Sie sind auf Seite 1von 294

Louisiana State University

LSU Digital Commons


LSU Historical Dissertations and Theses Graduate School

1997

Numerical Simulation of Geosynthetic Reinforced


Earth Structures Using Finite Element Method.
Surajit Pal
Louisiana State University and Agricultural & Mechanical College

Follow this and additional works at: https://digitalcommons.lsu.edu/gradschool_disstheses

Recommended Citation
Pal, Surajit, "Numerical Simulation of Geosynthetic Reinforced Earth Structures Using Finite Element Method." (1997). LSU
Historical Dissertations and Theses. 6587.
https://digitalcommons.lsu.edu/gradschool_disstheses/6587

This Dissertation is brought to you for free and open access by the Graduate School at LSU Digital Commons. It has been accepted for inclusion in
LSU Historical Dissertations and Theses by an authorized administrator of LSU Digital Commons. For more information, please contact
gradetd@lsu.edu.
INFORMATION TO USERS

This manuscript has been reproduced from the microfilm master. UMI
films the text directly from the original or copy submitted. Thus, some
thesis and dissertation copies are in typewriter face, while others may be
from any type o f computer printer.

The quality of this reproduction Is dependent upon the quality o f the


copy submitted. Broken or indistinct print, colored or poor quality
illustrations and photographs, print bleedthrough, substandard margins,
and improper alignment can adversely affect reproduction.

In the unlikely event that the author did not send UMI a complete
manuscript and there are missing pages, these will be noted. Also, if
unauthorized copyright material had to be removed, a note will indicate
the deletion.

Oversize materials (e.g., maps, drawings, charts) are reproduced by


sectioning the original, beginning at the upper left-hand comer and
continuing from left to right in equal sections with small overlaps. Each
original is also photographed in one exposure and is included in reduced
form at the back o f the book.

Photographs included in the original manuscript have been reproduced


xerographically in this copy. Higher quality 6” x 9” black and white
photographic prints are available for any photographs or illustrations
appearing in this copy for an additional charge. Contact UMI directly to
order.

UMI
A Bell & Howell Information Company
300 North Zeeb Road, Ann Arbor MI 48106-1346 USA
313/761-4700 800/521-0600

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
NUMERICAL SIMULATION OF GEOSYNTHETIC REINFORCED EARTH
STRUCTURES USING FINITE ELEMENT METHOD

A Dissertation

Submitted to the Graduate Faculty of the


Louisiana State University and
Agricultural and Mechanical College
in partial fulfillment of the
requirements for the degree of
Doctor of Philosophy

The Department of Civil and Environmental Engineering

by
Surajit Pal
B.C.E., Jadavpur University, Calcutta, India, 1987
M.S., State University of New York at Buffalo, Buffalo, 1993
December 1997

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
UMI Number: 9820741

UMI Microform 9820741


Copyright 1998, by UMI Company. All rights reserved.

This microform edition is protected against unauthorized


copying under Title 17, United States Code.

UMI
300 North Zeeb Road
Ann Arbor, MI 48103

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Dedicated to my wife, Gargi.

ii

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
ACKNOWLEDGMENTS
I wish to express my deepest gratitude to my advisor Dr. G. Wije Wathugala for his

guidance and encouragement throughout the course of this research. His constant financial
support and enthusiastic participation in the work are sincerely appreciated.
Special thanks are due to my committee members, Dr. Sukhamay Kundu, Dr. Mehmet

T. Tumay and Dr. Geotge Z. Voyadjis for their constant encouragement and helpful sugges­

tions during the course of this work. I would also like to express my respect and gratitutde
to the late Dr. Yalcin B. Acar whose sudden demise left a vacuum in the geotechnical

engineering community at LSU.

Thanks are also due to Dr. Khalid Farrag, Mr Hadi Shirazi and Mr. Ken Johnston of

Louisiana Transportation Research Ceneter (L T R C ). Their help in this work is deeply


appreciated.

In addition, I would like to thank my colleagues Dr. Rainer Echle, Dr. Hani Titi,

Mr. Fazle Rabbi, Mr. Baoshan Huang, Mr. Wenbing Zheng, Mr. Prasad Samarajiva, Mr.

Ananda Hirath and Ms. Senda Ozkan for their friendship and encouragement during my
stay at LSU.

Finally, I wish to express my gratitude to my mother, wife, sisters and friends for their
constant inspiration and support of this research work.

iii

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
CONTENTS

D E D IC A T IO N ............................................................................................................ ii

A C K N O W LED G M EN TS......................................................................................... iii

LIST OF T A B L E S ..................................................................................................... vii

LIST OF FIGURES ......................................................................................................viii

ABSTRACT...................................................................... xv

CHAPTER
1 IN T R O D U C T IO N ............................................................................................... 1
1.1 General........................................................................................................... 1
1.2 Proposed r e s e a rc h ........................................................................................ 4
1.3 Methodology................................................................................................. 4
1.4 Organization of various chapters.................................................................. 5

2 BACKGROUND OF THE S T U D Y ................................................................... 6


2.1 General........................................................................................................... 6
2.2 Current design m e th o d s .............................................................................. 6
2.3 Limitations of the current design m eth o d s.................................................. 10
2.4 Numerical simulation by F E M ..................................................................... 11
2.5 Constitutve models for soils and interfaces.................................................. 15

3 HiSS CONSTITUTIVE MODELS AND ITS IMPLEMENTATION IN FEM 19


3.1 In tro d u ctio n ................................................................................................. 19
3.2 HiSS constitutive models.............................................................................. 21
3.3 Incremental stress strain relatio n sh ip ......................................................... 24
3.4 Stress to strain a lg o rith m ............................................................................ 26
3.5 Strain to Stress Algorithm ............................................................................ 29
3.5.1 Introduction...................................................................................... 29
3.5.2 Non-Virgin L oading.......................................................................... 29
3.5.2.1 Linear E la s tic .................................................................... 29
3.5.2.2 N onlinear.......................................................................... 32
3.5.3 Virgin Loading ................................................................................ 35
3.5.3.1 Elastic Predictor-Plastic Corrector Method ................... 39
3.5.3.2 Comparison of Elastic Predictor and Plastic Predictor
M eth o d s............................................................................ 43
3.5.3.3 Implicit Integration M e th o d ............................................ 45
3.5.3.4 Modified Euler M e th o d ................................................... 49

iv

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
3.5.4 Subincrementation............................................................................ 51
3.6 Exam ples......................................................................................................... 53
3.6.1 Prediction of Simulated Triaxial T e s t s ............................................ 54
3.6.2 Footing P ro b le m s ............................................................................ 56
3.7 C onclusions.................................................................................................. 74

4 CALIBRATION OF HiSS CONSTITUTIVE MODELS USING GENETIC


ALGORITHMS .................................................................................................. 82
4.1 In tro d u ctio n .................................................................................................. 82
4.2 Traditional m eth o d ........................................................................................ 84
4.3 Coding and decoding of param eters............................................................ 8 6
4.4 Fitness f u n c tio n ........................................................................................... 87
4.5 Genetic algorithm { G A ) ............................................................................... 90
4.6 Exam ples......................................................................................................... 93
4.6.1 Simulated test d a ta ............................................................................ 93
4.6.2 Simulated cyclic t e s t .............................................................................102
4.6.3 Real test d a ta .........................................................................................102
4.7 Performance of two different cross-over s c h e m e s ......................................... 107
4.8 Discussion and conclusion................................................................................107

5 DISTURBED STATE CONCEPT BASED INTERFACE MODEL AND ITS


IMPLEMENTATION IN F E M ............................................................................ 110
5.1 In tro d u ctio n ..................................................................................................... 110
5.2 Disturbed state concept .................................................................................. I l l
5.3 Idealization of in te rfa c e .................................................................................. 113
5.4 Proposed m o d e l ............................................................................................... 114
5.5 Incremental stress-strain re la tio n s...................................................................120
5.6 Determination of material p aram eters............................................................ 124
5.7 Implementation in finite element method ...................................................... 127

6 INTERFACE MODEL VERIFICATION AND PULL-OUT TEST SIMULA­


TION ......................................................................................................................... 137
6.1 General...............................................................................................................137
6.2 Model v erification........................................................................................... 137
6.3 Pull-out sim ulation........................................................................................... 141
6.3.1 Modelling sa n d ..................................................................................... 152
6.3.2 Modelling reinforcement..................................................................... 152
6.3.3 Modelling interfaces............................................................................155
6.3.4 Pull-out simulation re s u lts ................................................................... 159
6.4 Summary and c o n c lu s io n ............................................................................... 159

7 NUMERICAL SIMULATION OF DENVER W A L L ......................................... 161


7.1 General...............................................................................................................161
7.2 Denver W a l l ..................................................................................................... 161
7.3 Material m o d e ls ...............................................................................................163

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
7.3.1 Backfill s a n d ....................................................................................... 164
7.3.2 Reinforcem ent....................................................................................165
7.3.3 Wood f a c in g ....................................................................................... 166
7.3.4 In te rfa c e ............................................................................................. 169
7.4 Finite element m o d el........................................................................................ 174
7.5 Result and d is c u s s io n .....................................................................................175
7.6 Summary...........................................................................................................188

8 NUMERICAL SIMULATION OF RMC W A L L ............................................... 194


8.1 General..............................................................................................................194
8.2 RMC W a ll........................................................................................................194
8.3 Material m o d e ls .............................................................................................. 197
8.3.1 Backfill s a n d .......................................................................................197
8.3.2 Reinforcem ent................................................................................... 198
8.3.3 F a c in g ................................................................................................200
8.3.4 Interfaces............................................................................................ 201
8.4 Finite element m o del.......................................................................................202
8.5 Result and d is c u s s io n ....................................................................................206
8 .6 Summary.......................................................................................................... 222

9 SUMMARY AND CONCLUSION .....................................................................228


9.1 Summary.......................................................................................................... 228
9.2 Conclusion.......................................................................................................229
9.3 Recommendations for future rese a rc h ...........................................................231

R E FE R EN C ES..............................................................................................................232

APPENDIX A: INCORPORATION OF HiSS ^ MODEL INTO ABAQUS . 244

APPENDIX B: DENVER W A L L ...........................................................................248

APPENDIX C: RMC W A L L ................................................................................. 264

V I T A ............................................................................................................................. 272

vi

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
LIST OF TABLES
3.1 Possible stress increment direction when current stress is on the yield sur­
face (Linear elastic non-virgin loading case). (After Wathugala, 1990) . . 36

3.2 Possible stress increment directions when the current stress is inside the
yield surface (Nonlinear non-virgin loading) (After Wathugala, 1990) . . . 37

3.3 Possible stress increment directions when the current stress is on the yield
surface (Nonlinear non-virgin loading case) (After Wathugala, 1990) . . . 38

3.4 Summary of comparisons of different methods ......................................... 80

4.1 Material Parameters for Simulated Triaxial Tests ...................................... 95

4.2 Material Parameters for Simulated Cyclic T e st............................................... 100

4.3 Material parameters obtained from real te s ts .................................................. 103

6.1 Material constants for interfaces..................................................................... 139

6.2 Material constants for H i S S Si model for s a n d ............................................ 154

6.3 Material constants for geogrids ..................................................................... 155


7.1 Material constants for H i S S Si model for granular ba ck fill......................... 164

7.2 Material constants for sand-geotextile interfaces............................................ 173

7.3 Material constants for sand-plywood in te rfa c e s............................................ 174

8.1 Material constants for H i S S Si model for granular b a ck fill......................... 198

8.2 Material constants for geogrid-soil interfaces ...............................................203

A. 1 Parameters to be defined in the input file for the HiSS £i m odel................ 245

A.2 Format of the input file insdv.in required for the subroutine SDVINI . . . 247

vii

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
LIST OF FIGURES
1.1 A schematic of reinforced earth w a ll........................................................... 2

2.1 Earth pressure distributions used in tied-back wedge methods (After Clay-
boumandW u, 1991)...................................................................................... 8

2.2 Vertical stress distributions at the base of the w a l l ....................................... 12

3.1 Shape of yield surfaces in - yJJ2D plane ................................................. 22

3.2 Shape of yield surfaces in triaxial p l a n e ....................................................... 23

3.3 Shape of yield surfaces in octahedral p l a n e ................................................. 24

3.4 Possible stress increment directions when current stress is on the yield
surface (Linear elastic non-virgin case)(After Wathugala, 1 9 9 0 )................ 30

3.5 Possible stress increment directions when current stress is inside the yield
surface (Linear elastic non-virgin case)(After Wathugala, 1 9 9 0 )................ 32

3.6 Possible stress increment directions when current stress is inside the yield
surface (nonlinear non-virgin case)(After Wathugala, 1 9 9 0 )....................... 33

3.7 Possible stress increment directions when current stress is on the yield
surface (Nonlinear non-virgin loading case)(After Wathugala, 1990) . . . 36

3.8 Schematic diagram for ideal predictor corrector algorithm (After Wathugala,
1990)............................................................................................................... 40

3.9 Schematic diagram for iteration procedure (After Wathugala, 1990) . . . . 43

3.10 Schematic representation of Implicit M ethod............................................. 46

3.11 Schematic representation of the Modified Euler m e th o d .......................... 52

3.12 Stress paths followed by different triaxial t e s t s .......................................... 54

3.13 Stress paths of the C T C drained test (Aemax = 10-4, T O L = 10” 5) . . 57

3.14 Stress paths of the C T C drained test (Aemai = 10-3, T O L = 10-2) . . 58

3.15 Stress paths of the C T C drained test (Aemax = 10-2, T O L = 1) . . . . 59

3.16 Stress paths of the T E drained test (Aemax = 10~4, T O L = 10-5) ... 60

3.17 Stress paths of the T E drained test (Aemax = 10~3, T O L — 10“ 2) ... 61

3.18 Stress paths of the T E drained test (Aemai = 10~2, T O L = 1 ) ............. 62

viii

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
3.19 Stress paths of the C T C undrained test (AemffiC = 1 0 “4, TOL = 1 0 -5) . 63

3.20 Stress paths of the C T C undrained test ( A e ,^ = 1 0 ~3, TOL = 1 0 ~2) . 64

3.21 Stress paths of the C T C undrained test (Aemai = 1 0 ~2, TOL = 1) . . . 65

3.22 C P U time taken by different algorithms for C D —C T C ...................... 66

3.23 C P U time taken by different algorithms for C D —T E ......................... 67

3.24 C P U time taken by different algorithms for C U —C T C ...................... 68

3.25 F E M mesh for the footing p ro b le m ......................................................... 69

3.26 Load-settlement curves for the rigid footing (Aemax = 10-2) ................. 71

3.27 Load-settlement curves for the rigid footing (Aemai = 10-3) ................. 72

3.28 Load-settlement curves for the rigid footing (Aemax = 10“4, T O L = 10-5) 73

3.29 Load-settlement curves for the flexible footing (Aemax = 10-2) .............. 75

3.30 Load-settlement curves for the flexible footing (Aemax = 10-3) ............. 76

3.31 Load-settlement curves for the flexible footing (Acmax = 10-4, T O L =


10~5) ........................................................................................................... 77

3.32 C P U time taken by different algorithms for the rigid f o o t i n g ................ 78

3.33 C P U time taken by different algorithms for the flexible fo o tin g ............. 79

4.1 Definition of fitness v a l u e ............................................................................. 92

4.2 Schematic diagram of cross-over.................................................................... 92

4.3 Flow chart of genetic a lg o rith m .................................................................. 94

4.4 Simulated C T C test (confining pressure 150 kPa) .................................... 96

4.5 Simulated C T C test (confining pressure 100 k P a ) .................................... 97

4.6 Simulated C T C test (confining pressure 200 kPa) .................................... 98

4.7 Simulated cyclic t e s t ......................................................................................... 101

4.8 Real C T C test (confining pressure 34.5 kPa) ................................................ 104

4.9 Real C T C test (confining pressure 89.6 kPa) ................................................ 105

4.10 Real C T C test (confining pressure 138 k P a ) ................................................ 106

4.11 Performance of one-point cross-over and six-point cross-over schemes . 108

ix

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
5.1 Disturbed state as a mixture of R I and F A s t a t e s ........................................ 112
5.2 Evolution of damage fu n c tio n ....................................................................... 113

5.3 Idealization of interface: (a) Natural and idealized interface, (b) Interface
at disturbed state, (c) Definition of stresses and displacements (After Desai
and Ma, 1 9 9 2 ).............................................................................................. 115

5.4 Yield surface for R I m a te ria l....................................................................... 117

5.5 Schematic representation of observed response: (a) strain softening and


(b) strain h a rd e n in g ....................................................................................... 1 2 1

5.6 Schematic representation of observed response: (a) compression and (b)


dilation............................................................................................................. 1 2 1

5.7 Elastic shear and normal stiffnesses: (a) Shear stiffness, K s, (b) Normal
stiffness, K n .................................................................................................... 125

5.8 Ultimate shear stress ru for (a) strain hardening response and (b) strain
softening resp o n se .......................................................................................... 128

5.9 Determination of Ti2 and 7 [ 1 —(n2 / n x ) ] ........................................................ 129

5.10 Determination of hardening parameters ai and 77! 129

5.11 Determination of critical state parameter M ................................................ 130

5.12 Determination of critical volume change parameters a and b .......................130

5.13 Determination of damage parameters k and A .............................................131

5.14 Thin-layer e le m e n t....................................................................................... 131

6.1 ln(r*/PZ) vs. ln(a/Pa) p lo t.......................................................................... 140

6.2 ln{pL) vs. ln(£v) p l o t .......................................................................................141

6.3 r c vs. a p l o t ..................................................................................................... 142

6.4 vc vs. ln(a/Pa)p l o t ..........................................................................................143

6.5 ln[—ln(l — vs. ln{^o) p lo t..................................................................... 144

6 .6 ln[—ln(l —^j-)] vs. ln(^D) p lo t..................................................................... 145

6.7 Test data and prediction of direct shear test at normal stress of 26.8 k P a:
(a) Shear stress vs. shear displacement and (b) Normal displacement vs.
shear displacement.......................................................................................... 146

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
6 .8 Test data and prediction of direct shear test at normal stress of 53.6 k P a :
(a) Shear stress vs. shear displacement and (b) Normal displacement vs.
shear displacement........................................................................................ 147

6.9 Test data and prediction of direct shear test at normal stress of 80.4 k P a :
(a) Shear stress vs. shear displacement and (b) Normal displacement vs.
shear displacement........................................................................................148

6.10 Test data and prediction of direct shear test at normal stress of 214.4 k P a :
(a) Shear stress vs. shear displacement and (b) Normal displacement vs.
shear displacem ent.....................................................................................149

6.11 Pull-out box (After Farrag, 1990)............................................................... 150

6.12 Finite element mesh for pull-out sim u la tio n ............................................ 151

6.13 Test data and prediction of H i S S model: (a) \J Jid v s . \JIid and (b)
y/IiD vs. —I \ ...............................................................................................153
6.14 Test data and prediction of uniaxial tensile test on T E N S A R geogrid . . 156

6.15 Test data and prediction of uniaxial tensile test on C O N W E D geogrid . 156

6.16 Test data and predictions of pull-out tests with T E N S A R geogrid . . . . 157

6.17 Test data and predictions of pull-out tests with CO A W E D geogrid . . . 158
7.1 Configuration of Denver Wall (After Wu, 1 9 9 2 )......................................... 162

7.2 Test data and prediction by H i S S <5t model for the granular backfill: \J J id
vs. v/ ^ d ........................................................................................................ 165
7.3 Test data and prediction by H i S S 8x model for the granular backfill: —I x
vs. I 2D ........................................................................................................ 166
7.4 Load-deformation behavior of geotextile reinforcement for Denver Wall
(After Wu, 1992)........................................................................................... 167

7.5 Test data and prediction of uniaxial tensile test on geotextile .................. 168

7.6 Attachment of plywood boards to timber logs (After Wu, 1 9 9 2 )............... 169

7.7 Bending test configuration of a typical timber facing (After Wu, 1992) . . 170

7.8 Load vs. displacement curve the timber facing unit (After Wu, 1992) . . . 171

7.9 Structural model of bending test on wood facing........................................ 171

7.10 Stress-strain curve of the wood facing (After Ling and Tatsuoka,1992) . 172

xi

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
7.11 Direct shear test on geotextile-sand in te rfa c e s .............................................173

7.12 FEM mesh for Denver W all............................................................................176

7.13 Facing movement of Denver W all.................................................................. 177

7.14 Top surface displacement profile of Denver W a ll.........................................178

7.15 Axial strain in reinforcement: Height = 0.1 5 H ............................................ 179

7.16 Axial strain in reinforcement: Height = 0 .5 2 H ............................................ 180

7.17 Axial strain in reinforcement: Height = 0 .8 8 H ............................................ 181

7.18 Maximum facing and top surface movement with su rc h arg e ......................182

7.19 Horizontal stress distribution at 103 kP a surcharge .................................. 183

7.20 Horizontal stress distribution at 186 k P a surcharge .................................. 184

7.21 Vertical stress contours at 103 kPa surcharge pressure............................... 185

7.22 Vertical stress contours at 186 kPa surcharge pressure...............................186

7.23 Vertical stress distribution at the base of the wall at 103 kPa surcharge . 187

7.24 Vertical stress distribution at the base of the wall at 186 kPa surcharge . 188

7.25 Field observation of vertical stress at base (After Ho, 1 9 9 3 ) ................... 189

7.26 Axial load in reinforcements: 103 kPa surcharge ..................................... 190

7.27 Axial load in reinforcements: 186 kP a surcharge ..................................... 191

7.28 Plastic failure of soil elements at 103 kP a surcharge...............................192

7.29 Plastic failure of soil elements at 186 kP a surcharge...............................193

8 .1 R M C retaining wall test facility (After Bathurst et al., 1 9 9 3 )..................... 195

8.2 R M C propped panel wall test (After Bathurst et al., 1 9 9 2 ) ........................ 196

8.3 Test data and prediction by H i S S 8\ model for the granular backfill: y/J^D
VS. y / l ^ D ............................................................................................................199

8.4 Test data and prediction by H i S S 5\ model for the granular backfill: —I\
VS. y j I i D ........................................................................................................... 200

8.5 100 hour isochronous load-strain behavior of geogrid reinforcement (After


Karpurapu and Bathurst, 1995)....................................................................... 201

xii

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
8 .6 Test data and predictions of direct shear tests: (a) Shear stress vs. shear
displacement and (b) Normal displacement vs. shear displacement . . . . 204

8.7 FEM mesh for R M C W all............................................................................ 205

8 .8 Facing movement of R M C W all...................................................................207

8.9 Facing movement at mid-height with surcharge.......................................... 208

8.10 Axial strain in geogrid, layer 1 ......................................................................209

8.11 Axial strain in geogrid, layer 2 ......................................................................210

8.12 Axial strain in geogrid, layer 3 ......................................................................211

8.13 Axial strain in geogrid, layer 4 ......................................................................212

8.14 Horizontal and vertical loads at t o e ............................................................... 213

8.15 Horizontal stress distribution at 38 kPa surcharge pressure ....................... 214

8.16 Horizontal stress distribution at 80kPa surcharge pressure ........................ 215

8.17 Vertical stress distribution at 38 kPa surcharge p ressu re............................. 216

8.18 Vertical stress distribution at 80 kPa surcharge p ressu re............................. 217

8.19 Vertical stress contours at 38 kPa surcharge pressure ................................ 218

8.20 Vertical stress contours at 80 kPa surcharge pressure ................................ 219

8.21 Vertical stress distribution on the foundation at 38 kPa surcharge . . . . 221

8.22 Vertical stress distribution on the foundation at 80 kPa surcharge . . . . 222

8.23 Axial load in reinforcements: 38 kPa surcharge......................................... 223

8.24 Axial load in reinforcements: 80 kPa surcharge......................................... 224

8.25 Plastic failure of soil elements at 38 kPa su rc h a rg e ................................... 225

8.26 Plastic failure of soil elements at 80 kPa su rc h a rg e ................................... 226

B. 1 Facing movement of Denver Wall (von Mises model for reinforcements,


no interfaces) ................................................................................................. 249

B.2 Top surface displacement profile of Denver Wall (von Mises model for
reinforcements, no interfaces)........................................................................250

B.3 Axial strain in reinforcement: Height = 0.15H (von Mises model for rein­
forcements, no interfaces) .............................................................................. 251

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
B.4 Axial strain in reinforcement: Height = 0.52H (von Mises model for rein­
forcements, no interfaces).............................................................................. 252

B.5 Axial strain in reinforcement: Height = 0.88H (von Mises model for rein­
forcements, no interfaces).............................................................................. 253

B.6 Facing movement of Denver Wall (Hyperbolic model for reinforcements,


no interfaces) .............................................................................................. 254

B.7 Top surface displacement profile of Denver Wall (Hyperbolic model for
reinforcements, no interfaces)..................................................................... 255

B.8 Axial strain in reinforcement: Height = 0.15H (Hyperbolic model for re­
inforcements, no interfaces)........................................................................ 256

B.9 Axial strain in reinforcement: Height = 0.52H (Hyperbolic model for re­
inforcements, no interfaces)........................................................................... 257

B.10 Axial strain in reinforcement: Height = 0.88H (Hyperbolic model for


reinforcements, no in te rfac es)..................................................................... 258

B. 11 Facing movement of Denver Wall (Hyperbolic model for reinforcements,


Goodman element for interfaces)..................................................................259

B. 12 Top surface displacement profile of Denver Wall (Hyperbolic model for


reinforcements, Goodman element for in te rfac es)......................................260

B.13 Axial strain in reinforcement: Height = 0.15H (Hyperbolic model for


reinforcements, Goodman element for in te rfa c e s).....................................261

B.14 Axial strain in reinforcement: Height = 0.52H (Hyperbolic model for


reinforcements, Goodman element for in te rfac es)......................................262

B.15 Axial strain in reinforcement: Height = 0.88H (Hyperbolic model for


reinforcements, Goodman element for in te rfac es)......................................263

C.l Facing movement of R M C Wall (E = .24 x 10- 7 kPa facing stiffness) . 265

C.2 Facing movement at mid-height with surcharge (E = .24 x 10“ 7 kPa


facing stiffn e ss)..............................................................................................266

C.3 Axial strain in geogrid, layer 1 (E = .24 x 10~ 7 kPa facing stiffness) . . 267

C.4 Axial strain in geogrid, layer 2 (E = .24 x 10- 7 kP a facing stiffness) . . 268

C.5 Axial strain in geogrid, layer 3 (E = .24 x 10- 7 kP a facing stiffness) . . 269

C .6 Axial strain in geogrid, layer 4 {E —.24 x 10- 7 kP a facing stiffness) . . 270

C.7 Horizontal and vertical loads at toe (E = .24 x 10~ 7 kPa facing stiffness) 271

xiv

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
ABSTRACT
Geosynthetics are widely used as reinforcements in various earth structures. Current
design methods are based on some simplified assumptions and are primarily modified
versions of limit equilibrium methods used in design of the unreinforced earth structures.
All of these design methods are conservative and result in uneconomical design.

The best method for verifying the design assumptions and studying the stress-deformation

behavior of geosynthetic reinforced earth structures is by constructing a series of large


scale test walls with adequate instrumentation and collecting data from them. But the

cost involved in such a scheme precludes undertaking this kind of study. Thus Numerical

simulation provides an alternative and cost effective means for such a study.

In the present research, a finite element model has been established for the numeri­

cal simulation of geosynthetic reinforced earth structures. The H i S S model has been
implemented in the finite element method for modelling granular backfill. A new calibra­
tion technique has been developed based on genetic algorithms (G A ). The new technique

allows one to calibrate a constitutive model even when all the test data needed for cal­
ibration are not available. An elasto-plastic constitutive model based on the disturbed

state concept (DSC) has been developed and implemented in the finite element method to

model soil/geosynthetic interfaces. The interface model has been validated by simulating
the large scale pull-out tests performed at the Louisiana Transportation Research Ceneter

(L T R C ), and the finite element model of the geosynthetic reinforced wall has been vali­
dated with the observed behavior of two large scale test walls which were constructed and

tested at the University of Colorado at Denver and the Royal Military College of Canada.
The finite element simulation is able to predict the measured behavior of these walls well

xv

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
and the results of the simulation show some discrepancies and conservativeness in the as­

sumptions made in the current design methods. The present finite element model can be
used for parametric study and for formulating a realistic design method.

xvi

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
CHAPTER 1
INTRODUCTION

1.1 General
Geosynthetics have been used in civil engineering applications since the mid 1960’s. Ini­

tially, they were used in low risk engineering work and to facilitate construction work
under difficult environmental conditions. With the success of their applications in low risk

engineering work based on simple empirical design methods and the advancement of poly­

meric science, geosynthetics were used in critical engineering applications. At present,

geosynthetics are used extensively as a reinforcement in various earth structures such as


retaining walls, embankments, levees, etc., replacing metal strips which were common

reinforcement elements in a reinforced earth structure before the advent of geosynthetic

materials.

However, the present practice for designing such structures is based on simplified as­

sumptions and design methods. Most of these design methods are modified versions of

the limit equilibrium method used for designing unreinforced embankments and retain­
ing walls. Provisions are made to take into account the effect of the reinforcement In
general, such design methods inherit some degree of uncertainty regarding failure mecha­
nism, and to lessen the effect of the uncertainty, a large factor of safety is used. This large

factor safety makes these design methods conservative which results in an increase in the
costs of reinforced soil structures. These design methods also do not provide information
regarding the stress-deformation behavior of such structures. They also fail to address

the interactions among different elements of a reinforced earth structure, such as facing,
reinforcement, foundation, backfill, etc. (Figure 1.1).

The best way of verifying the assumptions and the design methods and of understand­
ing the stress- deformation behavior is to construct and monitor a series of large scale field

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
2

backfill

retained soil
wail
facing

reinforcement

foundation

Figure 1.1: A schematic of reinforced earth wall

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
3

test walls. Unfortunately, the cost of performing and adequately monitoring a sufficiently

large number of full scale wall tests is so large that it is neither feasible nor practiced to
perform a detailed experimental study. An alternative way of carrying out such a study

inexpensively is by numerical simulation using the finite element method. The numeri­
cal simulation also helps in formulating a reliable and economical design method and in

understanding the stress-deformation behavior of such structures. It also allows the study

of stresses and strains everywhere in the structure, whereas in an experimental study of

the wall, stresses and strains can be measured at certain points from a limited number of
instruments.

In such a numerical simulation, the constitutive models which describe the stress-strain

relationships of various constituent materials of a reinforced earth structure, play the most

important role, and the reliability of the numerical simulation depends on how accurately
these constitutive models capture the behavior of respective materials, and also the algo­

rithm used to implement them into the finite element method. The current trend is to model

the granular backfill of a reinforced soil structure using the hyperbolic model (Duncan and

Chang, 1970) or Mohr-Coulomb model. But both of these models are incapable of cap­

turing all the essential characteristics of soil. Another important aspect of the numerical

simulation of a reinforced soil-structure is the modeling of the reinforcement-soil inter­


face. Another current trend is to model such an interface with the zero thickness element

(Goodman et al., 1968) which was originally developed for rock joints. However, such
elements have a kinematic deficiency caused by large off-diagonal terms in the element

stiffness matrix. This type of interface element describes the shear behavior of interface
in an approximate sense, and it does not properly consider the normal response of the
interface. Thus it ignores the coupling of the shear and normal behaviors of interfaces.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
4

1.2 Proposed research


The present research focuses on developing a finite element model for the numerical simu­

lation of geosynthetic reinforced earth structures. The H i S S 6Xmodel (Desai et al.,1986;

Frantziskonis et al.,1986) has been used for granular backfill. The H i S S 5X is a modem
but relatively simple elasto-plastic constitutive model which captures all the essential fea­
tures of granular soil behavior. A constitutive model based on the disturbed state concept

(DSC) [Desai, 1989] has been developed for modeling interface behavior. F E M imple­
mentation of the interface constitutive model has been done using a numerically stable thin
layer element (Sharma and Desai, 1992). The model is capable of representing most of the

important characteristics of the interface response such as dilation, strain hardening and
strain softening. The interface model has been verified with the direct shear test results

of geogrid-sand interfaces. Five pull-out tests with two types of geogrids have been simu­

lated in F E M using the model. Finally, two full-scale laboratory test walls (1) the Denver
wall (Wu, 1992), and (2) the Royal Military College (RMC ) wall (Bathurst et al., 1992)

have been simulated to verify the effectiveness of the constitutive models and to validate

the finite element model.

