm <8%
IV >1 >20 mm <25%
V >1 N.A. N.A.
516  Geomechanics in soil, rock, and environmental engineering

DEFECT SPACING
The terms relate to spacing of natural fractures in NMLC, NO, and HO diamond drill cores
and have the following definitions:
Defect spacing Terms used to describe defect
mm spacinga
>2000 Very widely spaced
600–2000 Widely spaced
200–600 Moderately spaced
60–200 Closely spaced
20–60 Very closely spaced
<20 Extremely closely spaced
a After ISO/CD 14689. 1995. Geotechnics in Civil Engineering:
Identification and classification of rock ISRM (International
Society for Rock Mechanics). 1978. International Journal of
Rock Mechanics and Mineral Science,Vol. 5, pp. 316–368.

ALLOWABLE SEAMS
Seams include clay, fragmented, highly weathered or similar zones, usually sub-parallel to
the loaded surface. The limits suggested in the tables relate to a defined zone of influence.
For pad footings, the zone of influence is defined as 1.5 times the least footing dimension.
For socketed footings, the zone includes the length of the socket plus a further depth equal
to the width of the footing. For tunnel or excavation assessment purposes, the defects are
assessed over a length of core of similar characteristics.

Table II  Design values for foundations on shale


Ultimate end Serviceability end bearing Ultimate shaft
Class bearinga (MPa) pressureb (MPa) adhesionc (kPa) Typical Efield MPa
I >120 Max. 8 1000 >2000
II 30–120 0.5qu 600–1000 700–2000
Max. 6
III 6–30 0.5qu 350–600 200–1200
Max. 3.5
IV >3 1.0 150 100–500
V >3 0.7 50–100 50–300
a Ultimate values occur at large settlements (>5% of minimum footing dimensions).
b End bearing pressure to cause settlement of <1% of minimum footing dimension.
c Clean socket of roughness category R2 or better.Values may have to be reduced because of smear.

Table III  Design values for vertical loading on sandstone


Ultimate end Serviceability end bearing Ultimate shaft
Class bearinga (MPa) pressureb (MPa) adhesionc (kPa) Typical Efield (MPa)
I >120 12 3000 >2000
II 60–120 0.5qu 1500–3000 900–2000
Max. 12
III 20–40 0.5qu 800–1500 350–1200
Max. 6
IV 4–1 5 0.5qu 250–800 100–700
Max. 3.5
V >3 1.0 150 50–100
a Ultimate values occur at large settlements (>5% of minimum footing dimensions).
b End bearing pressure to cause settlement of <1% of minimum footing dimension.
c Clean socket of roughness category R2 or better.
References

Aas, G., Lacasse, S., Lunne, I., and Hoeg, K. 1986. Use of in situ tests for foundation design in
clay, Proceedings, In Situ 86, American Society of Civil Engineers, ASCE Geotechnical Special
Publication 6, Blacksburg, VA, pp. 1–30.
ABACUS finite element software V6.14. 2014. Dassault Systèmes.
Abbott, T.M. and Corless, B.M. 1993. Use of clay and geomembrane liners as base seals for baux-
ite residue disposal, Geotechnical Management of Waste and Contamination, Eds. Fell, R.,
Phillips, T., and Gerrard, C., Balkema, Rotterdam, pp. 315–322.
Arai, K. and Tagyo, K. 1985. Determination of non-circular slip surface giving the minimum factor of
safety in slope stability analysis, Soils and Foundations, Vol. 25, No. 1, pp. 43–51.
AS 1170.4. 2007. Structural Design Actions. Part 4: Earthquake Actions in Australia. Standards Australia.
AS 1289.3.8.1. 2006. Methods of Testing Soils for Engineering Purposes. Soil Classification Tests –
Dispersion–Determination of the Emerson Class Number of a Soil.
AS 1289.6.2.2. 1998. Methods of Testing Soils for Engineering Purposes. Method 6.2.2: Soil Strength
and Consolidation Tests – Determination of the Shear Strength of a Soil–Direct Shear Test
Using a Shear Box. Standards Australia.
AS 1289.6.3.1. 2004. Methods of Testing Soils for Engineering Purposes. Soil Strength and
Consolidation Tests – Determination of the Penetration Resistance of a Soil – Standard
Penetration Test (SPT). Standards Australia.
AS 1289.6.4.2. 1998. Methods of Testing Soils for Engineering Purposes. Method 6.4.2: Soil Strength
and Consolidation Tests – Determination of Compressive Strength of a Soil–Compressive
Strength of a Saturated Specimen Tested in Undrained Triaxial Compression with Measurement
of Pore Water Pressure. Standards Australia.
AS 1289.6.5.1 1999. (R2013). Methods of Testing Soils for Engineering Purposes. Method 6.5.1 Soil
Strength and Consolidation Tests – Determination of the Static Cone Penetration Resistance
of a Soil–Field Test Using a Mechanical or Electrical Cone or Friction-Cone Penetrometer.
Standards Australia.
AS 1289.6.6.1. 1998. Methods of Testing Soils for Engineering Purposes – Soil Strength and
Consolidation Tests – Determination of the One-Dimensional Consolidation Properties of a Soil –
Standard Method. Standards Australia.
AS 1726. 1993. Geotechnical Site Investigations. Standards Australia.
AS 2159. 1995. Piling–Design and Installation. Standards Australia.
AS 2870. 2011. Residential Slabs and Footings. Standards Australia.
AS 4133.4.1. 2005. Methods of Testing Rocks for Engineering Purposes. Method 4.1 Rock Strength
tests – Determination of Point Load Strength Index. Standards Australia.
AS 4133.4.3.1. 2009. Methods of Testing Rocks for Engineering Purposes – Rock Strength Tests –
Determination of Deformability of Rock Materials in Uniaxial Compression – Strengths of
50 MPa and Greater. Standards Australia.
AS 4133.4.3.2. 2013. Methods of Testing Rocks for Engineering Purposes. Method 4.2 Rock Strength
Tests – Determination of Uniaxial Compressive Strength. Standards Australia.

517
518 References

AS 4133.4.3.2. 2013. Methods of Testing Rocks for Engineering Purposes – Rock Strength Tests –
Determination of the Deformability of Rock Materials in Uniaxial Compression – Rock Strength
Less Than 50 MPa. Standards Australia.
AS 4678. 2002. Earth Retaining Structures. Standards Australia.
Asaoka, A. 1978. Observational procedure of settlement prediction. Soils and Foundations, Vol. 18,
No. 4, pp. 87–101.
ASTM D1143-81. Re-approved 1994. Standard Test Method for Piles under Static Axial Compressive
Load.
ASTM D1586-11. 2011. Standard Test Method for Standard Penetration Test (SPT) and Split-Barrel
Sampling of Soils.
ASTM D2166/D2166M-13. 2013. Standard Test Method for Unconfined Compressive Strength of
Cohesive Soil.
ASTM D2434-68. 2006. Standard Test Method for Permeability of Granular Soils (Constant Head).
ASTM D2435/D2435M–11. 2011. Standard Test Methods for One-Dimensional Consolidation
Properties of Soils Using Incremental Loading.
ASTM D2435-04. 2004. Standard Test Methods for One-Dimensional Consolidation Properties of
Soils Using Incremental Loading.
ASTM D2845-08. 2008. Standard Test Method for Laboratory Determination of Pulse Velocities and
Ultrasonic Elastic Constants of Rock.
ASTM D2850-15. 2015. Standard Test Method for Unconsolidated-Undrained Triaxial Compression
Test on Cohesive Soils.
ASTM D3080/D3080M-11. 2011. Standard Test Method for Direct Shear Test of Soils under
Consolidated Drained Conditions.
ASTM D3966-90. Re-approved 1995. Standard Test Method for Piles under Lateral Loads.
ASTM D4428/D4428M-07. 2007. Standard Test Methods for Crosshole Seismic Testing.
ASTM D4645-08. 2008. Standard Test Method for Determination of In Situ Stress in Rock Using
Hydraulic Fracturing Method.
ASTM D4647/D4647M-13. 2013. Standard Test Methods for Identification and Classification of
Dispersive Clay Soils by the Pinhole Test.
ASTM D4719-07. 2007. Standard Test Methods for Prebored Pressuremeter Testing in Soils.
ASTM D4767-11. 2011. Standard Test Method for Consolidated Undrained Triaxial Compression
Test for Cohesive Soils.
ASTM D5084-10. 2010. Standard Test Methods for Measurement of Hydraulic Conductivity of
Saturated Porous Materials Using a Flexible Wall Permeameter.
ASTM D5578-99. 1999. Standard Test Method for Performing Electronic Friction. Cone and
Piezocone Penetration Testing of Soils.
ASTM D5731-08. 2008. Standard Test Method for Determination of the Point Load Strength Index of
Rock and Application to Rock Strength Classifications. American Society for Testing of Materials.
ASTM D5882-07. 2007. Standard Test Method for Low Strain Impact Integrity Testing of Deep
Foundations.
ASTM D6032-08. 2008. Standard Test Method for Determining Rock Quality Designation (RQD)
of Rock Core. American Society for Testing of Materials.
ASTM D6572-13e1. 2013. Standard Test Methods for Determining Dispersive Characteristics of
Clayey Soils by the Crumb Test. American Society for Testing Materials.
ASTM D6760-02. 2002. Standard Test Method for Integrity Testing of Concrete Deep Foundations
by Ultrasonic Crosshole Testing.
ASTM D7012-14. 2014. Standard Test Methods for Compressive Strength and Elastic Moduli of
Intact Rock Core Specimens under Varying States of Stress and Temperatures. American Society
for Testing of Materials.
ASTM D7181-1. 2011. Method for Consolidated-Drained Triaxial Compression Test for Soils.
ASTM D7400-08. 2008. Standard Test Methods for Downhole Seismic Testing.
Atkinson, J.H. 2007. The Mechanics of Soils and Foundations through Critical State Soil Mechanics,
2nd ed. Taylor & Francis, London.
References 519