1.3 Methodology
All the material models used in the present study have been implemented in the com­
mercial software ABAQUS, and this software has been used for the finite element anl-

ysis. A B A Q U S (HKS,Inc., 1996) has the options for using user defined subroutines for
material models, elements, initial stresses, and initial solution dependent variables. The
methodology of the proposed research is as follows:

Task 1: Implementation of the H i S S Sx constitutive model in A B A Q U S using different


algorithms and choosing the best algorithm based on their performances.

Task 2: Development of a constitutive model based on the disturbed state concept (DSC)
to describe interface behavior. Verification of the model with the laboratory test data.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
5

Task 3: Implementation of the D S C constitutive model in the finite element method in


association with the thin layer element

Task 4: Validation of the D S C model in the finite element method by simulation of large
scale pull-out tests and by comparing finite element results with the test data.

Task 5: Validation of the finite element model by simulating large scale geosynthetic re­
inforced laboratory test walls.

1.4 Organization of various chapters


After the introduction, which is included in Chapter 1, the background of the study and

the relevant literature review are described in Chapter 2. Chapter 3 gives an overview of
the H i S S model followed by a detailed description of its implementation in F E M .

Chapter 4 describes a new technique developed based on the genetic algorithm to calibrate

the H i S S 5i model. Chapter 5 describes the constitutive model based on the disturbed
state concept (DSC) developed for sand-geosynthetic interfaces and its implementation in

F E M . The model verification and its application in simulating pull-out tests are described
in Chapter 6 . Chapter 7 contains the results of the Denver wall simulation, and the results

of the R M C wall simulation are presented in Chapter 8 . Chapter 9 includes a summary

and conclusions.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
CHAPTER 2
BACKGROUND OF THE STUDY

2.1 General
In this chapter, a review of relevant literature concerning reinforced earth structures has

been presented to explain the background of the study. The discussion begins with a review

of the current design methods in practice for designing geosynthetic reinforced earth struc­

tures. It is then followed by a discussion on the limitations of the current design methods.

The subsequent discussion includes how a numerical simulation using F E M could help

in understanding the stress deformation behavior of geosynthetic reinforced earth struc­


tures. The defficiencies of the existing finite element simulations are also pointed out. The

chapter is concluded with a discussion on the need for reliable constitutive models for the

numerical simulation of geosynthetic reinforced earth structures.

2.2 Current design methods


Current design methods in practice for designing reinforced soil structures can be grouped
into two major categories: ( 1 ) limit equilibrium based methods and (2 ) strain compatibility

based methods. The limit equilibrium based methods can be grouped into two categories.
Methods in the first category use force equilibrium analysis where the horizontal forces

caused by lateral earth pressure are equilibriated with the tensile forces in the horizontal

reinforcements. The second category uses an approach similar to the conventional slope
stability analysis. These methods evaluate the force and/or moment equilibrium on an

assumed failure surface, taking care of the balancing force/moment provided by the rein­

forcement
Clayboum and Wu (1993) conducted a comprehensive evaluation and comparison of
six design methods based on the limit equilibrium approach. The six design methods are:

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
7

1. Forest Service method (Steward et aIM1977 (revised in 1983))

2. Broms method (Broms, 1978)

3. Collin method (Collin, 1986)

4. Bonaparte et al. method (Bonaparte et al., 1987)

5. Leshchinsky and Perry method (Leshchinsky and Perry, 1987)

6 . Schmertmann et al. method (Schmertmann et al., 1987)

The assumptions made in these methods regarding stress distributions, failure surfaces,

safety factors, and the inclination of the reinforcement at the failure surface vary from

each other. Methods 1 through 4 fall into the first category, i.e. force equilibrium analysis,
where the horizontal forces caused by lateral earth pressure are balanced by the tensile
forces in the horizontal reinforcement. The stresses considered in these methods are the
vertical and horizontal stresses in soil, the horizontal tensile stresses in the reinforcement,

and the horizontal resistance to pull-out provided by the portion of the reinforcement be­

hind a potential failure surface. Two independent factors of safety are used for each layer
of reinforcement: the factor of safety for reinforcement rupture, and the factor of safety

for pull-out The soil/reinforcement interface friction in the portion of the reinforcement
behind the failure surface provide the resistance against pull-out.

All these methods assume a failure plane through the reinforced soil mass which is

same as the failure plane in the Rankine active failure condition. The stresses on the failure

plane are not analyzed. The reinforcements extended beyond the assumed failure surface

are considered tension-resistant tiebacks for the assumed failure wedge. For this reason,
these methods are frequently referred to as “tied-back wedge" methods. Although there
are many similarities in these methods, they assume different earth pressure distributions

to be resisted as shown in Figure 1.1 (Clayboum and Wu, 1991). In addition to these

four tied-back wedge methods, there are some other methods, such as the Jewel method

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
8

(q=0 )

^ = 0.65^ ( 1.57 + 7 ^)

FOREST SERVICE BROMS

&h.{kPa) = 15.722(m)

(q=0 )
(Th.{kPa) = 3.14H(m)

<rK(kPa) = 2.38 H (m )

COLLIN (GEOTEXTILE) COLLIN (GEOGRID)

<7/i = Lateral earth pressure


H = Height o f wall
K~(yz+<i) 7 = Backfill unit weight
t . z+3q)(z/L)*
1"1" 3 ( 7f*+,) 7 = Surcharge at top o f wall
/Co = 1 —sin $
K a = tan2(45° — <p/2)
<j>= Friction angle o f backfill soil
BONAPARTE ET AL. z = Depth below top o f wall

Figure 2.1: Earth pressure distributions used in tied-back wedge methods (After Clayboum
and Wu, 1991)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
9

(Jewel, 1987), the Coherent Gravity method (Schlosser, 1990), the TC method (Tensar

Corp., 1986), the Murray method (Murray, 1980), the Simac et al. method (Simac et al.,
1990), which fall into this category.

Methods 5 and 6 use an approach similar to the conventional slope stability analysis

with some modifications to account for reinforcements. The Leshchinsky-Perry (1987)

approach is based on the variational limit equilibrium analysis of slope stability (Baker and

Garber, 1977). They consider only the log-spiral slip surfaces passing through the toe and

analyze the rotational failure mode. The soil is assumed to obey the Mohr-Coulomb failure
criterion, and each reinforcement sheet is assumed to develop tensile resistance of a known

magnitude. They also assume that at the instant of failure, the direction of tensile forces

in the reinforcement is inclined with respect to the horizontal, i.e. the reinforcements are

orthogonal to the radius of the log-spiral at their intersection. The Schmertmann et al.

method uses wedge failure models based on limiting equilibrium analysis. Both straight

line and bi-linear wedges are used in the analyses. The results of the wedge analysis were

modified by using extended versions of Bishop’s modified method and Spencer’s method

of slope stability analysis. The force in each layer of reinforcement is assumed to be

equal, which requires reinforcement spacing which decreases with depth. The surcharge

is assumed to be equivalent to some additional fill height

Sawicki and Lesniewska (1991) proposed a design method for reinforced soil based

on the plastic limit analysis. In this method the reinforced soil is treated as macroscopi-

cally homogeneous and anisotropic material, the gross behavior of which depends on the

mechanical properties and interactive contributions of the soil and the reinforcement A

yield condition for reinforced soil is proposed, and an associated flow rule is assumed.

With the help of the yield criterion, the equilibrium equations are solved by the method of

characteristics for given geometry of the slope and respective boundary conditions.

In addition to the abovementioned methods, a number of researchers have proposed


design methods based on strain compatibility (Gourc et al.,1990; Jewell and Milligan,

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
10

1989; Juran et al., 1990). Gourc et al. (1986) used the concept of an anchored membrane

to determine the inclination of the reinforcement at the failure surface to estimate the
tension in the reinforcement at different strain levels. Juran et al. (1990) extended this

analytical approach by including several factors, such as soil constitutive relationships,

reinforcement stress-strain relationships, soil/reinforcement interaction, strain path, and

construction effects on initial strains. Recently, Ehrlich and Mitchell (1994) proposed a

working stress design method based on a strain compatibility analysis. In this method, the

effects of reinforcement stiffness, relative movement between the soil and reinforcements,
and compaction-induced stresses were considered.

2.3 Limitations of the current design methods


The limitations of all of these design methods are that they are based on some simpli­
fied assumptions about the failure surface and maximum horizontal stress distribution on

which the maximum reinforcement tensions depend. Generally, these assumptions do not

consider explicitly all the factors that influence the horizontal earth stress in a reinforced

soil system. More importantly, the design methods mentioned above do not provide in­

formation concerning deformations or stress distributions in the reinforcement and soil.

These methods also do not consider the effects of wall facing and the foundation. These

conservative design methods increase costs of reinforced soil-structures. Moreover, they


predict embankment failure poorly. For example, limit equilibrium methods underesti­

mated the failure of two large scale geosynthetic reinforced soil walls using granular and
cohesive backfills which were tested at the University of Colorado, Denver (Wu, 1992).

In addition, research on some instrumented field reinforced soil walls show that there

are some discrepancies between the actual stress distributions in the reinforced soil wall

and the assumed stress distributions in the design. The data from an instrumented rein­

forced earth wall using Tensar geogrids (Fishman et al., 1993) show that the measured re­

sponse of the wall system (particularly near the wall face, in terms of stresses, strains, and

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
11

displacements) indicates nonlinear behavior affected by interaction effects such as stress

transfer, relative motions, and arching. The data also show that there is some amount of
lateral earth pressure against the wall facing which the design method did not predict Fur­

thermore, vertical stress distributions along the geogrids were also found to be nonlinear,
but linear distributions were assumed in the design. Bathurst et al. (1990) performed test­

ing of large scale reinforced soil walls in the laboratory. They found that the distribution

of vertical earth pressures below the reinforced soil mass was uniform except for the first
0.5 m behind the panel facing. At this location there was about a 30% reduction in earth
pressure based on soil self weight and uniform surcharge pressure. This is not reflected in

design assumptions where one of the three types of vertical stress distributions is gener­
ally assumed as shown in Figure 2.2. They also found that the distribution of strain in the

reinforcement at a surcharge load equivalent to about one half of the ultimate capacity of
the composite structure did not reflect the trends observed at failure. This difference in the

strain distributions has an implication in relating mechanisms at failure to wall behavior at


working load conditions. A simple factor of safety approach to relate failure conditions to

working load conditions may not be the right way for this type of structures (Bathurst et
al., 1990).

2.4 Numerical simulation by FEM


In addition to the limitations of the current design methods, i.e. they are based on some

simplified assumptions about the failure surface and maximum horizontal stress distribu­
tion, these methods also do not provide information concerning deformation or stress dis­
tributions in the reinforcement and soil. To understand these behaviors, some researchers

have started analyzing reinforced soil structures using the finite element method (F E M )

(e.g., Seed et al., 1990; Adib et al., 1990; Chew et al., 1990; Bathurst et al., 1992 & 1995;

Ho and Rowe, 1993 & 1994; Lee, 1993; Chou, 1992). A F E M analysis provides informa­
tion not only for the purposes of design, but also for use in developing/validating simple

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
12

T ? T T T T V V
Uniform

cr„ = 7 H+q

Trapezoidal
= (yH + tf) + K ofa H + 3q)(H /L)2

Vvimax) = iriH + g) - ^ra( 7 ^ + 3q)(H /L)2

2e

▼t t t t t t
Meyerhof
_ iH+g___
0 «(mai) - l-((/fa(7H4-3g)(/f/L)i )/3(7H+7))

e = (KaH 2( jH + 3q))/(6L('yH + q))

Figure 2.2: Vertical stress distributions at the base of the wall

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
13

design methods involving only hand calculations. However, the finite element method also
needs to be validated against the observed behavior.

One such validation has been done by Ho and Rowe (1994), who used F E M to an­

alyze two model soil walls, one constructed with non-woven geotextile, the other con­

structed with geogrid, and both tested in a centrifuge. They used the Mohr-Coulomb

model for the granular backfill and the interfaces have been modelled by springs with dif­

ferent normal and shear stiffnesses. The model tests were performed and reported by Jaber

(1989). Ho and Rowe found that the numerical model was successful in predicting most

features of the experimental data. This finding supports the credibility of this tool in the

investigation of reinforced soil structures. A similar result was furnished by Bathurst et

al.(1992). They used F E M to analyze a full scale geogrid reinforced soil wall tested at the
Royal Military College, Canada (Bathurst, 1990) using the hyperbolic model (Duncan and

Chang, 1970) for the granular backfill and the Mohr-Coulomb model for the interfaces.

They found that the numerical simulation accurately predicted the important quantitative

features of the wall at failure, including panel movement, location and magnitude of peak

strains in reinforcement layers, and location of internal failure surface. Lee (1993) used a

hypo-elastic constitutive model for granular backfill and validated the finite element model

against observed behavior of the Seattle Welded Wire Retaining Wall and the Denver Wall
with granular backfill.

The numerical simulation of a geosynthetics reinforced soil structure is also useful for
a parametric study (Ho and Rowe, 1993 & 1994). Chou did some parametric study for

the reinforced earth wall with cohesive backfill. He uses the Sekiguchi-Ohta model (1977)

and validated his finite element model against the observed behavior of the Denver Wall

with cohesive backfill. The parametric study helps in understanding the effects of various

components like wall facing, soil/facing interface friction behavior, reinforcement connec­

tion with the facing, and tensile strength of the reinforcement, in the overall behavior of

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
14

the reinforced soil composite. It also helps in performing sensitivity studies regarding key

assumptions adopted in a numerical model (Ho and Rowe, 1994).

From the above discussion, it is evident that the numerical simulation of reinforced

soil structures using the finite element method can successfully capture stress-deformation

behavior. The accuracy of the finite element analysis of a reinforced soil structure de­

pends on various factors. The most important ones are the constitutive model used for

soil, soil/reinforcement, and soil/facing wall interfaces. So far, most of the researchers

have used the hyperbolic model (Duncan and Chang, 1970) or the modified version of
this model to capture the dilation of the granular backfill. Others have used the Mohr-

Coulomb yield criterion based on elastic-perfecdy plastic theory. Some of them modified

it and used the nonassociative flow rule to capture the shear dilation. None of these models,

however, are “advanced" constitutive models (Duncan, 1994) which are based on elasto-

plastic theory with hardening. The hyperbolic model, which relates strain increments to

stress increments through the extended Hooke’s Law, has been fundamentally an elas­

tic stress-strain relationship though the model simulates nonlinear behavior, irrecoverable

strains, and stress dependent stiffness. This model can predict irrecoverable deformations,

or “inelastic" behavior by using different values of modulus for virgin loading, unloading

and reloading; but it cannot model the behavior of real soils near failure. The theory of

plasticity models the soil behavior at failure and after failure more realistically than the

theory of elasticity because it models a very important aspect of the stress-strain behavior
of real soils: Plastic strain occurs in the direction of stresses rather than the direction of

stress increments (Duncan, 1994). Though the Mohr-Coulomb model has the capability to

capture soil behavior at failure, it overestimates dilation. Also, this model is incapable of

modeling K 0-consolidation.

Another important aspect of numerical simulation is the modeling of interfaces be­

tween soil and reinforcement and between soil and wall. The zero thickness Goodman

element (Goodman et al., 1968) has been used to model these interfaces in most research

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
15

(Mird and Kwok, 1989; Bathurst et al., 1992; Ho and Rowe, 1994). The shear behavior

has been defined by linear/non-linear elastic models, and the onset of sliding has been
defined using a limiting shear stress criterion as Mohr-Coulomb. The normal behavior

has usually been defined empirically for the bonded and debonded states. Tavassoli and
Bakeer (1994) have used the hyperbolic model (Clough and Duncan, 1971) to model the

material of the interface. They assumed that the bond strength of the interface was gov­

erned by the Mohr-Coulomb criterion, and once the bond strength is exceeded, the shear

stiffness of the interface was automatically reduced to a value close to zero to model a
slip condition. Matsui and San (1992) have used interface elements for the prediction of

the “Denver Wall" based on restrained dilation derived from elasto-plastic material be­
havior. They have been successful in predicting the behavior of the “Denver Wall", with

the granular backfill very close to the observed behavior. Gens et al. (1993) have used

a non-linear elastic relationship for defining normal stress behavior and an elasto-plastic

hardening model for defining shear stress behavior.

However, the Goodman element possesses a kinematic deficiency due to the specific
form of the element stiffness matrix. As a result, when this element is subjected to a tan­

gential force applied at one of the four nodes in a single element “patch", the remaining

free node displaces in a direction opposite to the first. This response leads to unreasonable
tangential force oscillations under simulated pull-out conditions (Li and Kaliakin, 1994).

Other researchers tried to model the interface behavior in a gross way by putting normal
and shear springs between the nodes of two materials originally occupying the same posi­
tion to accommodate slip between them.

2.5 Constitutve models for soils and interfaces


In the present research, the hierarchical single surface (H iS S ) Si constitutive model de­

veloped by Desai and co-workers [Desai (1980), Desai et al. (1986)] has been used for
modelling granular backfill. The reasons for choosing the H iS S Si constitutive model

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
16

are that it is simple, requires fewer material parameters, and has a single surface which

is easier for computer programming. The model is capable of capturing all the important
features of soil behavior. The other popular model which has been based on elasto-plastic

theory and which has been used for granular material is the Cap model. The Cap model,

being a volumetric hardening model, shows dilation only near the failure. In real material,

dilation occurs much sooner than the failure. The H iS S Si model, which is nonassocia-

tive, hardens depending on the total plastic strain and, thus, captures dilation properly.
Modeling reinforcement/soil interface behavior is one of the most important aspects

of the numerical simulation of reinforced soil structures. The present trend is to model
such interfaces using the zero thickness element (Goodman et al., 1968). However, such

an element, as mentioned earlier, suffers ill conditioning due to large off-diagonal terms in

the stiffness matrix (Wilson, 1977). Ghaboussi et al. (1973) and Wilson (1977) advocate
the use of relative displacement as a nodal variable to avoid such ill conditioning. But use

of relative displacement as a nodal variable requires modification of the adjacent elements

so that they use the same nodal variable. This makes incorporation of such elements
into a finite element program more complex. Desai et al. (1979, 1986) have found that

although the zero thickness element provides satisfactory prediction of shear behavior, it

does not provide the realistic behavior of normal stress in soil-structure interaction and

joint problems. Zienkiewicz et al. (1970) proposed using a solid element as an interface
element Howeverjiot many critical and systematic studies and implementations of the

concept are available in the published literature. In recent times, Desai et al. (1984) and
Sharma and Desai (1992) have proposed a thin 2D joint/interface element called a thin
layer element for modeling various interfaces. The thin layer element has been formulated

as a continuum element; but its constitutive response has been defined differently from

that of the adjoining solid elements. The constitutive response has been defined only
in terms of the normal and shear components of the behavior, based on shear tests on

planar interfaces or joints. It is also shown (Sharma and Desai, 1992) that the thin layer

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
17

formulation includes the zero thickness formulation as a special case as the thickness,

t —> 0. One of the major improvements provided by the thin layer element is that it

provides consistent and satisfactory computations of stresses in the interfaces themselves,

which are very difficult to obtain with the zero thickness element.

In the present research, the thin layer element has been used to model various inter­

faces. Since the formulation of the thin layer element is essentially the same as other

solid elements, it is easier to implement into a computer program. For the representation

of interface behavior, hyperbolic models and elasto-perfectly plastic models with Mohr-
Coulomb friction law have been used by many investigators. The shortcomings of these

models are that they describe shear behavior of interfaces in an approximate sense and

that they do not properly consider the normal response of the interface. The coupling of

the shear and normal behavior of interfaces is also ignored. These models do not consider

the dilating behavior of interfaces. In the present research a constitutive model based on

the disturbed state concept (D SC) has been developed for describing interface behavior.

The motivation for developing a DSC-based model is derived from the success of some

researchers in modelling rock joints and interfaces based on the D S C concept (D e sai and

Ma, 1992; Ma, 1990). DSC is based on the idea that the response of a material can be

related to and expressed as the responses of the reference states (Desai, 1989; Desai et al.,

1990). As a result, the observed behavior of a material is thus treated as the disturbance

to the behavior of the reference states. This disturbance can be viewed as a change in the
physical properties such as density and the structure of the material, which is analogous

to the concept of damage caused by micro-cracking and fracture. Non-associativeness,

anisotropy, and strain softening can be expressed as disturbances with respect to the refer­

ence states.

In the D SC approach the observed material behavior is considered to be composed

of two reference states called intact state and disturbed state. The intact state behaves

as a continuously hardening material without disturbance. The disturbed state may be

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
18

assumed to be an invariant state during the stress deformation process such as the ulti­

mate, failure, and zero shear stress and/or volume change condition. The progressive dam­
age model for concrete developed by Frantziskonis and Desai (Frantziskonis and Desai,

1987; Frantziskonis, 1986) can be treated as a special case of the disturbed state modeling
method. Wathugala and Desai (1987 & 1988) have proposed the use of the critical state as

the disturbed or fully adjusted state for granular materials. The disturbed state modelling
method has been used by Armaleh (1990) and Katti (1991) for modelling sand and clay re­

spectively. This approach has also been used to model other materials, such as electronic
packaging (Basaran, 1994). The proposed model is a specialized version of the model
proposed by Wathugala and Desai (1987 & 1989) for granular materials. The constitutive

model based on D S C can allow proper modeling of shear transfer, volumetric behavior

and localized slip in the interface zone. In F E M the new elasto-plastic constitutive model

will be implemented with the thin layer element described earlier.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
CHAPTER 3
HiSS CONSTITUTIVE MODELS AND ITS
IMPLEMENTATION IN FEM

3.1 Introduction
The stress-strain behavior of soils is nonlinear in nature. Unlike metals, the volumetric
behavior of soils show shear dilation, a phenomenon observed in the behavior of most ge­

ologic materials. These characteristics of soils make all the advanced constitutive models

developed for them very complex. The complexity of these constitutive models prevents
development of analytical solutions for boundary value problems. To be useful in solving

practical problems, these models need to be implemented into numerical solution tech­
niques. Thus, developing efficient and robust algorithms for implementation of constitu­

tive models in computer procedures is very important.

Simulation of stress-controlled laboratory tests and the stress-based finite element

method (F E M ) requires an algorithm to integrate the constitutive model, which provides

the incremental strain corresponding to a given stress increment This algorithm is referred
to here as stress to strain algorithm. Similarly, simulation of strain-controlled laboratory

tests and displacement based F E M requires an algorithm to integrate the constitutive


model which provides a stress increment corresponding to any given strain increment.

This algorithm is referred to as strain to stress algorithm. A very accurate and efficient
stress to strain algorithm which provides a numerically exact solution for stress controlled
laboratory tests has been presented. This numerically exact solution can be used to ana­

lyze the accuracy of the strain to stress algorithm. This allowed us to evaluate the accuracy
of the iterative strain to stress algorithm for different stress paths, which is otherwise not
possible under available methods (Wathugala, 1990).

19

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
20

Most of the existing nonlinear finite element procedures for geotechnical engineering

problems are based on the displacement approach. These procedures require a strain to
stress algorithm for integrating the constitutive model. In this chapter, such procedures are

described for the H iS S family of constitutive models (6q,5\ and £q) for different geologic
materials. Four different algorithms have been implemented for virgin loadings. They are

(1) elastic predictor-plastic corrector, (2) plastic predictor-plastic corrector, (3) implicit

integration and (4) Modified Euler method. Special procedures have been described to
take care of nonlinear unloading and reloading of the H iS S model.
Wathugala (1990) has developed numerical procedures based on the elastic predictor

- plastic corrector, and the plastic predictor-plastic corrector methods to implement the
H iS S SQ, 6 1 , Sq, and 5%models in the finite element method that have a number of new

features and incorporate the best properties of available methods such as the subincrement
method (Faruque and Desai, 1985), the drift correction procedure (Potts and Gens, 1985)

and the elastic predictor plastic corrector method (Ortiz and Simo, 1986). Here, these

algorithms have been further improved and implemented in A B A Q U S in addition to the


implicit integration method and the Modified Euler method. These algorithms are more

general in nature, and includes special procedures to improve robustness, accuracy, and
efficiency.

The procedures have been verified by evaluating stresses from strains for three different

triaxial tests. All the procedures have been incorporated into the commercial F E M pack­
age ABAQ U S, and two footing problems, (1) rigid and (2) flexible, have been analyzed
using F E M . The accuracy and the efficiency of these procedures have been compared.

Most of the derivations in this chapter only assume a stress-space-plasticity based consti­
tutive model. However, specific comments related to H iS S models are given wherever

appropriate.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
21

3.2 HiSS constitutive models


The H iS S constitutive models are elasto-plastic constitutive models, and they share the

same yield surface. The yield surface F is defined in terms of stress invariants J\, the first

invariant of the stress tensor, cry, JiD, the second invariant of the deviatoric stress tensor,
and J zdj the third invariant of the deviatoric stress tensor as

F s (jr)~ F >F - =0 <31>


where pa is the atmospheric pressure, and

where aps is the hardening or growth function. The function Fb describes the shape of the

yield surface in Ji — y/J^D space. 7 and n are material parameters. The parameter n (2 <

n < 00) is related to the phase change point, where material changes from contractive
to dilative behavior. The function, Fs, describes the shape of the yield surface in the
octahedral plane and is given by

Fs = (1 - 0 S r)m (3.3)

where Sr is defined as a stress ratio, and given by the following equation

v/27
Sr = - J zd J-id (3-4-)

P (0 < P < 0.77) and m (= -0.5) are material parameters. The shape of the yield surface

for a typical soil in different stress spaces is shown in Figures 3.1 to 3.3. When a ps = 0,

the yield surface becomes the ultimate surface, which envelops all the yield surfaces.
The hardening function OpS can be a function of various internal variables related to
the plastic deformations. The hardening function of 50 and ^ models is defined as:

^ (3-5)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
22

30.0

Ultimate Surface

20.0
Phase Change Line

Yield Surfaces, F
Q

>
10.0

0.0
0.0 10.0 20.0 30.0 40.0 50.0 60.0 70.0
J \, p si

Figure 3.1: Shape of yield surfaces in Ji - \ / J id plane

whereas the hardening function for the 6g model is defined as:

h
a p s ~
( <fvr + h 3^ j

where a\, rji, hi, h.2 , h3and h 4 are material parameters. For clays,h3 is equal tozero. The
incrementsof trajectoriesof total, volumetric, and deviatoricplasticstrains (fy and £d)
are defined as:

= (* ? ■ & ? ) i (3.7)

<*?d = (<*<$*?,)* (3.8)

and

d tv = (d^y) (3.9)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
23

250.0

200.0

150.0

100.0

50.0 Yield Surface


HC Line

0.0I ,-------1-------,-------1-------,-------1____ ,____ i_______ i . i


0.0 50.0 100.0 150.0 200.0 250.0 300.0 350.0
V ^o -2 = V 2 a 3

Figure 3.2: Shape of yield surfaces in triaxial plane

where de? is the incremental deviatoric plastic strain tensor, and defined as

(3.10)

day is the incremental volumetric plastic strain due to virgin loading.

For nonassociative plasticity, the yield function and the plastic potential function are
not the same, and in that case, the plastic potential function is defined as:

where otq is defined as:

olq = dps + k (qq - aps) (1 - r„) (3.12)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
24

Figure 3.3: Shape of yield surfaces in octahedral plane

ac is a nonassociative material parameter, a 0 is the value of a ps at the beginning of shear


loading, and r„ = where £ = trajectory of plastic strain and = trajectory of volu­

metric plastic strain. The H iS S 50 and Sq models are associative, but Si is nonassociative,
and the abovementioned potential function is used for this model.

3.3 Incremental stress strain relationship


In this section, general assumptions made about the constitutive models used in the nu­
merical algorithms are presented. The details of the H iS S models can be found in Desai

et al.(1986), Desai and Wathugala (1987), Wathugala (1990), and Wathugala and Desai
(1991a, 1991b, 1993). General incremental stress-strain relationship for any loading may

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
25

be given by

dcfij — C i]kld£kt (3.13)

or

fe ii = D ijkld(Jkl (3.14)

The superscript * can be V L, RL or UL, depending on virgin loading, reloading or


unloading, respectively. and D ^kl are elasto-plastic constitutive stiffness and com­

pliance tensors, respectively. General form of these tensors are given by Wathugala(1990)
as

(3.15)

and
i j n m 'Ln m t'crp^'opkl
(3.16)
H* +
where Cfjk[ and Dfjkl are elastic constitutive stiffness and compliance tensors respec-
tively. and n® are unit normal tensors to reference Goading) surface R and potential

surface Q respectively, and H* is the plastic modulus. For virgin loading, H*(= HVl )

is found from the consistency condition in the theory of plasticity. For nonlinear reload­
ing and unloading as used in the 6q model, H* is found from an interpolation rule. The

reference surface R is defined as

R = R(aij, a r) = 0 (3.17)

where a r is a parameter defined so that the current stress point lies on the R surface. When

the material experiences virgin loading, the yield surface F coincides with R (F = R),

and a r becomes the hardening function aps. When the incremental stress tensor is pointing
outward the reference surface (i.e. (dR/d<Jij)dcrij > 0 ), the material experiences loading

(virgin loading if the current stress point is on the yield surface and reloading otherwise)

and when it is directed inward to the reference surface (i.e. (dR/d<Tij)daij < 0) the
material experiences unloading. The situation when {d R f doij)daij = 0 is defined as the

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
26

neutral loading. The hardening functions a-p, can be defined as a general case as:

OCps — ^D) £v) (3.18)

The potential function is defined as a deviation of the loading function, and in general it is
defined as:

Q = <*q ) (3.19)

where qq is defined as a function of o;r (or a ps in the case of virgin loadings) and some
other history parameters such as induced anisotropy. For associative plasticity models,
potential function is the same as the loading function.

3.4 Stress to strain algorithm


The details of this algorithm can be found in Wathugala (1990), and Wathugala and Desai

(1993). A brief description is given here for completeness. The incremental strain tensor
dek[ can be calculated from Equation 3.14. Since D ^kl is a function of stress, it is not

constant during a stress increment, and according to the mean value theorem, D*jkl should

be calculated at an intermediate point during the stress increment in order to obtain the
correct strain increment. However, the exact point at which it should be calculated depends

on the amount of nonlinearity in D ^kl and is usually not known. For virgin loading,
strain increments calculated using Equation 3.14 would produce a change in the hardening

parameter so that the new yield surface passes through the new stress point. However,

due to the accumulation of errors during different stress increments, the stress point drifts
away from the yield surface. The following algorithm is developed to avoid this situation.

Consider the case where the material experiences virgin loading during the stress in­
crement, dxiij. Let the initial stress point be A and the final stress point be B. The current

yield surface passes through point A, and the next one should pass through point B. From
the consistency condition

j OCpg) — 0 (3.20)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
27

where aps is the hardening parameter and F is the yield function. Equation 3.20 may
be solved for a fs for the given stress of- and may be expressed in the form

< = /(* $ (3-21)

where the function / is derived from the yield function F. Similarly, the hardening
parameter at B, a is given by

< = H<r§) (3-22)

SuperscriptsA and B in Equations 3.21 and 3.22 refer to quantities at stress points A

and B respectively. Now the change in the hardening parameter, daps,may be found from

daps = o £ - a fs (3.23)

Alternatively, daps may be found by differentiating the general expression for the hard­
ening function given in Equation 3.18 as

da* = +^ d tv (3.24)
d£D o^ v

where d f, dfo, and d£v are increments of the trajectories of total,deviatoric and volu­
metric plastic strains, respectively. They are defined in Equations 3.7, 3.8, and 3.9 respec­
tively.

Note that the second and the third terms on the R H S of Equation 3.24 are zero for the

S0 and 6 1 models. Incremental plastic strains, cfe?-, may be given by the flow rule as

cfc?- = dXnfj (3.25)

where dA is a scalar of proportionality and is the unit tensor in the direction of the
incremental plastic strains. It can be defined through a potential function (Equation 3.19)

for nonassociative plasticity, such as the <Ji model. Otherwise yield function can be used

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
28

to compute nfj. Substituting Equations 3.10 and 3.25 in Equations 3.24, 3.7,3.8, and 3.9
yields

doipg do?] dotps < nkk >


daps — + d\ (3.26)
d£v v/3 J
where is the deviatoric part of n®. By combining Equations 3.23 and 3.26, dA can be
expressed as

qB —q A
ps ps
dX = (3.27)
d a ia i aaP>(v Q nQ i 9a„,
ae + d^D \ n D i j n D ij) + gW jf-

Now incremental plastic strains, efc?-, for any given stress increment which results
in the virgin loading may be calculated using Equations 3.25 and 3.27. The incremental
elastic strain tensor, efc?- for the given stress increment may be calculated from the Hooke’s
law as

deekl = Deim d<Jij (3.28)

where Dfjkl is the elastic compliance tensor. Now the total incremental strain, d£ki,
may be found by just adding elastic and plastic components as

feki = deli + depkl (3.29)

The proposed procedure eliminates the yield surface drift and employs direct computa­

tions rather than the iterative methods used by Galagoda(1986). Therefore, it provides ac­
curate and efficient computation of a strain path corresponding to a given stress path. The
correct stress point where the direction of incremental plasticstrain tensor, n$, should be

calculated is not known and could introduce a small error to the computations when large
stress increments are used. It is found that this error is negligible when small stress incre­
ments or small subincrements are used. Therefore, by using small stress increments, it is

possible to obtain an exact strain path for a given stress path using the proposed algorithm.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
29

This numerically exact solution is used to evaluate the accuracy of the iterative strain to
stress algorithms.

3.5 Strain to Stress Algorithm


3.5.1 Introduction

In general, the incremental stress, doij, corresponding to a given strain increment, de^,

may be found from Equation 3.13. However the expression for C*jkl is different for loading
and unloading. Therefore, first we need to identify whether the given strain increment
produces unloading, reloading, or virgin loading, or a combination of them. Though most

of the elasto-plastic constitutive models consider unloading and reloading as linear elastic

like 5o and <5i models, the Sq model used nonlinear behavior for unloading and reloading.

When the material experiences virgin loading, the plastic strains which develop change
the hardening parameters and, in turn, change the yield surface (prestress surface). The

linearising errors in Equation 3.13 accumulate, and the stress point starts to drift away

from the yield surface if Equation 3.13 is used for virgin loading situations. Different
algorithms for the virgin loading cases which do not have this problem are described later
in Section 3.5.3.

3.5.2 Non-Virgin Loading

3.5.2.1 Linear Elastic

Most of the constitutive models used for engineering materials assume linear elastic be­
havior for non-virgin loadings (un/reloadings). SQand Si models (Desai et al., 1986) in

the Hierarchical Single Surface (H iS S ) modelling approach are among them. In this sec­

tion, the strain to stress algorithm for this class of models is described. The first step in
the algorithm is to determine the location of the current stress point with respect to the
current yield surface. It could be inside or on the yield surface. According to the theory of

plasticity, it is not possible to have the stress point out side the yield surface. If the stress

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
30

point is found to be outside the yield surface, then that is due to numerical errors, and it
should be moved back to the yield surface by correcting the stress point or yield surface
or both. Traditionally, the location of the current stress (tr^) point is found by evaluating
the yield function, F. If the stress point is on the yield surface, F(<7 ,y, ap3) = 0, and if it is

inside, F(<Xy, a ps) < 0. When the stress point is on the yield surface, the second step is to

determine whether the strain increment will cause virgin loading or unloading. Figure 3.4

illustrates all the possible types of stress increments. These possibilities and some of their

160.0

120.0

80.0

40.0

0.0
0.0 40.0 80.0 120.0 160.0 200.0
\ / 2 cr2 = x / 2 a 3

Figure 3.4: Possible stress increment directions when current stress is on the yield surface
(Linear elastic non-virgin case)(After Wathugala, 1990)

properties are given in Table 3.1.


The exact direction of the stress increment, da^, for a given strain increment, dekl,
is not known in advance. Therefore an approximate direction is evaluated by assuming

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
31

elastic behavior. This stress increment is commonly known as the elastic predictor stress
increment, dafj. It may be calculated from

do\j = Ctjudea (3.30)

Now this stress could be used to identify the loading direction as illustrated in Ta­

ble 3.1. It should be noted here that the correct stress increment direction, daij, for the

virgin loading case (OA) is not in the direction of dcr?-. However, for the virgin loading

case, both da^ and dafj point outward from the yield surface, and therefore, it is suffi­
cient to calculate dafj to determine whether deki will cause loading or unloading. For
the stress increment OC, daij = dafj, and no further calculations are required. For the

stress increment OE, the point D is found by solving the following equation for k using
the Newton-Raphson iterative technique.

F(aij + kdafj, aps) = 0 (3.31)

From the positive solution of k, the point D is located. The strain for this stress increment,
dek[D, can be calculated from

* 8 ° = D k „ iK D 0-32)

where

da?0 = kda\j (3.33)

Now the remainder of the strain, deklE, can be calculated from

dsj?tE = dekl - d£°tD (3.34)

Stress increment for this strain increment is found using the algorithm presented in Sec­

tion 3.5.3. When the stress point is inside the yield surface, the possible stress increment
types are illustrated in Figure 3.5. As in the earlier case, the elastic predictor stress incre­

ment, dafj, is evaluated from Equation 3.30. If the elastic predictor a?- = crtJ- + daf}- stress
lies inside the yield surface (OA), the elastic predictor gives the correct stress increment

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
32

160.0

120.0

b 80.0

40.0

0.0
0.0 40.0 80.0 120.0 160.0 200.0
y f a * — y/2a3

Figure 3.5: Possible stress increment directions when current stress is inside the yield
surface (Linear elastic non-virgin case)(After Wathugala, 1990)

and therefore no further processing is required. If the elastic predictor stress lies outside

the yield surface (OC), it could be handled similarly to the stress increment OE in the
earlier case.

3.5.2.2 Nonlinear

The S* series of Hierarchical Single Surface (H iS S ) models (Wathugala, 1990; Wathugala

and Desai, 1991a, 1991b, 1993) treat unloading and reloading as nonlinear, and therefore
the procedure for computer implementation is more complex than the linear elastic case.
Here it is necessary to distinguish between unloading and reloading so that the correct

constitutive matrix can be selected. For a stress point inside the yield surface, the possible

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
33

types of stress increments are illustrated in Figure 3.6. The directions are marked using

200.0

160.0

120.0

80.0

40.0

0.0
0.0 40.0 80.0 120.0 160.0 200.0 240.0
V 2a2 — v^ 3

Figure 3.6: Possible stress increment directions when current stress is inside the yield
surface (nonlinear non-virgin case)(After Wathugala, 1990)

the elastic predictor stresses. Similar to the linear elastic case, elastic predictor stress

can be calculated from Equation 3.30. Some properties of these elastic predictor stresses
for different stress directions are described in Table 3.2. Since non-virgin loadings are
considered to be nonlinear, the stress path corresponding to a given strain increment in

general is not straight line. The paths plotted in Figure 3.6 using the elastic predictor
stresses are only approximate. First the quantity dcr? n£- is calculated. If it is less than zero,
that means unloading takes place at the beginning, which could be followed by reloading
and virgin loading. Now, the point where unloading changes to reloading (neutral loading

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
34

point) is found as the stress point with maximum 0^. Here a r is defined as cer = R- l (tr,j),
where R = R(<Jij, aT) = 0, a reference function which is algebraically similar to F.
Along the elastic predictor stress path, we want to calculate a stress point where slope

of Or changes sign. This means we need a solution for t where

Equation 3.35 is solved by Newton’s iteration. It is possible that there is no solution


for t in the range 0 < t < 1. In that case, the whole strain increment is unloading,

and the corresponding constitutive matrix is used for the integration. When a solution is
found, the strain increment part t x cfey is considered as unloading, and the remaining

part (1 — t) x dsij is considered as reloading or reloading followed by virgin loading.


If F{<7ij + dofj, a ps) <= 0, then it is simply reloading and, if F(<jy + dafj, aps) > 0,

then it may be reloading followed by virgin loading. For nonlinear reloading, the initial

elastic predictor stress is, in general, larger than the actual stress increment The point
where the stress path changes from reloading to virgin loading can be determined using
the procedure described in Section 3.5.2.1.

The integration of the constitutive equation for reloading and unloading is performed

using the Forward Euler method with subincrementation. This subincrement technique

has been used by many researchers to calculate stress increments corresponding to strain
increments in nonlinear constitutive models (Faruque and Desai, 1985). In these methods,

the strain increment, de^i, is subdivided into n equal parts, where n is a predetermined
value. In this procedure, n does not depend on the size of cfc*/ and therefore will subdivide
even small strain increments. When the strain increment, deki, is very large, this method

could result in unacceptably large subincrements. Therefore, to avoid these situations, the
following method is proposed by Wathugala (1990) to calculate n value.

n = largest integer part of (3.36)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
35

where dema* is the largest subincrement permitted. Now the subincrement strain, £efcf,
may be calculated from

(3.37)
n
After calculating 8eki, the corresponding stress increment Saij is calculated from

5 & ij — C j j k[5 £ k l (3.38)

where C*jk[ is the corresponding constitutive tensor. This procedure is performed for n

steps. At each step, C'jki and <7ij are updated. If the stress point lies on the yield surface,

the possible types of stress increments are illustrated in Figure 3.7. Similar to the last case,

possible elastic predictors and their properties are listed in Table 3.3. For the stress paths

OA and O B , strain increments cause virgin loading, and therefore, the algorithm given in

Section 3.5.3 is used. For all the other strain increments, the subincrementation method
described above is used.

3.5.3 Virgin Loading

The basic problem in calculating stress increment, da^ corresponding to a given strain

increment, deki which causes virgin loading, may be expressed as follows:


Given: erg, a°s, , dekl and

F(ffg,o&) = 0 (3.39)

Find: afj and so that

F ( o § .o g ) = 0 (3.40)

where and are found from

deki = deh + deh (3.41)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
36

200.0

160.0

120.0

80.0

40.0

0.0
0.0 40.0 80.0 120.0 160.0 200.0
\/2<72 — y/2a*

Figure 3.7: Possible stress increment directions when current stress is on the yield surface
(Nonlinear non-virgin loading case)(After Wathugala, 1990)

Table 3.1: Possible stress increment direction when current stress is on the yield surface
(Linear elastic non-virgin loading case). (After Wathugala, 1990)

Stress Increment dofjnfi F (oeip ocps) Description

OA > 0 > 0 Virgin Loading


OB = 0 > 0 Neutral Loading

OC < 0 < 0 Non-virgin Loading

OD < 0 = 0 Non-virgin Loading

OE < 0 = 0 Non-virgin Loading


followed by Virgin Loading

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
37

Table 3.2: Possible stress increment directions when the current stress is inside the yield
surface (Nonlinear non-virgin loading) (After Wathugala, 1990)

Stress Increment <4 Description

O A >0 <0 O r > Ol% > 0 Reloading


O B >0 = 0 olt > aj : = 0 Reloading
O C >0 >0 Of ^ ^ 0 Reloading followed

by Virgin Loading

= 0 <0 Neutral Loading

o
OD V

V
A
followed by Reloading

O E ,O J <0 <0 a* > a T > 0 Unloading


o
u l-

Unloading followed
A

<0 = 0
&

O F
II

by Reloading

OG < 0 > 0 ocr ^ 0Cpg ^ 0 Unloading followed

by Reloading and

Virgin Loading
OH < 0 <0 OiT > O r* > 0 Unloading followed

by Reloading

01 <0 <0 a* > Or > 0 Unloading followed

by Reloading

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
38

Table 3.3: Possible stress increment directions when the current stress is on the yield
surface (Nonlinear non-virgin loading case) (After Wathugala, 1990)

Stress Increment d(Jfjnij Ops) <*r Description

OA > 0 > 0 a* < ups Virgin Loading


OB = 0 > 0 <*r < <V Neutral Loading

OC,O G < 0 < 0 a* > <V Unloading

OD < 0 = 0 Ur —Ups Unloading followed


by Reloading
OE < 0 > 0 u ev < Ups Unloading followed
by Reloading and

Virgin Loading
OF < 0 < 0 Ur ^ Ups Unloading followed
by Reloading

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
39

daij — Cfjkideh (3.42)

a g = erg + (3.43)

(3.44)

(3.45)

and

°S = /.« ? ) (3.46)

where superscripts O and C refer to initial and final converged quantities. F is the yield

function, and f a is the hardening function. Total, elastic, and plastic strain increments are

given by dekl, dekl, and depkl respectively. C?jk[ is the elastic constitutive tensor, and
are different trajectories of plastic strains such as £, and f y. The functions, relate

incremental plastic strains to incremental trajectories. If the exact decomposition of the


incremental strain tensor, deki, into its elastic and plastic components is known, erg,

and apS can be calculated from Equations 3.43,3.45,3.46 respectively. Unfortunately, this

decomposition of the incremental strain tensor is not known in advance, and most of the

iterative schemes for solution of the above system start with a trial decomposition of the
incremental strain tensor.

3.5.3.1 Elastic Predictor-Plastic Corrector Method

The elastic predictor plastic predictor method (Ortiz and Simo, 1986) uses dskl = 0 as

the trial solution. Let us assume that these trial solutions move stress to an intermediate
state I, and a single correction moves it to the final converged solution at C. This is

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
40

schematically illustrated in Figure 3.8. If the trial solution assumes the decomposition for

40.0

30.0

Predictor

20.0 Corrector

10.0 = 0

0.0
0.0 100.0 200.0 300.0 400.0
Ji

Figure 3.8: Schematic diagram for ideal predictor corrector algorithm (After Wathugala,
1990)

the strain tensor to be

dekl = deekf l + depkf I (3.47)

Then the quantities at the intermediate state, I, may be found from

erg = erg + dog/ (3.48)

(3.49)

and

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
41

< = /*(£/) (3.50)

Where dcr°r = Cfjk[d£k ° r and d £ fl = /,(d e^ 0 /). Superscript 0 1 refers to the quan­

tities associated with the stress path 01. If the errors in the assumed incremental elastic
and plastic strain tensors are d£eklrc and dek\lc , they are related as follows:

dekl = d e l f 1 + del)(c + d ^ f 1 + d ^ IC (3.51)

Subtracting Equation 3.47 from Equation 3.51 yields

d£ek)IC + (Hkl)rc = 0 (3.52)

Now quantities at the converged state C may be calculated from

(3.53)

ic (3.54)

= /« (€ f) (3-55)

Where d a [f = Cfjkldeklic and d£[c = / t(cfe(f). Substitution of Equations 3.53, 3.54


and 3.55 into Equation 3.40 yields

F ( 4 + d T f P , / (1(?/+ < ie/o)) = o (3.56)

Taylor’s series expansion of Equation 3.56 around the point I is given by

df d F df*
0 = F ( 4 , /«(£/)) + d^(c + higher order terms (3.57)
dcxps d£i

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
42

For most of the constitutive models in practice, d a ff and d£[c can be expressed in terms
of a single variable as follows. From the flow rule in plasticity

d 4 - r = dX’c n% (3.58)

Substituting Equation 3.58 into Equation 3.44 yields

d iic = /j(<iA'c ng) (3.59)

For most of the hardening functions in use, including the ones used in the H iS S mod­

els, Equation 3.59 may be simplified to

# ' c = d \ ' c li( n l) (3.60)

Substituting Equations 3.52 and 3.58 into Equation 3.42 yields

d a g = dX'c C ‘ijkln% (3.61)

Now substituting Equations 3.60 and 3.61 into Equation 3.57 yields

a/ dF d f a
0 = F(a'ip / „ « / ) ) + C ijk ln kl + dX'c +
da.y dccps d£i
higher order terms of d \ IC (3.62)

By neglecting higher order terms of dXrc, Equation 3.62 may be solved for dXlc and is
given by

dXIC = - F ( 4 . M (!)) (3.63)

Now all three quantities to be found at the state C can be calculated as follows, a
can be calculated from Equations 3.63, 3.61, and 3.53. can be calculated from Equa­

tions 3.63,3.60, and 3.54, and then can be calculated from Equation 3.55. Since higher

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
43

order terms in Equation 3.62 have been neglected, the solution obtained here would not
satisfy the yield function given by Equation 3.40. The quantities obtained here actually
also refer to an intermediate state closer to the final converged solution. Substituting the

solution obtained here as the intermediate state in Equation 3.63, the procedure is repeated

until d \ lc or F(<r(j,aj)S) is less than a prescribed tolerance. This procedure is schemati­


cally illustrated in Figure 3.9.

40.0

30.0
Predictor

Corrector
S 20.0

10.0

0.0
0.0 100.0 200.0 300.0 400.0
Ji

Figure 3.9: Schematic diagram for iteration procedure (After Wathugala, 1990)

3.5.3.2 Comparison of Elastic Predictor and Plastic Predictor Methods

The two most popular methods in evaluating the intermediate stress, (a) elastic predictor
(b) plastic predictor, are compared here. In the plastic predictor-plastic corrector method,

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
44

the predictor stress increment dag' is calculated using elasto-plastic constitutive matrix,

Cijki* corresponding to the initial stress point O instead of elastic constitutive matrix Cfjkl.
Substituting Equation 3.63 into Equation 3.61 yields

(3.64)

For the elastic predictor plastic corrector method, dek ° ' = deki anddejj!}0/ = 0. There­
fore a!pS = ctpS. The yield function, F, at the point I may be expanded using the Taylor’s
series around the point O as

dF
F ( 4 , a 's) = F (ag + dag', o £ ) = F (ag + o £ ) + d o g '+ O \d o gOI\
l ) (3.65)
daij}

Since d e f f ' = dekl for this method, from Equation 3.42

dag' = C eijkldskl (3.66)

Substituting Equations 3.39, 3.66 into Equation 3.65 and rearranging after neglecting
higher order terms, yields

dF
Cfjkldeki (3.67)
dO{j _
The stress change from O to C, a g ° may be given by

d ovf f = da?/ + daip


*3
(3.68)

Substituting Equations 3.64,3.66 and 3.67 into Equation 3.68 yields

dagc = d£kl (3.69)

The term inside the square bracket [ ] is similar to the expression for the Cg*kl in Equa­
tion 3.16. The only difference is that all the terms in Cg?kl are calculated at a single point,

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
45

whereas in Equation 3.69, different terms are calculated at different points as indicated.

In the drift correction method (Potts and Gens, 1985), C*?kl is calculated at the point O.
For small strain increments, both methods gave the same answer, but the elastic predictor
method took one more iteration to achieve the same accuracy. Since the elastic predictor
method does not require the calculation of C*?kl, both methods require equal numerical
effort to achieve the same convergence limits for small increments.

3.5.3.3 Implicit Integration Method

The implicit integration method used here is based on the Backward-Euler integration

scheme where the normal to the potential surface, nk[, at the final stress state is used for
the plastic correction. This method has been used by various researchers (Simo et al., 1986

& 1987; Borja and Lee, 1990; Jeremic and Sture, 1995; Macari et al., 1997) to integrate

the elasto-plastic constitutive models.

The Backward-Euler integration scheme can be expressed as:

aS = aij ~ XCm n<ki (3-7°)

where a? is the final stress state, is the elastic prediction, and nk ° is the normal to the

plastic potential surface at the final stress state C. Generally an initial estimate of does

not satisfy the yield criteria as well as Equation 3.70 and an iterative scheme is necessary
to bring the stress on the yield surface.

In order to derive such an iterative scheme, a residual stress tensor is defined as:

where represents the difference between estimated current state crij (point E in Fig­
ure 3.10) and the implicit stress state of Backward-Euler scheme (crL — ACfjkln kl) which
represents the error in the current estimate of stress Equations 3.70 and 3.71 are

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
I, Elastic Predictor

F c o n to u r
Current Estimate

C, Final converged point

Figure 3.10: Schematic representation of Implicit Method

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
47

schematically illustrated in Figure 3.10. Suppose at a stress state cn ,

A") # 0 (3.72)

Suppose a correction stress da-j brings the stress state to cN + 1 where the residual stress is

rij+l(°ij+l’ " S ( ^ +l. < +I). AN+1) = 0 (3.73)

where, a g +1 = + d<rg, t f +l = + d£tN and A"+l = A* + dXN. A Taylor series


expansion of Equation 3.73 yields

r iV + l _ r N + d r ij d , ( ^ L d c N , d n kl d N \ , 0 ? J L d \ N i
r iJ + d a » n d a m n + d n Q { d t i 6 + d a m ndam n) + d \ dX +

higher order terms = 0 (3 .7 4 )

Neglecting higher order terms and keeping the elastic predictor stress cr/y fixed, Equa­
tion 3.74 can be written as

dn?N
r ” +l = r " + d<rmn + d \ NCtm n $ + = 0
tj ' ij ^ ' m da„
(3.75)
where d£^ = d \ N f t( n ^ ) . Solving Equation 3.75 for da^n yields

d°L = ~ ( r ij + d A C Z u Z S H A & n )-1 (3.76)

where A,%„ = StmSnj + Z& = n « " + A'v ^ / . ( n ‘f ) and =


^ { d jk S j i -I- S u S k j ) .

The new value of the yield function, F N+l, at the final stress state due to change in

O ij and can be obtained by a first order Taylor series expansion of the yield function F
about the final stress cr,y+ l:

r\ pi
F N+l = F n + n^da^n + + higher order terms (3.77)

where n £ n is the normal to the yield surface. The solution for dX can be obtained by
setting F N+l = 0 and substituing dXNf ^ n ^ ) for d^^ in Equation 3.77:

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
48

FK+nL(r% +dV%uZS)(A$mJ-' + ^ c iA 'V r f) = 0 (3.78)

or,

d \" = (3.79)
rCCZuZSWj^ ) - 1- ^/<(ng")
Now, substituting d \ N in Equation 3.76, we get the final expression for da*n:

da — (r I nmnrij (-^tjmn) x-,e Z N\ ( A N )_l (3.80)


a a mn ~ I r ij + F /^ e 7 * * ( A N \ - i U i j k l * k l I lA j m r J
\ m n ^ i j k l kl K ^ ijm n ) J
The iteration continues until the yield criterion is satisfied (or F N+l = 0 and r ^ +l = 0).

In a finite element procedure, usually Newton-Raphson iteration is used for the global

equlibrium. In that situation the continuum tangent stiffness tensor C*?kl destroys the
quadratic rate of asymptotic convergence. A consistent tangent stiffness tensor (Simo and
Taylor, 1985; Runesson and Samuelson, 1986) preserves the quadratic rate of asymptotic

convergence. The consistent tangent stiffness is derived as follows:


Differentiating Equation 3.70, we obtain

damn + CfjUng + ACS dn?i


ijkl fi(nkl) dX —Cfjkld€ki — 0
db
(3.81)
The consistency condition of plasticity gives:

dF
n m n .d (7Tnn + ~ Q ^ d ^ f i { n m .n ) = 0 (3.82)

From Equations 3.81 and 3.82, the expression for dX can be found as:

dX = (3.83)
< nC?jklZ klN j mn + § / t(ng)
Substituting Equation 3.83 into Equation 3.81 gives:

n RpqklZkl'R'ijRijmn
damn — ftmnkl n*^ ~ nrp . , deki (3.84)
n otRotpq% kl + Q^i fi{nkl)
where Rmnki is the reduced stiffness tensor defined as Rmnki = ^ijmn^ijki-

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
49

3.5.3.4 Modified Euler Method

This method has been proposed by Sloan (1987) based on the modified Euler scheme.

The main features of this scheme are that it controls the error in the integration of the
constitutive model by adjusting the size of each substeps automatically and it does not
require the drift correction generally used to make sure that the stress point lies on the
yield surface. The detailed description of the method is given by Sloan (1987). Here, a

brief description is given for the sake of completeness.

The Forward Euler method gives the stresses and hardening parameters at time t n+l

according to

a ^ l = a ^ + C ^kldekl (3.85)

£tn+l = £* + dXM nZ) (3.86)

where C ^ kl and n kl are evaluated at the stress state cr," at time tn and AT, the increment

in the dimensionless time T — (t — tn) /( tn+l — t n) where tn < t < tn+l, is assumed to be

unity. This procedure is accurate only for very small time steps, and the accuracy of the

procedure is improved if the time step AT = 1 is broken up into a number of smaller time
steps of equal size (Nayak and Zienkiewicz, 1972). If the time increment is divided into
N number of substeps, then the stresses and the hardening parameters can be evaluated by

the following expressions:

= 4 + C$u del, (3.87)

f?+L = ? ,- + dA /i(ng‘ ) (3.88)

where dekl = ATkdekiandAT* = l / N and k = 1 ,2 ,., N . In general,the number


of substepsare calculated empirically (Nayak and Zienkiewicz, 1972; Owen and Hinton,

1980) and the errors in the stresses are not controlled direcdy. Sloan (1987) has proposed
a scheme based on the modified Euler method where errors in stresses are controlled in-

directiy by adjusting the size of substeps. In this method, the solution is obtained by the

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
50

modified Euler scheme as:

°y+l = ai i + ^ ( Acri + A4 ) ( 3 .89)

^ + l = ^ + ^ ( A^ + Aa (3.90)

where Act,!- = C ^klA £kl and A f/ = A Afi(n kl) as determined by the Forward-Euler

method and Ae*/ = AT*Ae*/. The quantities Ct^ , A A, and n^( arc calculated at the

stress state cr£-. Act?- and A £ 2 are evaluated in a similar way, but the quantities C ^kl, A A,
and njy are calculated at the stress state (a 4 - Acr^-).

For a given strain increment Ae**, the Euler method has a local truncation error of
order 0 (A T 2), and the local truncation error for the modified Euler solution is 0 (A T 3).
Therefore, an estimate of the local truncation error in is defined as:

A+i (3.91)
ij

which is only accurate to 0 (A T 2). The global error is monitored indirectly by controlling
the relative error for each substep which is defined as:

II ejr II
II II (3.92)

The size of each substep is continually updated during the integration procedure so that
Rk+i < T O L where T O L is small number.

The integration procedure is begun by assuming a value for AT*. If it satisfies Equa­
tion 3.92, then the integration for that step is complete. If Equation 3.92 is not satisfied,
then size of AT* is reduced and the calculation is repeated. Step sizes are reduced or in­

creased depending on the estimated value of R using a local extrapolation technique. If


the current substep size is AT*, then the next substep size is given by:

AT*+t = toAT* (3.93)

where m is given by

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
51

1
TO L 5
( 3.94)

To make m conservative, the above equation is modified as:

(3.95)

Sloan (1987) has proposed the use of the following limits on the size of each new substep

so that extrapolation is not carried too far.

0.1ATk < ATk+l < 2ATk (3.96)

The same limits are used in the present study. A schematic representation of the Modified
Euler method is shown in Figure 3.11.

3.5.4 Subincrementation

All the four methods use subincrementation for better accuracy. In the Modified Euler
method, the size of the subincrementation is controlled automatically depending on the

accuracy of the solution as described above. The other methods, E P —PC, P P —PC, and

Implicit, use fixed size subincrementation where the strain increment dekl is subdivided

into n equal parts. The value of n is calculated using Equation 3.36. The subincrement
strain 5eki can be calculated from Equation 3.37. For each subincrement strain 5eki, the

above mentioned algorithms are used to compute the stress increment £cr,y.

For the Implicit method, the formulation of the consistent tangent stiffness (Equa­

tion 3.84) is valid for the strain increment without any subincrementaion. To calculate the
consistent tangent stiffness for the subincrementation scheme, the following procedure is
used.
For each subincrement, the consistent tangent stiffness Cfj'hl is calculated using Equa­
tion 3.84, where i denotes the increment number. Since the size of subincrement strain 5ekl

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
52

tangent at C

tangent at A correct stress-strain curve

bd
td

\ parallel to the tangent at C

E = estimate of the current stress


BE= ED= 0.5BD= estimated error
EC= actual error

Figure 3.11: Schematic representation of the Modified Euler method

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
53

is same in all the subincrements, the total stress increment for a given strain increment can

be written as

daij = [Cfjkl + Cfjkl + ..... + C?kl] 8eki (3.97)

Using Equation 3.37, Equation 3.97 can be written as

dati = + ......+ (3.98)

where (Cglt + C ^h + ......+ CfjlL)/n represent the consistent tangent stiffness tensor for
n subincrementations.

3.6 Examples
In this section, the performances of the algorithms described here have been evaluated by

predicting three simulated triaxial test results and comparing them with the input data. All

of the four methods have been implemented in the finite element procedure. To evaluate

the performances of these procedures in F E M , a flexible and a rigid footing problems


have been analyzed. Material parameters for a typical clay has been used for all the anal­
yses here (E = 11032 kP a, u = 0.35, 7 = 0.047, /? = 0., m = —0.5, n = 2.8,

h\ = 1.0 x 10-4, /12 = 0.78, hz = 0., and /14 = NA). Since our objective is to compare
the algorithms for virgin loading, all the numerical examples have been designed to ex­

perience virgin loading only. Therefore, non-virgin loading parameters do not affect our
results. In the present study, linear elsatic non-virgin loading has been assumed. However,

the algorithm presented for non-virgin loading is still necessary to improve the robustness
of the algorithm (Wathugala and Pal, 1996). In the following discussion, the elastic pre­

dictor - plastic corrector method, the plastic predictor - plastic corrector method and the

modified Euler method are referred to as E P — PC method, P P — P C method and M E

method respectively.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
54

3.6.1 Prediction o f Simulated IHaxial Tests

Three different triaxial tests, ( 1 ) drained conventional triaxial compression (C T C ), (2)


drained triaxial extension (T E ), and (3) undrained conventional triaxial compression {CTC),
have been used to evaluate the performances of the algorithms. The stress paths followed

by the abovementioned triaxial tests are defined in Figure 3.12. The methodology used to
evaluate the performance is as follows:

CTC
CTC (Undrained)

TE

Figure 3.12: Stress paths followed by different triaxial tests

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
55

For the drained triaxial tests, stress paths are defined first Then the stress to strain

algorithm is used to predict the strain paths using very small stress increment size ( 1 kPa).
This algorithm predicts a very accurate strain path for a given stress path (Wathugala,

1990). Now, the strain to stress algorithms described here are used to predict the stress
paths using the strain paths obtained by the stress to strain algorithm. These predicted
stress paths and the actual stress paths are compared to evaluate the performances of the

algorithms.

For the undrained test, the strain path is defined first, and the corresponding stress path
is obtained by using the algorithms described here. The ‘exact’ stress path is obtained by

using a strain to stress algorithm with very small strain increment steps. Comparisons of

the predicted stress paths and the ‘exact’ stress paths are used to evaluate the performances
of the algorithms.

Figures 3.13 to 3.15 show the stress paths of the C T C drained test for the strain incre­

ment sizes of 10- 4 ,1 0 -3, and 10- 2 respectively. Note that the strain increment size does

not affect the result of the Modified Euler method. The accuracy of the Modified Euler
method is affected by the tolerance (TOL), and three TO L values of 1 , 1 0 -2, and 1 0 -5

are used in the present problems. These TO L values are chosen because in terms of the
total number of subincrementations, these values match the virgin strain increment sizes
of 10- 2 ,1 0 -3, and 10- 4 respectively. It can be observed in Figures 3.14 and 3.15 that for
the strain increment sizes 10-3, and 10- 2 the results of the Implicit method deviate near
failure and the result becomes unstable for the strain increment size of 10-2. On the other
hand, method 1 and method 2 show only a little deviation near failure. For the strain in­
crement size of 10-4, all the methods yield stress paths very similar to the actual one. The

stress path obtained from method 4 with TO L = 10~ 5 is also very similar to the actual
stress path. However, the stress path predicted by the M E method with T O L = 1 crosses
the critical state line near failure (Figure 3.15).

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
56

The stress paths for the drainde T E test are shown in Figures 3.16,3.17 and 3.18 for
different strain increment sizes. In this case all the algorithms have predicted identically
same stress paths for the strain increment size of 10-4. However, the predictions are little

different from the original stress path. At large strain increment size (10-2), the stress
paths have deviated from the original path by about 0.3%. This deviation is maximum for
the Implicit method. Predictions of the E P — PC and P P — P C methods have moved

away from the original stress path around \] J id = 60, but they predict the correct stress at
failure. All the methods are stable in these predictions.
The predictions for the C T C undrained test are shown in Figures 3.19,3.20 and 3.21

for different strain increment sizes. The trend observed in this case is similar to the trend

observed for the drained C T C test All the methods have predicted the same stress path

for small strain increment size (10-4). For large strain increment size (10-2), the predicted
stress path for the M E method has crossed the phase change line and the Implicit method

becomes unstable near failure.