Atkinson, J.L. and Bransby, P.L. 1978. The Mechanics of Soils: An Introduction to Critical State Soil
Mechanics. McGraw-Hill, London.
Avalle, D.L. 2004. Improving pavement subgrade with the “square” roller. Proceedings of the 23rd
Southern African Transport Conference (SATC 2004), CSIR International Convention Centre,
Pretoria, South Africa, 12–15 July, pp. 44–54.
Baker, H.W. (ed.). 1985. Issues in dam grouting. Proceedings of Geotechnical Engineering Division,
ASCE Convention, April 30, Denver, Colorado, 174pp.
Balaam, N.P. and Booker, J.R. 1981. Analysis of rigid rafts supported by granular piles, International
Journal of Numerical Methods in Engineering, Vol. 5, pp. 379–403.
Balaam, N.P. and Poulos, H.G. 1978. Methods of analysis of single stone columns, Proceedings
of Symposium on Soil Reinforcing and Stabilising Techniques, NSW Inst. Tech., Sydney,
October 16–19, pp. 497–512.
Baldi, G., Bellotti, R., Ghionna, V., and Jamiolkowski, M. 1981. Cone resistance of a dry medium
sand. Proceedings of the 10th International Conference on Soil Mechanics and Foundation
Engineering, Stockholm, Vol. 2, pp. 427–432.
Banerjee, P.K. 1978. Analysis of axially and laterally loaded pile groups, Chapter 9, Developments in
Soil Mechanics, Ed. Scott, C.R., Applied Science, London, England.
Barton, N.R. and Choubey, V. 1977. The shear strength of rock joints in theory and practice, Rock
Mechanics, Vol. 10, No. 1–2, pp. 1–54.
Barton N., Lien R., and Lunde J. 1974. Engineering classification of rock masses for the design of
tunnel support, Rock Mechanics, Vol. 6, No. 4, pp. 189–236.
Barton, N., Løset, F., Lien, R., and Lunde, J. 1980. Application of the Q-system in design decisions.
Subsurface Space, Ed. Bergman, M., Vol. 2, Pergamon, New York, pp. 553–561.
Bazaraa, A.R.S.S. 1967. Use of the standard penetration test for estimating settlement of shallow
foundations on sand. PhD thesis, University of Illinois.
Beer, E. and Carpentier, R. 1977. Methods to estimate lateral force acting on stabilising piles,
Discussion, Soils and Foundations, Vol. 17, No. 1, pp. 68–82.
Bergado, D.T. and Lam, F.L.L. 1987. Full scale load test of granular piles with different densities and
different proportions of gravel and sand on soft Bangkok clay, Soils and Foundations, Vol. 27,
No. 1, pp. 86–93.
Berry, A., Visser, A., and Rust, E. 2004. A simple method to predict the profile of improvement
after compaction using surface settlement, International Symposium on Ground Improvement,
September 9–10, Laboratoire central des ponts et chaussées, Paris.
Bieniawski, Z.T. 1973. Engineering classification of jointed rock masses, Transactions of the South
African Institution of Civil Engineers, Vol. 15, pp. 335–344.
Bieniawski, Z.T. 1975. The point load test in geotechnical practice, Engineering Geology,
Vol. 9, pp. 1–11.
Bieniawski, Z.T. 1989. Engineering Rock Mass Classifications. Wiley, New York.
Bilotta, E., Caputo, V., and Viggiani, C. 1991. Analysis of soil–structure interaction for piled rafts.
Deformation of soils and displacements of structures, Proceedings of the Tenth European
Conference on Soil Mechanics and Foundation Engineering, Florence, May 1991.
Birmingham, P., Ealy, C.D., and White, J.K. 1994. A comparison of Statnamic and static field
tests at seven FHWA sites, International Conference on Design and Construction of Deep
Foundations, FHWA, Orlando, Vol. 2, pp. 616–630.
Bishop, A.W. 1955. The use of the slip circle in the stability analysis of earth slopes, Géotechnique,
Vol. 5, No. 1, pp. 129–150.
Bishop, A.W. and Morgenstern, N. 1960. Stability coefficients for earth slopes, Géotechnique,
Vol. 10, No. 4, pp. 129–150.
Bjerrum, L. 1963. Discussion on Proceedings of the European Conference on Soil Mechanics and
Foundation Engineering, Norwegian Geotechnical Institute Publication, Oslo, Norway,
Vol. III, No. 98, pp. 1–3.
Bjerrum, L. 1972. Embankments on soft ground, Proceedings of the ASCE Conference on Performance
of Earth-Supported Structures, Purdue University, Lafayette, Indiana, 11–14 June, Vol. 2, pp. 1–54.
520 References

Bjerrum, L. and Eide, O. 1956. Stability of strutted excavations in clay, Géotechnique, Vol. 6,
No. 1, pp. 32–47.
Booker, J.R. 1991. Analytic methods in geomechanics, Computer Methods and Advances in
Geomechanics, Proceedings of the 7th International Conference on Computer Methods and
Advances in Geomechanics, Eds. Beer, G., Booker, J.R., and Carter, J.P., Vol. 1, Balkema,
Cairns, Australia, 6–10 May, pp. 3–14.
Booker, J.R. and Davis, E.H. 1972. A note on a plasticity solution to the stability of slopes in homo-
geneous clays, Géotechnique, Vol. 22, pp. 509–513.
Booker, J.R. and Rowe, R.K. 1987. One dimensional advective–dispersive transport into a deep layer
having a variable surface concentration, International Journal for Numerical and Analytical
Methods in Geomechanics, Vol. 11, No. 2, pp. 131–142.
Booker, J.R. and Small, J.C. 1975. An investigation of the stability of numerical solutions of Biot’s
equations of consolidation, International Journal of Solids and Structures, Vol. 11, pp. 907–917.
Booker, J.R. and Small, J.C. 1983. The analysis of liquid storage tanks on deep elastic foundations,
International Journal for Numerical and Analytical Methods in Geomechanics, Vol. 7, pp. 187–207.
Booker, J.R. and Small, J.C. 1987. A method of computing the consolidation behaviour of lay-
ered Soils using direct numerical inversion of Laplace transforms, International Journal for
Numerical and Analytical Methods in Geomechanics, Vol. 11, pp. 363–380.
Boone, S.J. 1995. Ground-movement-related building damage, Journal of Geotechnical Engineering,
ASCE, Vol. 122, No. 11, pp. 886–896.
Borges, J.L. 2004. Three-dimensional analysis of embankments on soft soils incorporating vertical
drains by finite element method, Computers and Geotechnics, Vol. 31, No. 8, pp. 665–676.
Bouazza, A. and Avalle, D.L. 2006. Verification of the effects of rolling dynamic compaction using a
continuous surface wave system, Australian Geomechanics, Vol. 41, No. 2, pp. 101–108.
Boutwell, G. and Tsai, C.N. 1992. The two stage field permeability test for clay liners, Geotechnical
News, Vol. 10, No. 2, pp. 32–34.
Bradbury, C.E. 1990. Harris dam slurry trench: Design and construction, Australian Geomechanics,
December, No. 19, pp. 22–27.
Brady, B.H.G. and Brown, E.T. 2006. Rock Mechanics for Underground Mining, 3rd ed. (reprinted).
Springer, Dordrecht, 626p.
BRE. 2004. Working Platforms for Tracked Plant: Good Practice Guide to the Design, Installation,
Maintenance and Repair of Ground-Supported Working Platforms, British Research
Establishment, Garston, Watford.
Brett, D.M. 1990. Duncan colliery tailings disposal – Filter dams, Proceedings of the International
Symposium on Safety and Rehabilitation of Tailings Dams, ICOLD, Sydney, May 23, pp. 296–305.
British Standards Institution BS EN ISO 22476-5. 2012. Geotechnical Investigation and Testing –
Field Testing – Part 5: Flexible Dilatometer Test. British Standards Institution.
Britto, A.M. and Gunn, M.J. 1987. Critical State Soil Mechanics via Finite Elements. Ellis Horwood,
Chichester, ISBN 0-470-20816-3, 488pp.
Broms, B.B. 1964a. Lateral resistance of piles in cohesive soils, Journal of the Soil Mechanics and
Foundations Division, ASCE, Vol. 90, SM2, pp. 27–63.
Broms, B.B. 1964b. Lateral resistance of piles in cohesionless soils, Journal of the Soil Mechanics and
Foundations Division, ASCE, Vol. 90, SM3, pp. 123–156.
Brown, P.T. 1969. Numerical analyses of uniformly loaded circular rafts on elastic layers of finite
depth, Géotechnique, Vol. 19, No. 2, pp. 301–306.
Brown, P.T. 1975. Strip footing with concentrated loads on deep elastic foundations, Geotechnical
Engineering, Vol. 6, pp. 1–13.
Brown, P.T. and Booker, J.R. 1985. Finite element analysis of excavation, Computers and Geotechnics,
Vol. 1, No. 3, pp. 207–220.
BS 5930. 1981. Code of Practice for Site Investigations. British Standards Institution.
BS 8006-1. 2010. Code of Practice for Strengthened/Reinforced Soils and Fills. British Standards
Institution.
BS EN 1926. 2006. Natural Stone Test Methods. Determination of Uniaxial Compressive Strength.
British Standards Institution.
References 521

BS EN ISO 22476-1. 2012. Geotechnical Investigation and Testing. Field Testing. Electrical Cone and
Piezocone Penetration Test. British Standards Institution.
BS EN ISO 22476-3. 2005 + A1, 2011. Geotechnical Investigation and Testing. Field Testing. Standard
Penetration Test. British Standards Institution.
BS EN ISO 22476-4. 2012. Geotechnical Investigation and Testing – Field Testing – Part 4: Menard
Pressuremeter Test. British Standards Institution.
Burd, H.J. and Frydman, S. 1997. Bearing capacity of plane-strain footings on layered soils, Canadian
Geotechnical Journal, Vol. 34, No. 2, pp. 241–253.
Burland, J.B., Broms, B.B., and de Mello, V.F.B. 1977. Behaviour of foundations and structures:
State-of-the-Art Review, IX International Conference on Soil Mechanics and Foundation
Engineering, Tokyo, Vol. 2, pp. 495–546.
Bustamante, M. and Gianeselli, L. 1982. Pile bearing capacity prediction by means of static pen-
etrometer CPT, Proceedings of ESOPT II, Amsterdam, Vol. 2, pp. 492–500.
Butterfield, R. and Banerjee, P.K. 1971. The elastic analysis of compressible piles and pile groups,
Géotechnique, Vol. 21, No. 1, pp. 43–60.
Callanan, J.F. and Kulhawy, F.H. 1985. Evaluation of Procedures for Predicting Foundation
Uplift Movements. Report, Electric Power Research Institution, No. EPRI EL-4107, Cornell
University.
Campanella, R.G. 2008. The Third James K. Mitchell Lecture: Geo-environmental site char­
acterization, Geomechanics and Geoengineering: An International Journal, Vol.  3, No.  3,
pp. 155–165.
Campanella, R.G. and Robertson, P.K. 1988. Current status of the piezocone test, Invited Lecture,
First International Conference on Penetration Testing, March 20–24, ISOPT-1, Disney World,
A.A. Balkema, Orlando, Florida, pp. 93–116.
Canadian Geotechnical Society. 2006. Canadian Foundation Engineering Manual, 4th ed., Canadian
Geotechnical Society, Calgary, 488pp.
CANMET. 1977. Pit Slope Manual Chapter 9: Waste Embankments, Canadian Centre for Mineral
and Energy Technology, Report 77-01.
Caquot, A.I. and Kerisel, J. 1948. Tables for the calculation of passive pressure, active pressure, and
bearing capacity of foundations, Libraire du Bureau des Longitudes, de L’école Polytechnique,
Paris Gauthier-Villars, Imprimeur-Editeur, 120.
Carillo, N. 1942. Simple two and three-dimensional cases in the theory of consolidation of soil,
Journal of Mathematical Physics, Vol. 21, No. 1, pp. 1–5.
Carter, J.P. and Kulhawy, F.H. 1988. Analysis and Design of Drilled Shaft Foundations Socketed into
Rock, Report El-5918, Electric Power Research Institute, Palo Alto, CA.
Carter, J.P., Randolph, M.F., and Wroth, C.P. 1979. Stress and pore pressure changes in clay during
and after the expansion of a cylindrical cavity, International Journal of Numerical Methods in
Engineering, Vol. 3, pp. 305–322.
Chapuis, R.P. 1988. Two case histories of major frost heaving in refrigerated buildings:
Thermal analyses, repairs and prevention, Canadian Geotechnical Journal, Vol. 25, No. 3,
pp. 535–540.
Chen, J.J., Zhang, L., Zhang, J.F., Zhu, Y.F., and Wang, J.H. 2013. Field tests, modification, and appli-
cation of deep soil mixing method in soft clay, Journal of Geotechnical and Geoenvironmental
Engineering, Vol. 139, No. 1, pp. 24–34.
Chen, R. and Chameau, J. 1983. Discussion three-dimensional limit equilibrium analysis of slopes,
Géotechnique, Vol. 33, No. 1, pp. 215–216.
Chen, T.-J. and Fang, Y.-S. 2008. Earth pressure due to vibratory compaction, Journal of Geotechnical
and Geoenvironmental Engineering, ASCE, Vol. 134, No. 4, pp. 437–444.
Cheng, Y. and Yip, C. 2007. Three-dimensional asymmetrical slope stability analysis exten-
sion of Bishop’s, Junbu’s, and morgenstern–price’s techniques, Journal of Geotechnical and
Geoenvironmental Engineering, ASCE, Vol. 133, No. 12, pp. 1544–1555.
Cheung, Y.K. 1976. Finite Strip Method in Structural Mechanics. Pergamon Press, Oxford.
Cheung, Y.K., Tham, L.G., and Guo, D.J. 1988. Analysis of pile group by infinite element layer
method, Géotechnique, Vol. 38, No. 3, pp. 415–431.
522 References