The C P U times taken by all these analyses on a I B M R S /6000 model 355 worksta­
tion are shown in Figures 3.22,3.23 and 3.24 for the C T C drained, T E drained and C T C

undrained tests respectively. It can be observed in these figures that the Implicit method
is much more computationally expensive than the other methods. The CPU times taken

by the other three methods are more or less in the same range.

3.6.2 Footing Problems

All the four methods have been implemented in the commercial finite element package
ABAQUS. To evaluate their performances in FEM, two footing problems, (1) rigid and

(2) flexible, have been analyzed. The finite element mesh and the boundary conditions are

shown in Figure 3.25. The problems have been considered as plain strain problems with
strip footings. The element used in both the analyses is the 8 -node plain strain quadrature
element with the reduced integration scheme (element type C P E 8 R in ABAQ U S). The

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
57

100.0

80.0
Phase Change Line.

60.0
IQ
k
40.0

“C orrect”
o— o E P -P C
20.0 ■>— o P P -P C
♦ - —♦Im plicit
*-----♦ Modified Euler

0.0
400.0 600.0 800.0 1000.0
Ji

(a) Full view

95.0

90.0
Phase Change Line.

85.0
IQ
k
80.0

“C orrect”
a o E P -P C
75.0 >—♦ P P -P C
•— ♦ Implicit
♦-----♦ Modified Euler

70.0
725.0 750.0 775.0 800.0 825.0
Ji

(b) Zoom view

Figure 3.13: Stress paths of the C TC drained test (A e ^ ,. = 10-4, TO L = 10-5)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
58

100.0

80.0
Phase Change Line.

60.0
IQ
k
40.0
“Correct”
o o EP-PC
20.0 w — m PP-PC
♦ - —♦Implicit
♦-----♦ Modified Euler

0.0
400.0 600.0 800.0 1000.0
Ji

(a) Full view

95.0

90.0
Phase Change Line...

'/ /
85.0
IQ
k
80.0

“Correct”
EP-PC
75.0 PP-PC
Implicit
Modified Euler

70.0
725.0 750.0 775.0 800.0 825.0
h

(b) Zoom view

Figure 3.14: Stress paths of the C T C drained test (&EmaX = 10-3, T O L = 10-2)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
59

100.0

80.0
Phase Change Line..

60.0
I3
k
40.0

“Correct”
20.0 a o EP-PC
m-— m PP-PC
♦ - —♦Im plicit
*-----♦ Modified Euler
0.0
400.0 600.0 800.0 1000.0
Ji

(a) Full view

95.0

90.0
P hase Change

85.0

80.0

“C orrect”
75.0 o o E P-PC
» — m P P -P C
♦ - —♦Im plicit
a— a Modified Euler
70.0 —
725.0 750.0 775.0 800.0 825.0
Jx

(b) Zoom view

Figure 3.15: Stress paths of the C T C drained test (A£max = 10-2, T O L = 1 )

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
80.0

P h ase Change Lipe-

60.0

K 40.0

“C orrect”
20.0 o E P -P C
P P -P C
— Implicit
i— a Modified Euler
0.0
500.0 600.0 700.0 800.0

(a) Full view

75.0 “I __ 1-

Phase Change Line


70.0

65.0

60.0

“ C orrect”
a o E P -P C
55.0 B— « P P -P C
«— ♦ Implicit
e a Modified Euler

50.0
600.0 610.0 620.0 630.0 640.0 650.0
Ji

(b) Zoom view

Figure 3.16: Stress paths of the T E drained test (Aemax = 10-4, T O L = 10“

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
80.0

Phase Change Line



60.0 1 >

K 40.0

“C o rrec t”
20.0 o E P -P C
P P -P C
— ♦ Implicit
i a M odified Euler
0.0
500.0 600.0 700.0 800.0
Ji

(a) Full view

75.0

P hase Change Line..


70.0

65.0

60.0

“ C o rrect”
E P -P C
55.0 — m P P -P C
—♦ Im p lic it
— * M odified Euler
50.0
600.0 610.0 620.0 630.0 640.0 650.0
Jx

(b) Zoom View

Figure 3.17: Stress paths of the T E drained test (A emax = 10-3, TO L = 10'

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
62

80.0

P h ase Change Lipe-

60.0

•-? 40.0

11 “C orrect”
20.0 a o E P -P C
11 * - - e P P -P C
e— Implicit
11 a— 4 Modified Euler
0.0
500.0 600.0 700.0 800.0
Ji

(a) Full view

75.0

Phase Change Line...


70.0

65.0
I^
K
60.0
“C orrect”
E P-PC
55.0 P P -P C
Implicit
Modified Euler

50.0
600.0 610.0 620.0 630.0 640.0 650.0
Jx

(b) Zoom view

Figure 3.18: Stress paths of the T E drained test (Aemoa. = 10-2, T O L = 1)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
63

100.0

80.0 Phase Change Line

60.0
IP
k
40.0

“ Correct”
20.0
o o E P-PC
* — « p p -P C
♦ Implicit
4-----4 Modified Euler
0.0
400.0 500.0 600.0 700.0 800.0 900.0
Ji

(a) Full view

70.0

65.0 P hase Change Line

60.0
IQ
k
55.0
“C orrect”
a E P -P C
50.0 * — « pp-pc
♦— ♦ Implicit
a 4 Modified Euler

45.0
560.0 570.0 580.0 590.0 600.0 610.0
Ji

(b) Zoom view

Figure 3.19: Stress paths of the C T C undrained test (Aemax = 10-4, T O L = 10-5)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
64

100.0

80.0 Phase Change Line

60.0
Q
es

40.0

“Correct”
20.0 a o EP-PC
* — a PP-PC
•— ♦ Implicit
a * Modified Euler
0.0
400.0 500.0 600.0 700.0 800.0 900.0
■h

(a) Full view

70.0

65.0 Phase Change Line

60.0

k
55.0
"C orrect”
e o E P-PC
50.0 — « P P -P C
♦ — ♦Im plicit
a a Modified Euler

45.0
560.0 570.0 580.0 590.0 600.0 610.0
Ji

(b) Zoom view

Figure 3.20: Stress paths of the C T C undrained test (Aemax = 10-3, T O L = 10-2)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
65

100.0

80.0 Phase Change Line

60.0

40.0

f “ C orrect”
o E P -P C
20.0
* — m P P -P C
♦— ♦ Im p lic it
♦ a Modified Euler

0.0
400.0 500.0 600.0 700.0 800.0 900.0
Ji

(a) Full view

70.0

65.0 P h ase Change Line

60.0
Q
>
55.0
“C orrect”
E P -P C
P P -P C
50.0 Implicit
Modified Euler

45.0
560.0 570.0 580.0 590.0 600.0 610.0
Ji

(b) Zoom view

Figure 3.21: Stress paths of the C T C undrained test (Aemai = 10-2, T O L = 1 )

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
66

CPU time, seconds

■ 1.0E-02
□ 1.0E-03
« 1 .0E-04

EP-PC PP-PC Implicit Mod.


Euler

Figure 3.22: CPU time taken by different algorithms for CD-CTC

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
67

0.5
CPU time, seconds

0.4
0.3 ■ 1.0E-02
□ 1.0E-03
0.2
M 1.0E-04
0.1

EP-PC PP-PC Implicit Mod.


Euler

Figure 3.23: CPU time taken by different algorithms for CD-TE

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
68

CPU time, seconds

■ 1.0E-02
□ 1.0E-03
m 1.0E-04

EP-PC PP-PC Implicit Mad.


Euler

Figure 3.24: CPU time taken by different algorithms for CU-CTC

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
69

0.25 m.

2 m.

Figure 3.25: F E M mesh for the footing problem

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
70

number of elements were made small in order to facilitate the large number of analyses in

the study. Initial conditions such as stresses, and hardening parameters were computed by
assuming normally consolidated clay at K 0 (=.52) stress conditions. In order to minimize

numerical problems due to zero effective stress state on the surface, a surcharge of 2 0 kPa
was applied on the surface of the clay for all the problems. This may be interpreted as a
soil of about 1 m deep without any shear strength.

In the first problem, a rigid strip footing is considered, and 200 displacement incre­
ments of equal size are imposed on the footing. The final displacement induces a state of
failure in the soil mass. Figures 3.26, 3.27, and 3.28 show footing load vs. settlement
curves for three different strain increment sizes and TO L values. The M E method yields
a convergence problem when a TO L value of larger than 10- 5 is used. Therefore, only

one curve is shown in Figure 3.28 which corresponds to T O L = 10-5. Figures 3.26,3.27,

and 3.28 show that the strain increment size does not affect the results for the first three

methods. Figure 3.28 shows that the final load observed for the M E method is higher
than the final loads obtained by the other methods. The M E method also fails to show

the failure state at the final displacement, though the other three methods have captured

that. A comparison of values of F and the stress ratio \JJ id I at gauss points show that
the stress points for E P — P C method are on the yield surface or inside it as expected.
However, for the M E method, several stress points are outside the yield surface. This is
caused by the accumulation of errors in each small step. It should be noted that the M E
method does not perform drift correction.

In the second problem, a flexible strip footing is considered, and a pressure of 135 kPa

is applied to the footing in increments. The final footing pressure induces a state of failure

in the soil mass. For the first three methods, the footing pressure has been applied in 135

increments. For the Implicit method, this increment size produces convergence problem
for the strain increment size of 10-2. Another problem was run with smaller increment size
(135 kPa in 400 increments)using the Implicit method and it had convergence problem

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
71

0.0

- 20.0
Settlement, mm

- 40.0

- 60.0

G o EP-PC
PP-PC
- 80.0
e - —♦Im plicit
* ----- ▲“Correct”

- 100.0
0.0 20.0 40.0 60.0 80.0 100.0 120.0
Footing Load, kN/m

Figure 3.26: Load-settlement curves for the rigid footing = 10“2)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
72

0.0

- 20.0
Settlement, mm

- 40.0

- 60.0

Q o EP-PC
■ PP-PC
- 80.0
e - —♦Im p licit
a a “Correct”

- 100.0
0.0 20.0 40.0 60.0 80.0 100.0 120.0
Footing Load, kN/m

Figure 3.27: Load-settlement curves for the rigid footing (Aemai = 1 0 ~3)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
73

0.0

- 20.0
Settlement, mm

- 40.0

- 60.0

g o EP-PC
* — M PP-PC
- 80.0
❖ -—♦Im p lic it
a a Modified Euler

- 100.0
0.0 20.0 40.0 60.0 80.0 100.0 120.0
Footing Load, kN/m

Figure 3.28: Load-settlement curves for the rigid footing (AcmaI = 10-4, T O L = 10-5)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
74

before reaching the failure state too. This case is designated as ‘smaller increment’ in

Figure 3.29. The footing pressure has been applied in 400 equal increments for the M E
method because this method had convergence problem when the footing pressure was

applied in 135 increments. Figures 3.29, 3.30 and 3.31 show footing load vs. settlement

curves obtained from the four methods respectively. The trends observed here are similar
to the trends observed for the rigid footing. As observed in the case of rigid footing,

the first three methods give almost same results, and the results obtained from the M E
method deviate from the other results as they approach failure because of the drifting of
stress points away from the yield surface.

Figures 3.32 and 3.33 show the CPU time taken by each analysis on an I B M R S / 6000

Model 355 workstation for the rigid and flexible footings respectively. Similar trend as for

the triaxial stress paths has been observed here. The Implicit method is the most expensive

one in terms of C P U time. For a given accuracy and stability, E P —PC and P P — PC
are the least C P U time consuming.

3.7 Conclusions
An efficient and accurate stress to strain algorithm can be very useful for various reasons.
Apart from its use in the stress-based finite element method and back predicting stress con­

trolled laboratory tests, it allows us to evaluate the accuracy of strain to stress algorithms.

Most of the finite element procedures used for solving geotechnical engineering problems
are displacement-based and require a strain to stress algorithm for integrating constitutive
models. Therefore, an efficient and accurate strain to stress algorithm is required for nu­

merically solving boundary value problems using the finite element method. Among the

four algorithms implemented in the present study, the elastic predictor - plastic corrector
and the plastic predictor - plastic corrector methods are found to be the most accurate, and

efficient in terms of C P U time. Though the implicit method produces accurate results, it
consumes much more C P U compared to the other methods. The Modified Euler method,

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
75

0.0

- 20.0
Implicit (large)
Settlement, mm

Im plicit (small)
-4 0 .0

Q o HP—PC
* — -■ PP-PC
-6 0 .0
o - — Implicit
v — v Imlicit (smaller increment)
a a "Correct”

-8 0 .0
0.0 40.0 80.0 120.0 160.0 200.0
Footing Load, kN/m

Figure 3.29: Load-settlement curves for the flexible footing (Aemax = 1 0 -2 )

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
76

0.0

- 20.0
Settlement, mm

-4 0 .0

-6 0 .0 o o EP-PC
* — « PP-PC
♦ Implicit
it ▲"Correct

-8 0 .0
0.0 40.0 80.0 120.0 160.0 200.0
Footing Load, kN/m

Figure 3.30: Load-settlement curves for the flexible footing (A^maI = 10-3)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
77

0.0

- 20.0

E
E
c
| -4 0 .0
0)

a
0)

-6 0 .0 0 G EP-PC
PP-PC
«— ^ Implicit
▲----- ▲Modified Euler

-8 0 .0
0.0 40.0 80.0 120.0 160.0 200.0
Footing Load, kN/m

Figure 3.31: Load-settlement curves for the flexible footing (Aemax = 10~4,T O L =
10 “ 5)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
78

500
CPU time, seconds

■ 1.00E-02
□ 1.00E-03
H1.00E-04
■ 1.00E-05

EP-PC PP-PC Implicit Mod.


Euler

Figure 3.32: CPU time taken by different algorithms for the rigid footing

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
79

250
CPU time, seconds

200
■ 1.0E-02
150 □ 1.0E-03
1 0 0 0 1.0E-04

50
0
it H
EP-PC PP-PC Implicit Mod.
■1.0E-05

Euler

Figure 3.33: CPU time taken by different algorithms for the flexible footing

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
80

Table 3.4: Summary of comparisons of different methods

Test Method Largest Aemai or Largest A e max Comments or


T O L for stability TO L for accuracy remarks

CTC E P -P C > 10~2 > 10"2 Implicit method

drained P P -P C > 10"2 > 10“ 2 unstable near

Implicit > 10“ 3 > 10~3 failure for

ME > 1 > 10"2 A s max = 10

TE E P -P C > 10"2 > 10~3 Implicit method


drained P P -P C > 10"2 > 10~3 deviates

Implicit > 10"2 > 10"3 most for

ME > 1 > 10"2 As-max — 1 0


CTC E P -P C > 10"2 > 10"2 Implicit method

undrained P P -P C > 10“ 2 > 10“2 unstable near

Implicit > 1 0 -3 > 10"3 failure for

ME > 1 > 1 A s max — 1 0

Rgid E P -P C > 10"2 > io ~ 2 Modified Euler


footing P P -P C > 10"2 > 10“ 2 method predicts

Implicit > 10"2 > 10"2 higher footing

ME > 10“ 5 > 10"5 load

Flexible E P -P C > 10 " 2 > 10~2 Modified Euler

footing P P -P C > 10"2 > 10"2 method predicts

Implicit > 1 0 -3 > 10-3 higher footing

ME > 10~5 > 10"5 load

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
81

though consumes less C P U time, shows deviation from the correct solution as the ma­
terial approaches failure. The elastic predictor - plastic corrector method and the plastic
predictor - plastic corrector method are very similar. The difference is in the stress points

used to calculate certain functions and derivatives. These two methods give similar results

with the same numerical effort for small strain increments. For large strain increments, the
stress point used to evaluate certain functions and derivatives becomes important It has

been observed that calculating the return path direction, nfjy at the initial stress point leads
to a more accurate solution. For the first three methods, it has been found that the subin­
crementation technique improves the accuracy of the solution considerably. The elastic

predictor - plastic corrector method has been chosen for all finite element analyses in the

present study. A summary of the results is given in Table 3.4.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
CHAPTER 4
CALIBRATION OF HiSS CONSTITUTIVE MODELS
USING GENETIC ALGORITHMS

4.1 Introduction
Any statically indeterminate load-deformation problem requires constitutive models for
the stress-strain behavior of all the materials involved. Before these models can be used to
solve the problem, whether analytically or by using the finite element method, the models

need to be calibrated based on laboratory tests. The calibration of a constitutive model can

be defined as the determination of appropriate values for the material parameters so that
the observed stress-strain behavior of the material matches the stress-strain behavior pre­

dicted by the constitutive model. The constitutive models for most engineering materials,

especially those which are geologic, have been developed based on the theory of elasticity
and plasticity. As mentioned in Chapter 3, the theory of plasticity requires definitions of

a yield surface, a plastic potential surface (which is the same as yield surface when the

flow rule is associative), and a hardening law. In most cases, definitions of a yield surface,

a plastic potential surface, and a hardening law require quite a few material parameters,
which makes the calibration of the model quite complex.
The traditional approach of calibration is to find certain well defined states in certain

laboratory tests where the behavior of a material is controlled by only a couple of material
parameters. Then the stress and strain tensors and other history parameters at these states

can be used to find those material parameters. Sometimes it is found that adding certain

material parameters to a model improves the prediction capability of the model substan­

tially; however, it is not possible to find an easy way to find these parameters using labo­

ratory tests. The effects of those parameters may be for the whole stress-strain response,

82

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
83

and in that case, we may not be able to isolate certain states where these parameters fully
control the response.

In the present work, how a GA can be used to calibrate constitutive models has been

shown. The method has been applied to the H iS S Si model described in Chapter 3.
Since a genetic algorithm {GA) can use the whole stress-strain response of all the tests
simultaneously, any material parameter can be used in the model without worrying about

how it can be determined from laboratory test data. Therefore, we have more flexibility in

developing constitutive models when we can use a GA to calibrate them.


The basic structure of a GA is founded on the basic mechanism of natural evolution. It

is a selective, efficient random global search algorithm, and it is widely used in optimiza­

tion and machine learning problems in Artificial Intelligence, commerce and engineering.
GAs have been used to solve various problems in different fields of Civil Engineering.

Simpson and Priest (1993) applied genetic algorithms to identify the maximum disconti­
nuity frequency in a complex rock structure. Wu and Chow (1995) formulated a technique

using genetic algorithms for discrete optimization of trusses. Koumousis and Georgiou

(1994) used genetic algorithms for discrete optimization of steel truss roofs, whereas Adeli
and Cheng (1994) applied genetic algorithms for optimizing space structures. The topo­

logical optimization of trusses using genetic algorithms was carried out by Hajela and Lee

(1995). Tesar and Drzik (1995) used genetic algorithms for resonance tuning and the dy­

namic balance of structures. Chakroborty et al. (1995) applied genetic algorithms in the
field of transportation engineering where they optimized urban transit systems. Genetic
algorithms were also used in road maintenance planning by Chan et al. (1994) and Fwa et
al. (1994).

A genetic algorithm employs the Darwinian survival-of-the-fittest theory to promote


the best characteristics among a population of partial (semi-optimal) solutions (De Jong,

1975; Goldberg, 1989; Goldberg and Samtani, 1988; Syswerda, 1989; Davis, 1991; Ha­
jela, 1990). It performs a random information exchange (cross-over and mutation) to

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
84

create new potential solution points, or offspring. There are three fundamental steps asso-

dated with a genetic algorithm. The first step is the coding of the parameters by binary
strings, called chromosomes. One can also use direct coding with real numbers (Eshelman

and Schaffer, 1995; Wright, 1991). The second step is defining an evaluation function

which gives the fitness value for each set of parameters (chromosomes). The third step

is the search algorithm for finding the optimum solution (chromosomes) in which one

population of chromosomes is considered at a time and new populations are created by


“cross-over" (and other operations) on the chromosomes in the previous population. The
details of each of these steps are given later.

4.2 Traditional method


As described in Chapter 3, there are seven material parameters, in addition to two elastic
parameters, associated with the H iS S Si constitutive model: 7 , /?, m, n, a: , rji and k. The
description of the traditional method (Desai and Wathugala, 1987) for determining these
material parameters is given below. After that the proposed GA is described.

Ultimate parameters ( 7 , P and m ) :


The value of m for many geologic materials is found to be equal to -0.5. The param­

eters P and 7 are evaluated as follows: At ultimate state a ps = 0 and from Equations 3.1
to 3.4, the following relation can be derived;

(4 .1 )

If the ultimate stress for at least two stress paths with different Sr (i.e. compression and
extension) arc available, Equation 4.1 can be solved for 7 and p. If ultimate stresses for

more than two stress paths are available, an optimization scheme such as least squares
(Desai and Siriwardene, 1984) may be used to evaluate 7 and >3. If all the test data have
the same ST (i.e., triaxial compression only), then an additional assumption, such as the

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
85

friction angle in compression, <\>c, and friction angle in extension, (j>s, are the same, is used
(Desai and Wathugala, 1987; Wathugala, 1990).

Phase change param eter (n):

The value of n is determined at the state of stress at which the plastic volume change

(efep) is zero. Stress points where this phase change occurs for associative materials lie
on a straight line passing through the origin. This straight line is called the “phase change

line", and the slope of this line is found as (Wathugala, 1990):

(4.2)

Using the known values of 7 and /3 and the stresses at which the plastic volume change
(de%) is zero, n can be determined from the above expression (Equation 4.2). Assumption
of the associative flow rule in Equation 4.2 could lead to some error for material with a

high value of k .

Hardening param eters (ai and rji):


Taking the natural logarithm of both sides of Equation 3.5 yields

In aps = In ai —rjiln £ (4.3)

Using the consistency condition in the theory of plasticity, ap3 for any point in a stress
control laboratory test can be found from

£=f = /(dZijdZij)1^2 can also be computed from the laboratory test data. Substituting
the values of aps and f in Equation 4.3 leads to the number of simultaneous equations
equal to the number of points in the observed data. These equations can be solved for
ln(ai) and 771 using the least square procedure.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
86

Nonassoriative parameter ( k ):
From Equation 3.12, k can be expressed as

* = («„ 1 « , ! ) ( ! (4-5)

or<3 which is a function of (a, s, /?, 7 , n ) (Equation 3.12), can be calculated at each point of
the observed curve. Then k can be determined at each point from Equation 4.5. Generally,

the portion of ev - £\ or - \JI%D curve near the ultimate condition is used for determining
k because the deviation (ocq - aps) is the greatest in the ultimate zones.

4.3 Coding and decoding of parameters


We use here a simple binary coding of the parameter values. The length of binary string to

be used to code a parameter depends on the required accuracy. A real variable X whose

range is a < X < b can be discretized and coded using a binary string of length L as:

C = a+ - a) (4.6)

where C = value of the parameter the string represents and B = the decimal integer value
of the binary string. As an example, for L = 6 the string 000000 corresponds to the lower

bound a, and the string 111111 corresponds to the upper bound b. In this case, C can
be one of the 2s = 64 values equally spaced between the lower bound, a, and the upper
bound, 6 . If a = 0, 6 = 10, and the binary string is 010110, the decimal value of the string
B = 22 and C = 3.492. Note that a more general scheme would be to use a nonlinear

discretization where a table of size 2 L can be used to translate each binary string to a real
number in the range a < X < b. Such methods are useful when different points of the

range a < X < b need to be searched or examined at different resolutions. In general, a

constitutive model always has more than one parameter. To code the multiple parameters,
binary strings representing the parameters are simply concatenated in a certain order, and
a single string is generated which represents all the parameters. For example, if we have
six parameters A u A 2 , A3, A*, A 5 and A 6 and each one is coded with six bits (L = 6 ), the

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
87

final string will contain 36 bits. Suppose A \ = 110001; A 2 =010101; A$ = 101010; A \ =


111000; A$ = 000111; and A 6 = 100110; then the string becomes

110001 010101 101010 111000 000111 100110

Ai A2 A3 A .i >15 Aq

One can, of course, use a different string length (L ) for different parameters depending

on the required precision. To decode such a string, it is decomposed into components of

appropriate length and then decode each part using Equation 4.6. There are seven material

parameters associated with the H iS S Si model in addition to the two elastic parameters E

(Young’s modulus) and u (Poisson’s ratio). The determination of elastic parameters is very

simple for the H iS S model because the model uses linear elastic relationship. Therefore,
they were not used in our Gi4-based search method. Out of these seven parameters, the
parameter m is taken as a constant (= -0.5) for every material. This leaves six parameters:

ultimate parameters 7 and /?; phase change parameter n; hardening parameters a Land 771;

and non-associative parameter k. Since the range of hardening parameter a t is very large,

a binary string to code ln(a 1 ) is used instead of ai. In the present study, each material
parameter has been represented by 10 bits; therefore, each string contains a total of 60 bits,

giving a search space of 260 possible combinations of the values of six parameters. In this

study, the maximum number of generations used is 500, and the size of the population used
is 50. Therefore, 25000 sets of parameters are examined out of 260 parameter sets in the

search space. Computing the fitness value for a given parameter set takes approximately
0.3 second of CPU time on a IBM RS/6000-355 computer.

4.4 Fitness function


In our problem, a binary string, or chromosome, represents a set of material parameters.
The fitness function evaluates a scalar value called fitness value for each set of parameters
or chromosomes. The fitness value represents the quality of the binary string as a solution,

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
88

and the goal of a genetic algorithm is to minimize or maximize this fitness value depend­

ing upon the problem. In the calibration problem, our goal is to find a set of material
parameters for the constitutive model which predicts the measured stress-strain behavior

of laboratory tests as closely as possible. So the fitness value could be obtained by com­
paring the observed stress-strain behavior of laboratory tests with the behavior predicted
by the constitutive model using the set of material parameters represented by a string.

In the present study, a strain path is predicted for the stress path used in the laboratory

test using the H i S S model. We use three types of stress-strain plots. They are I x vs. J x,

\Zho vs. y/JiDi and \JIid v s. -I\ where I\ = first invariant of the strain tensor, / 2D =
second invariant of the deviatoric strain tensor, Jx = first invariant of the stress tensor, and

J -id - second invariant of the deviatoric stress tensor. These three curves are used because
they represent the stress-strain behavior of a material in full stress and strain spaces. For
each stress-strain curve, a fitness value is defined as the ratio of the area between the

predicted curve and the laboratory test curve to the area of the rectangle generated by the
maximum and the minimum values of stresses and strains of the laboratory test ( area of

rectangle = ((maximum stress - minimum stress) x (maximum strain - minimum strain))

as shown in Figure 4.4. Note that the ratio is independent of the scales used for stress and

strain. We denote this fitness value as F{ , where i = 1, 2 ,3 representing the three curves.
For cyclic tests, this definition is valid for each cycle, and the rectangle which contains the
cycle is used in the denominator. It should be noted here that since we are interested in
the minimization of the area-ratio, the higher the value of fitness function, the lower the

fitness of the parameter set. In short, the fitness value represents the degree of “unfitness"
in this problem.

Now, the ultimate strain values of the predicted curves are checked and a penalty de­

pending on the deviation of the predicted ultimate strain values from the observed ultimate

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
89

strain values is added. A quadratic penalty function is used and defined as:

Penalty = + 72 (l - ) (4.7)
\ \J 2D ,u lt / \

where y/I2D,rdt - Predicted ultimate value of y/I2 V^ 2 D,uit = Observed ultimate value
of \ / I 2d , / £ * = Predicted ultimate value of I x, = Observed ultimate value of I x,

and 7 1 and 7 2 are two scalar multipliers. A similar penalty function is used by Adeli and
Cheng(1993, 1994) to optimize space structures. The penalty is related to the horizontal

distance between the points B and D in Figure 4.4 and forces these points to be close

to each other (in addition to minimizing the shaded area). The deviation of the ultimate

strain value of the predicted test result from the real test result indicates that the set of
parameters corresponds to a stronger or weaker material than the real one, and by adding

a penalty depending on that, eliminates such bad parameter sets, i.e. strings, from the

population. It has been observed that adding the penalty improves the accuracy of the
solution considerably.

Therefore, the final fitness value, / , for each test becomes:


2 2
ult ' (4.8)
\ j ^2D ,u l t j V L l'ult

where i = 1, 2,3. For multiple test data, the fitness values of all the tests are averaged and

that average value is taken as the final fitness value. Finally, the fitness value is multiplied

by 100 to represent it as a percentage. Our goal is to find the optimal combination of the 6
parameter values that minimizes this final fitness value, using a genetic algorithm.

From many laboratory test results, the stress tensors associated with the stress path and
the corresponding strain tensors can be evaluated. Thus a genetic algorithm can be used
to evaluate material parameters from many conventional laboratory tests. Therefore, the

restriction on using certain types of tests to find material parameters by traditional meth­
ods can be eliminated, resulting in flexibility in the development of constitutive models.

However, it should be noted here that if a parameter has little affect on the behavior of a

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
90

material for a certain stress path, then the accuracy of that parameter will be much less
certain.

4.5 Genetic algorithm (G A )


We are using here a simple form of a genetic algorithm. There are more sophisticated
variations available (Goldberg, 1989; Davis, 1991). The three basic operations for creating
a new population from an old population are: (1) reproduction, (2) cross-over, and (3)
mutation. Reproduction is a process in which a decision is made on the number of copies

of each string (chromosome) that will go into the mating pool. This decision is based on

the fitness value of the string. In a maximization problem, strings with a higher fitness
value have a higher probability of survival and consequently,proportionately more copies
are made in the mating pool to create the next generation. Similarly, in a minimization

problem such as the one considered here, a string with less fitness value has a higher
probability of survival, and more copies are produced in the mating pool. For that reason,
the fitness value of each chromosome is redefined as

fi = /max - Si (4.9)

where f max is the maximum fitness value in the population. For a population of size n, the
probability of the i-th string to survive is:

Pi = (4.10)
‘-‘Ji

where /, is the fitness value of i-th string. The number of copies of the z-th string to be
made for the mating pool is given by (approximately, differing at most by 1 ),

7ii=nxpi (4.11)

This process is analogous to natural selection in the real world. In a more complex form of
GA, a certain percentage of the population may be copied directly to the next generation

in order to make sure that some of the best chromosomes survive and are not destroyed by

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
91

mutation or cross-over. The fitness values are rescaled so that the best chromosome will

have a probability to survive by a factor, called M A X - M I N factor, times the survival


probability of the worst chromosome. The M A X - M I N factor induces a relative measure;

how good the best chromosome is compared to the worst chromosome. In the present
study, a M A X - M I N factor of 2 has been used for all the runs. The next generation is
created by cross-over from the mating pool and followed by mutation. This is the same as

mutation first and then cross-over. Cross-over is a random and relatively simple process

in which no prior knowledge about the fitness value of an individual string is necessary.
In this process, two mating strings swap a certain number of bits and produce two off­
spring which have new characteristics derived from their parent strings. There are many

ways a cross-over operation can be implemented in a GA. In the present work, cross­
over is performed as follows. First, two participating parent strings are chosen at random
from the mating population. Then a position (the same position for both the strings) in

the strings is chosen at random about which bits are exchanged after the crossover point

between the two parent strings. The process is depicted in Figure 4.2 Much of the power
of a genetic algorithm relies on reproduction together with cross-over.

In the present problem, each chromosome has six components representing six material

parameters. The cross-over in the present study has been implemented by choosing a
random position for each component and swapping bits of that component of the two
parent chromosomes; i.e. cross-over takes place at six points in a chromosome. Figure 4.2

can be considered as the representative of one component The mutation operation is used
mainly to prevent loss of some characteristics of the population and to reduce duplicate
members in the population. The rate of mutation is generally low, a fact which conforms

to the real world where mutation takes place only occasionally (De Jong, 1975).

After reproduction, if some bits have the same value (0 or 1) for all the strings in the
mating population, then the same would be true for all off-spring and for all successive

generations if mutation is not used. Mutation, which is a process in which some bits,

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
92

curve obtained from laboratory stress-strain measurement

Area ODB
Fitness value -
Area OABC
>
a
curve predicted by GA

Invariant of strain

Figure 4.1: Definition of fitness value

Parents:
0 1 1 0 0 0
Chosen randomly from
the mating population 0 10 1 1 0

Random position, about which


bits are exchanged

Off-spring: 0 1 1 0 10

0 1 0 1 0 0

Figure 4.2: Schematic diagram of cross-over

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
93

depending on the mutation rate, are changed in all the strings, acts as a safeguard for that

situation. It could also make two strings different which were similar before the mutation
operation. Though mutation is a secondary operation in a GA, sometimes it plays a vital

role in the search operation, especially in the situations mentioned above. A mutation rate

of 2% is used in the present study. The cross-over rate used is 100%, i.e. each and every

string (chromosome) in a population participates in the cross-over operation. The initial

population is created randomly. We tried with different initial populations, and as one

might expect, the results do vary from one run to the other. In our case, except for k, all
other parameters showed very small variations (< 5%); but k varied between 10% and

40%. A flow chart of genetic algorithms is shown in Figure 4.3.

4.6 Examples
In this section, three numerical examples of the calibration of the H i S S Si model using

GA are presented. In the first example, three simulated conventional laboratory test data

have been used to calibrate the model. The reason for using the simulated test data is that
the answer is known in this case. Thus, it is possible to compare the parameters obtained
by GA with the input parameters and to test the robustness and accuracy of the algorithm.

In the second example, simulated cyclic test data have been used. In the traditional method,

the material parameters cannot be evaluated from a cyclic test result. Here, the cyclic test
data have been chosen to test the usefulness of the present approach in a non-traditional

situation. The reason for choosing the simulated data is the same as before. In the last
example, three conventional real test data have been used to show the applicability of the
algorithm to a real problem.

4.6.1 Simulated test data

Three triaxial tests were simulated: (1) a triaxial compression (TC, Ji = constant, ct\ in­

creasing, 02 = 0 3 decreasing) test at a confining pressure of 100 kPa, (2) a conventional

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
94

INITIAL
POPULATION

FITNESS
EVALUATION

MATING
POOL

CROSS-OVER

MUTATION

NEW GENERATION

Figure 4.3: Row chart of genetic algorithm

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Reproduced
with permission
of the copyright owner.

Table 4.1: Material Parameters for Simulated Triaxial Tests

Material parameters and Parameter Parameter values Parameter values

their ranges used in GA values used obtained by using obtained by traditional

for simulated GA and percentage method and percentage


Further reproduction

test difference difference

7(0.04-0.11) 0 .1 0.103 (+3.0%) 0.098 (-2.0%)

P (0.2-0.7) 0.4 0.375 (-6.25%) 0.428 (+7.0%)


n (2.1-4.0) 2.5 2.627 (+5.08&) 2.442 (-2.32%)
prohibited without p e r m is s io n .

In at (-11.513 to -6.908) -8 . 1 1 2 -7.862 (-3.07%) -6.941 (-14.42%)

771 (0.4-1.0) 0.9 0.99 (+10.0%) 0.725 (-19.44%)

k (0-0.5) 0.05 0.0425 (-15%) 0.072 (+44.0%)

VO
96

500-,

400-

300-

200 -

100 Sim ulated T e st Data


Traditional Method
G enetic Algorithm

0.0 0.1 0.2 0.3 0.4

IL

(a) y/IiD - \ /h D curves

0 . 1-1

0.0
0 0 0.1 0.2 0.3 0.4
- 0. 1-

-0 .2 - S im ulated T e s t D ata
L
Traditional M ethod
-0.3- G enetic Algorithm

-0.4-

-0.5-

-0 .6 -

(b) \J I i d - h curves

Figure 4.4: Simulated C T C test (confining pressure 150 k P a )

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
97

110i

100 -

90-
80-

20- S im ulated T est D ata


Traditional Method
10- G enetic Algorithm

0.00 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09

(a) y j l w - y/JiD curves

0 .02-1

0.00
0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09
- 0 . 02 -

-0.04-
/, -0.06-

-0.08-

- 0 . 10 -

-0 .1 2 - Sim ulated T est D ata


Traditional Method
-0.14- G enetic Algorithm

-0 .16 J

(b) y / h o - h curves

Figure 4.5: Simulated C T C test (confining pressure 100 k P a )

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
98

180-1

160-

140-

120 -

100 -
(kPa)
80-

60

40- Sim ulated T est D ata


Traditional Method
20-
G enetic Algorithm

0.00 0.02 0.04 0.06 0.08 0.10 0.12 0.14 0.16 0.18 0.20 0.22

(a) y/h D - \JJ id curves

0.1 -|

0.0
P4 0.06 0.08 0.10 0.12 0.14 0.16 0.18 0.20 0.22

E2D
I -0 - ' -

- 0.2 -

Simulated T est Data


Traditional Method
-0.3- Genetic Algorithm

(b) y j h o - h curves

Figure 4.6: Simulated C T C test (confining pressure 200 kPa)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
99

triaxial compression (C T C , a\ increasing, cr2 = 0 3 constant) test at a confining pressure of

200 kPa, and (3) a triaxial extension (TE, J x = constant, <rx decreasing, cr2 = cr3 increas­
ing) test at a confining pressure of 200 kPa. Later, the data from these simulated tests were

used to evaluate material parameters using GA. The purpose of this study is to evaluate

the performance of the algorithm, i.e. to see whether GA is able to give back the original

parameters used to simulate the tests. The multipliers 7 1 and 72 for the penalty function
were taken as 10. Material parameters by conventional approach were also evaluated us­
ing this simulated test data. Table 4.6.1 shows the material parameters used to create the
simulated test data, those obtained from GA, and the material parameters determined by
the traditional approach (Desai and Wathugala, 1987). Note that the nearest values the
parameters can take on in GA, due to discretization, match with the real values up to third

decimal place. Therefore, they are not shown in this table. It can be observed from this ta­

ble that the material parameters obtained using GA are closer to the actual parameters used
to simulate the triaxial compression tests. Figures 4.4 to 4.6 show the predicted curves by

GA and the traditional method together with the simulated test data. It can be observed
from Figures 4.4 to 4.6 that in general, the GA predicted strains near failure better than the

traditional method. For example, in each of Figures 4.4 to 4.6, the termination points of
the curves for simulated test data and those obtained by GA are very close to each other.

However, for some of the predictions, the traditional method had predicted better at low
strains. In GA, by using a suitable fitness function, it is possible to give weightage to any
property of the stress-strain curve. In the present study, more weightage has been given for

the ultimate strains through the penalty function in Equation 4.8. Overall, both the meth­
ods gave the acceptable results in this example. However, it was found that for materials

with higher k values, GA gives better results than the traditional method which is reflected
in the results for the real test data described latter.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
100

Table 4.2: Material Parameters for Simulated Cyclic Test

Material parameters and Parameter Parameter values

their ranges used in GA values used obtained by using


for simulated GA and percentage

test difference

7(0.04-0.11) 0.089 0.092 (+3.37%)

0 (0.2-0.7) 0.442 0.413 (-6.56%)

n (2.1-4.0) 3.0 2.751 (-8.3&)

In ai (-11.513 to -6.908) -8.634 -8.202 (-5.0%)


rh (0.4-1.0) 0.850 0.970 (+14.12%)

k (0-0.5) 0.251 0.327 (+30.28%)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
101

700-1

600-

500-1

40i

00 -
Simulated Test Data
Genetic Algorithm

-0 . 0 1 0 . )0 0.01 0.03 0.04 0.05 0.06


-100

^ZOi Si - 8

(a) - e2 - cn —rr2 curves

0 . 02-1

0 .0 1 -

- 0.01 0.02 0.03 0 .0 4 0.05 0.06

- 0 .0 2 -
% - Simulated Test Data
Genetic Algorithm
-0.03 J

(b) E\ —£ 2 - £v curves

Figure 4.7: Simulated cyclic test

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
102

4.6.2 Simulated cyclic test

In the traditional calibration method, we cannot use a single cyclic test to find material

parameters. However, GA could be used even for this case. A cyclic triaxial test (<jx cyclic,

<72 = 0 3 = constant) was simulated with the material parameters given in Table 4.6.1. The
test was not run up to failure. Table 4.6.2 also shows the material parameters predicted
by GA. The multipliers 71 and 72 were taken as 20 for this run. These values were set

because we want to match the ultimate strain values more closely since the test was not

run up to failure. The plots of a x —0 2 and ex —e2 were used instead of y/J2D and y/T2D to
capture the sign (positive or negative) of these quantities. Figure 4.7 shows the simulated

test data and the predictions by GA. Material parameters obtained from GA are close to

the material parameters used to simulate the test except for rji and k; the deviation for k
is especially large. The reason for this large deviation for k could be that the cyclic test

was not run up to failure and the effect of k in the behavior of the material is maximum

near failure. The deviation of rji could be a shortcoming of the fitness function. Despite

the fact that only one simulated non-traditional test result was used, and that was also not
run up to failure, GA was able to estimate the parameters satisfactorily.

4.6.3 Real test data

Three laboratory triaxial tests on Leighton Buzzard Sand (Hashmi, 1986) were used in
this example. These three tests were two C T C tests at confining pressures of 90 kPa (13

psi) and 34.5 kP a (5 psi) and one triaxial extension (TE) test at a confining pressure
of 138 kPa (20 psi). The multipliers 71 and 72 for the penalty function were taken as

10 in this case. Table 4.6.3 shows the material parameters obtained using GA and those

obtained by traditional approach. To compare how well GA evaluates material parameters


compared to the traditional approach, predictions have been made for all the tests using

the material parameters obtained by the traditional method and those obtained by GA.
Figures 4.8 to 4.10 show plots of predicted curves together with laboratory test data. It is

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Reproduced
with permission
of the copyright owner.

Table 4.3: Material parameters obtained from real tests

Material parameters and Parameter values obtained Parameter values obtained


Further reproduction

their ranges used in GA by using GA by traditional method

7(0.04-0.11) 0.098 0.089

P (0.2-0.7) 0.45 0.442

n (2.1-4.0) 2.56 3.0


prohibited without p e r m is s io n .

ai (1 . 0 x 1 0 " 5 - 1 .0 x 1 0 " 3) 1 .0 x 10"4 1.78 x 10" 4

rji (0.4-1.0) 0.981 0.852

K (0-0.5) 0.356 0.251


104

90-1

80-

70-

60-

50-

40-

30-

20-
Genetic Algorithm
Traditional Method
Test Data

0 .0 0 0.01 0.02 0.03 0.04 0.05

tJf 2D

(a) y / h o - V J 2 D curves

0.002

0 .0 0 0
0.02 0.03 0.04 0.05
- 0.002-

t -0.004-
il
-0.006-

-0.008-

- 0. 0 10-
Test Data
- 0. 0 12- Traditional Method
Genetic Algorithm
-0.014 J

(b) y/l-io - h curves

Figure 4.8: Real C T C test (confining pressure 34.5 k P a )

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
105

220-|
200-

180-
160-
140-
120 -
ikPa)
100-

80-
60-
40- Test Data
Traditional Method
20 Genetic Algorithm

0.00 0.01 0.02 0.03 0.04 0.05 0.06 0.07

IL

(a) y/ho - y/J-io curves

0.0 0 4 -

0 .0 0 0
0.30 0.0 0.02 0.03 0.05 0.06 0.07

-0.004-
/,
-0.008-

- 0 . 0 12 -
T est Data
Traditional Method
-0.016- Genetic Algorithm

- 0 . 0 2 0 -1

(b) y / h o - h curves

Figure 4.9: Real C T C test (confining pressure 89.6 kPa)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
106

110-.
100 -

90-
80-

50-
40-
30
20- T est D ata
Traditional M ethod
G enetic Algorithm

0.00 0.01 0.02 0.03 0.04

IL

(a) y / h o - \ / J id curves

0 .01-1

0.00
0.02 0.03 0.04

/,

- 0 . 02 -

T e st Data
-0.03- Traditional Method
Genetic Algorithm

-0.04-*

(b) \fliD - h curves

Figure 4.10: Real C T C test (confining pressure 138 kPa)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
107

observed from these figures that the curves predicted by GA are closer to the laboratory
test data than the curves predicted by the traditional method, except for the lower portion

of the curves in Figures 4.8(a), 4.9(a), and 4.10(a). Note that in these figures, the deviation
of the upper portion of the curves predicted by the traditional method is larger than the

deviation in the lower portion showed by GA. Also in Figures 4.8(b), 4.9(b), and 4.10(b),

the deviation of the curves predicted by the traditional method is significant, whereas the
curves predicted by GA are very close to the actual material behavior. Considering the

overall behavior, the material parameters obtained by GA are more acceptable than those
obtained by the traditional method.

4.7 Performance of two different cross-over schemes


In conventional G A , cross-over is done with respect to one point To compare the per­

formance of one-point cross-over with six-point cross-over, one problem was run with the
simulated test data using one-point cross-over. Figure 4.11 shows the plot of the best fit­
ness value in a generation vs. that generation number for both cross-over methods. It can

be observed that the six-point cross-over improves the convergence rate. It also achieves a

lower fitness value which results in more accurate material parameters. It should be noted
that the set of parameters corresponding to the lowest fitness value in the whole genera­

tion is reported as the best set. not the best parameter set in the last generation. For the

one-point cross-over scheme, the best parameter set was achieved in the vicinity of the
generation 400 (Figure 4.11).

4.8 Discussion and conclusion


In the present study, a random search technique, genetic algorithm (GA), is used to cali­

brate the Hierarchical Single Surface (H i S S ) <Ji model. The results of this study demon­
strate how a modem and relatively simple random search algorithm, GA, can be used to
determine material parameters for a constitutive model. The main advantage of using GA

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
108

1500.0

Six point cross-over


O ne point cross-over

1000.0
Best Fitness Value

500.0 -

M ail ii. . I , . , I .|L It l i t 4 .1 .. • 1 . td llM a . I ll . II

0.0
0 100 200 300 400 500
Generation

Figure 4.11: Performance of one-point cross-over and six-point cross-over schemes

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
109

over the traditional method is that it takes care of the global characteristics of the test re­

sults, i.e. behavior at each and every point in a stress or strain path, not the characteristics
of the test result at some specific points or states, such as ultimate, phase change, etc.

Apart from that, GA can be used to find material parameters when only non- traditional

test results are available, in which case a traditional method cannot be used. The tradi­
tional method of calibration is sequential; i.e. it finds one or more parameters first and

later uses these values to find the other parameters. If any error occurs in the parameters

found initially, the parameters found later become affected. Sometimes, two or more pa­

rameters are interdependent, and there are no explicit relations available to solve for all
the parameters. Some assumptions are generally made to handle such situations.

For example, the value of the phase change parameter, n, must be known in order

to find the value of the nonassociative parameter, k. However, the assumption that the
volumetric plastic strain changes sign at the phase change line is not true for the nonas­
sociative model. The higher nonassociativeness may give an erroneous n value which
eventually triggers an erroneous k value. In such situations GA works better than the tra­

ditional method. The way in which the fitness value of a string is evaluated in the present
study is capable of capturing the overall trend of the laboratory test results. The penalty,

depending on the ultimate strain values of the predicted result, and the real test result im­

proves the accuracy of the solution considerably. For the present problem, this deviation
indicates that the set of parameters corresponds to a material stronger or weaker than the
real one, and by adding a penalty depending on that, eliminates such bad parameter sets,
i.e. strings, from the population. The penalty function used in the present study may not
be the best one. It should be noted that one may wish to consider the behavior of a material

at specific states (e.g. phase change, ultimate, etc.) In that case, an appropriate penalty
function can be added to the definition of fitness function to capture that behavior.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
CHAPTER 5
DISTURBED STATE CONCEPT BASED INTERFACE
MODEL AND ITS IMPLEMENTATION IN FEM

5.1 Introduction
Interfaces between two dissimilar materials are very common in many geotechnical engi­
neering problems. Building foundations have interfaces between structural elements and

adjoining soils. Interfaces exist between reinforcement materials and adjoining soils in

reinforced earth structures. Geosynthetic materials, such as geogrids and geotextiles, are
used extensively as reinforcements in earth embankments. To study the stress-deformation

behavior of such structures through numerical simulation, it is very important to model the

interfaces accurately. The present trend of modelling the interfaces between geosynthetic
reinforcement and soil is to use linear or nonlinear elastic models where the onset of slid­

ing is defined using a limiting shear stress criterion such as Mohr-Coulomb (Bathurst et

al., 1992; Ho and Rowe, 1994). The behavior in normal direction is generally defined em­
pirically for the bonded and debonded states. The hyperbolic model (Clough and Duncan,

1969) has been used by some researchers (Tavassoli and Bakeer, 1994) to model the in­
terface between reinforcement and soil. It is assumed in this model that the bond strength
is governed by the Mohr-Coulomb criterion, and once the bond strength is exceeded, the
shear strength of the interface is automatically reduced to a value close to zero to model a
slip condition.

The shortcomings of these models are that they describe the shear behavior of the in­

terface in an approximate sense and do not consider the coupling of normal and shear be­

havior. The models are incapable of capturing restrained dilation, which is one of the gov­
erning factors for pull-out load capacity of a reinforcement (Farrag et al., 1993). Recently,

110

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
I ll

some researchers have proposed and used elasto-plastic models for the geosynthetic-soil
interface (Matsui and San, 1989; Gens et al., 1993).

In the present study, an elasto-plastic constitutive model has been developed based on

the disturbed state concept (D SC) (Desai, 1992) for the geosynthetic-sand interface. A
similar concept has been used to develop constitutive models for interfaces (Navayogara-
jah et al., 1992), rock joints (Ma, 1992; Desai and Ma, 1992), and solids (Wathugala and

Desai, 1987; Wathugala and Desai, 1989; Katti, 1991; Katti and Desai, 1995; Armaleh,
1991; Armaleh and Desai, 1994; Basaran, 1994). The proposed model has been imple­
mented in the finite element method and used to simulate large scale pull-out tests with
Tensar and Conwed geogrids.

5.2 Disturbed state concept


The disturbed state concept is based on the idea that the response of a material can be

related to and expressed as the responses of the reference states (Desai, 1987). As a result,

the observed behavior of a material is thus treated as a disturbance to the behavior of the
reference states. This disturbance takes place in the process of certain physical changes:

alteration in density and micro-structural changes. The latter is analogous to the concept

of damage caused by micro-cracking and fracture. Nonassociativeness, anisotropy and


strain softening can be expressed as disturbances with respect to the reference states. In
the D S C approach, the observed material behavior is considered to be composed of two

reference states called relative intact (RI) and fully adjusted (FA). The materials in the
R I state behave as continuously hardening materials without disturbance. The F A state
may be assumed to be an invariant state during the stress deformation process such as the

ultimate, failure, and zero shear stress and/or volume change conditions. The progressive

damage model for concrete developed by Frantziskonis and Desai(1987) can be treated
as a special case of the disturbed state modelling method. The proposed model can be
considered as a specialized version of the damage based model for granular materials

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
112

proposed by Wathugala and Desai (1987, 1989). The constitutive model based on the
D S C can allow proper modeling of shear transfer, volumetric behavior and localized slip
in the interface zone.
With the onset of the deformation process in a soil mass, the affected soil particles

change their structural arrangement, which initiates the F A state. Soil particles affected
in such a maimer are assumed to be in a critical state (Schofield and Wroth, 1968). The

remaining part of the material is assumed to be in the R I state, as shown in Figure 5.1.
The disturbance in the soil mass is defined as
/ FA material

RI material

Figure 5.1: Disturbed state as a mixture of R I and F A states

where M is the mass of solids in the material, and M c is the mass of solids in the critical
state, and D represents the extent of disturbance in the material. Initially, D = 0 and it

can attain a maximum value of 1, which represents that the whole material is in F A state.
Frantziskonis and Desai (Frantziskonis and Desai, 1987; Desai et al., 1986) have proposed
the following expression to describe the disturbance

D = Du[l —exp(—A£i))\ (5.2)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
113

where Du is the ultimate or critical value of D, A and k are material parameters, and £d
is the trajectory of the deviatoric plastic strain. A schematic diagram of the disturbance
function is shown in Figure 5.2.

Co

Figure 5.2: Evolution of damage function

5.3 Idealization of interface


To describe the behavior of an interface, we need to define some variables in terms of
which the behavior is described. For the description of load - deformation behavior of
an interface, the variables are stresses and displacements. An interface is the contact area

of two dissimilar materials, and its physical properties are determined by the frictional
properties of the contact surface as well as physical properties of both the materials. To be
able to describe the behavior of an interface, idealization of an interface has to be made.

The interface is generally considered as a planar surface. The behavior of the interface can
be defined by a coordinate system where the direction within the interface plane is taken as
tangential direction and the direction normal to the interface plane is taken as the normal

direction. The schematics of a natural and idealized interface between two contacting

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
114

bodies are shown in Figure 5.3(a). During shear transfer from body A to body B, relative

slip may take place along the interface plane. This relative slip causes the particles in the
vicinity of the interface to displace and reorient themselves. Thus the interface can be

treated as an equivalent thin zone with a small finite thickness, t. According to the D S C ,
the interface zone consists of a mixture of R I and FA state materials, Figure 5.3(b).
Unlike solids, the interface is considered as two dimensional, and the deformation of

an interface is derived in terms of displacements rather than strains. The displacements at

the interface can be decomposed into shear displacement, u, and normal displacement, v
along the plane of interface and normal to the plane respectively. Similarly, the stresses

are decomposed into shear stress, r, and normal stress, cr, and defined as

rr — K
a - A (5.3)

where T and N are shear and normal forces acting on the interface and A is the area of the
interface plane.

5.4 Proposed model


The disturbed state concept approach of modelling is based on the hypothesis that the
observed behavior can be related to and expressed as the behavior of the reference states.
The material at two reference states, the R I state and the F A state, behaves differendy.

The material at the R I state has its own mechanical behavior, whereas the F A state is an

invariant state where shear stresses and normal displacements are stabilized. The observed
material behavior is the combination of behavior at R I and F A states.
The material behavior at the R I state can be described by the theory of elasticity,
plasticity or visco-plasticity. Since the geosynthetic/sand interface behavior is a nonlinear
strain hardening type, an elasto-plastic model would be most appropriate. In the proposed
model, the R I behavior is defined by an elasto- plastic volumetric hardening associative

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
115

Natural interface

Critical material

Intact material

(b) (c)

Figure 5.3: Idealization of interface: (a) Natural and idealized interface, (b) Interface at
disturbed state, (c) Definition of stresses and displacements (After Desai and Ma, 1992)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
116

constitutive model. The yield surface of this model is a specialized two dimensional form
(Desai and Fishman, 1991) of the H iS S 5$ model (Wathugala, 1990; Wathugala and De­
sai, 1991; Wathugala and Desai, 1993) for clays which is defined as:

where, r = shear stress

a - normal stress
pa = atmospheric pressure

a = hardening function

rii, n 2 , 7 = material parameters

The hardening function, a, is defined as:

<* = fs vh (5-5>

where, = trajectory of volumetric plastic strain


ai, 771 = material parameters

In the H iS S 6 q model, n 2 is taken as 2. In this model, the material parameter ni is called


the phase change parameter which governs the point of transition from compressive to

dilative volume change behavior. Since this is a volumetric hardening model, the material
fails when it reaches the phase change line, i.e. the shear stiffness of the material becomes
zero. The phase change line is the locus of the crest of yield surfaces, and it is nonlinear in

this model. When n 2 = 0, the phase change line becomes a straight line. The expression
for the phase change line can be derived as follows. Let us assume that

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
117

At the phase change line, Fb is maximum. Therefore, we can write

dFb crni_1 rU2 —l


_ = - a n i —------ (- 7n2— = 0 (5.7)
da Pani Pan*

Equation 5.7 can be rewritten as:

/ a \n 2 -m
(5.8)
° = l/W Vn i

From Equations 5.4 and 5.8, we obtain the expression for the phase change line as:

n2
1 - (5.9)
Pa \P a ) ntJ

A schematic figure of the yield surface is shown in Figure 5.4. Because of volumetric

Phase change line

CO Yield surface

CO

Normal stress, o

Figure 5.4: Yield surface for R I material

hardening, the elasto-plastic model described here does not show dilative behavior. The

dilative behavior can be viewed as caused by disturbance, and it is incorporated in the

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
118

present D S C model through the critical behavior. In a similar way, strain softening is also
incorporated in the present model.

The materials at the F A state in the proposed model are assumed to behave similarly

to those in a critical state in which no change in shear stress or volume occurs for a given

normal stress. Therefore, we need to relate the critical shear stress and the critical normal
displacement with the normal stress at the interface. The shear stress at the critical state is
related to the normal stress by Equation 5.10.

t c = Ma (5.10)

where M = slope of the critical state line. The critical normal displacement is related to
the normal stress by the following expression.

(5.11)

where a and b are material parameters. Equation 5.11 is similar to the pressure void ratio
relationship for sands.

To describe the average or observed behavior of interfaces, we assume that the interface

material is a mixture of material at the R I and F A states and the mixture is homogeneous

(Figure 5.3). The total area, A, can be divided into two parts, the R I part A* and the
F A part A c. Similarly, the total thickness, t, can be separated into two parts, tl and tc,
corresponding to the R I and F A parts respectively.

The shear resistance is provided by both R I and F A states, and the total shear resis­
tance is the sum of the shear resistance by the R I and FA states. Therefore, we can write
the following expression.
A r a = A ' t 1 + Act c (5.12)

Dividing both sides of Equation 5.12 by the total area A, we get

(5.13)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
119

Now, we define a disturbance function Ds as the ratio of the area of the FA state to the

total area as:

Ds = ^ <5-14>

Then, A*f A becomes


4* A — Ac
-a = - T ~ = l ~ D* (5’15)
since A = A1 + Ac. Therefore, Equation 5.13 can be written as:

r “ = (1 —D , y + D 3 t c (5.16)

The disturbance function can be evaluated by using an expression sim ilar to Equation 5.2.

Ds = Dtt[l - e x p ( - A s^ ) ] (5.17)

where D u, As and k3 are material constants.

Similar to Equation 5.12, the normal strains at different states can be related as:

tea
v = £‘4 + t cecv (5.18)

The above expression can be rewritten as:

teS = 7 fe* + 7 teS (5-19)

The quantities te", telv and tecv give average, R I, and FA displacements respectively.
If we define another disturbance function Dv as:

Dv = j (5.20)

Equation 5.19 can be written as:

va = (1 - Dv)vl + Dvvc (5.21)

noting that t = tx + tc. The disturbance function D„ can be expressed as:

Dv = D„[l - e a :p (-A „ ^ )] (5.22)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
120

where Du, A v and kv are material constants.


The normal stress and shear displacement are taken as the same at the three states
which can be expressed as:
aa = <rl = ac (5.23)

ua = uf = uc (5.24)

The use of two damage functions, in terms of area and thickness, can be viewed as that
the mixture of theR I and FA states is anisotropic. Desai and Ma(1992) have also used
different damage functions for their D SC model forrock joints. A schematic representa­

tion of the stress-strain responses of RI, F A and average states is shown in Figures 5.5
and 5.6.

5.5 Incremental stress-strain relations


For the verification of the model, the incremental stress-strain relation is necessary so that

the laboratory test results can be backpredicted. The incremental stress-strain relations are

also required for the implementation in the finite element method. First, the incremental
stress-strain formulation for the R I material is attempted, followed by the incremental
formulation of average material behavior.

The stress and displacement vectors of the R I material can be expressed as:

= (5.25)

and f \
i
= (5.26)

Since the plastic displacement does not cause any stress increment, the incremental stress
can be related with the increment in elastic displacement as:

{da1} = [C']e{dui}e (5.27)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
121

RI response FA response

Observed response
Observed response
RI response

FA response
00 00

Shear displacement, u Shear displacement, u


(a) (b)

Figure 5.5: Schematic representation of observed response: (a) strain softening and (b)
strain hardening

*S
c<o e<D
E
<
oo uo£
‘333
FA response
TWS* Q . 'n,
09
Observed response

O Shear displacement, u ear displacement, u


z Z .2
RI response RI response
S
s.
Observed response uo
FA response

Figure 5.6: Schematic representation of observed response: (a) compression and (b) dila­
tion

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
122

where [C7]e is the elastic stiffness matrix and can be expressed as:

Ks 0
[C]“ = (5.28)
0 Kn

K s and K n are elastic shear and normal stiffnesses, and shear and normal behavior is
considered uncoupled. Since the total displacement is the sum of elastic displacement and
plastic displacement, the elastic displacement can be written as:

{du’}e = {du1} - { d v t y (5.29)

By substituting Equation 5.29 into Equation 5.27, the following expression can be derived

{da1} = [C]e ({du1} - {du*}p) (5.30)

Using the flow rule of plasticity, the plastic displacement can be expressed as:

W V = A{!£}
where A is a scalar multiplier, and it can be obtained by using the consistency condition of
the yield function which is
T
f3Fr dF
o (5.32)

where

d£v = (dviPdv iP ) 2 (5.33)

By substituting Equation 5.31 into Equation 5.33, we get

*-({£}'{s: (5.34)

From Equations 5.32 and 5.34, the expression for A can be obtained as:

A= (5.35)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
123

Now, the incremental stress-displacement relationship can be obtained from Equa­


tions 5.30,5.31 and 5.35 as:

/ , , , .-r \
{da1} = [q«- p r { & } { < & } 'ter {du‘> (5.36)

Equation 5.36 can be written as:

{da1} = [C]ep{dtzi} (5.37)

where

[C]* = [C]e - w { & } { & } Tw (5.38)


12

The partial derivatives in Equation 5.38 can be evaluated from Equations 5.4 and 5.5 as
given below.
dF aI'M- ,in2 -
= otni —---- ~ 7^2- (5.39)
da1 Pani Pa" 2
5P _ 2r 1
(5.40)
dr* Pa2
d /r ni
= - a i Vi ( t t ] (m 11 (5.41)
~ 11 L

The incremental stress-strain relations for observed or average behavior can be ob


tained from Equations 5.16 and 5.23 as:
' 1 r * * *
dra dr 1 drc TC_ Ti
* > = ( l - D s)< ' + Ds i >+ - SdD (5.42)
doa da 1 1 dac 0
< 4 k 4 k 4
Since F A is at the critical state, we have

<Lt c = M dac = M do{ (5.43)


r x

drc 0 M f dri ) 0 M ( dul


or, < ►= [cp (5.44)
=
dac 0 1 0 1
< 4 I .

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
124

Now, from Equations 5.42 and 5.44, we can obtain


* ' t * r ’
dr° dux du 1 T c — T*
►= (i - d ,)[ c t < + DS[C)C<
► ►+ < dD
► (5.45)
daa dv* dv 1 0
* k 4 . 4 * 4

where,
0 M ep
[C]c = [Cl
0 1

Now, Equation 5.45 can be rewritten as:

[da}a = [L] {duY + {a }d D s (5.46)


r
dra
where, {da} = <
daa

[£] = ( l- D ,) [ C T + DJ[C]‘
f \
dul
{du}* = <
dv1

{a} =

5.6 Determination of material parameters


The elastic shear and normal stiffnesses, K s and K n, can be found from the unloading

slopes of the r vs. u and v vs. u curves as shown in Figures 5.7(a) and 5.7(b) respectively.

In the present study, K s is determined from the initial slope of the r vs. u curve. K n is
derived from the stiffness of soils because the normal behavior is mainly governed by the
soil around the geosynthetics. The vaule of K n is taken in such a way that the ratio of K n

to the normal stiffness of solid soil in a plane strain condition is the same as the ratio of
K s to the shear stiffness of solid soil in a plain strain condition.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
125

C/3
C/3

C/3

Shear displacement, u

(a)

o
2

Normal displacement, v

(b)

Figure 5.7: Elastic shear and normal stiffnesses: (a) Shear stiffness, K s, (b) Normal stiff­
ness, K n

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
126

Intact model: According to the model of R I material, the material fails when it reaches
the phase change line. The ultimate constants 7 and n2, and the phase change parameter
rxx, can be found using the equation for the phase change line (Equation 5.9). The value of

the ultimate shear stress, ru, is assumed to be asymptotic, and it is about 10 to 25 percent

higher than the peak stress if the material shows strain softening and lower than the peak
stress if the material shows strain hardening (Figure 5.8). From the plot of in(r„/p„) vs.

ln(a/pa), n 2 and 7 (1 — (n 2 /ni)] can be determined (Figure 5.9). From the parametric
study it is found that ni —n 2 « 0.01 works satisfactorily for this model. Therefore, for
all practical purposes, n\ can be taken a sn 2 + 0.01. With ni and n 2 , 7 can be determined
using the value of 7 [ 1 —(n 2 / n x)].

The hardening parameters, ai and 7^ , can be found from the plot of ln(a) vs. ln(£v)
(Figure 5.10). or can be found for a number of points on a given stress-displacement

response using Equation 5.4. Since the R I model produces only compressive normal

displacements, it is not possible to calculate £v from the volume change curve of any
tests. However, it can be obtained from the stress path and the shear displacement as

follows. Calculate d F /d r at a number of points on a stress-displacement response, and

from u at each point, calculate A A as A A = , where A up = A u — Now, vp can

be computed as A v p = A A |£, and A£y can be computed as A£^ = (dvpdv p ) l / 2 at each


point.