Cheung, Y.K. and Zienkiewicz, O.C. 1965. Plates and tanks on elastic foundations – An application
of finite element methods, International Journal of Solids and Structures, Vol. 1, pp. 451–461.
Chin, F.K. 1970. Estimation of the ultimate load of pile from tests not carried to failure, Proceedings
of the 2nd Southeast Asian Conference on Soil Engineering, Singapore, 11–15 June, pp. 81–92.
Chow, Y.K. 1986. Analysis of vertically loaded pile groups, International Journal for Numerical and
Analytical Methods in Geomechanics, Vol. 10, pp. 59–72.
Christoulas, S. and Frank, R. 1991. Deformation parameters for pile settlement. In Proceedings,
10th European Conference on Soil Mechanical and Foundation Engineering, Florence, Vol. 1,
pp. 373–376.
Christian, J.T. and Carrier, W.D. 1978. Janbu, Bjerrum and Kjaernsli’s chart reinterpreted, Canadian
Geotechnical Journal, Vol. 15, pp. 123–128.
CIRIA. 1995. Remedial treatment for contaminated land, In Situ Methods of Remediation, Eds.
Harris, M.R., Herbert, S.M., and Smith, M.A., Construction Industry Research Information
Association, London, Special Publication No. 109, Vol. IX.
Clancy, P. and Randolph, M.F. 1993. An approximate analysis procedure for piled raft founda-
tions, International Journal for Numerical and Analytical Methods in Geomechanics, Vol. 17,
No. 12, pp. 849–869.
Clough, G.W. and O’Rourke, T.D. 1990. Construction induced movements of in situ walls, Design
and Performance of Earth Retaining Walls, ASCE, Geotechnical Special Publication No. 25,
pp. 439–470.
Collin, J.G. 2004. Column supported embankment design considerations, Proceedings of the 52nd
Annual Geotechnical Conference, University of Minnesota, Mineapolis, MN, February 27, 2004.
Collin, J.G., Watson, C.H., and Han, J. 2005. Column-supported embankment solves time constraint
for new road construction, ASCE Geo-Frontiers Conference, Austin, Texas, January 2005.
Corless, B.M. and Glenister, D.J. 1990. The rehabilitation challenges of bauxite residue disposal
operations in Western Australia, Proceedings of the International Symposium on Safety and
Rehabilitation of Tailings Dams, ICOLD, Sydney, May 23, pp. 326–339.
Coulomb, C.A. 1776. Essai sur une application des règles des maximis et minimis à quelques prob-
lèmes de statique relatifs à l’architecture, Mémoires de Mathématiques et de Physique Présentés
à l’Académie Royale des Sciences par Divers Savants, et lus dans ses Assemblées, Paris, Vol. 7,
pp. 343–382.
CRISP. 2013. SAGE CRISP, The CRISP Consortium Limited, http://www.sagecrisp.net.
Cundall, P. 1976. Explicit finite-difference methods in geomechanics, Second International
Conference on Numerical Methods in Geomechanics, Ed. Desai, C.S., June 1976, Virginia
Polytechnic Institute and State University, Blacksburg, Virginia, USA. Vol. 1, pp. 132–150.
D’Andrea, R.A. and Sangrey, D.A. 1982. Safety factors for probabilistic slope design, Journal of the
Geotechnical Engineering Division, ASCE, Vol. 108, GT9, pp. 1101–1118.
Daniel, D.E. 1989. In situ hydraulic conductivity tests for compacted clay, Journal of Geotechnical
Engineering, ASCE, Vol. 115, No. 9, pp. 1205–1226.
D’Appolonia, D.J., D’Appolonia, E., and Brissette, R.F. 1970. Settlement of spread footings on
sand, Journal of the Soil Mechanics and Foundations Division, ASCE, Vol. 96, No. SM2,
pp. 754–762.
Davis, E.H. 1980. Some plasticity solutions relevant to the bearing capacity of rock and fissured clay,
J.C. Jaeger Memorial Lecture, Proceedings of the Third Australia–New Zealand Conference
on Geomechanics, Wellington, New Zealand, May 12–16, Vol. 3, pp. 27–36.
Davis, E.H. and Booker, J.R. 1971. The bearing capacity of strip footings from the point of view of plas-
ticity theory, Proceedings of the First Australia–New Zealand Conference on Geomechanics,
Melbourne, Vol. 1, pp. 276–282.
Davis, E.H. and Booker, J.R. 1973. The effect of increasing strength with depth on the bearing capac-
ity of clays, Géotechnique, Vol. 23, No. 4, pp. 551–563.
Davis, E.H. and Poulos, H.G. 1968. The use of elastic theory for settlement prediction under three-
dimensional conditions, Géotechnique, Vol. 18, No. 1, pp. 67–91.
Davis, E.H. and Poulos, H.G. 1972. Rate of settlement under three-dimensional conditions,
Géotechnique, Vol. 22, No. 1, pp. 95–114.
References 523

de Sanctis, L. and Mandolini, A. 2006. Bearing capacity of piled rafts on soft clay soils, Journal
of Geotechnical and Geoenvironmental Engineering Division, ASCE, Vol. 132, No. 12,
pp. 1600–1610.
Décourt, L. 1994. Prediction of Settlement of Spread Footings, Eds. Briaud, J.-L. and Gibbons,
R.M., ASCE, New York, ASCE Geotechnical Special Publication No. 41, pp. 210–213.
Décourt, L. 1995. Prediction of load–settlement relationships for foundations on the basis of
the ­SPT-T. Ciclo de Conferencias Internacionales “Leonardo Zeevaert”, UNAM, Mexico,
November 1995, pp. 85–104.
Décourt, L., Belicanta, A., and Quaresma Filho, A.R. 1989. Brazilian experience on SPT, Proceedings
12th International Conference on Soil Mechanics and Foundation Engineering, Rio de Janeiro,
Supplement, Contributions by Brazilian Society for Soil Mechanics, pp. 49–54.
Deere, D.U. and Deere, D.W. 1988. The rock quality designation (RQD) index in practice. Rock
Classification Systems for Engineering Purposes, Ed. Kirkaldie, L., ASTM Special Publication
No. 984, American Society for Testing of Materials, Philadelphia, pp. 91–101.
Deere, D.U., Hendron, A.J., Patton, F.D., and Cording, E.J. 1967. Design of surface and near surface
construction in rock. Failure and Breakage of Rock, Proceedings of the 8th U.S. Symposium on
Rock Mechanics, Ed. Fairhurst, C., Soc. Min. Engrs, Am. Inst. Min. Metall. Petroleum Engrs,
New York, pp. 237–302.
Deere, D.U. and Miller, R.P. 1966. Engineering classification and index properties of rock. Technical
Report No. AFNL-TR-65–116. Air Force Weapons Laboratory, Albuquerque, NM.
Desai, C.S., Johnson, L.D., and Hargett, C.M. 1974. Analysis of pile supported gravity lock, Journal
of the Geotechnical Engineering Division, ASCE, Vol. 100, No. GT9, pp. 1009–1029.
DIN 4150. 3 February 1999. 02 Structural Vibration. Part 3: Effects of Vibration on Structures.
DIPS. 2014. Stereonet plotting software Rocscience, https://www.rocscience.com.
Douglas, B.J. and Olsen, R.S. 1981. Soil classification using electric cone penetrometer, Symposium
on Cone Penetration Testing and Experience, Geotechnical Engineering Division, ASCE,
St. Louis, October 1981, pp. 209–227.
Dunbavan, M. and Carter, J.P. 1994. Response of a composite stone column-clay foundation system,
Proceedings of the Seminar on Ground Improvement Techniques – Featuring the Nelson Point
Ore Handling Facility at Point Hedland, May 19, Ed. Jewell, R.J., Curtin Printing Services,
Perth, Australia, pp. 113–126.
Duncan, J.M. and Buchignani, A.L. 1976. An Engineering Manual for Settlement Studies.
Department of Civil Engineering, University of California, Berkeley.
Duncan, J.M. and Seed, R.B. 1986. Compaction-induced earth pressures under K0-conditions,
Journal of Geotechnical Engineering, ASCE, Vol. 112, No. 1, pp. 1–22.
Dunnicliffe, J. 1993. Geotechnical Instrumentation for Monitoring Field Performance. John Wiley
& Sons, New York.
Eurenius, J. 1990. Long-term studies and design of tailings dams, Proceedings of the International
Symposium on Safety and Rehabilitation of Tailings Dams, ICOLD, Sydney, May 23, pp. 306–315.
Eurocode 7. 2004. Geotechnical Design – Part 1: General Rules, January.
Federation of Piling Specialists. 2006. Handbook on Pile Load Testing.
Fell, R., MacGregor, J.P., Williams, J., and Searle, P. 1987. Hue Hue road landslide, Wyong, Soil Slope
Instability and Stabilisation, Eds. Walker, B. and Fell, R., Balkema, Rotterdam, pp. 315–324.
Fellenius, W. 1927. Erdstatische Berechnungen mit Reibung und Kohaesion, Ernst, Berlin.
Fick, A. 1855. Über diffusion, Poggendorff’s Annalen der Physik und Chemie, Vol. 94, pp. 59–86,
and (in English) Philosophical Magazine S.4, Vol. 10, pp. 30–39.
Fleming, W.G.K., Weltman, A.J., Randolph, M.F., and Elson, W.K. 1992. Piling Engineering, 2nd
ed. Halsted Press, New York.
Focht, J.A., Khan, F.R., and Gemeinhardt, J.P. 1978. Performance of one shell plaza deep mat
foundation, Journal of the Geotechnical Engineering Division, ASCE, Vol. 104, No. GT5,
pp. 593–608.
Forrest, M.P., Connell, A.C., and Scheuring, J. 1990. Reclamation planning of a tailings impound-
ment, International Symposium on Safety and Rehabilitation of Tailings Dams, ICOLD,
Sydney, May 23, Vol. 1, pp. 316–325.
524 References