Critical State: The value of M can be found from the slope of r c vs. a plot (Figure 5.11).
The values of a and b can be determined from the plot of vc vs. ln(a/pa) (Figure 5.12).

Disturbance Function: The maximum value, Du, of the disturbance function is taken as

0.9. Taking logarithms twice in Equation 5.17 leads to the following expression:

ln[—ln( 1 - — )\ = l n k a + A 3 In (5.47)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
127

The value of D. can be calculated from

r* — ra
Ds = ----- (5.48)
r* — t c

From the plot of ln[—ln( 1 —j^)] vs. ln(£D), k 3 and A s can be determined (Figure 5.13).

Similarly, kv and Av can be determined from the plot of ln[—ln( 1 —^-)] vs. /n(£o) where
Dv is computed from
v%—va
D« = V - 'jiC
/»* — r <5-49)

5.7 Implementation in finite element method


The proposed interface model has been implemented in the finite element method using
the thin layer element proposed by Sharma and Desai (1992). The details of this element

can be found in several studies (Desai et al., 1984; Sharma and Desai, 1992). However, a

brief description is provided for completeness.

The formulation of the thin layer element has been developed as a continuum element,
whereas its constitutive response has been defined differently from that of the neighboring

solid elements. The response has been defined only in terms of the normal and shear

components of the behavior, based on shear tests on planar interfaces. The inplane strain,
ex, will generally involve, with bilinear shape function, the ratio t /B , while those related
to ey and ~fxy will involve the ratio l / t (Figure 5.14). As a consequence, as t —> 0, the

inplane strain ex 0, which may be considered negligible. The inplane stress, ax, will

also be small, particularly when Poisson’s ratio, v, is small. As a result, the strain and
stress components in the interface can be expressed as:

ex — f ( t / B) « 0, ax « 0

£ y = £ = /( I /O = f, ay = a (5.50)
Jxy = 7 = /( I /O = 7, Tx y = r

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
128

Measured response

RI response

Shear displacement, u

(a)

RIresponse

Measured response

Shear displacem ent, u

(b)

Figure 5.8: Ultimate shear stress ru for (a) strain hardening response and (b) strain soften­
ing response

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
129

/n(r„/j%)

Figure 5.9: Determinatioii of n 2 and 7 [ 1 — (n2/rii)\


M a)

ln{£v)

Figure 5.10: Determination of hardening parameters a t and r)x

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
130

o
t-

Figure 5.11: Determination of critical state parameter M

iI

ln(o/pa)

Figure 5.12: Determination of critical volume change parameters a and b

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
131

e
K

Figure 5.13: Determination of damage parameters k and A

x,u

>

c = cosd
s = sind

X, u

Figure 5.14: Thin-layer element

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
132

The relative displacement can be expressed in terms of strain components as:


r i f 5
du t 0 d'y
i ►= < * (5.51)
dv 0 t de
k 4

Now, the incremental stress-strain relation can be written as

dr kga
'•ss k.sn
n t 0 d'y
(5.52)
da kns knn 0 t de

The global stresses and strains can be transformed into the normal and shear components
in the local coordinate system using Equations 5.53 and 5.54.

* *
d'y y _ —2 cs 2 cs c2 —s2
cv (5.53)
de (f —cs
k 4
7 xy

r \

dr —cs cs c2 — s2
(5.54)
II

s2 c2 —CS
x /
J3)
J

where, c = cosd and s = sinO as shown in Figure 5.14.

From a parametric study, Sharma and Desai (1992) proposed the following criteria to
decide on a value of t.
B ( E NGN
_ r < 104 (5 55)
t t(knks)« - ^
and
- < 0.01 (5.56)

where E n and G n are the Young’s modulus and shear modulus of the neighbouring ma­
terial respectively.
The finite element method is a numerical technique used to solve the governing dif­

ferential equation of a problem by discretization. Here, we try to satisfy the governing

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
133

differential equation on an average within a discretized element callled a finite element,

rather than at each and every point in a body. The force-displacement equilibrium equation

is derived by applying the principle of virtual work, which states that for any compatible,

small virtual displacements imposed on a body, the total internal virtual work is equal to
the total external virtual work (Bathe, 1982).

f v {6e}T({o‘ } + {oS})dV = j { 6 U } T{T}dS + f v {6U}T{X}dV +

£ { « / } r {P} (5.57)

where {6 e} is virtual strain vector, {oa} is average stress vector, {erg} is initial average

stress vector, {SU} is virtual displacement vector, {T} is surface traction vector, {X} is
body force vector, and {P} is nodal load vector. The left side of Equation 5.57 gives the

internal virtual work, whereas the right side gives the external virtual work.

In the finite element method, the continuum body is approximated as an assemblage

of discrete finite elements. In that case, Equation 5.57 represents the sum of integrations

over the volume and area of all finite elements and can be rewritten as

T. = E f s { m T{T }ds + Y : f v {su}T{ x } d v +

(5-58)

The boundaryof each discretefinite element is defined by nodes,and elements are in­

terconnected at nodes. The displacement field withineach element is assumed to be a

function of displacements at nodes and can be expressed as:

{iU} = [ M W (5.59)

where {U} is the displacement vector within an element, [N] is the displacement interpo­

lation matrix, and {?} is the vector of nodal displacements.

The strain within an element can be expressed in terms of nodal displacements as:

W = [B ]M (5.60)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
134

where the vector {e} represents strains within an element, [B] is the strain-displacement

tranformation matrix which can be derived from the displacement interpolation matrix [A/]
(Bathe, 1982).

Substituting Equations 5.59 and 5.60 into Equation 5.58, we obtain

W r £ f v lB}T{ ^ } d V = {i«}T[ £ f s m Td S + £ Jv [N]T{ X } d V + £ { P } -

£ ^ [ B ] TK } d V ] (5.61)

where [Ns] is a interpolation matrix for surface displacements.

Canceling out {6 q}T from both sides of Equation 5.61 and expressing the right side as

the generalized load vector {Q}, we obtain

£ j v [ B f {a“} d V = { Q } (5.62)

Equation 5.62 represents the equilibrium equation of finite elements.

For the nonlinear constitutive model, we apply loads in steps which are commonly

referred to as time steps rather than load steps. The equilibrium equation at time step n

can be written as:

£ / v;[B]{<}<iV = {Q„} (5.63)

For a nonlinear problem, Equation 5.63 will not be satisfiedat the first evaluation, and we

need an iterative procedure to find the correct solution. The most popular such iterative

scheme is the Newton-Raphson iterative procedure. According to this procedure, let’s say

that at iteration k, the equlibrium is not satisfied. Therefore, we have

W ( { ‘ ? » = £ f v m {*<} d v - / o (5.64)

where {'&} is the residual load vector and is the function of nodal displacement vector {g}.
Now, let

{*}({‘+l«}) = + {*<*?}) = 0 (5.65)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
135

Taylor series expansion of Equation 5.65 gives

« ( { * « } ) +* { H I {**<!} = 0 (5.66)

neglecting higher order terms. Note that the term is the tangential stiffness matrix.
Now, Equations 5.46,5.51,5.64 and 5.66 lead to

£ / v [B]r (‘ [£W {‘ & } ‘ + {‘<f„}‘ (iD„)Ii y = {‘ - ' Q „ } - S j v { B Y { ^ } d V (5.67)

or, Y . ! v m Tkm A B \ { kd l} ‘i V = { k- lQ n } - E j v [B]T {ko i } d V -

£ IjB f dD„dV (5.68)

where {dq }* is the vector of nodal displacement. The use of R I nodal displacement vector

in F E M is an approximation to make the algorithm stable. If we want to use the average

nodal displacement vector in F E M , the incremental stress-displacement relation can be


expressed as given below following the procedure given in Desai and Ma(1992).
r \
Is

Ka dua
i (5.69)
II

doa 1 Kn dva
8 / < 4
i

where.

dr° Kk *vrm
D sda° + ^ i 1
~K*nn + (1 Dvy K'nn

+(1 - D,) — K\nK


sn k (5.70)
33 Knn

D — 4-
3 da* ^ \ L
(I —D
sn 7-j 3o<= 4. (i _ n \_1_ (5.71)
^ da* ^ v 1 v )FEn

ns (5.72)
A,f£ + (1-

■n n (5.73)
D«)w~
A nn

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
136

K'ss, K lm , and K lnn are the corresponding terms in the elasto-plastic constitutive ma­

trix of R I material. When this formulation was tried in F E M using average nodal dis­
placement vector, the procedure encountered convergence problem specially in the soften­
ing zone.

The displacement vector for solid elements in F E M represents the average or ob­
served response of the material. Whereas the displacement vector for interface elements

represents R I behavior of interfaces. Therefore, it is necessary to relate the average dis­


placement vector of solid elements and R I displacement vector of interface elements at
the nodes shared by both the elements. In the present procedure we assume that both the

displacement vectors are same. This approximation induces error in the normal displace­

ment only since the shear displacement of R I and average states are same. Also, it should
be noted that the thickness of an interface element is very small compared to the adjoining

solid elements, and the small error in the displacement along the normal direction does
not affect the global nodal displacement vector much. It should also be noted that there

is no error introduced in the stress field by this approximation. The stiffness matrix ob­
tained fron Equation 5.68 is nonsymmetric due to the nonsymmetric constitutive matrix

[L\. A similar method has been used by Desai and Woo(1993) for a three-dimensional

D SC model.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
CHAPTER 6
INTERFACE MODEL VERIFICATION AND
PULL-OUT TEST SIMULATION

6.1 General
In this chapter the verification of the model developed in Chapter 5 has been presented.
First, the model has been verified by predicting direct shear test results of sand-geogrid
interfaces and comparing the predictions with the test data. This way we can verify the

capability of the model to capture the essential characteristics of the interface behavior.

Later, the model has been verified by simulating large-scale pull-out tests with two types

of geogrids, T E N S A R and C O N W E D , using the finite element method. The pull-


out simulation establishes the capability of the present model, capturing the behavior of

geosynthetic reinforcements under the pull-out condition which is commonly present in


real geosynthetic reinforced earth walls. The material parameters of the model have been
computed following the method described in Chapter 5.

6.2 Model verification


The proposed model has been verified by predicting the behavior of the interface between

sand and geogrid (geosynthetic reinforcement). The model has been used in the finite
element simulation of five pull-out tests. In the pull-out condition, it is found that the slip

occurs in a plane between sand particles in the vicinity of geogrids instead of along the

physical interface between geogrid and sand (Johnston and Romstad, 1989; Karpurapu and
Bathurst, 1995). In view of this observation, the direct shear tests have been performed

on sand only instead of on the sand-geogrid interface. Though it has been observed that
the peak shear strength of interfaces between well compacted sand and geogrids is 0 —

10% higher than that of the soil alone (Farrag, 1990; Koemer; Ingold, 1983), these peak

137

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
138

shear strengths are assumed to be the same in the present study. It has been found in

Khalid (1990) that the pattern of the stress-displacement curves for sand-geogrid and sand-
sand interfaces are similar. To obtain the shear and volumetric behavior of geogrid-sand

interfaces, large shear box is necessary because of the geometry of geogrids. However,
in large shear box, it is extremely difficult to measure the dilative behavior of interfaces.
Because of the large thickness of the sand layer, the top surface displacement does not

reflect the dilative behavior at interfaces. The reasonableness of the assumption that the

behavior of sand-geogrid interfaces can be estimated by sand only direct shear test, has
been proved by the prediction of the pull-out tests using F E M .

Four direct shear tests have been performed at normal stresses of 26.8 kPa, 53.6 kPa,

80.4 kPa and 214.4 kPa. The model has been calibrated with the results obtained from

the direct shear tests. The material parameters obtained from these tests are shown in
Table 6.1. The procedure described in Chapter 5 has been followed to calculate these

parameters. Figure 6.1 shows the data points and the linear fit to obtain n 2 and 7 [1 —
(n 2 /ni)]. The hardening parameters are computed from the linear fit of the data points

in ln(a) vs. ln(£v) plot as shown in Figure 6.2. The data points and the linear fits for

computing the critical state parameters m, a, and b are shown in Figures 6.3 and 6.4. The

parameters for the damage functions Ds and Dv are obtained from the linear fits of the data
points in ln[—ln( 1 —^ ) ] vs. I n ^ o ) plots as shown in Figures 6.5 and 6 .6 respectively.

Figures 6.7, 6 .8 , 6.9, and 6.10 show the plots of shear stress vs. shear displacement
and normal displacement vs. shear displacement for test data as well as for predictions.

It can be observed in these figures that the predictions for the tests at normal stresses of

26.8 kPa, 53.6 kPa and 80.4 k P a agree very well with the test data. The D S C model
predicted the strain softening for the first three tests very well. The prediction showed

strain hardening for the last test which is evident in the test data. Therefore, the claim
that has been made in Chapter 5 that the DSC based constitutive model can represent
strain hardening as well as strain softening, has been reflected in the present results. The

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
139

Table 6.1: Material constants for interfaces

Elastic constants Kn 550 k P a /m m

Ks 88.83 k P a /m m
Intact state 7 95.7

ril 1.4
U<1 1.395
ai 96.9324

»7i 0.01965
Critcal state M 0.456
a -0.2635
b 0.0444

Disturbed state ks 2.2935

As 0.0896

ky 1.17

Ay 1.1158

Du 0.9

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
140

0.0

- 1.0

- 2.0

• Data points
- Linear fit
-3.0

-4.0
- 2.0 - 1.0 0.0 1.0 2.0
ln (a/P a

Figure 6.1: ln(r*/P%) vs. ln(a/P a) plot

dilation in the volumetric behavior has been predicted well too. The elasto-plastic model
used for the intact state does not show strain hardening, neither does it show dilation
because of volumetric hardening. The combination of the intact behavior and the critical
behavior captures strain softening and dilation. The predictions for the test at a normal
stress of 214.4 k Pa are slightly off from the test data. Considering that the same material

parameters are used to predict the tests at normal stresses in the wide range of 26.8 kP a

to 214.4 kPa, the performance of the proposed model can be considered satisfactory.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
141

4.570

4.565 Data points


Linear fit

4.560

4.555

4.550
0.0 0.2 0.4 0.6 0.8 1.0
1 .0 1.2
1 .2 1.4 1.6
1 .6
ln (£ v)

Figure 6.2: ln(a) vs. ln(£v) plot

6.3 Pull-out simulation


Five large scale pull-out tests (Farrag, 1990) performed at the Louisiana Transportation

Research Center (L T R C ) have been simulated in F E M using the proposed interface

model. Two of these tests were performed with T E N S A R — SR 2 geogrid, and the other
three tests were conducted with the C O N W E D X 3022 geogrid. The pull-out box is 60

in long (152.4 cm.), 36 in (91.44 cm.) wide and 36 in (91.44 cm.) high (Figure 6.11).

The box is equipped with a hydraulic loading system which is capable of performing pull-

out under a constant displacement rate or under a constant pull-out load. All the pull-out
tests reported here were performed at the constant displacement rate of 6 mm/min. The

finite element mesh used for the simulation is shown in Figure 6.12. The finite element

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
142

100.0

80.0

60.0

40.0

Data points
Linear fit
20.0

0 .0
0 .0 100.0 200.0 300.0
a

Figure 6.3: r c vs. o plot

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
143

- 0.10

- 0.20
O

-0.30

Data points
Linear fit
-0.40

-0.50
-3.0 - 2 .0 - 1 .0 0 .0 1 .0 2 .0 3.0
l n ( a / P a)

Figure 6.4: v° vs. ln (a fP a) plot

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
144

0.5
- ^ )]

0 .0
1
ln [-ln (

-0.5

Data points
- 1 .0
Linear fit

-1.5

- 2 .0
0.4 0 .6 0 .8 1 .0 1 .2 1.4 1 .6

Figure 6.5: ln[-ln{ 1 —-§^)] vs. ln(£D) plot

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
145

3.0

2 .0

1 .0

0 .0
■--o
I
- 1.0
Data points
Linear fit
- 2 .0

-3 .0
- 2 .0 - 1 .0 0 .0 1 .0 2 .0
ln(£D)

Figure 6 .6 : ln[—ln( 1 — )] vs. Zn(fo) plot

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
146

30.0

Normal stress = 26.8 kP a

20.0
CO

• Test data
Prediction

o.o
o.o 1.0 2.0 3.0 4.0
Shear displacement, u

(a)

0.50

0.40 Normal stress = 26.8 kP a

S
S
<D 030

%
■q.
CO
0.20

^ 0.10
'S • Test data
£^ Prediction
0.00

- 0.10
o.o 1.0 3.0 4.0
Shear displacement, u

(b)

Figure 6.7: Test data and prediction of direct shear test at normal stress of 26.8 kPa: (a)
Shear stress vs. shear displacement and (b) Normal displacement vs. shear displacement

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
147

50.0

Normal stress = 53.6 kPa


40.0

£ 30.0

S 20.0
43
co
Test data
10.0 Prediction

o.o
0.0 1.0 2.0 3.0 4.0
Shear displacement, u

(a)

0.50 I ' ~~l ' I " , ——1

0.40 - Normal stress = 53.6 kPa



0.30 •

0.20
/ •
0.10 “
• Test data
' ---- Prediction
0.00 i __ •

n 1n I — —L . 1
0.0 1.0 2.0 3.0 4.0
Shear displacement, u

(b)

Figure 6 .8 : Test data and prediction of direct shear test at normal stress of 53.6 kPa: (a)
Shear stress vs. shear displacement and (b) Normal displacement vs. shear displacement

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
148

60.0

50.0 Normal stress = 80.4 k P a

40.0

3 30.0

55 2o.o
• Test data
Prediction
10.0

o.o
o.o 1.0 3.0 4.0
Shear displacement, u

(a)

0.50

0.40 Normal stress = 80.4 k P a

g
5u °-30
6
'H.
CO
0.20

S 0.10
t: • Test data
Prediction
o.oo

- 0.10
0.0 1.0 2.0 3.0 4.0
Shear displacement, u

(b)

Figure 6.9: Test data and prediction of direct shear test at normal stress of 80.4 kP a : (a)
Shear stress vs. shear displacement and (b) Normal displacement vs. shear displacement

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
149

140.0

120.0 Normal stress = 214.4 k P a


100.0

CO
§3 80.0
is
CO

e3 60.0
u
JS
CO
40.0 • Test data
Prediction
20.0

o.o
0.0 1.0 2.0 3.0 4.0 5.0
Shear displacement, u

(a)

0.50

0.40 Normal stress = 214.4 k P a


cn>
g 0.30
<D
M
Q, 0.20
co
"3
”g 0.10
'£ • Test data
£ Prediction
0.00

- 0.10
o.o 1.0 2.0 3.0 4.0
Shear displacement, u

(b)

Figure 6 .10: Test data and prediction of direct shear test at normal stress of 214.4 kP a : (a)
Shear stress vs. shear displacement and (b) Normal displacement vs. shear displacement

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
(a)

4*«4'

1/2* Stiff* Back wall n

M M 4*14'

(b)

Figure 6 . 1 1 : Pull-out box (After Farrag, 1990)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
151

sleeve

o\ pull-out displacement

geogrid

interface

60 in. (152.4 cm.)

Figure 6 .12: Finite element mesh for pull-out simulation

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
152

mesh consists of 40 four-node isoparametric elements to model soil, 16 thin layer elements

to model interfaces, and 8 two-node bar elements to model geogrid reinforcements. The
bottom nodes are restraint from both horizontal and vertical movements, and the front and

back nodes are restraint from horizontal movement, but they are allowed to move in the

vertical direction. The nodes around the sleeve are restrained from vertical movement

6.3.1 Modelling sand

The sand used in the pull-out test is a uniform blasting sand. It is placed in the pull-out

box in four layers of 6 in. each at the density of 17 k N / m z from an elevated hopper.

The hopper is moved during pouring to ensure constant height of fall. After pouring, each

layer is compacted manually by a vibrating electric hammer, and the density is measured

by a nuclear density gauge. The sand is modelled in the finite element analysis with

the hierarchical single surface (H i S S ) S\ model described in Chapter 3. This model is

an elasto-plastic nonassociative model, and there are nine material parameters associated
with this model: elastic constants E and u; ultimate parameters 7 , f3 and m; phase change

parameter n; hardening parameters and rji; and nonassociative parameter k . Three

triaxial compression tests (C T C ) have been performed at the confining pressures of 20,
40, and 60 psi (138, 276, and 414 kPa respectively) to calibrate the H iS S Si model. The

material parameters are computed following the procedure described in Chapter 4 using
the results of these triaxial tests and are shown in Table 6.2. The test data and the model

predictions are shown in Figure 6.13. It can be observed in this figure that the H i S S Si
model characterizes the behavior of sand very well.

6.3.2 Modelling reinforcement

Two different types of geogrids, (a) TENSAR SR2 and (b) CONWED X3022 were used
in the pull-out tests. Each geogrid specimen had a length of 3 f t (91.44 cm) and a width
of 1 ft. (30.48 cm) which allowed it to keep its edge at 1 f t (30.48 cm) from the side

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
153

800.0

Oj = 414 k P a
JOOOOOOOOOOOOOOO

600.0

<t.j = 276 A:P a


jjOOOOOOOOOOOOOOOOOOOOOj

400.0

(Ta = 138 kP a
>0000000000

200.0
a Test data
- Prediction

0.0 *—
0.000 0.020 0.040 0.060

(a)

0.020

0.015

0.010

0.005

° Test data (138 kPit)


0.000 ° Test data (276 kPu)
4 Test data (414 kPu)
- Prediction (138 kPa)
- Prediction (276 kPu)
-0.005
- Prediction (414 kPa)

- 0.010 —
0.000 0.020 0.040 0.060

(b)

Figure 6.13: Test data and prediction of H iS S model: (a) y/J^D vs. y/Tw and (b) y/Ho
vs. —I x

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
154

Table 6.2: Material constants for H i S S Si model for sand

Elastic constants E 2.258 x 104 kPa


V 0.17

Ultimate 7 0.07

0 0.7
m -0.5

Phase change n 2 .1

Hardening ai 8.867 x 10~ 3

m 0.257

Nonassociative K 0.756

walls of the box. A sleeve of 1 ft. (30.48 cm) length was used around the clamping plates

at the front wall to reduce the effect of front rigid boundary on test results. A hyperbolic

model(Chou, 1992) is used to describe the load-deformation behavior of the geogrids. The
model can be described as:

where, E{ = initial tensile stiffness

Tuu = ultimate tensile force per unit width

Equation 6 .1 can be transformed into a linear relationship similar to the hyperbolic model
for sand (Clough and Duncan, 1969), and the transformed equation is given below.

T ~ E, + T*. ( )

From uniaxial tensile test data, E{ and Tuit can be calculated using Equation 6.2.
Some geosynthetics have different load-elongation properties depending on whether
they are tested in the confinement of soil or tested in isolation (McGown et al., 1982;

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
155

Table 6.3: Material constants for geogrids

Geogrid Ei T u it TTL\ TTt2

T E N S A R SR 2 1197.6 k N /m 97.6 k N /m

C O N W E D X3022 951.8 k N /m 51.3 k N /m 76.2 m 0.198 m

Siel et al., 1987; Wu, 1991; Ling et al., 1992). In confinement, both Tu/t and Et- increase
with increasing normal stress. A simple linear relationship can be used to account for the

confinement effect (Chou, 1992).

Ei — E 0 + anm i (6.3)

Tuit = T 0 + anm 2 (6.4)

where, Eq = initial stiffness in unconfined condition


T0 = ultimate load in unconfined condition
m i = rate of increase in E, with confining pressure

m 2 = rate of increase in Tu(t with confining pressure

Though the Tensar geogrid does not show any change in Ioad-elongation behavior under

confinement, load-elongation behavior of the conwed geogrid is affected by confinement.


Table 6.3 shows the values of the material constants for the geogrids used in this study.

Figures 6.14 and 6.15 show the test data and predictions of uniaxial tensile tests on these
geogrids.

6.3.3 Modelling interfaces

The interfaces between geogrid and sand are modelled by the thin layer element in associ­
ation with the proposed model described in Chapter 5. The sand used in the pull-out tests
is the same sand used in the direct shear tests to verify the model. The material parameters
given in Table 6.1 are used in the finite element simulations.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
° Test data
- Prediction

O.o ----------- 1------- '------- ‘--- -------- 1


0.0 5.0 10.0 15.0 20.0
Axial Strain, %

Figure 6.14: Test data and prediction of uniaxial tensile test on T E N S A R geogrid

40.0

30.0
6

15 20.0
o
-I

#s o Test data
< Prediction
10.0

0.0
0.0 5.0 10.0 15.0
Axial Strain, %

Figure 6.15: Test data and prediction of uniaxial tensile test on C O N W E D geogrid

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
157

80.0

60.0
k N / m
Pull-out load,

40.0

• Test data (crn = 3 4 kPa)


20.0 ■ Test data (<jn = 4 8kPa)
- Prediction (<rn = 3 4 kPa)
- Prediction (<7n = 4 8kPa)

0.0
0.0 20.0 40.0 60.0
Front displacement, m m

Figure 6.16: Test data and predictions of pull-out tests with T E N S A R geogrid

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
158

120.0

100.0
k N / m

80.0
^ H’ ♦
♦ _______
Pull-out load,

♦ ♦♦ ♦
60.0

40.0 Test data (an = 48kPa)


■ Test data (an = 96kPa)
♦ Test data (crn = 1 4 5 kPa)
20.0 Prediction (<rn = 4SkPa)
Prediction (an = 96kPa)
Prediction (<jn = 1 4 5 A ;P a )
0.0
0.0 10.0 20.0 30.0
Front displacement, m m

Figure 6.17: Test data and predictions of pull-out tests with C O N W E D geogrid

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
159

6.3.4 Pull-out simulation results

Figures 6.16 and 6.17 show the results of the simulations of five pull-out tests using the
finite element method. The pull-out load vs. pull-out displacement agrees well with the
test data for each simulation. The predictions of the pull-out load tests with Tensar

geogrids show a small strain softening when the proposed model for interfaces is used.

The similar predictions for Conwed geogrids do not show any strain softening though the

pull-out test with this geogrid at a normal stress of 145 kP a show a decrease in pull-out
load after attaining a peak value. Since the test data for lower confining stresses do not
show any strain softening, it is very unlikely that the strain softening observed in the test
data for 145 kPa confining stress is caused by the frictional behavior of the interfaces.

This apparent strain softening could be caused by the rupture of the geogrid. In the finite

element analysis, this is evident from the large strain developed in the first element of the

geogrid compared to the second element At the front displacement of 20 mm, the first

element developed an axial strain of 1 0 %, whereas the element next to it had a strain of
2%. The high normal stress prevents the pull-out displacement from propagating along

the geogrid. As a consequence, most of the displacement is carried by the front portion
of the geogrid which ruptures after certain point From Figures 6.16 and 6.17, it can be

concluded that the proposed D S C based interface model is capable of describing pull-out
behavior of geosynthetic reinforcements in the finite element simulation.

6.4 Summary and conclusion


An elastoplastic constitutive model based on the disturbed state concept is presented for

sand-geosynthetic interfaces. The model is capable of characterizing dilation, hardening

and softening responses of interfaces. The proposed model is verified with a series of
direct shear test data, and the model predictions are found to be highly satisfactory. The
model is implemented in the finite element method together with the thin layer element
for interfaces and joints. Five pull-out tests on geogrids have been simulated using the

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
160

proposed model, and the simulation results are found to agree well with the test data. The

proposed model can be used in the numerical simulation of large-scale reinforced soil
walls. Such simulation is very useful in understanding the stress-deformation behavior of

reinforced soil walls. In Chapter 8 , the numerical simulation of a full-scale laboratory test

wall conducted at the Royal Military College of Canada using the proposed D SC based
interface model has been described.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
CHAPTER 7
NUMERICAL SIMULATION OF DENVER WALL

7.1 General
Before a finite element simulation can be used for a stress-deformation study, the finite
element model needs to be validated against observed behavior. The comparison of the nu­

merical simulation results and the observed behavior shows the performance of the various

constitutive models used to represent different elements in the structures. In the present

study, two full-scale laboratory test walls have been chosen for the numerical simulation

and the validation of the finite element models. The walls were tested at the University

of Colorado, Denver, Colorado, and at the Royal Military College, Kingston, Canada, and

they are referred to as the Denver Wall and the RMC Wall respectively. In this chapter the

numerical simulation of the Denver Wall is presented. The simulation of the RMC Wall
has been presented in Chapter 8 .

7.2 Denver Wall


The Denver Wall is a 3.05 m (10 ft) high geotextile reinforced soil retaining wall with

granular backfill. The wall was constructed and tested at the University of Colorado,
Denver as a part of the International Symposium on geosynthetic-reinforced soil retaining
walls held in Denver, Colorado, in 1991. The details of the test and the predictions by var­

ious researchers are given in Wu (1992). The important features of the test are described

here for completeness. The purpose of that research was to develop guidelines for design

and construction of geosynthetic reinforced soil retaining walls. Figure 7.1 depicts the
configuration of the Denver Wall and the loading facility. For better control of the test

conditions, the wall was built within a rigid loading facility in the laboratory. The wall

facing was constructed using timber logs inter-connected by plywood board. The side and

161

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
162

d rs /iN /IN
Air Bags

Surcharge San*

UK

5 ft < in.

Figure 7.1: Configuration of Denver Wall (After Wu, 1992)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
163

back panels were heavily reinforced to make them behave as rigid. The side panels were
lined with 0 .0 2 mm latex membrane and lubricated with silicon grease to minimize the

induced shear stress in the plane of side walls, thus ensuring plane-strain condition. The
back panel was lined with a smooth, high density polyethylene (H D P E ) sheet The bot­
tom of the wall was made rough by gluing course aggregate to the steel floor. A specially
designed hopper was used to place the granular backfill by an air-pluviation method where

the sand was allowed to fall a constant height for all lifts to maintain uniform density.

The advantage of using this method is that it avoids the large horizontal stress caused by
vibratory compaction. The wall was reinforced with 12 layers of nonwoven polypropy­
lene heat-bonded geotextile. At the facing, each reinforcement was nailed between the
plywood board and the logs and was folded flat toward the back of the wall.

The wall was loaded in an increment of 3 psi (20.7 kPa) on the top surface of the

backfill using an air bag. The wall was allowed to creep for 100 hours at 15 psi (103.4

kPa) surcharge. After that the wall was loaded until it failed at 29 psi (200 kP a) surcharge
load. The measurement of the facing movement, movement of the top surface, and the

strains in geosynthetic reinforcements at three different levels, 0.15/7,0.52/7 and 0.8877,


where 77 being the height of the wall, were reported for the following loading conditions:

(a) end of construction (E O C ), (b) 15 psi (103.4 kP a ) surcharge, and (c) 27 psi (186
kPa) surcharge.

7.3 Material models


In any finite element simulation, the material models play the most important role. There­
fore, it is very important to select appropriate material models to represent various ele­

ments of a structure. The proper calibration of these models using different laboratory
test data is equally important In this section, the material models used to represent the

different elements of the Denver Wall and their calibration have been described.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
164

Table 7.1: Material constants for H i S S Si model for granular backfill

Elastic constants E 1.14 x 105 kPa


V 0.3

Ultimate 7 0.05

0 0.7

m -0.5

Phase change n 2.95

Hardening ai 9.66 x 10“s

V\ 0.54

Nonassociative K 0.233

7.3.1 Backfill sand

The hierarchical single surface (H iS S ) Si model described in Chapter 3 has been used to

represent the granular backfill material. For the calibration of this model, it is desirable
to have a few tests in compression, extension and simple shear stress paths though only

one conventional triaxial compression test (C T C ) can produce fairly good approximate

values for the material constants. Three C T C tests results were reported for the granular

backfill used in the Denver Wall (confining pressures are 68.95 kP a (10 psi), 206.85 kP a
(30 psi) and 344.75 kPa (50 psi)). Out of these three C T C tests, the volumetric strain vs.
axial strain results were not reported for the 68.95 kP a (10 psi) C T C test Therefore, this
test result was excluded from use in material parameter determination. The procedure for

determining the material constants can be found in Chapter 4. The material parameters for

the H i S S <Ji model used in this study are given in Table 7.1.
The stress-strain behavior and the volume change behavior of the material have been

back predicted using the material parameters shown in Table 7.1. Figures 7.2 and 7.3 show

the J\ vs. Ii, \fJiD vs. y /h o and h vs. y / h o plots of the CTC tests for the predictions

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
165

as well as the test data. The agreement between the predictions and the test data can be
considered satisfactory.