Francescon, M. 1983. Model pile tests in clay: Stresses and displacements due to installation and axial
loading. PhD Thesis, Cambridge University.
Frank, R. and Magnan, J.-P. 1995. Cone penetration testing in France: National report, Proceedings,
CPT ‘95, Swedish Geotechnical Society, Linkoping, October 4–5, Vol. 3, pp. 147–156.
Franke, E. 1991. Measurements beneath piled rafts, Keynote Lecture ENPC Conference, Paris. pp. 1–28.
Franke, E., Lutz, B., and El-Mossallamy, Y. 1994. Measurements and numerical modelling of high-rise
building foundations on Frankfurt clay, Vertical and Horizontal Deformation of Foundations
and Embankments, ASCE, Geotechnical Special Publication No. 40, Vol. 2, pp. 1325–1336.
Fraser, R.A. and Wardle, L.J. 1976. Numerical analysis of rectangular rafts on layered foundations,
Géotechnique, Vol. 26, No. 4, pp. 613–630.
Frechette, R.J. 1994. Construction of access ramps and reclamation covers upon tailings impound-
ment surfaces, Tailings and Mine Waste ‘94, A.A. Balkema, Rotterdam, pp. 195–204.
Gemperline, M.C. 1988. Centrifuge modelling of shallow foundations, Proceedings of the ASCE
Spring Convention, ASCE Press, Reston, VA, pp. 45–70.
GEOrient. 2014. V9.x http://www.holcombe.net.au.
Geotechnical Engineering Office, Geoguide 1. 2000. Guide to Retaining Wall Design, Civil Engineering
Department, The Government of Hong Kong Special Administrative Region.
Geotechnical Manual for Slopes. 2011. 5th reprint, Geotechnical Engineering Office, Civil Engineering
and Development Department, The Government of the Hong Kong Special Administrative
Region, Hong Kong.
Gibbs, H.J. and Holz, W.G. 1957. Research on determining the density of sands by spoon penetration
testing, IV International Conference on Soil Mechanics and Foundation Engineering, London,
Vol. 1, pp. 35–39.
Gibson, R.E. and Anderson, W.F. 1961. In situ measurement of soil properties with the pressureme-
ter, Civil Engineering and Public Works Review, Vol. 56, No. 658, pp. 615–618.
Goodman, R.E. 1989. Introduction to Rock Mechanics, 2nd ed. Wiley, New York.
Grant, R., Christian, J.T., and Vanmarcke, E.H. 1974. Differential settlement of buildings, Journal of
the Geotechnical Engineering Division, ASCE, Vol. 100, No. GT9, pp. 973–991.
Greenwood, D.A. 1970. Mechanical improvement of soils below ground surface, Proceedings of
Ground Engineering Conference, London, ICE, June, Paper 2, pp. 9–20.
Griffiths, D.V. and Lane, P.A. 1999. Slope stability analysis by finite elements, Géotechnique, Vol. 49,
No. 3, pp. 387–403.
Griffiths, D.V. and Marquez, R.M. 2007. Three-dimensional slope stability analysis by elasto-plastic
finite elements, Géotechnique, Vol. 57, No. 6, pp. 537–546.
Grimstad, E. and Barton, N. 1993. Updating of the Q-System for NMT, Proceedings of the
International Symposium on Sprayed Concrete – Modern Use of Wet Mix Sprayed Concrete for
Underground Support, Fagernes, Eds. Kompen, R., Opsahl, O.A., and Berg, K.R., Norwegian
Concrete Association, Oslo, pp. 46–66.
Guido, V.A., Knueppel, J.D., and Sweeny, M.A. 1987. Plate loading tests on geogrid-reinforced earth
slabs. Proceedings of the Geosynthetic ‘87 Conference, New Orleans, USA, pp. 216–225.
Hackney G.A. 2002. Excavation induced vibrations in Sydney sandstones, Proceedings 5th Australian
New Zealand Young Geotechnical Professional Conference, Ed. Davies, T., Rotorua, Australian
Geomechanics Society, New Zealand, March 13.
Hain, S. and Lee, I.K. 1978. The analysis of flexible raft-pile systems, Géotechnique, Vol. 28, No. 1,
pp. 65–83.
Hanna, A.M. and Meyerhof, G.G. 1980. Design charts for ultimate bearing capacity of foundations
on sand overlying soft clay, Canadian Geotechnical Journal, Vol. 17, pp. 300–303.
Hansbo, S. and Källström, R. 1983. A case study of two alternative foundation principles, Väg-och
Vattenbyggaren, Vol. 7–8, pp. 23–27.
Hansen, J.B. 1970. A revised and extended formula for bearing capacity, Danish Geotechnical
Institute Bulletin, No. 28, pp. 5–11.
Hatanaka, M. and Uchida, A. 1996. Empirical correlation between penetration resistance and inter-
nal angle of friction of sand, Soils and Foundations, Vol. 36, No. 4, pp. 1–10.
References 525

Hayhoe, H.N. and Balchin, D. 1990. Field frost heave measurement and prediction during periods of
seasonal frost, Canadian Geotechnical Journal, Vol. 27, pp. 393–397.
Hemsley, J.A. (ed.). 2000. Design Applications of Raft Foundations. Thomas Telford, London.
Hill, R. 1983. The Mathematical Theory of Plasticity. Oxford Science Publications, Oxford.
Hirayama, H. 1991. Pile group settlement interaction considering soil nonlinearity. Computer
Methods and Advances in Geomechanics, Eds. Beer, G., Booker, J.R., and Carter, J.P.,
A.A. Balkema, Rotterdam, Vol. 1, pp. 139–144.
Hird, C.C., Pyrah, I.C., Russell, D., and Cinicioglu, F. 1995. Modelling the effect of vertical
drains in two-dimensional finite element analyses of embankments on soft ground, Canadian
Geotechnical Journal, Vol. 32, No. 5, pp. 795–807.
Hirschfeld, R.C. and Poulos, S.J. 1972. Embankment-Dam Engineering, Casagrande Volume, John
Wiley & Sons, New York.
Hoek, E. and Bray, J.W. 1981. Rock Slope Engineering, 3rd ed. The Institution of Mining and
Metallurgy, London, UK.
Hoek, E. and Brown E.T. 1980a. Underground Excavations in Rock. Institution of Mining and
Metallurgy, London.
Hoek, E. and Brown, E.T. 1980b. Empirical strength criterion for rock masses, Journal of the
Geotechnical Engineering Division, ASCE, Vol. 106, No. GT9, pp. 1013–1035.
Hoek, E. and Brown, E.T. 1997. Practical estimates of rock mass strength. International Journal
of Rock Mechanics and Mining Sciences & Geomechanics Abstracts, Vol. 34, No. 8,
pp. 1165–1186.
Hoek, E. and Franklin, J.A. 1968. A simple triaxial cell for field and laboratory testing of rock,
Transactions of the Institution of Mining and Metallurgy, Vol. 77, pp. A22–A26.
Hoek, E., Carranza-Torres, C., and Corkum, B. 2002. Hoek-Brown failure criterion – 2002 edition,
Mining and tunnelling innovation and opportunity, Proceedings of the 5th North American
Rock Mechanics Symposium and 17th Tunnelling Association of Canada Conference, Eds.
Hammah, R., Bawden, W., Curran, J., and Telesnicki, M., University of Toronto, Toronto,
July 7–10, Vol. 1, pp. 267–273.
Hooper, J.A. 1973. Observations on the behaviour of a piled-raft foundation on London Clay,
Proceedings of The Institution of Civil Engineers, Vol. 5, No. 2, pp. 855–877 (Discussion 1974
Vol. 57, No. 2, pp. 547–552).
Horikoshi, K. and Randolph, M.F. 1997. Optimum design of piled raft foundations, Proceedings of
the 14th International Conference on Soil Mechanics and Foundation Engineering, Hamburg,
Vol. 2, pp. 1073–1076.
Houlsby, A.C. 1990. Construction and Design of Cement Grouting: A Guide to Grouting in Rock
Foundations, John Wiley & Sons, New York.
Hsi, J.P. and Small, J.C. 1992. Analysis of excavation in an elasto plastic soil involving drawdown of
the water table, Computers and Geotechnics, Vol. 13, No. 1, pp. 1–19.
Hsi, J.P. and Small, J.C. 1993. Application of a fully coupled method to the analysis of an excavation,
Soils and Foundations, Vol. 33, No. 4, pp. 36–48.
Hubbert, M. K. and Willis, D.G. 1957. Mechanics of hydraulic fracturing, Transactions of AIME,
Vol. 210, pp. 153–166.
Hughes, J.M.O. and Withers, N.J. 1974. Reinforcing of soft cohesionless soils with stone columns,
Ground Engineering, Vol. 7, No. 3, pp. 42–49.
Hughes, J.M.O., Withers, N.J., and Greenwood, D.A. 1975. A field trial of reinforcing effects of stone
columns in soil, Géotechnique, Vol. 25, No. 1, pp. 31–44.
ICOLD Bulletin 51. 1985. Filling Materials for Watertight Cut-Off Walls. Commission Internationale
des Grands Barrages, Paris.
Imai, T. and Tonouchi, K. 1982. Correlation of N value with S-wave velocity and shear modu-
lus, Proceedings, European Symposium on Penetration Testing, ESOPT-II, A.A. Balkema,
Amsterdam, Vol. 1, pp. 67–72.
Ingold, T.S. 1979. The effects of compaction on retaining walls, Géotechnique, Vol. 29, No. 3,
pp. 265–283.
526 References