120.0

100.0

80.0
•cS>
CO

Q 60.0

>
40.0
Test data ( 0 3 = 30 p s i )
*□ O.
Test data (<73 = 50 psi)
20.0 CD Prediction (<73 = 30 p si).
Prediction ( 0 3 = 50 psi)

0.0
0.00 0.01 0.02 0.03 0.04 0.05

V h D

Figure 7.2: Test data and prediction by FTiSS 6 i model for the granular backfill: yJJ^D
VS. \/I<i d

7.3.2 Reinforcement

The geotextile reinforcement has been represented by the hyperbolic model described in
Section 6.3.2. An analysis was performed using the von Mises model for geotextile rein­

forcements (E = 192800 kPa, K = 11131 kPa). The results of that analysis arc given
in Appendix B. Figure 7.4 shows the force-elongation curve of the reinforcement. The

material parameters and Tu[t obtained for this model are 70.6 k N /m and 6.7 k N /m
respectively. Since the load-elongation behavior of the geotextile is not influenced by the

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
166

0.040

0.020

OO
°o o
0 .0 0 0
Test data (<r3 = 30 p s i )
Test data (<r3 = 50 p si )
Prediction (a3 = 30 psi).
Prediction (a3 = 50 psi)

- 0 . 0 2 0 L—
0.00 0.01 0.02 0.03 0.04 0.05

y /h o

Figure 7.3: Test data and prediction by H iS S model for the granular backfill: - I x vs.
y /? 2 d

confining pressure as it is evident in Figure 7.4, the material constants m x and m 2 are taken

as zero. Figure 7.5 shows the test data and the prediction of the hyperbolic model, and it

can be seen that the prediction agree very well with the test data.

7.3.3 Wood facing

The facing elements were composed of timber logs (10.16 cm (4 in.) width and 13.97

cm (5.5 in.) height) interconnected by 1.27 cm ( | in.) thick plywood boards as shown
in Figure 7.6. Each plywood board was 27.94 cm (11 in.) wide and was nailed to three

timber logs in such a way that one log was positioned along the center of the board and
half of the other two logs were on both sides of the center log.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
167

(00 UncanfimO IIn-UrI


(U pal

CbnOmO in Soil (u pd pnssuKl


400
4#
A

5
S
| «o

120

(0

Axial SCrain, 1

Figure 7.4: Load-deformation behavior of geotextile reinforcement for Denver Wall (After
Wu, 1992)

Figure 7.7 shows the test set up used for testing a typical timber facing unit consist­

ing of five timber blocks and two plywood forming elements. The relationship between

total applied load versus vertical movement of the loading piston is shown in Figure 7.8.
A bilinear elastic model has been used for the facing elements. The material parameters

have been calculated for this model following the procedure described by Ling and Tat-

suoka(1992) from the bending test result. The bending test on the timber/plywood facing
unit can be modelled by a continuous beam as shown in Figure 7.9. The deflection, 6 , at
point C or D can be expressed as:

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
168

8.0

6.0
£
sr
-se
« 4.0
o
J
H
'x
<
o Test data
2.0
- Prediction

0.0
0.0 20.0 40.0 60.0 80.0
Axial Strain, %

Figure 7.5: Test data and prediction of uniaxial tensile test on geotextile

where E is the Young’s modulus and I is moment of inertia of the beam. For the given
geometries, the Young’s modulus is an unique function of the applied load P and the de­
flection S. The Young’s modulus at different loads can be determined using Equation 7.1.

Now, using the Young’s modulus, moment of inertia and bending moment at the mid­
section of the beam, the average stress at the mid-section can be determined. The corre­

sponding strain can be computed by dividing the Young’s modulus with the average stress.

The plot of stress versus strain (Figure 7.10) shows that the material can be characterized
by a bilinear elastic model with the elastic and plastic moduli of 26050 kP a and 12260
kPa respectively, and the yield stress is 256 kPa.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
169

Figure 7.6: Attachment of plywood boards to timber logs (After Wu, 1992)

7.3.4 Interface

The D S C based interface model developed in Chapter 5 needs volumetric behavior of


interfaces to calibrate. Unfortunately, no volumetric behavior was reported for sand-

geotextile interfaces of the Denver Wall. Only shear stress versus horizontal displace­

ment of direct shear tests was reported for sand-geotextile interfaces. A hyperbolic model
(Clough and Duncan, 1969; Acar et al., 1982) has been used to represent the interfaces
between sand and geotextile. The description of this model is given below.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
170

f itlKM TtfU

aioc mw

1 1 1
1
0 • «
1
• • •
v T
•»
1 1 1
1 1 1
i
to* mew

Figure 7.7: Bending test configuration of a typical timber facing (After Wu, 1992)

The nonlinear shear stress-displacement relationship can be expressed as a hyperbola

as
— = a + bu (7.2)
r
where r is shear stress, u is shear displacement, and a and b are interface constants which

depend on the roughness and normal stress, an. 1 / a gives the initial tangent shear stiffness,

G{, and 1/6 gives the ultimate shear stress, ru/t. Since the shear strength, 7 /, is reached
before the ultimate shear stress, a failure ratio, R j, is defined as:

Rf = -2 - (7.3)
Tuft

The initial tangent shear stiffness depends on the normal stress, and this dependency can
be accounted for using the following relation (Janbu, 1963).

= (7.4)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
171

Note: Defections ore measured from


dial guoge on loading tooie

300 - -

200 -*

100 --

0 .0 0 .2 0 .4 0.6 o.s
D e f l e c t i o n (in)

Figure 7.8: Load vs. displacement curve the timber facing unit (After Wu, 1992)

L — 2£i

Figure 7.9: Structural model of beading test on wood facing

R e p r o d u c e d w ith p e r m is s io n o , t h e c o P y rig h , o w n e r . P u d h e r r e p r o d u c tio n p r o v e d w ith o u t p e r m is s io n .


172

-too

o iE tm o u tn

1770 p d

« * ■ M W

j l' - r ^ V « > *«■*%

j_
O.OI 0.02 0.03 0.04 0.05
STMM

Figure 7.10: Stress-strain curve of the wood facing (After Ling and Tatsuoka, 1992)

where Pa is the atmospheric pressure, K is the shear stifftiess at an = Pa, j w is the unit

weight of water, and n is a material constant

At any level of shear stress or displacement, the nonlinear tangent shear stiffness can
be represented by

g, - au
£ - g, 1 - (7.5)
T~ult-

Using Mohr-Coulomb strength criterion and Equation 7.3, Equation 7.5 can be rewritten
as:
Rf
Gt = Gi 1 - (7.6)

where (f>is the angle of friction.

The variation of (f>with the normal stress can be accounted for using the following
relation.

0 = & - A to o tfio ( p 1) (7 .7 )

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
173

Table 7.2: Material constants for sand-geotextile interfaces

Modulus number K 2.2545 x 10“ k N /m 3


Modulus exponent n 0.246
Fiction angle fa 26.18
parameters A <f> -3.972

Failure ratio Rf 0.89

where fa is the value of 0 at crn = Pa and A(f>is the decrease in $ for a ten-fold increase
in the normal stress, crn. The material parameters used for the sand-geotextile interfaces
are shown in Table 7.2. The test data and the predictions are shown in Figure 7.11.

200

345 kPa

207 kPa

69 kPa Test Data


Prediction

o.ooo 0.001 0.002 0.003 0.004


Horizontal Displacement, m

Figure 7.11: Direct shear test on geotextile-sand interfaces

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
174

Table 7.3: Material constants for sand-plywood interfaces

Modulus number K 2.2545 x 104 k N / m 3


Modulus exponent Tl 0.246
Fiction angle 00 26.18
parameters A <0 -3.972

Failure ratio Rf 0.89

No test was reported to determine the frictional behavior of the interface between sand
and plywood facing. Acar et al. (1982) performed some direct shear tests on wood/sand
interfaces and reported the variation of different parameters of the hyperbolic model with

a void ratio of sand. The material parameters for plywood/sand interfaces have been taken

from the chart provided by Acar et al. (1982) corresponding to the void ratio at which the

backfill was poured. These material parameters are given in Table 7.3. An analysis was
performed without using any interfaces. The results of that analysis are given in Appendix
B.

7.4 Finite element model


The wall was modeled as a plane-strain, two dimensional problem for the finite element
analysis. Figure 7.12 shows the finite element mesh used for the analysis. The mesh con­

sists of 850 nodes, 330 four-node quadrature elements to represent the soil, 169 two-node
bar elements to represent the reinforcements, and 32 two-node beam elements to represent
the wall facing. The surcharge has been modeled by 60 four-node quadrature elastic ele­

ments which have the same bulk modulus as the soil, but a very low shear modulus so that
the surcharge elements will not put any shear resistance on the outward movement of the
soil elements underneath. In reality, there was a slip between the surcharge and the backfill

because the drop gate holds the surcharge as the wall deforms. To take care of this slip,

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
175

1 m m thick thin layer elements (described in Chapter 5) have been used between the sur­

charge and the granular backfill. Similar elements have been used between the wall facing
and the adjoining soil elements and the geosynthetic reinforcements and the adjoining soil

elements to take care of the effects of the slip between two dissimilar materials. There are
total of 367 such elements in the model. An analysis was carried out using the Goodman
element for interfaces. The results of that analysis are given in Appendix B.

The nodes along the back of the wall are restricted to move along the horizontal direc­

tion, but they are free to move along the vertical direction. The nodes at the bottom of the
wall are restricted in both the directions since the bottom of the wall was made rough by
gluing course aggregate to the rigid floor. To simulate the construction sequence, initially

the nodes along the wall facing were supported along the horizontal direction; then the

supports were released gradually from the bottom to top. A similar method was used by
Yeo et al.(1992).

In the first load step, the self weights of the backfill and the surcharge and a 20 kPa
pressure on the top of the surcharge were applied. After that, the applied 20 kPa pres­
sure was increased to 200 kPa in one hundred load steps. The results of the analysis are
furnished in the following section.

7.5 Result and discussion


In this section the results of the F E M analysis are presented and compared with the

measured behavior. The following measurements were reported for the loading conditions;
end of construction (E O C ), 103 kPa (15 psi) surcharge, and 186 kPa (27 psi) surcharge.
The wall underwent a very large deformation for the load increment from 186 kPa (27 psi)

to 200 kP a (29 psi). Therefore, the surcharge load 186 kPa (27 psi) can be taken as the

ultimate load capacity of the wall.


Figure 7.13 shows the predicted and measured facing movement profiles for the wall
at 103 k P a and 186 kP a surcharge pressures. In the test, the facing movement profiles

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
176

Figure 7.12: FEM mesh for Denver Wall


u.

m
2

r *1 ’
|_

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Ill

were measured by the paper targets. It can be observed from the figure that the present
predictions agree well with the measured facing movements at 103 kPa and 186 kPa
surcharge loads.

300.0

200.0

100.0

• Measured (103 kPa)


■ Measured (186 kPa)
“ FEM Prediction (103 kPa)
~ FEM Prediction (186 kPa)

0 .0
0 .0 2 .0 4.0 6 .0 10.0
8 .0 12.0 14.0
Wall Movement, cm

Figure 7.13: Facing movement of Denver Wall

Figure 7.14 illustrates the top surface displacement profiles at 103 kPa and 186 kPa

surcharge pressures. It can be observed in this figure that the F E M analysis underpredicts

the top surface displacements. The measured displacement profiles show no dilation of
the backfill material, whereas the finite element prediction shows approximately 0 .8 %
dilation. In Figure 7.3, it can be observed that the material model predicts approximately
1% dilation at failure. Therefore, there could be two reasons for this discrepancy between

the measured and predicted top surface displacement profiles, ( 1 ) the dilative behavior of

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
178

0.0

£ -4-0
o
a
£
o<o
IL - 8 .0
S
73
o
'5
£ - 12.0 • Measured (103 kPa)
■ Measured (186 kPa)
~ FEM Prediction (103 kPn)
~ FEM Prediction (186 kPa)

-16.0
0 .0 50.0 100.0 150.0 200.0 250.0
Horizontal Distance from the Facing, cm

Figure 7.14: Top surface displacement profile of Denver Wall

the backfill material in the wall is different from the dilative behavior in the triaxial tests,

and (2 ) the density of the backfill material in the wall was different from the density at
which the triaxial tests were conducted.

The axial strain distribution in the reinforcement at three different heights- 0.15H,
0.52H, and 0.88H (measured from the base of the wall where H is the total height)- are

shown in Figures 7.15 to 7.17 for three different loading conditions. It is seen in these

figures that the FEM results are in agreement with the measured behavior. The mea­
sured distributions immediately behind the wall facing were not furnished because of the

malfunction of strain gauges caused by excessive strains in that area. The FEM results
show that the maximum strain in the reinforcement occurs within a small distance from

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
179

the wall facing. This large strain may have caused the malfunction of strain gauges near
the facing. A similar trend has been observed by Ho and Rowe (1993) and Fishman et
al.(1993). A sadle-shaped strain distribution is observed in Figure 7.17 for the top most

layer at a surcharge pressure of 186 kPa. A similar trend was observed by Karpurapu and
Bathurst(1995).

15.0

• Measured ( E O C )
■ Measured (103 kPn)
10.0 A Measured (186 kPa)
FEM Prediction ( EOC)
FEM Prediction (103 kPa)
c — — FEM Prediction (186 kPa)
*faa 5 0
co Cl,u
*a
‘x
<

0 .0

-5.0
0.0 50.0 100.0 150.0 200.0
Horizontal Distance from the Facing, cm

Figure 7.15: Axial strain in reinforcement: Height = 0.15H

The maximum facing movement with surcharge pressure and the maximum vertical
displacement of the top surface with surcharge pressure are illustrated in Figure 7.18. A

failure load is not observed in this figure and the probable cause could be that the failure
of the wall is not a wedge failure type. It is a progressive failure and is manifested in the

failure pattern discussed later.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
180

15.0

— • Measured (EOC)
■ Measured (103 kPa)
* Measured (186 kPa)
FEM Prediction {EOC)
x 100 FEM Prediction (103 kPn)

G FEM Prediction (186 kPa)
•a
is
C/3

'5
< 5.0

0 .0
0 .0 50.0 100.0 150.0 200.0
Horizontal Distance from the Facing, cm

Figure 7.16: Axial strain in reinforcement: Height = 0.52H

The horizontal soil stress distributions behind the facing, at a vertical plane through

the middle of the reinforcement and at a vertical plane at the end of the reinforcement,
are shown in Figures 7.19 and 7.20 at 103 kPa and 186 kPa surcharges respectively.
It can be seen that at the middle part, the horizontal stress is lower than that for active

condition. This happens because the vertical stresses also get reduced in the vicinity of
the facing. The horizontal stress at the toe is larger than the theoretical at rest pressure

due to the constraint provided by the rigid foundation and the toe. The horizontal stress
distribution behind the facing is oscillating in nature. This oscillating nature of horizontal

stress distribution has been observed by Andrawes et al. (1990) in a field experiment and

Ho (1993) in a numerical study. This oscillation may be caused by the local variation

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
181

8.0

“• Measured (EOC)
“■ Measured (103 kPa)
6 .0 -*■ Measured (186 kPa)
~ FEM Prediction (EOC)
~ FEM Prediction (103 kPn)
ss — FEM Prediction (186 kPa)
•a
B 4^-u
0
co
*«3
<

2 .0

0 .0
0.0 50.0 100.0 150.0 200.0
Horizontal Distance from the Facing, cm

Figure 7.17: Axial strain in reinforcement: Height = 0.88H

in vertical stresses due to the interactions among the reinforcements, backfill and facing

as observed in Figures 7.21 and 7.22. Ho and Rowe (1993) have also reported the same
observation. The increase in horizontal stress at the upper portion of the wall beyond the

at-rest pressure is quite unusual though some field measurements show larger horizontal

stress at the upper portion of the wall than that observed in the middle portion (Fishman
et al., 1993). However, in those cases, horizontal stress didn’t increase beyond the at rest

earth pressure. One reason for this increase in horizontal stress could be the horizontal

restraint provided by the drop gate used for applying surcharge on the wall. This horizontal

restraint is evident in the facing deformation where the top of the facing underwent very

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
182

20.0

Facing
g Top surface
5 15.0
eD
<
s
<
CJu
ea
£ 10.0
s
e
3
s
| 5.0

0 .0
0 .0 100.0 200.0 300.0
Surcharge Pressure, kPa

Figure 7.18: Maximum facing and top surface movement with surcharge

small horizontal displacement compared to the maximum displacement the middle portion
of the facing encountered.

To design the geosynthetic reinforced vertical wall, the tied back wedge methods are

generally used. In these methods, several assumptions are made regarding the vertical
stress distribution at the foundation and the horizontal stress distribution in order to cal­
culate loads on reinforcements. The shape and the location of the failure plane are also

assumed in order to compute the length of reinforcements providing pull-out resistance.


The vertical stress distribution is used in the design to check the safety against bearing

capacity failure and the failure caused by sliding. Figures 7.23 and 7.24 show the vertical
stress distributions at the base of the wall at 103 kPa and 186 k P a surcharge pressures

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
183

300.0

200

J3
00
'53
X • Behind facing
100 • Mid-section of reinforcement -
♦ End of reinforcement
■ Active pressure
* At rest pressure

50.0 100.0 150.0


Horizontal Stress, kPa

Figure 7.19: Horizontal stress distribution at 103 kPa surcharge

respectively. It can be observed in this figure that the vertical soil stress is less than the the­
oretical value (i.e., av = 'yh) close to the facing, except at the toe, due to partial transfer of
vertical stress from the fill to the facing through the fill/facing friction and reinforcements.

Whereas at the toe, the vertical stress is more than the theoretical value due to the rigid

foundation which prevents relative settlement between the fill and the facing. The vertical
stress increases in the region away from the facing, exceeding the theoretical value. A.

similar trend has been observed in the field, as shown in Figure 7.25. Though this figure

does not show the vertical stress immediately beside the toe, the increase in vertical stress
at the toe has been reported by Bolton and Pery (1980) based on a laboratory experiment

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
184

300.0

200.0

oo
Behind facing
100.0 Mid-section of reinforcement
End of reinforcement
Active pressure
At rest pressure

0 .0
0 .0 50.0 100.0 150.0 200.0
Horizontal Stress, kPa

Figure 7.20: Horizontal stress distribution at 186 kPa surcharge

and Ho and Rowe (1993) based on a finite element analysis. The vertical stress distribu­

tions observed in Figures 7.23 and 7.24 do not conform to the trapezoidal distribution nor
do they conform to Meyerhof distribution. As opined by Ho (1993), the assumption of a
uniformly distributed vertical stress is more suitable.

To design the spacing of the reinforcement and to choose a reinforcement type of


adequate strength, the tied-back wedge methods assume different horizontal stress distri­

butions to be carried by the reinforcement as shown in Figure 2.2. Figures 7.26 and 7.27

illustrate the maximum tensile forces in the reinforcements obtained from the finite ele­
ment analysis and those calculated from the Broms method and Bonaparte et al. method

for the surcharge pressures of 103 kP a and 186 kPa respectively. One can see in these

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
185

kPa
300

0 20 40 60 80 100 120 140 160 180 200


Distance from the facing, cm

Figure 7.21: Vertical stress contours at 103 kPa surcharge pressure

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
186

kPa
SOO

250 -

200 -

150 -
Height, cm
100 -

50 -

~i i i I I I I i r
20 40 60 80 100 120 140 160 180 200
Distance from the facing, cm

Figure 7.22: Vertical stress contours at 186 kPa surcharge pressure

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
187

600.0
Bottom most layer
- Theoretical vertical stress

400.0
<3 Meyerhof

CO
ia 200.0
co
73
o
•*«*
t:
2
0 .0

Trapezoidal

- 200.0
0 .0 50.0 100.0 150.0 200.0
Horizontal Distance from the Facing, cm

Figure 7.23: Vertical stress distribution at the base of the wall at 103 kP a surcharge

figures that the forces obtained from the finite element analysis are much less than the

forces computed from those methods, which implies that those traditional methods are
highly conservative in nature.

The present finite element analysis doesn’t show the development of shear band, but

if we look at the elements which are in plastic failure state, as shown in Figures 7.28
and 7.29, a pattern can be seen. Instead of reaching the failure state in a narrow shear

band, all the elements above a plane reach the plastic failure state. The state of failure is

checked by comparing the ratio of \JJ id to J\ with the value of y/yPg which represents

the state of plastic failure for the H i S S 6i model. Superimposed on these figures are the

Rankine failure surface and the failure surface assumed in the Coherent Gravity method

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
188

800.0

Meyerhof
600.0

(5 400.0
-42

£s 200.0
T3
o
* 0.0

-• Bottom most layer


- 200.0 - Theoretical vertical stress Trapezoidal

-400.0
0.0 50.0 100.0 150.0 200.0
Horizontal Distance from the Facing, cm

Figure 7.24: Vertical stress distribution at the base of the wall at 186 kP a surcharge

(Schlosser, 1990). The failure plane assumed in the Coherent Gravity method resembles

the observed failure pattern closer than the Rankine failure plane. In the tied-back wedge
methods, the length of reinforcement extending beyond the failure plane is considered to
provide resistance against pull-out The identification of the actual failure plane in the

geosynthetic reinforced soil wall is important for that reason.

7.6 Summary
A large scale geotextile reinforced laboratory test wall, the Denver Wall, has been simu­

lated using the finite element method. The granular backfill of the wall has been modelled

by the H iS S S\ constituive model. The interfaces between the reinforcements and backfill

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
189

L —
1

Uniform
TfTrrm
Trapezoidal Meyerhof
2 e

rein .
0.4 s o il
<rv(calculQted)=7 ( H - h ) + q
TJ w all
o 0 .6 HX
3O
aQ
> 0.8

-o
•u
3n
a
E
b>
0.0 0 .2 0.4 0 .6 0.8
x/L
RMC Incremental wall, q=0, H=3m, h/H = 0
RMC propped wall, q=:50kPa, H=6.1m, h /H = 0
Tuscon wall (Desert E arth Eng., 1987) H=4.57m, h/H =0.33
Lithonia wall (Watson,1986), H=2.44m, h /H = 0
RMC w rap-around wall (Lescoutre, 1986), H=1.9m, h/H = 0

Figure 7.25: Field observation of vertical stress at base (After Ho, 1993)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
190

3.0 -i----- r ---- ,-----r ' "- | ' i1 i , |

• ♦ ■

• ♦ ■

• ♦ ■
8 2.0
<D • ♦ ■
■B
B . • ♦ ■

S • ♦ ■
<
0t>
1 i.o - • ♦ ■ -
• FEM
Q • ♦ ■
■ Broms
• ♦ ■ ♦ Bonaparte e t al.

• ♦ ■
0.0 . i i .i___ .___ i___ i i . i
0.0 2.0 4.0 6.0 8.0 10.0 12.0 14.0 16.0
Axial Load, kN/m

Figure 7.26: Axial load in reinforcements: 103 kPa surcharge

and the facing and backfill have been represented by the thin-layer element using a hyper­
bolic model. The measured behavior of the wall has been compared with the prediction
of the finite element analysis in terms of facing movement, strains in the reinforcements,

and the top surface displacement The measured facing movement and the reinforcement

strains agree well with the predictions. The present analysis underpredicts the top surface
displacement

A verification of the assumptions made in the current design methods reveal some

mismatches between the assumptions and the results of the present analysis. The verti­
cal stress distributions at the foundation shows nonlinearity, but it doesn’t conform to the

trapezoidal or Mayerhof type stress distribution assumed in some design methods. The

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
191

3.0 - 1—1 I - 1 1 1 1 1 1 1

• ♦ ■

• ♦ ■

• ♦ ■
8 2.0
CD • ♦ ■
5
e • ♦ ■

6 • ♦ ■
<D
a
1.0 - • ♦ ■ _
CD
s • ♦ ■


FEM
Broms
■ • ♦ ■ ♦ Bonaparte et al.

• ♦ ■
o.o 1 . 1 1 i
0.0 4.0 8.0 12.0 16.0 20.0 24.0
Axial Load, kN/rn

Figure 7.27: Axial load in reinforcements: 186 k P a surcharge

current design methods overpredict the maximum load in the reinforcements. The present
study doesn’t show any definite failure plane. However, the elements, which are in plastic
failure state, show a resemblance to the failure plane assumed in the Coherent Gravity

method. The horizontal stress distribution on the facing is oscillating in nature, and the
magnitudes of the horizontal stresses are less than the Rankine active earth pressure in
most of the region except near the base and the top of the wall. The increase in horizontal

stresses near the base is due to the restraint provided by the rigid foundation and the toe
of the wall, whereas the increase in horizontal stresses near the top could be due to the re­

straint provided by the drop gate used to place the surcharge. Because of this restraint, the

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
192

Elements in plastic failure


Coherent Gravity

Rankrne

Figure 7.28: Plastic failure of soil elements at 103 kPa surcharge

behavior of the Denver Wall may not reflect the behavior of a real geosynthetic reinforced
field wall.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
193

— Elements in plastic failure


Coherent Gravity

— Rankine
2.97 m

H/2

Figure 7.29: Plastic failure of soil elements at 186 kPa surcharge

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
CHAPTER 8
NUMERICAL SIMULATION OF RMC WALL

8.1 General
In this chapter, the finite element simulation of a full-scale geosynthetic reinforced test

wall which was built and tested at the Royal Military College of Canada has been de­

scribed. The detail of the test has been given in Bathurst et al. (1990). The test was carried

out in the R M C Retaining Wall Test Facility (Figure 8 .1) which consists of six rigid rein­
forced concrete counterfort cantilever wall modules and may contain a block of soil up to

3.8 m high, 6 m long and 2.4 m wide. The R M C Wall is different from the Denver Wall

in number of ways. A full height aluminium facing was used in the R M C Wall which

was much more rigid than the plywood facing used in the Denver Wall. The R M C Wall
was reinforced with geogrids which are stronger than the geotextiles used in the Denver
Wall. The main difference in the testing conditions is that the drop gate used in the Denver

Wall to hold the surcharge provided considerable restraint in wall deformation contrary to
reality, where such a restraint is hardly present in geosynthetic reinforced earth walls. For

the R M C wall, no such unrealistic restraint at the top of the wall was present.

8.2 RMC Wall


The R M C Wall is a 3 m high, 6 m long and 2.4 m wide model reinforced wall. A uni­

formly graded sand was used as the backfill. The average density of the sand backfill was
18 k N / m 3. The wall facings were constructed with 0.75 m high panels bolted together to

make a full height panel. The facing consisted of three columns of panels. The two outer

columns were 0.7 m wide and the central instrumented column was 1 m wide. Each panel

was connected to a strip of geogrid reinforcement extending 3 m into the soil backfill. The

reinforcements were placed at a vertical distance of 0.25 m, 1.0 m, 1.75 m and 2.5 m from

194

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
195

HOLLOW STRUCTURAL
SECTION
2 .4

MODULE
NUMBER

4m

6 m
FRONT EDGE
OF T E ST
FACILITY
ANCHOR
BOLTS

Figure 8.1: R M C retaining wall test facility (After Bathurst et al., 1993)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
196

the pin connection at the base of the wall. There was a 0.5 m thick well compacted sand

layer between the base of the wall and the base slab of the test facility. A composite of

plywood, plexiglass and plyethylene sheeting was used to reduce the side wall friction for
ensuring plain strain condition.

The wall was constructed as a full height panel wall where the full facing was supported

externally during the fill placement The supports were removed after the full height of fill
had been placed.

The surcharge load on the wall was applied through airbags confined between the
backfill and top of the test facility. The surcharge load was applied in several increments

until the wall failed. The wall was allowed to creep for at least 100 hours at each load
increment A schematic configuration of the wall and the instrumentaion are shown in
Figure 8.2.

displacement
potentiometer
surcharge

I
3m
load ring . '» ■ ■
Layer 4
■ - -----
earth pressure cell
, strain gage ^ 3
geosynthetic
^ reinforcement Layer 2

1
0.5m
T
MM-4C
f
r extensometer -
Layer 1
earth pressure cell

Figure 8.2: R M C propped panel wall test (After Bathurst et al., 1992)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
197

8.3 Material models


As it has been observed for the Denver Wall, a number of different constitutive models are

required to represent the various components of a reinforced wall for the finite element
modelling. Before those constitutive models can be used in the finite element analysis,
they need to be calibrated against laboratory test data. In this section, the material models
used to represent the diffrenet elements of the R M C Wall and their calibration have been
described.

8.3.1 Backfill sand

The hierarchical single surface (H i S S ) di model described in Chapter 3 has been used to
represent the granular backfill material. The H iS S model is an elasto-plastic constitu­

tive model and is capable of characterizing most of the important behavior of sand, such
as strain hardening, shear dilation, stress dependent shear strength, etc.

The triaxial compression tests on the granular backfill reported by Bathurst et al.(1988)

were carried out at densities lower than the density at which the backfill was placed in the

R M C Wall. Because of that, the material parameters computed from those tests would
not produce the actual behavior of the backfill in the wall. Karpurapu and Bathurst(1995)

computed the material parameters for the hyperbolic model by trial and error. They con­
sidered the triaxial tests as a starting point and then adjusted the parameters based on the

results of direct shear tests carried out on sand specimens prepared at the same density
at which the backfill had been placed. For the calibration of the H i S S Si model, first,

three triaxial compression tests have been simulated using the hyperbolic model and the

material parameetrs reported by Karpurapu and Bathurst(1995). Then the deviatoric stress
and axial strain part of the simulated results and volumetric strain and axial strain part of

the triaxial tests have been used to calibrate the H iS S 5i model. The material parameters
computed from the above procedure are shown in Table 8.1. The model predictions are
shown in Figures 8.3 and 8.4. It can be observed in Figure 8.4 that the prediction shows

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
198

Table 8.1: Material constants for H iS S Si model for granular backfill

Elastic constants E 4.97 x 104 kPa

V 0.328

Ultimate 7 0.096

P 0.7

m -0.5

Phase change n 2.5

Hardening ax 8.56 x 10" 6

Vi 1 .2

Nonassociative K 0.384

more dilation than that shown by the triaxial test data. Since the density of the backfill in

the wall is higher than the density at which the triaxial tests were carried out, the higher
dilation produced by the material parameters can be taken as acceptable.

8.3.2 Reinforcement

Unlike the geo textile used in the Denver Wall, the geogrid reinforcement used in the R M C

Wall showed creep deformation. The creep hehaviorhas been accounted for in the present
study by using the isochronous load-strain-time data derived from constant load creep test

results using the method proposed by McGown et al.(1984). The results of creep tests
and the corresponding 100 hour isochronous curve for the geosynthetic reinforcements
are shown in Figure 8.5. The reason for using a 100 hour isochronous curve is that the
wall was allowed to creep for 100 hours under each surcharge increment

A nonlinear relation has been used based on the isochronous load-strain-time data
(Karpurapu and Bathurst 1995). In this model, the axial load, T, in the reinforcement is

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
199

300.0

200.0
/♦
CO
a
£$ ♦/

s
100.0 Test data (a3 = 30fcPa) _
Test data (<r3 = 45kPa)
Test data (a3 = 60kPa)
Predictioa (cr3 = 30k Pa)
Prediction (<x3 = 45fcPa)
Prediction (cr3 = 60kPa)

0.0 *—
0 .0 0 0 0.005 0.010 0.015 0.020 0.025
y h .d

Figure 8.3: Test data and prediction by H iS S 6i model for the granular backfill: yJJ^D
vs. -v/T^

related to the axial strain, e, in the reinforcement as:

T = Ae + Be2 (8.1)

where A and B are the material constants. The tangent stiffness, K t, can be obtained as:

Kt = ^ = A + 2Be (8.2)
OE

With A = 60 and B = —126, the nonlinear approximation matches well with the isochronous
curve as shown in Figure 8.5.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
200

0.010

0.008

0.006

0.004

0.002

0 .0 0 0

- 0.002
, f ♦♦ • Test data (cr3 = 30kPa)
■ Test data (<r3 = 45kPa)
-0.004 ♦ Test data (cr3 = 60k Pa)
— Prediction (cr3 = 30fcPa)
-0.006 Prediction (<r3 = 45k Pa)
Prediction (cr3 = 60/:Pa)
-0.008 —
0 .0 0 0 0.010 0.020 0.030
yhx)

Figure 8.4: Test data and prediction by H iS S S\ model for the granular backfill: — vs.
\Zho

8.3.3 Facing

The aluminium facing units have been represented as linear elastic. Karpurapu and Bathurst

(1995) didn’t report the stiffness parameters of the facing unit though, in their paper, it is

mentioned that stiffness parameters are found from a bending test on a facing unit Sharma
et al. (1994) have used E = 0.24 x 107 kPa and v = 0.17 in the finite element simu­

lation of a wall with similar facing units. After some sensitivity analyses, it is found that
E = 0.24 x 106 kPa and u = 0.17 give a similar deformation shape of the wall as ob­

served in the experiment The E values in the range of 0.24 x 106 k P a to 0.24 x 107

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
201

P
30 6
I

test data
non-fine ar
approxim ation
0
0 200 400 600 800 1000 0 0.05 0.10 0.15
100 time ( hrs) strain e (mm/mm)
a) constant load (creep) test data b) curve-fit for 100 hour
isochronous test data

Figure 8.5: 100 hour isochronous load-strain behavior of geogrid reinforcement (After
Karpurapu and Bathurst, 1995)

kPa do not change the stress condition in the soil and the maximum deformation of the
facing much, but they influence the deformed shape of the facing and the strains in the fac­

ing/reinforcement connections. In the present study, E = 0.24 x 106 kP a and v = 0.17

have been used. An analysis was carried out using E = 0.24 x 106 and u = 0.17 for the
facing. The results of the analysis are given in Appendix C.