Iskander, M., Omidvar, M., and Elsherif, O. 2013. Conjugate stress approach for Rankine seismic
active earth pressure in cohesionless soils, Journal of Geotechnical and Geoenvironmental
Engineering, ASCE, Vol. 139, No. 7, pp. 1205–1210.
ISRM (International Society for Rock Mechanics). 1978. ISRM Commission on Standardization of
Laboratory and Field Tests. Suggested methods for the quantitative description of disconti-
nuities in rock masses. International Journal of Rock Mechanics and Mineral Science, Vol. 5,
pp. 316–368.
ISO/CD 14689. 1995. Geotechnics in Civil Engineering: Identification and classification of rock.
ISO/TS 17892-5. 2004. Ed1. Geotechnical Investigation and Testing – Laboratory Testing of Soil –
Part 5: Incremental Loading Oedometer Test.
ISO/TS 17892–9. 2004. Ed1. Geotechnical Investigation and Testing – Laboratory Testing of Soil –
Part 9: Consolidated Triaxial Compression Test on Water Saturated Soil.
ITASCA Consulting Group, Inc. 2014. FLAC – Fast Lagrangian Analysis of Continua. Minneapolis,
MN. http://itascacg.com/software/flac3d.
Ito,T. and Matsui, T. 1975. Methods to estimate lateral force acting on stabilising piles, Soils and
Foundations, Vol. 15, No. 4, pp. 43–59.
Ito,T., Matsui, T., and Hong, W.P. 1982. Extended design method for multirow stabilising piles
against landslide, Soils and Foundations, Vol. 22, No. 1, pp. 1–13.
Jamiolkowski, M. and Garassino, A. 1977. Soil modulus for laterally loaded piles, Proceedings of
the 9th International Conference on Soil Mechanics and Foundation Engineering, Tokyo,
Specialty session 10, July 1977, pp. 43–58.
Jamiolkowski, M., Ghionna, V., Lancellotta, R., and Pasqualini, E. 1988. New correlations of
penetration tests for design practice. Proceedings ISOPT-I, Vol. 1, pp. 263–296.
Jamiolkowski, M., Ladd, C.C., Germaine, J.T., and Lancellotta, R. 1985. New developments in
field and laboratory testing of soils, Theme Lecture, 11th International Conference on Soil
Mechanics and Foundation Engineering, San Francisco, Vol. 1, pp. 57–153.
Janbu, N. 1973. Slope stability computations, Slope Stability Computations in Embankment Dam
Engineering, Eds. Hirschfield, R.C. and Poulos, S.J., John Wiley, New York.
Katzenbach, R., Arslan, U., Gutwald, J., and Holzhuser, J. 1997. Soil–structure-interaction of
the  300 m high Commerzbank tower in Frankfurt am Main. Measurements and numerical
­studies, Proceedings of the 14th International Conference on Soil Mechanics and Foundation
Engineering, Hamburg, Vol. 2, pp. 1081–1084.
Katzenbach, R., Arsalan, U., and Holzhäuser, J. 1995. Geotechnische Meßüberwachung des 300 m
hohen Commerzbank-Hochhauses in Frankfurt am Main, Pfahl-Symposium ‘95, Mitteilung
des Instsituts für Grundbau und Bodenmechanik, TU Braunschweig, Vol. 48, pp. 233–244.
Katzenbach, R., Arslan, U., and Moorman, C. 2000. Piled raft foundation projects in Germany,
Chapter 13, Design Applications of Raft Foundations, Ed. Hemsley, J.A., Thomas Telford, London.
Katzenbach, R. and Reul, O. 1997. Design and performance of piled rafts, Proceedings of the 14th
International Conference on Soil Mechanics and Foundation Engineering, Hamburg, Vol. 4,
pp. 2253–2256.
Kay, J.N. 1976. Safety factor evaluation for single piles in sand, Journal of the Geotechnical
Engineering Division ASCE, Vol. 102, No. GT10, pp. 1093–1108.
Kelly, R., Small, J.C., and Wong, P. 2008. Construction of an embankment using vacuum con-
solidation and surcharge fill, Proceedings of Selected Sessions of GeoCongress 2008:
Geosustainability and Geohazard Mitigation, Annual Congress of the Geo-Institute of
ASCE, New Orleans, Louisiana, USA, March 9–12, 2008, Geotechnical Special Publication
No. 178, pp. 578–585.
Kishida, H. and Nakai, S. 1977. Large deflection of a single pile under horizontal load, Proceedings
of the 9th International Conference on Soil Mechanics and Foundation Engineering, Tokyo,
Specialty Session 10, July 1977, pp. 87–92.
Kitiyodom, P, Matsumoto, T., and Kanefusa, N. 2004. Influence of reaction piles on the behaviour of
a test pile in static load testing, Canadian Geotechnical Journal, Vol. 41, No. 3, pp. 408–420.
Konrad, J.M and Morgenstern, N.R. 1982. Prediction of frost heave in the laboratory during tran-
sient freezing, Canadian Geotechnical Journal, Vol. 19, No. 3, pp. 250–259.
References 527

Kulhawy, F.H. and Mayne, P.W. 1990. Manual on Estimating Soil Properties for Foundation Design.
Report EPRI-EL 6800, Electric Power Research Institute, Palo Alto, 306 p.
Kumar J. and Rao S. 1997. Passive pressure coefficients, critical failure surface and its kinematics
admissibility, Géotechnique, Vol. 47, No. 1, pp. 185–192.
Kuwabara, F. 1989. An elastic analysis for piled raft foundations in a homogeneous soil, Soils and
Foundations, Vol. 29, No. 1, pp. 82–92.
Ladanyi, B. and Roy, A. 1971. Some aspects of bearing capacity of rock mass, Proceedings of the
Seventh Canadian Symposium on Rock Mechanics, Edmonton, Alberta, March 1971, pp.
161–190.
Ladd, C.C. 1991. Stability evaluation during staged construction, (The 22nd Karl Terzaghi lecture),
Journal of Geotechnical Engineering, ASCE, Vol. 117, No. 4, pp. 540–615.
Ladd, C.C. and Foott, R. 1974. New design procedure for stability of soft clays, Journal of the
Geotechnical Engineering Division, ASCE, Vol. 100, No. GT7, pp. 763–786.
Ladd, C.C., Foote, R., Ishihara, K., Schlosser, F., and Poulos, H.G. 1977. Stress-deformation and
strength characteristics, Proceedings of the 9th International Conference on Soil Mechanics
and Foundation Engineering, Tokyo, Specialty Session 10, July 1977, Vol. 2, pp. 421–494.
Lancellotta, R. 2002. Analytical solution of passive earth pressure. Géotechnique, Vol. 52, No. 8,
pp. 617–619.
Lav, M.A., Carter, J.P., and Booker, J.R. 1995. The effect of fissures on the bearing capacity of clay,
Proceedings of the 14th Australasian Conference on the Mechanics of Structures and Materials,
Eds. Beasley, A.J., Foster, C.G., and Melerski, E.S., December 11–13, Hobart, pp. 38–43.
Leshchinsky, D. 2015. Bearing capacity of footings placed adjacent to c′–ϕ′ slopes, Journal of Geotechnical
and Geoenvironmental Engineering, ASCE, Vol. 141, No. 6, pp. 1, 13, paper: 04015022.
Leshchinsky, D., Baker, R., and Silver, M.L. 1985. Three-dimensional analysis of slope stability,
International Journal for Numerical and Analytical Methods in Geomechanics, Vol. 9, No. 3,
pp. 199–223.
Li, J., Cameron, D.A., and Mills, K.G. 1996. Numerical modelling of covers and slabs subject to
seasonal surface suction changes, 7th Australia–New Zealand Conference on Geomechanics,
Adelaide, July 1–5, 1996.
Lin, D.G., Bergado, D.T., and Balasubramanium, A.S. 1999. Soil–structure interaction of piled raft foun-
dation in Bangkok subsoil, Eleventh Asian Regional Conference on Soil Mechanics and Founda­tions
Engineering, Ed. Hong, Sung-Wan, August 16–20, A.A. Balkema, Seoul, Korea, pp. 183–187.
Liu, H.Y. Small, J.C., and Carter, J.P. 2008. Full 3D modelling for effects of tunnelling on existing
support systems in the Sydney region, Tunnelling and Underground Space Technology, Vol. 23,
No. 4, pp. 399–420.
Luo, Z., Chen, Y., and Liu, Y. 2000. Monitoring the dynamic characteristics of tall buildings by GPS
technique, Geospatial Information Science, Taylor & Francis, Vol. 3, No. 4, pp. 61–66.
Madhyannapu, R.S., Puppala, A.J., Nazarian, P.E., and Yuan, D. 2010. Quality assessment and
quality control of deep soil mixing construction for stabilizing expansive subsoils, Journal of
Geotechnical and Geoenvironmental Engineering, ASCE, Vol. 136, No. 1, pp. 119–128.
Mana, A.I. and Clough, G.W. 1981. Prediction of movements for braced cuts in clay, Journal of the
Geotechnical Engineering Division, ASCE, Vol. 107, No. GT6, pp. 759–777.
Mandel, J and Salençon, J. 1969. Force portante d’un sol sur une assise rigide, Proceedings of the 7th
International Conference on Soil Mechanics and Foundations Engineering, Vol. II, pp. 157–164.
Marchetti, S. 1997. The flat dilatometer: Design applications, Proceedings of the Third International
Geotechnical Engineering Conference, Keynote lecture, Cairo University, Cairo, January 5–8,
421–448.
Marchetti, S. 1980. In situ tests by flat dilatometer, Journal of the Geotechnical Engineering Division,
ASCE, Vol. 106, No. GT3, pp. 299–321.
Marchetti, S. and Crapps, D.K. 1981. Flat Dilatometer Manual. GPE Inc., Gainesville, FL, USA.
Martin, C.M. 2004. Program ABC (Analysis of Bearing Capacity) V1.0. http://www.eng.ox.ac.uk/
civil/people/cmm/software.
Martin, C.M. 2005. Exact bearing capacity calculations using the method of characteristics,
Proceedings of the 11th International Conference of IACMAG, Turin, Vol. 4, pp. 441–450.
528 References

Matar, M. and Salençon, J. 1977. Capacité portante d’une semelle filante sur sol purement cohérent
d’épaisseur limitée et de cohésion variable avec la profondeur, Annales de l’Institute Technique
du Batiment et des Travaux Publics, No. 352, Serie: Sols et Fondations, Vol. 143, Juillet-Août,
pp. 95–107.
Matlock, H. 1970. Correlations for the design of laterally loaded piles in soft clay, Proceedings of the
Second Offshore Technology Conference, Houston, Texas, Vol. 1, pp. 577–594.
Mayne, P.W. and Rix, G.J. 1993. G max -qc relationships for clays. Geotechnical Testing Journal,
ASTM, West Conshohocken, PA, Vol. 16, No. 1, pp. 54–60.
Mayne, P.W. Jones, J.S., and Dumas, J.C. 1984. Ground response to dynamic compaction, Journal of
Geotechnical Engineering, ASCE, Vol. 110, No. 6, pp. 757–774.
Mazindrani, Z.H. and Ganji, M.H. 1997. Lateral earth pressure problem of cohesive backfill with
inclined surface, Journal of Geotechnical and Geoenvironmental Engineering, ASCE, Vol. 123,
No. 2, pp. 110–112.
Mindlin, R.D. Force at a point in the interior of a semi-infinite solid, Journal of Applied Physics,
Vol. 7, No. 5, pp. 195–202.
MELT. 1993. Règles techniques de conception et de calcule des fondations de ouvrages de genie civil.
CCTG, Fascicule No. 62, Titre V, Min. de L’Équipement du Lodgement et de Transport, Paris.
Merifield, R.S. and Sloan, S.W. 2006. The ultimate pullout capacity of anchors in frictional soils,
Canadian Geotechnical Journal, Vol. 43, No. 8, pp. 852–868.
Merifield, R.S., Sloan, S.W., and Yu, H.S. 1999. Rigorous plasticity solutions for the bearing capacity
of two-layered clays, Géotechnique, Vol. 49, No. 4, pp. 471–490.
Merifield, R.S., Sloan, S.W., and Yu, H.S. 2001. Stability of plate anchors in undrained clay,
Géotechnique, Vol. 51, No. 2, pp. 141–153.
Mesri, G. 1973. Coefficient of secondary compression, Journal of the Soil Mechanics and Foundations
Division, ASCE, Vol. 99, No. SM1, pp. 123–137.
Mesri, G. and Godlewski, P.M. 1977. Time and stress–compressibility interrelationship, Journal of
the Geotechnical Engineering Division, ASCE, Vol. 103, No. GT5, pp. 417–430.
Meyerhof, G.G. 1953. The bearing capacity of foundations under eccentric and inclined loads,
Proceedings of the Third International Conference on Soil Mechanics and Foundation
Engineering, Zürich, Vol. 1, pp. 440–445.
Meyerhof, G.G. 1956. Penetration tests and bearing capacity of cohesionless soils, Journal of Soil
Mechanics and Foundation Engineering, ASCE, Vol. 82, No. SM1, pp. 1–19.
Meyerhof, G.G. 1963. Some recent research on the bearing capacity of foundations, Canadian
Geotechnical Journal, Vol. 1, No. 1, pp. 16–26.
Meyerhof, G.G. 1974. Ultimate bearing capacity of footings on sand layer overlying clay, Canadian
Geotechnical Journal, Vol. 11, No. 2, pp. 223–229.
Meyerhof, G.G. and Hanna, A.M. 1978. Ultimate bearing capacity of foundations on layered soil
under inclined load, Canadian Geotechnical Journal, Vol. 15, No. 4, pp. 565–572.
Michalowski, R.L. and Shi, L. 1995. Bearing capacity of footings over two-layer foundation soils,
Journal of the Geotechnical Engineering Division, ASCE, Vol. 121, No. GT5, pp. 421–428.
Middendorp, P., Birmingham, P., and Kuiper, B. 1992. Statnamic load testing of foundation piles,
Proceedings of the 4th International Conference on the Application of Stress Wave Theory to
Piles, Ed. Barends, F.B.J., September 21–24, The Hague, Balkema, pp. 581–588.
Middendorp, P. and van Weele, P.J. 1986. Application of characteristic stress wave method in offshore
practice, Proceedings of the 3rd International Conference on Numerical Methods in Offshore
Piling, Nantes, France, May 21–22, Supplement, pp. 6–18.
Mitchell, J.K., Hooper, D.R., and Campanella, R.G. 1965. Permeability of compacted clay, Journal of
the Soil Mechanics and Foundations Division, ASCE, Vol. 91, No. SM4, pp. 41–65.
Mitchell, J.K. and Katti, R.K. 1981. Soil improvement: State-of-the-art report. Proceedings of the
10th International Conference on Soil Mechanics and Foundation Engineering. Volume for
General Reports, State-of-the-art Reports and Lectures, June, A.A. Balkema, Stockholm,
Vol. 4, pp. 261–317.
Mitchell, P.W. and Avalle, D. 1984. A technique to predict expansive soil movements, Fifth
International Conference on Expansive Soils, Adelaide, May 21–23, pp. 124–130.
References 529

Mononobe, N. and Matsuo, O. 1929. On the determination of earth pressure during earthquakes,
Proceedings of the World Engineering Conference, Tokyo, Vol. 9, pp. 179–187.
Morgenstern, N. 1967. Stability charts for earth slopes during rapid drawdown, Géotechnique,
Vol. 17, No. 2, pp. 121–131.
Morgenstern, N.R. and Price, V.E. 1965. The analysis of the stability of general slip surfaces,
Géotechnique, Vol. 15, pp. 79–93.
Moyle, R. 2013. Remediation of deep uncontrolled fill using dynamic compaction. PhD thesis, The
University of Sydney.
Nguyen, V.U. 1985. Determination of critical slope failure surfaces, Journal of Geotechnical
Engineering, ASCE, Vol. 111, No. 2, pp. 238–250.
Nixon, J.F. 1991. Discrete ice lense theory for frost heave in soils, Canadian Geotechnical Journal,
Vol. 28, No. 6, pp. 843–859.
Ogata, A. and Banks, R.B. 1961. A solution of the differential equation of longitudinal dispersion in
porous media, U.S. Geological Survey, Professions Paper 411-A.
Ohsaki, Y. and Iwasaki, R. 1973. On dynamic shear moduli and Poisson’s ratio of soil deposits. Soils
and Foundations, Vol. 13, No. 4, pp. 61–73.
Okabe, S. 1924. General theory on earth pressure and seismic stability of retaining wall and dam,
Journal of the Japanese Society of Civil Engineering, Vol. 10, No. 6, pp. 1277–1323.
O’Neill, M.W. and Poormoayed, N. 1980. Methodology for foundations on expansive clays, Journal
of the Geotechnical Engineering Division, ASCE, Vol. 106, No. GT12, pp. 1345–1367.
OPTUMG2. 2014. Finite element code for geotechnical problems. http://www.optum.com.
Osterberg, J. 1989. New device for load testing driven piles and drilled shafts separates friction and
end bearing, Proceedings of the International Conference on Piling and Deep Foundations,
A.A. Balkema, London, May 1989, Vol. 1, pp. 421–427.
Ottaviani, M. 1975. Three-dimensional finite element analysis of vertically loaded pile groups,
Géotechnique, Vol. 25, No. 2, pp. 159–174.
Ou, C.-Y., Kao, C.-C., and Chen, C.-I. 2009. Performance and analysis of artificial ground freezing
in shield tunnelling, Journal of Geoengineering, Vol. 4, No. 1, pp. 29–41.
Oweidat, L.M. and Day, S.R. 1988. Installation of a composite slurry wall to contain mine tailings,
Tailings and Mine Waste’98, Balkema, Rotterdam, pp. 421–429.
Palmerston, A. and Broch, E. 2006. Use and misuse of rock mass classification systems with particular
reference to the Q-system, Tunnels and Underground Space Technology, Vol. 21, pp. 575–593.
Parry, R.G.H. 1977. Estimating bearing capacity in sand from SPT values, Journal of the Geotechnical
Engineering Division, ASCE, Vol. 103, No. GT9, pp. 1014–1019.
Peck, R. 1969. Deep excavation and tunnelling in soft ground, Proceedings of the Seventh
International Conference on Soil Mechanics and Foundation Engineering, Mexico City, State-
of-the-Art Volume, pp. 225–290.
Peck, R.B., Hanson, W.E., and Thornburn, T.H. 1974. Foundation Engineering, 2nd ed. Wiley,
New York.
Pells, P.J.N. 2004. Substance and mass properties for the design of engineering structures in the
Hawkesbury sandstone, Australian Geomechanics, Vol. 39, No. 3, pp. 1–21.
Pells, P.J.N. and Bertuzzi, R. 2007. Limitations of rock mass classification systems for tunnel support
designs, Tunnels and Tunnelling International, April 2007 issue.
Pells, P.J.N., Douglas, D.J., Rodway, B., Thorne, C.P., and McMahon, B.R. 1978. Design loadings for
shales and sandstones of the Sydney region. Australian Geomechanics Journal, G8, pp. 31–39.
Pells, P.J.N., Mostyn, G., and Walker, R.F. 1998. Foundations on sandstone and shale in the Sydney
region, Australian Geomechanics, No. 33, Part 3, pp. 17–20.
Pells, P.J.N., Poulos, H.G., and Best, R.J. 1991. Rock reinforcement design for a shallow large span
cavern, 7th International Congress on Rock Mechanics, Ed. Wittke, W., A.A. Balkema, Aachen,
Septenber 1991, pp. 1193–1198.
Phase2 Finite Element Software. 2014. Version 8.0, Rocscience. https://www.rocscience.com/products/3/
Phase2.
Pickersgill, G. 1994. Mine residue management – Rehabilitation and decommissioning strategies,
Short Course on Mine Residue Management, Australian Centre for Geomechanics.
530 References

PLAXIS2D. 2014. Finite element software. http://www.plaxis.nl.


Polshin, D.E. and Tokar, R.A. 1957. Maximum allowable non-uniform settlement of struc-
tures, Proceedings of the 4th International Conference on Soil Mechanics and Foundation
Engineering, Vol. 1, pp. 402–405.
Potts, D.M. and Zdravkovic, L. 1999. Finite Element Analysis in Geotechnical Engineering – Theory.
Thomas Telford, London.
Potulski, B.C. 1990. Harris Dam slurry trench: Slurry mix design, Australian Geomechanics,
December, No. 19, pp. 28–32.
Poulos, H.G. 1973. Load-deflection prediction for laterally loaded piles, Australian Geomechanics
Journal, Vol. G3, No. 1, pp. 1–8.
Poulos, H.G. 1988. Cyclic stability diagram for axially loaded Piles. Journal of Geotechnical
Engineering, ASCE, Vol. 114, No. 8, pp. 877–895.
Poulos, H.G. 1989. Pile behaviour-theory and application. Géotechnique, Vol. 39, No.3, pp. 365–415.
Poulos, H.G. 1991. Analysis of piled strip foundations, Proceedings of the 7th International
Conference on Computational Methods and Advances in Geomechanics, Cairns, Eds. Beer,
G., Booker, J.R., and Carter, J.P., Balkema, Rotterdam, May 6–10, 1991, Vol. 1, pp. 183–191.
Poulos, H.G. 1993. Settlement of bored pile groups. Proceedings BAP II, Ghent, Ed. van Impe W.F.,
Balkema, Rotterdam, pp. 103–117.
Poulos, H.G. 1994. An approximate numerical analysis of pile-raft interaction, International Journal
for Numerical and Analytical Methods in Geomechanics, Vol. 18, No. 2, pp. 73–92.
Poulos, H.G. 1995. Design of reinforcing piles to increase slope stability, Canadian Geotechnical
Journal, Vol. 32, No. 5, pp. 808–818.
Poulos, H.G. 2001. Pile foundations, Geotechnical and Geoenvironmental Engineering Handbook,
Chapter 10, Ed. Rowe, R.K. Klewer Academic Publishers, Boston, pp. 261–304.
Poulos, H.G. 2009. Tall buildings and deep foundations – Middle East challenges, Terzaghi Oration,
Proceedings of the 17th International Conference on Soil Mechanics and Geotechnical
Engineering, Alexandria, Egypt, October 5–9, 2009, Vol. 4, pp. 3173–3205.
Poulos, H.G. and Davis, E.H. 1974. Elastic Solutions for Soil and Rock Mechanics. John Wiley &
Sons, New York.
Poulos, H.G. and Davis, E.H. 1980. Pile Foundation Analysis and Design, John Wiley & Sons,
New York.
Priebe, H.J. 1995. The design of vibro-replacement, Ground Engineering, Vol. 28, No. 10, pp. 31–37.
Randolph, M.F. 1981. Response of flexible piles to lateral loading, Géotechnique, Vol. 31, No. 2,
pp. 247–259.
Randolph, M.F. 1983. Design of piled raft foundations, CUED/D, Soils TR 143. Cambridge University.
Randolph, M.F. 1994. Design methods for pile groups and piled rafts, Proceedings of XIII ICSMFE,
New Delhi, January 5–10, pp. 61–82.
Randolph, M.F. and Wroth, C.P. 1978. Analysis of deformation of vertically loaded piles, Journal of
the Geotechnical Engineering Division, ASCE, Vol. 104, No. GT12, pp. 1465–1488.
Rankine, W. 1857. On the stability of loose earth, Philosophical Transactions of the Royal Society of
London, Vol. 147, pp. 9–27.
Rausche, F. Goble, G.G., and Likins, G. 1985. Dynamic determination of pile capacity, Journal of
Geotechnical Engineering, ASCE, Vol. 111, No. 3, pp. 367–383.
Reddy, N.S.C., Dewaikar, D.M., and Mohapatra, G. 2013. Computation of passive earth pressure
coefficients for a horizontal cohesionless backfill using the method of slices, International
Journal of Advanced Civil Engineering and Architecture Research, Vol. 2, No. 1, pp. 32–41.
Reese, L.C., Cox, W.R., and Koop, F.D. 1974. Analysis of laterally loaded piles in sand, Proceedings
of the Sixth Offshore Technology Conference, Houston, TX, May 6–8, Paper OTC 2080,
pp. 473–483.
Reese, L.C. and O’Neill, M.W. 1988. Drilled Shafts: Construction Procedures and Design Methods.
Pub. No. FHWA-H1-88-042, U.S. Dept. Transportation.
RMS QA Specification R57. 2012. Design of Reinforced Earth Walls. Roads and Maritime Services
of New South Wales, Sydney, Australia.
References 531

Robertson, P.K. 1990. Soil classification using the cone penetration test, Canadian Geotechnical
Journal, Vol. 27, No. 1, pp. 151–158.
Robertson, P.K. 2010. Estimating in situ soil permeability from CPT & CPTu, International
Symposium on Cone Penetration Testing, 2, CPT’10, Huntington Beach, CA, May 9–11, 2010,
Vol. 2, pp. 535–542.
Robertson, P.K. and Campanella, R.G. 1983. Interpretation of cone penetration tests – Part I (Sand),
Canadian Geotechnical Journal, Vol. 20, No. 4, pp. 718–733.
Robertson, P.K., Campanella, R.G., Gillespie, D., and Grieg, J. 1986. Use of piezometer cone data.
Proceedings of In Situ 86, ASCE Specialty Conference on Use of In Situ Tests in Geotechnical
Engineering, Blacksburg, Virginia. Geotechnical Special Publication 6, ASCE, Reston, VA,
pp. 1263–1280.
Roscoe, K.H. and Burland, J.B. 1968. On the generalised stress-strain behaviour of “wet” clay,
Engineering Plasticity, Cambridge Univ. Press, Cambridge, pp. 535–609.
Roscoe, K.H. and Schofield, A.N. 1963. Mechanical behaviour of an idealised “wet” clay, 2nd
ECSMFE, Wiesbaden, Vol. 1, pp. 47–54.
Rowe, R.K. and Armitage, H.H. 1983. Design of piles socketed into weak rock, Research Report
GEOT-11–84, University of Western Ontario, London, Ontario.
Rowe, R.K. and Booker, J.R. 1986. A finite layer technique for calculating three-dimensional pollut-
ant migration in soil, Géotechnique, Vol. 36, No. 2, pp. 205–214.
Rowe, R.K., Quigley, R.M., and Booker, J.R. 1997. Clayey Barrier Systems for Waste Disposal
Facilities. E & FN Spon, London.
Russo, G.V., Abagnara, V., Poulos, H.G., and Small, J.C. 2012. Re-assessment of foundation
settlements for the Burj Khalifa, Dubai, ACTA Geotechnica, Springer, Berlin, Published online
October 30, 2012.
Santos, J.A.D. and Correia, A.G. 2001. Reference threshold shear strain of soil. Its application
to obtain an unique strain-dependent shear modulus curve for soil, Proceedings of the 15th
International Conference on Soil Mechanics and Geotechnical Engineering, Istanbul, Vol. 1,
pp. 267–270.
Sarma, S.K. 1973. Stability analysis of embankments and slopes, Géotechnique, Vol. 23, No. 3,
pp. 423–433.
Sarma, S.K. 1979. Stability of embankments and slopes, Journal of the Geotechnical Engineering
Division, ASCE, Vol. 105, No. GT12, pp. 1511–1524.
Scheurenberg, R.J. 1982. Experiences in the use of geofabrics in underdrainage of residue deposits,
2nd International Conference on Geotextiles, Las Vegas, Vol. 1, pp.199–204.
Schmertmann, J.H. 1970. Static cone to compute static settlement over sand, Journal of the Soil
Mechanics and Foundations Division, ASCE, Vol. 96, No. SM3, pp. 1011–1043.
Schmertmann, J.H., Hartman, J.P., and Brown, P.R. 1978. Improved strain influence factor diagrams,
Journal of the Geotechnical Engineering Division, ASCE, Vol. 104, No. GT8, pp. 1131–1135.
Schnaid, F. 2009. In Situ Testing in Geomechanics–The Main Tests. Taylor & Francis Group, London,
327p.
Schofield, A. and Wroth, P. 1968. Critical State Soil Mechanics. McGraw Hill, London.
Schubert, W. 1996. Dealing with squeezing conditions in alpine tunnels, Rock Mechanics and Rock
Engineering, Vol. 29, No. 3, pp. 145–153.
Seed, H.B. 1979. Considerations in the earthquake resistant design of earth and rockfill dams,
Géotechnique, Vol. 92, No. 3, pp. 13–41.
Seed, H.B. and Sultan, H.A. 1967. Stability analysis for a sloping core embankment, Journal of the
Soil Mechanics and Foundations Division, ASCE, Vol. 93, No. SM4, pp. 69–83.
Seed, H.B. and Idriss, I.M. 1971. Simplified procedure for evaluating soil liquefaction poten-
tial, Journal of the Soil Mechanics and Foundations Division, ASCE, Vol. 97, No. SM9,
pp. 1249–1273.
Seed, H.B., Tokimatsu, K., Harder, L.F., and Chung, R.M. 1985. Influence of SPT procedures in
soil liquefaction resistance evaluations, Journal of Geotechnical Engineering, ASCE, Vol. 111,
No. 12, pp. 1425–1445.
532 References

Selvadurai, A.P.S. 1979. Elastic Analysis of Soil-Foundation Interaction. Elsevier Publishing


Company, New York.
Settlement Analysis. 1994. Technical Engineering and Design Guides as Adapted from the U.S.
Army Corps of Engineers, No. 9. ASCE Press, New York.
Sharma, P.V. 1997. Environmental and Engineering Geophysics. Cambridge University Press,
Cambridge.
Sherard, J.L. and Dunnigan, L.P. 1989. Critical filters for impervious soils, Journal of Geotechnical
Engineering, ASCE, Vol. 115, No. 7, pp. 927–947.
Sherard, J.L., Dunnigan, L.P., Decker, R.S., and Steele, E.F. 1976. Pinhole test for identifying
dispersive soils, Journal of Geotechnical Engineering, ASCE, Vol. 102, No. GT1, pp. 63–80.
Shiau, J.S., Augarde, C.E., Lyamin, A.V., and Sloan, S.W. 2008. Finite element limit analysis of pas-
sive earth resistance in cohesionless soils, Soils and Foundations, Vol. 48, No. 6, pp. 843–850.
Shields, D., Chandler, N., and Garnier, J. 1990. Bearing capacity of foundations in slopes, Journal of
Geotechnical Engineering, ASCE, Vol. 116, No. 3, pp. 528–537.
Simpson, B., O’Riordan, N.J., and Croft, D.D. 1979. A computer model for the ground movements in
London Clay, Géotechnique, Vol. 29, No. 2, pp. 149–175.
Sinha, J. 1997. Piled raft foundations subjected to swelling and shrinking soils. PhD thesis, University
of Sydney, Australia.
Sinha, J. and Poulos, H.G. 1996. Behaviour of stiffened raft foundations, 7th Australia–New Zealand
Conference on Geomechanics, Adelaide, July 1–5, 1996.
Skempton, A.W. 1957. Discussion: “The planning and design of the new Hong Kong airport,”
Proceedings, The Institution of Civil Engineers, London, Vol. 7, No. 4, pp. 305–307.
Skempton, A.W. and MacDonald, D.H. 1956. The allowable settlement of buildings, Proceedings of
the Institution of Civil Engineers, Part III, The Institution of Civil Engineers, London, England,
Vol. 5, pp. 727–768.
Sloan, S.W. 2013. Geotechnical stability analysis, Géotechnique, Vol. 63, No. 7, pp. 531–572. http://
dx.doi.org/10.1680/geot.12.RL.001.
Sloan, S.W. and Randolph, M.F. 1982. Numerical prediction of collapse loads using finite element
methods, International Journal for Numerical and Analytical Methods in Geomechanics,
Vol. 6, pp. 47–76.
Sloan, S.W. and Yu, H.S. 1996. Rigorous plasticity solutions for the bearing capacity factor Nγ,
Proceedings of the 7th Australia–New Zealand Conference on Geomechanics, Adelaide, July
1–5, 1996.
SLOPE/W. 2012. Slope stability software. http://www.geo-slope.com/products/slopew.aspx.
Small, J.C. 2012. Consolidation of embankments on soils with anisotropic permeability, 11th
Australia–New Zealand Conference on Geomechanics, Ground Engineering in a Changing
World, Crown Conference Centre, Melbourne, Australia, July 15–18, 2012, Paper 1.3.6,
pp. 179–184.
Small, J.C. 2013. FEAR User’s Manual, Centre for Geotechnical Research, The University of Sydney,
http://sydney.edu.au/engineering/civil/research/scigem/software/index.shtml.
Small, J.C. and Booker, J.R. 1979. Analysis of the consolidation of layered soils using the method of
lines, Proceedings of the 3rd International Conference on Numerical Methods in Geomechanics,
Ed. Wittke, W., Aachen, April 2–6, pp. 201–211.
Small, J.C. and Booker, J.R. 1984. Finite layer analysis of layered elastic materials using a flexibility
approach. Part 1 – Strip loadings, International Journal for Numerical Methods in Engineering,
Vol. 20, pp. 1025–1037.
Small, J.C. and Booker, J.R. 1986. Finite layer analysis of layered of layered elastic materials using
a flexibility approach. Part 2 – Circular and rectangular loadings, International Journal for
Numerical Methods in Engineering, Vol. 23, pp. 959–978.
Small, J.C. and Booker, J.R. 2014. Program FLEA (Finite Layer Elastic Analysis). Sydney Centre
in Geomechanics and Mining Materials, The University of Sydney. http://sydney.edu.au/
engineering/civil/research/scigem/software/index.shtml.
Small, J.C. and Poulos, H.G. 2007. A method of analysis of piled rafts, Proceedings 10th ANZ
Conference Geomechanics, Brisbane, Australian Geomechanics Society, Vol. 2, pp. 602–607.
References 533

Soubra A.H. and Macuh, B. 2002. Active and passive earth pressure coefficients by a kinemati-
cal approach, Proceedings of the Institute of Civil Engineers and Geotechnical Engineering,
Vol. 155, No. 2, pp. 119–131.
Spencer, E. 1967. A method of analysis of the stability of embankments assuming parallel interslice
forces, Géotechnique, Vol. 17, pp. 11–26.
Stain, R.T. and Williams, H.T. 1991. Interpretation of sonic coring results – A research project,
Proceedings of the 4th International Conference on Piling and Deep Foundations, Stresa, Italy,
Balkema, Rotterdam, April 7–12, Vol. 1, pp. 633–640
Stark, T.D. and Olsen, S.M. 1995. Liquefaction resistance using CPT and field case histories, Journal
of Geotechnical Engineering, ASCE, Vol. 121, No. 12, pp. 856–869.
Stokoe, K.H. and Hoar, R.J. 1978. Variables affecting in situ seismic measurement, Proceedings,
Earthquake Engineering and Soil Dynamics, ASCE, Pasadena, CA, June 19–21, pp. 919–938.
Stokoe, K.H. and Woods, R.D. 1972. In situ shear wave velocity by cross-hole method, Journal of the
Soil Mechanics and Foundations Division, ASCE, Vol. 98, No. SM5, pp. 443–460.
Stroud, M.A. 1974. The Standard Penetration Test in insensitive clays and soft rocks, Proceedings of
the European Symposium on Penetration Testing, ESOPT, Stockholm 1974. National Swedish
Building Research, Vol. 2.2, pp. 367–375.
Subba Rao, K.S. and Choudhury, D. 2005. Seismic passive earth pressures in soils, Journal of
Geotechnics and Environmental Engineering, ASCE, Vol. 131, No. 1, pp. 131–135.
Sullivan, W.R., Reese, L.C., and Fenske, C.W. 1979. Unified method for analysis of laterally loaded
piles in clay, Numerical Methods in Offshore Piling, London, May 22–23, pp. 135–146.
Sultan, H.A. and Seed, H.B. 1967. Stability of sloping core earth dams, Journal of the Soil Mechanics
and Foundations Division, ASCE, Vol. 93, No. SM4, pp. 45–67.
SVSLOPE-2D and SVSLOPE-3D. 2014. Soil Vision Systems. http://www.soilvision.com/subdomains/
svslope.com/index.shtml.
Ta, L.D. and Small, J.C. 1996. Analysis of piled raft systems in layered soils, International Journal
for Numerical and Analytical Methods in Geomechanics, Vol. 20, pp. 57–72.
Talbot, A. 1979. The accurate numerical inversion of Laplace Transforms, Journal of the Institute of
Mathematics and its Applications, Vol. 23, pp. 97–120.
Tan, Y. and Wang, D. 2013a. Characteristics of a large-scale deep foundation pit excavated by
the Central-Island Technique in Shanghai soft clay I: Bottom-up construction of the central
cylindrical shaft, Journal of Geotechnical and Geoenvironmental Engineering, ASCE, Vol. 139,
No. 11, pp. 1875–1893.
Tan, Y. and Wang, D. 2013b. Characteristics of a large-scale deep foundation pit excavated by the
Central-Island Techniquein Shanghai soft clay II: Top-down construction of the peripheral rect-
angular pit, Journal of Geotechnical and Geoenvironmental Engineering, ASCE, Vol. 139,
No. 11, pp. 1894–1910.
Tavenas, F. and Leroueil, S. 1980. Behaviour of embankments on clay foundations, Canadian
Geotechnical Journal, Vol. 17, No. 2, pp. 236–260.
Tchepak, S. 1998. Pile testing, Mini-Symposium on Recent Developments in Piling Practice in
Sydney, Australian Geomechanics Society, Sydney Chapter.
Terzaghi, K. 1923. Die Berechnung der Durchlassigkeitziffer des Tones aus dem Verlauf der hydrody-
namischen Spannungserscheinungen. Original paper published in 1923 and reprinted in From
Theory to Practice in Soil Mechanics. John Wiley & Sons, New York, pp. 133–146.
Terzaghi, K. 1943. Theoretical Soil Mechanics. John Wiley & Sons, New York.
Terzaghi, K. and Peck, R.B. 1967. Soil Mechanics in Engineering Practice. John Wiley & Sons,
New York.
Thaher, M. and Jessberger, H.L. 1991. The behaviour of pile-raft foundations investigated in centri-
fuge model tests, Centrifuge 91, Boulder, Colorado, USA, 1991.
Thompson, P.M., Corthesy, R., and Leite, M.H. 1997. Rock stress measurements at great depth using
the modified doorstopper gauge, Proceedings of the International Symposium on Rock Stress,
Japan, Eds. Sugawara, K. and Obara, Y., AA Balkema, Rotterdam, pp. 59–64.
Thornburn, S. 1975. Building structures supported by stabilized ground, Géotechnique, Vol. 25,
No. 1, pp. 83–94.
534 References

Timoshenko, S. and Woinowski-Krieger, S. 1959. Theory of Plates and Shells. McGraw-Hill,


New York.
Tomlinson, M.J. 1995. Foundation Design and Construction, 6th ed. Longman Scientific and
Technical, Harlow, UK.
UDEC. 2014. Universal Distinct Element Code V6.0. ITASCA Consulting Group. http://www.­
itascacg.com/software/udec.
U.S. Army Corps of Engineers. 2004. General Design and Construction for Earth and Rock-fill
Dams, EM1110-2-2300, July 30.
Vesic, A.S. 1967. A study of bearing capacity of deep foundations, Final report Project B-189, School
of Civil Eng., Georgia Inst. Tech., Atlanta, Georgia.
Vesic, A.S. 1972. Expansion of cavities in infinite soil mass, Journal of the Soil Mechanics and
Foundations Division, ASCE, Vol. 98, No. SM3, pp. 265–290.
Vesic, A.S. 1973. Analysis of ultimate loads of shallow foundations, Journal of the Soil Mechanics
and Foundations Division, ASCE, Vol. 99, No. SM1, pp. 45–73.
Vesic, A.S. 1975. Bearing capacity of shallow foundations, Chapter 3, Foundation Engineering
Handbook, Ed. Winterkorn, H.F. and Fang, H.-Y., Van Nostrand Reinhold, New York.
Vick, S.G. 1983. Planning, Design and Analysis of Tailings Dams, Wiley-Interscience, John Wiley &
Sons, New York.
Vidic, R.D. and Pohland, F.G. (eds.). 1995. Innovative technologies for site remediation and hazard-
ous waste management, Proceedings of the National Conference, Environmental Engineering
Division ASCE, Pittsburgh, July 26.
Viggiani, C. 1981. Ultimate lateral load on piles used to stabilise landslides, Proceedings of the 10th
International Conference on Soil Mechanics and Foundation Engineering, Stockholm, Vol. 3,
pp. 555–560.
Viggiani, C. 1998. Pile groups and piled raft behaviour, Deep Foundations on Bored and Augered
Piles, BAP III, Eds. van Impe, W.F. and Haegman, W., Balkema, Rotterdam, pp. 77–90.
Vijayvergia, V.N. and Focht, J.A. Jr. 1972. A new way to predict the capacity of piles in clay, 4th
Annual Offshore Technology Conference, Houston, Vol. 2, pp. 865–874.
WALLAP. 2013. Retaining Wall Analysis Program, www.geosolve.co.uk.
Wan, T.-Y. and Mitchell, J.K. 1976. Electro-osmotic consolidation of soils, Journal of the Geotechnical
Engineering Division, ASCE, Vol. 102, No. GT5, pp. 414–432.
Wickham, G.E., Tiedemann, H.R., and Skinner, E.H. 1972. Support determination based on
geological predictions, Proceedings of the Rapid Excavation and Tunneling Conference,
AIME, New York, pp. 43–64.
Williams, A.F. and Pells, P.J.N. 1981. Side resistance of rock sockets in sandstone, mudstone and
shale, Canadian Geotechnical Journal, Vol. 18, No. 4, pp. 502–513.
Wiss, J.F. 1981. Construction vibrations: State-of-the-art, Journal of the Geotechnical Engineering
Division, ASCE, Vol. 107, No. GT2, pp. 167–181.
Wong, P.K. 2007. Preload design to reduce post construction creep settlement. Soft Soils and Ground
Improvement Workshop. Proceedings of the 10th Australia New Zealand Conference on
Geomechanics, Brisbane, Australia, October 21–24, pp. 23–31.
Wong, P.K. 2010. Comparison of experimental methods for assessing secondary consolidation settle-
ment of soft soils after surcharging, Australian Geomechanics Society Seminar on Ground
Improvement, Perth, WA, June 11–12.
Worotnicki, G. and Walton, R.J. 1976. Triaxial hollow inclusion gauges for determination of rock
stresses in situ, Supplement to Proceedings ISRM Symposium on Investigation of Stress in
Rock, Advances in Stress Measurement, Sydney, The Institution of Engineers, Australia,
May 11–13, Suppl. pp. 1–8.
Wroth, C.P. 1984. The interpretation of in situ soil tests, Géotechnique, Vol. 34, No. 4,
pp. 449–489.
Wu, T.H. 1976. Soil Mechanics. Allyn & Bacon, Boston.
Wyllie, D.C. and Mah, C.W. 2004. Rock Slope Engineering, Civil and Mining, 4th ed. Spon, Oxford.
XSLOPE. 2014. Slope stability software for two-dimensional analysis. http://sydney.edu.au/
engineering/civil/research/scigem/software/xslope/index.shtml.
References 535

Zhang, H.H. and Small, J.C. 2000. Analysis of capped pile groups subjected to horizontal and verti-
cal loads, Computers and Geotechnics, Vol. 26, No. 1, pp. 1–21.
Zhang, L. and Einstein, H. 1998. End bearing capacity of drilled shafts in rock, Journal of
Geotechnical Engineering, ASCE, Vol. 124, No. 7, pp. 574–584.
Zhang, L., Ernst, H., and Einstein, H.H. 2000. Non-linear analysis of laterally loaded rock-socketed
shafts, Journal of Geotechnical and Geoenvironmental Engineering, ASCE, Vol. 126, No. 11,
pp. 955–968.
Zhemochkin, B.N. and Sinitsyn, A.P. 1962. Practical Methods of Designing Foundation Beams
and Slabs Resting on an Elastic Foundation, 2nd ed. State Publishing House for Literature on
Structures, Architecture and Structural Materials, Moscow.
Zhuang, G.M., Lee, I.K., and Zhao, X.H. 1991. Interactive analysis of behaviour of raft-pile foun-
dations, Proceedings of The International Conference on Geotechnical Engineering for
Coastal Development – Theory and Practice on Soft Ground – GEO-COAST ‘91, Yokohama,
September 3–6, 1991, Vol. 1, pp. 759–764.
Zienkiewicz, O.C. 1977. The Finite Element Method, 3rd ed. McGraw-Hill, London.
This page intentionally left blank

Вам также может понравиться