8.3.4 Interfaces

The interfaces between the reinforcements and soil have been represented by the D SC
based constitutive model described in Chapter 5. It has been observed by many researchers
that the pull-out load capacity of a reinforcement is governed by the shear dilation at
interface between the reinforcement and the adjoining soil. The D S C based constitutive
model is capable of capturing the shear dilation in addition to strain hardening and strain

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
202

softening. The pull-out test simulation described in Chapter 6 shows the ability of the

D S C based constitutive model to predict pull-out load capacity correctly.


As it is mentioned in Chapter 6 that the slip between reinforcements and adjoining soil

layers does not occur at the physical interface between reinforcements and soil. The slip

generally occurs between the soil particles in the vicinity of the physical interface between
reinforcements and soil. In addition to that, it has also been found by other researchers that

the increase in shear strength of sand/geogrid interfaces is 0 —1 0 % of the shear strength of

sand under direct shear condition. In view of this observation, the D S C based constitutive

model has been calibrated using the results of direct shear tests carried out on the backfill

sand. Table 8.2 shows the material parameters obtained from the calibration. The test data
and the predictions of the D SC based model for the results of direct shear tests are shown
in Figure 8 .6 .

The surface of the aluminium facing unit is smooth compared to other facing types

(concrete, wood etc.). Therefore, the interfaces between the facing and the backfill have
been modelled as elastic with low shear stiffness.

8.4 Finite element model


The wall was modelled as a plane-strain two dimensional problem for the finite element
analysis. The finite element mesh used for the numerical simulation is shown in Figure 8.7.

The finite element mesh consists of 1574 nodes, 1008 four-node quadrature elements to

represent the soil, 72 two-node bar elements to represent the reinforcements, 26 four-node
quadrature elements, and 2 three-node triangular elements to model the facing and the
hinge connection beneath the facing. The thin layer elements (Sharma and Desai, 1992)
have been used to represent the interfaces between reinforcements and soil and between

soil and facing elements. There are 244 such elements used in the finite element mesh.
The nodes along the back of the wall are restricted to move along the horizontal di­
rection, but they are free to move along the vertical direction. The nodes at the bottom of

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
203

Table 8.2: Material constants for geogrid-soil interfaces

Elastic constants Kn 220 k P a /m m

Ks 57 k P a /m m

Intact state 7 102.19


ni 1.75

7^2 1.71
Oi 1 2 1 .2 1

V\ 0.0528

Critcal state M 0..922


a -1.703
b 0.171
Disturbed state K 2.2298

As 0.0206

kv 0.9

Av 0.5

Du 0.9

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
204

100.0
• Test data ux„ = 47 k P a )
■ Test data urn = 32.5 kPa)
• Test data (<rn = 17.5 kPa)
Prediction (an = 47 kPa)
Prediction (<r„ = 32.5 kPa)
Prediction (er„ = 17.5 kPa)

S 40.0

10.0
Shear displacement, u

(a)

3.0

55
,-r 2.0
c -

<u
E
a>

"a.
09
•3
13 • Test data (<rn = 47 kPa)
u • Test data lan = 32.5 kPa)
• Test data (<rn = 17.5 kPa)
Prediction (ct„ = 47 kPa)
Prediction (crn = 32.5 kPa)
Prediction (<t„ = 17.5 kPa)
- 1.0
0.0 2.0 4.0 6.0 8.0 10.0
Shear displacement, u

(b)

Figure 8 .6 : Test data and predictions of direct shear tests: (a) Shear stress vs. shear
displacement and (b) Normal displacement vs. shear displacement

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
205

Figure 8.7: FEM mesh for RMC Wall


e
VO

f
> <

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
206

the wall are restricted in both the directions beacuse the foundation of the wall is rough.

To simulate the propped construction of the wall, first, the equlibrium iteration has been
done, keeping the wall facing restrained in the horizontal direction. Once the equilibrium

has been reached, the horizontal reaction forces along the facing have been decreased to

zero, gradually, during subsequent steps. The surcharge has been placed after releasing
the horizontal force fully.

Initially, the self weight of the backfill and a 10 kPa pressure on the top have been
applied. After that, the 10 kPa surcharge pressure has been increased to 80 k P a in two
hundred load steps. The results of the analysis are furnished in the following section.

8.5 Result and discussion


The performance of a numerical model can be evaluated by comparing the measured quan­

tities with the predicted ones. In this section different measured results of the R M C Wall
have been compared with the predicted results obtained from the finite element analysis.

Some analysis has been done of the assumptions made in the current design methods based
on the findings of the present study.

Figure 8 .8 shows the measured and predicted lateral displacement profiles of the fac­

ing at 80 kPa surchrage prior to the failure. The finite element model estimates the lateral
displacement of the facing well. It should be noted that it is difficult to predict the col­
lapse state because the geogrid layers showed tertiary creep and the topmost geogrid layer

ruptured along 80% of its width at the collapse state. In the present analysis the creep
behavior is approximated by taking a 100 hour isochronous curve which does not include

tertiary creep. Therefore, the measured behavior prior to the failure at 80 kP a surcharge
has been used here and in all subsequent comparisons.

The measured and predicted lateral displacements at the mid-height of the wall during

surcharge steps are shown in Figure 8.9. The finite element prediction slightly overesti­

mates the lateral displacements, and it fails to capture the failure because the failure of

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
207

3.0

2.0

Measured
FEM Prediction

0.0
0.0 20.0 40.0 60.0 80.0 100.0
Wall Movement, m m

Figure 8 .8 : Facing movement of R M C Wall

the wall was initiated by a large creep in the reinforcements followed by rupture of the
top most reinforcement. The 100 hour isochronous curve used in the present study is not
able to capture the creep prior to the failure. As far as the working load is concerned, the
present prediction can be considered satisfactory.

Figure 8.10 to 8.13 illustrate the measured and predicted strains in the reinforcements.
Elevated strain levels at the reinforcement/facing connections arc observed in the finite el­

ement results. A similar trend has been observed by other researchers, especially when the

propped wall construction method is used and reinforcements are firmly connected with
the facing (Ho, 1993). It should be noted that the measured strains arc reported as average

strains along a certain length of reinforcements. They does not reflect true strain level

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
208

100.0

Measured
FEM Prediction
80.0

s
5 60.0
c
<u
i
f 40.0
V3
Q

20.0

0.0
0.0 20.0 40.0 60.0 80.0 100.0
Surcharge Pressure, kPa

Figure 8.9: Facing movement at mid-height with surcharge

at reinforcement/facing connections. If the strains are averaged along the corresponding

length, the strain distributions are fairly close to the measured strain distributions. Also
Karparapu and Bathurst (1995) reported that the accuracy of the measured strains in the
RMC Wall is ±1%. It can be noted in Figures 8.10 through 8.13 that the finite element
model has predicted well the attenuation of the grid strains at the end portion of the rein­

forcements. This observation implies that there is perfect bonding between the reinforce­

ments and the backfill along the end portion of the reinforcements which was manifested
in grid rupture at collapse.

The variation of horizontal and vertical force components at the hinge connection at the

base of the facing are shown in Figure 8.14. The predictions agree well with the measured

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
209

4.0

Measured
FEM Prediction (average)
3.0 w FEM Prediction
\
\
\
I
I
• I 2 -° h »
\
5 \

1.0

0.0
0.0 1.0 2.0 3.0
Distance from the facing, m

Figure 8.10: Axial strain in geogrid, layer 1

values. The additional vertical force carried by the facing is responsible for reduction in
vertical soil stresses near the facing. Though many investigators stated that the transfer of
vertical load to the facing is performed by the friction between the facing and the adjacent
fill (Ho, 1993; Karpurapu and Bathurst, 1995), the transfer of vertical load also takes place

through the reinforcements, especially when they are firmly connected with the facing.
This is evident in the present study where the shear strength of the interfaces between
the aluminium facing and the adjoining backfill is taken as small because of the smooth

surface of the aluminium facing.

Figures 8.15 and 8.16 illustrate the horizontal soil stress distributions behind the facing,

at a vertical plane through the middle of the reinforcements, and at a vertical plane at the

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
210

6.0 \ r - - i r-

-------- Measured
------- FEM Prediction (average)
-------- FEM Prediction
A
4.0 - \ _

\
$ \
c \
•a I
is \
C/3
\
\
2 .0 — V —
\
-------
\
*
s
/
f

1
/
/

1
1
1
1
1

1
1
1
1
1
1
1
1
J

I
-

0.0
0.0 1.0 2.0 3.0
Distance from the facing, m

Figure 8.11: Axial strain in geogrid, layer 2

end of the reinforcements. Except at the bottom, the horizontal soil stresses are smaller

than the horizontal stress for active condition. Like the Denver Wall, the horizontal stress
near the toe is larger than the theoretical at-rest pressure due to the constraint provided by

the toe against lateral movement. Because of the absence of any restraint against lateral

movement at the top of the wall, as was the case for the Denver Wall, the horizontal

stresses at the top portion of the wall are also lower than the theoretical active pressure.
Therefore, the explanation given for the larger horizontal stresses at the top portion of

the Denver Wall seems reasonable. The horizontal stress distribution behind the facing

is oscillating in nature. This trend is similar to the variation of vertical stresses near the

facing as shown in Figures 8.17 and 8.18. This variation in the vertical stress distributions

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
211

8 .0

— Measured
X FEM Prediction (average)
6 .0 FEM Prediction
Strain, %

4.0 \

K
• \
'--V
2 .0 \

0 .0
0 .0 1 .0 2 .0 3.0
Distance from the facing, m

Figure 8.12: Axial strain in geogrid, layer 3

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
212

15.0 *-- ■ r - ---f ---

— Measured
------- FEM Prediction (average)
\ ------- FEM Prediction
\
10.0 - \
\
Strain, %

\
I
I
\

5.0 " V \ * _

\
\
V

0 .0 1
0 .0 1.0 2.0 3.0
Distance from the facing, m

Figure 8.13: Axial strain in geogrid, layer 4

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
213

50.0

Horizontal (measured)
40.0 Horizontal (FEM)
Vertical (measured)
Vertical (FEM)
Toe Load, kN/m

30.0

20.0

10.0

0 .0
0 .0 20.0 40.0 60.0 80.0
Surcharge Pressure, kPa

Figure 8.14: Horizontal and vertical loads at toe

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
214

4.0

* Behind facing
■* Middle of reinforcements
3.0 ♦ End of reinforcements
- Active pressure
- At rest pressure

0 .0
0 .0 20.0 40.0 60.0 80.0 100.0
Horizontal Stress, kPa

Figure 8.15: Horizontal stress distribution at 38 kPa surcharge pressure

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
215

4.0

Behind facing
Middle of reinforcements
3.0 End of reinforcements
Active pressure
At rest pressure

0 .0
0 .0 50.0 100.0 150.0
Horizontal Stress, kPa

Figure 8.16: Horizontal stress distribution at 80 kP a surcharge pressure

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
216

4.0
— • Behind facing
Middle of reinforcements
— End of reinforcements
7h + q
3.0

2 .0
'53
X reinforcement layer

1 .0

0 .0
0 .0 50.0 100.0 150.0 200.0
Vertical Stress, kPa

Figure 8.17: Vertical stress distribution at 38 kPa surcharge pressure

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
217

4.0
— • Behind facing
Middle of reinforcements
— End of reinforcements
7/1 + q
3.0
Height, m

2 .0

reinforcement layer

1.0

0.0
0.0 100.0 200.0 300.0
Vertical Stress, kPa

Figure 8.18: Vertical stress distribution at 80 kP a surcharge pressure

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
218

kPa

Distance from the facing, cm

Figure 8.19: Vertical stress contours at 38 kPa surcharge pressure

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
219

kPa

Distance from the facing, cm

Figure 8.20: Vertical stress contours at 80 kPa surcharge pressure

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
220

below the reinforcements near the facing will be less and results in a reduction in vertical
stress in those elements.

The vertical soil stress distribution at the base of the wall is shown in Figures 8.21

and 8.22. Similar to the nature of stress distribution observed in the Denver Wall, the

vertical soil stress is less than the theoretical value (i.e., av = 7 h) close to the facing due

to a partial transfer of vertical stress from the fill to the facing through the backfill/facing

friction and the reinforcements. The exception is observed near the toe where the vertical
stress shoots up beyond the theoretical value due to the rigid foundation which prevents
relative settlement between the fill and the facing. The vertical stress increases in the re­

gion away from the facing and stabilizes at almost theoretical value contrary to the Denver

Wall where the vertical stress increases beyond the theoretical value. One explanation of

this behavior could be the effect of the back wall in the case of the Denver Wall because

it is about 30 cm away from the edge of the reinforcements, whereas in the R M C Wall,

the back wall is 3 m away from the edge of the reinforcements. The field observations, as
shown in Figure7.25, show that in almost all cases the vertical stress decreases near the fac­

ing, but away from the facing, the vertical stress stabilizes above the theoretical value for

some cases and below the theoretical values for others. From Figures 8.21 and 8.22, it can

be said that for checking against bearing capacity, Meyerhof distribution may be suitable,

but for computing a factor of safety against sliding, a uniform vertical stress distribution
seems reasonable.

To design the spacing of the reinforcement and to choose a reinforcement type of

adequate strength, the tied-back wedge methods assume different horizontal stress distri­

butions to be carried by the reinforcement as shown in Figure 2.2. Figures 8.23 and 8.24

illustrate the maximum tensile forces in the reinforcements obtained from the finite ele­

ment analysis and those calculated from the Broms method and Bonaparte et al. method

for the surcharge pressures of 38 kPa and 80 kPa respectively. One can easily see in

these figures that the forces obtained from the finite element analysis are much less than

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
221

250.0

FEM

200.0 — Meyerhof
— Trapezoidal
53
a,
tCA! - 15°-°
S
CO
1 100.0

50.0

0.0
0.0 2.0 4.0 6.0
Distance from the facing, m

Figure 8.21: Vertical stress distribution on the foundation at 38 kPa surcharge

the forces computed from those methods, which implies that those semi-empirical design

methods are highly conservative in nature.


The present finite element analysis doesn’t show the development of shear band, but

if we look at the elements which are in plastic failure state, as shown in Figures 8.25
and 8.26, a pattern can be seen. Instead of reaching the failure state in a narrow shear

band, all the elements above a plane reach the plastic failure state. The state of failure is

checked by comparing the ratio of yJJ%o to J\ with the value of y /jF s which represents
the state of plastic failure for the H iS S model. Superimposed on these figures are the

Rankine failure surface and the failure surface assumed in the Coherent Gravity method.

The failure plane assumed in the Coherent Gravity method resembles the observed failure

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
222

250.0
+ FEM

200.0 Meyerhof
Trapezoidal
<3

tf 150.0
ca
s
00
1 100.0
t:
2

50.0

0.0
0.0 2.0 4.0 6.0
D istan ce fro m th e facing, m

Figure 8.22: Vertical stress distribution on the foundation at 80 kP a surcharge

pattern closer than the Rankine failure plane. In the tied-back wedge methods, the length
of reinforcement extending beyond the failure plane is considered to provide resistance
against pull-out. The identification of the actual failure plane in the geosynthetic reinforced
soil wall is important for that reason.

8 .6 S u m m ary

A large scale geogrid reinforced laboratory test wall, the R M C Wall, has been simulated

using the finite element method. The granular backfill of the wall has been modelled by

the H iS S S\ constitutive model. The interfaces between the reinforcements and backfill,
have been modelled by the DSC-based interface model described in Chapter 5. The finite

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
223

8 2.0
■*-»
-

CD
♦ ■
B
5
6
CD
0
1 1.0 -
• FEM
■ Broms
♦ Bonaparte et al.
■ ♦

0.0
0.0 2.0 4.0 6.0 8.0 10.0 12.0
A xial Load, kN/m

Figure 8.23: Axial load in reinforcements: 38 kP a surcharge

element representation of the interfaces has been made by the thin-layer element. The
measured behavior of the wall has been compared with the prediction of the finite element
analysis in terms of facing movement, strains in the reinforcements, and the horizontal and
vertical load at the toe. The measured facing movement and the horizontal and vertical

loads at the toe agree well with the predictions. The predicted average strain distributions
agree well with the measured strain distributions, though the actual predicted distributions

show some disagreement with the measured one. The attenuation of reinforcement strains

at the end portions of the reinforcements is predicted well by the present analysis.
A verification of the assumptions made in the current design methods reveals some
mismatches between the assumptions and the results of the present analysis. Like the

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
224

3.0

S
8 2.0
■4—i
<u ♦ ■
■S
s
Js
0<D
1 1.0
O • FEM
■ Broms
♦ Bonaparte et al.

0.0
0.0 5.0 10.0 15.0 20.0
A xial L oad, kN/m

Figure 8.24: Axial load in reinforcements: 80 kPa surcharge

Denver Wall, the vertical stress distribution at the foundation shows nonlinearity, and it

doesn’t conform to the trapezoidal or Mayerhof type stress distribution assumed in some
design methods. The current design methods overpredict the maximum load in the re­

inforcements. The present study doesn’t show any definite failure plane. However, the
elements, which are in plastic failure state, show a resemblance to the failure plane as­

sumed in the Coherent Gravity method as observed in the Denver Wall. The horizontal

stress distribution on the facing is oscillating in nature, and the magnitudes of the horizon­

tal stresses are less than the Rankine active earth pressure in most of the region except near

the base. The increase in horizontal stresses near the base is due to the restraint provided

by the rigid foundation and the toe of the wall. The oscillation of the horizontal stresses

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
225

Elements in plastic failure


Coherent Gravity
Rankine
3.5 m

H/2

I I I I I
r r r rr
i i i i i

3m

Figure 8.25: Plastic failure of soil elements at 38 kP a surcharge

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
226

— Elements in plastic failure


Coherent Gravity
Rankine

At i i i 1 1 1 1 i i ( i

I.U-U- j __c___ L_ -J_ _ i__a _ _ i__ i__i__ ___ __


iL - L . j. _ J __L _ . L_ _ i_- _i__i_ .1__ i - __ j . - -J__
i< » 1 1 1 i i i i i t i
frr-r-i 1
p > 1 1 1 i i 1 i x
3.5 m

H/2

. 1. i_
1I 1
i
. _ _l
L - J
(_
_ I__I . _L_ _
I « J
1 - -1 I__1 . - _ J

I _ J

II -- T
T * 1 i r
r- t

3m

Figure 8.26: Plastic failure of soil elements at 80 kP a surcharge

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
227

behind the facing is similar to the variation of vertical stresses near the facing, and could
be due to the interactions among the reinforcements, facing and backfill.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
CHAPTER 9
SUMMARY AND CONCLUSION

9.1 Summary
For a reliable and economical design of geosynthetic reinforced earth structures, it is

important to understand the stress-deformation behavior of geosynthetic reinforced earth

structures. The best way of understanding the stress-deformation behavior and verifying
the assumptions made in the current design methods is by the construction amd monitor­
ing of a large series of large scale field test walls. But the cost involved in such a scheme

prohibits performing a detailed experimental study. Numerical simulations provide an al­

ternative and cost effective means of performing such a study. In the present study, a finite
element model for numerical simulation of geosynthetic reinforced earth structures has

been established. In this numerical simulation the individual components of the geosyn­
thetic reinforced earth structure have been represented by different constitutive models
and these constitutive models together are able to capture the macroscopic behavior of the

geosynthetic reinforced earth structure.

In the present study, the granular backfill material is represented by the hierarchical
single surface (H iS S ) ^ model which is a relatively simple elasto-plastic constitutive

model, but is capable of capturing most of the important characteristics of a granular ma­
terial. The model has been implemented in the finite element procedure using four differ­
ent integration algorithms: (a) elastic predictor - plastic corrector, (b) plastic predictor -

plastic corrector, (c) Implicit integration method, and (d) Modified Euler method. Out of

these four methods, the plastic predictor - plastic corrector method has been found to be
the most accurate, and this method has been used in the present numerical simulation. A

new calibration technique has been developed using genetic algorithms (G A ). The new

228

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
229

technique is helpful when all the laboratory test data needed to calibrate the model are not

available.
To represent the interfaces between the reinforcement and adjoining soil, an elasto-

plastic constitutive model based on the disturbed state concept (D SC ) has been developed.

The model has been verified with the direct shear test results and the performance of the

model has been found to be highly satisfactory. The finite element implementation of the

model has been done in association with the thin layer element, and the implementation

has been validated by simulating large scale laboratory pull-out tests.

All the constitutive models representing the different components of the geosynthetic

reinforced earth structures have been implemented in the commerical finite element soft­

ware A BA Q U S. The finite element model of the geosynthetic reinforced earth structure
has been validated against observed behavior of two large scale geosynthetic reinforced

laboratory test walls: (a) the Denver Wall - a geotextile reinforced wall, and (b) the R M C

Wall - a geogrid reinforced wall. The two walls are different in their sizes, the ways they

are constructed, the types of facing used and the number of reinforcement layers and their

strength. The predictions of the numerical simulations agree well with the observed be­

havior of both the walls.

9.2 Conclusion
From the present study, the following conclusions can be derived:

I. The finite element simulation of the geosynthetic reinforced earth structure adopted

in this study provides good agreement between the observed and the predicted be­

havior of two large scale laboratory test walls. It provides an economical and re­

liable means of performing stress-deformation studies of geosynthetic reinforced

earth structures.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
230

2. The D SC based interface model developed in the present study is capable of char­
acterizing most of the important behavior of geosynthetic-soil interfaces, such as
strain softening, strain hardening, shear dilation, etc.

3. The D S C based interface model together with the thin layer element is capable

of capturing the pull-out condition of reinforcements in the finite element method.


This is reflected in its prediction of pull-out load vs. displacement for the large scale
laboratory pull-out tests.

4. The elastic predictor - plastic corrector method provides the most accurate integra­
tion of the H iS S Si model.

5. The stress distributions in the geosynthetic reinforced walls obtained from the nu­

merical simulations reveal that some stress distributions assumed in the current de­
sign methods may be wrong.

6 . The horizontal stress distribution on the facing is lower than the theoretical active

earth pressure except near the base where the value of the horizontal stress is larger

than the theoretical at-rest earth pressure due to the constraint provided by the rigid

foundation.

7. The vertical stress distribution near the facing is lower than the theoretical value
(i.e., 7 h) due to the transfer of load to the facing through the reinforcements. An

exception is observed near the toe where the vertical stress is higher than the theoret­
ical value due to the resistance provided by the rigid foundation against settlement
From the nature of the vertical stress distribution, it can be said that a Meyerhof

type distribution may be used for calculating factor of safety against bearing capac­
ity, but for computing factor of safety against sliding, the assumption of a uniform

stress distribution is more suitable.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
231

8. The finite element results do not show any definite failure plane. However, the

elements, which are in plastic failure state, show some resemblance to the failure
plane assumed in the Coherent Gravity method.

9. The finite element predictions of the strain distributions in the reinforcements cannot

be claimed to agree well with the measured ones. However, given consideration to
the accuracy of the measured data, the predictions can be considered satisfactory.

9.3 Recommendations for future research


The present research can be furthered by conducting parametric studies for better under­
standing of the effects of the different elements in a geosynthetic reinforced wall. The

results of the parametric study can be used to formulate a realistic design method.

The present finite element model has been validated against measured behavior of

laboratory test walls. The model needs to be validated against observed behavior of ac­
tual field walls. The creep behavior of reinforcements has been modelled in by taking
an isochronous curve. This can be improved by incorporating a proper creep model for
geosynthetic reinforcements.

In the present study, the backfill materials considered are dry sand. Further research

can be done to simulate geosynthetic reinforced soil walls with wet and cohesive backfills.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
REFERENCES
• Adeli, H. and Cheng, N.T.(1993)."Integrated Genetic Algorithm for Optimization of
Space Structures", J. Aerosp. Engrg., ASCE, 6(4), pp. 315-328.
• Adeli, H. and Cheng, N.T.(1994)."Concurrent Genetic Algorithm for Optimization
of Large Structures," Journal o f Aerospace Engineering, v 7, n 3, pp. 276-296.

• Adeli, H. and Cheng, N.T.(1994)."Augmented Lagrangian Genetic Algorithm for


Structural Optimization", J. Aerosp. Engrg., ASCE, 7(1), pp. 104-118.

• Adib, M., Mitchell, J. and Christopher, B.(1990)."Finite element modelling of re­


inforced soil walls and embankment," Proc. Design and Performance o f Earth Re­
taining Structures, ASCE, Cornell University, June 18-21, pp. 409-423.

• Andrawes, K.Z., Yogaraja, I., Yeo, K.C., and Saad, M.A.(1994)."An Experim ental
and Finite Element Study of Reinforced Soil Wall Behavior," 13th. Int. Conf. Soil
Mechanics and Foundation Engineering, New Delhi, India, pp. 1262-1264.

• Armaleh, S.H.(1991)."Modelling including testing of cohesionless soils using dis­


turbed state concept", Ph.D. Dissertation, The University of Arizona, Tucson, Ari­
zona.

• Armaleh, S.H. and Desai, C.S.(1994)."Modelling and testing of a cohesionless mate­


rial using the disturbed state concept", Journal o f Mechanical Behavior o f Materials,
Vol. 5, No. 3,279-295.

• Baker, R. and Garber, M.(1977)."Variational approach to Slope Stability," Proceed­


ings o f 9th International Conference o f Soil Mechanics and Foundation Engineering,
Tokyo, Japan, Vol. 2, pp.2-12.
• Basaran, C.(1994)."Finite element thermomechanical analysis of electronic packag­
ing problems using disturbed state constitutive models", Ph.D. Dissertation, Uni­
versity of Arizona, Tucson, Arizona.
• Bathurst, R.J. and Benjamin, D.J.(1990)."Failure of a geogrid reinforced soil wall,"
Transportation Research Record, No. 1288, Washington, DC, pp. 109-116.
• Bathurst, R.J., Jaret, P.M. and Benjamin, D.J.R.S.(1993)." A database of results from
an incrementally constructed geogrid-reinforced soil wall test", Proc. o f Soil Rein­
forcement: Full Scale Experiment o f 80 ‘s, ISSMFE/ENPC, Paris, France, pp. 401-
430.
• Bathurst, R.J., Karpurapu, R. and Jaret, P.M.(1992)."Finite element analysis of a
geogrid reinforced soil wall", Proc. Grouting, Soil Improvement and Geosynthetics,
ASCE Geotechnical Special Publication No. 30, R.H. Borden, R.D. Holtz, and I.
Juran, eds., Vol. 2,1213-1224.

232

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
233

• Bonaparte, R., Holtz, R.D. and Giroud, J.P.(1987)."Soil reinforcement design using
geotextiles and geogrids," Geotextile Testing and Design Engineer, ASTM STP 952,
ed. J.E. Fluet, Jr. American Society for Testing and Materials, Philadelphia, PA, pp.
69-116.

• Boija, R.L and Lee, S.R.(1990).”Cam-clay plasticity: implicit integration of elasto-


plastic constitutive relations," Computer Methods in Applied Mechanics and Engi­
neering, Vol. 78, pp. 49-72.

• Broms, B.B.(1978)."Design of Fabric Reinforced Retaining Structures," Proceed­


ings o f the Symposium on Earth Reinforcement, ASCE, Pittsburgh.

• Chakroborty, P., Deb, K. and Subrahmanyam, P.S.(1995)."Optimal Scheduling of


Urban Transit System Using Genetic Algorithm," Journal o f Transportation Engi­
neering, v 121, n 6 , pp. 544-553.

• Chan, W.T., Fwa, T. and F.Tan, C.Y.(1994)."Road-Maintenance Planning Using Ge­


netic Algorithms: I," Journal o f Transportation Engineering, v 120, n 5, pp. 693-
709.

• Chew, S.H., Schmertmann, G.R. and Mitchell, J.K.(1990)."Reinforced soil wall de­
formations by finite element method," Performance o f Reinforced Soil Structures,
British Geotechnical Society, pp. 35-40.

• Chou, N.N.S.(1992),"Performance of geosynthetic reinforced soil walls" JPh.D. Dis­


sertation, University of Colorado at Boulder, Colorado, USA.

• Clayboum, A.F. and Wu, J.T.H.(1993). "Geosynthetic-Reinforced Soil Wall Design,"


Geotextiles and Geomembranes, 12(1993), pp. 707-724.

• Clough, G.W. and Duncan, J.M.(1969)."Finite element analyses of Port Allen and
Old River Locks", Report No. TE-69-3, U.S. Army Engineers, Waterways Experi­
ment Station, University of California, Berkeley, CA.

• Clough, G.W. and Duncan, J.M.(1971)."Finite element analyses of retaining wall


behavior," Journal o f Soil Mechanics and Foundation Div., ASCE, Vol. 97, No.
SM12.

• Collin, J.G.(1986)."Earth Wall Design," Ph.D. Dissertation, University of California


at Berkeley, California.

• Davis, L.(1991). Handbook o f Genetic Algorithms, Van Nostrand Reinhold, New


York.

• DeJong, K.A.(1975)."An Analysis of the Behavior of a Class of Genetic Adaptive


System", PhD Thesis, University of Michigan, Ann Arbor, Michigan, USA.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
234

• Desai, C.S.(1979)."Some aspects of constitutive laws of geologic media," in W. Wit-


tke (Ed.), Proc. 3rd International Conference on Numerical Methods in Geome­
chanics, Aachen, W. Germany.

• Desai, C.S.(1980)."A General Basis for Yield, Failure and Potential Functions in
Plasticity," International Journal of Numerical and Analytical Methods in Geome­
chanics, Vol. 4, pp. 361-375.

• Desai, C.S.(1987). "Further on unified hierarchical models based on alternative cor­


rection or damage approach", Report, Dept of Civil Eng. and Eng. Mech., The
University of Arizona, Tucson, Arizona.

• Desai, C.S.(1992)."The disturbed state as a phase transformation through self - ad­


justment concept for modeling of mechanical response of materials and interfaces",
Report, Department of Civil Engineering and Engineering Mechanics, University of
Arizona.

• Desai, C.S. and Fishman, K.L.( 1991)."Plasticity-based constitutive model with as­
sociated testing for joints", International Journal o f Rock Mechanics and Mining
Sciences, Vol. 28, No. 1, 15-26.

• Desai, C.S., Galagoda, H.M. and Wathugala, G.W.( 1987)."Hierarchical modelling


for geologic materials and discontinuities-joints and interfaces", Proc. o f Second
International Conference on Constitutive Laws fo r Engineering Materials: Theory
and Applications, Tucson, Arizona, Desai et al. (Eds), Elsevier, New York, Vol. 1,
81-95.

• Desai, C.S. and Ma, Y.(1992)."ModeIling of joints and interfaces using the disturbed
state concept", International Journal fo r Numerical and Analytical Methods in Ge­
omechanics, Vol. 16, 623-653.

• Desai, C.S, Sharma, K.G., Wathugala, G.W., and Rigby, D.B.(1991). "Implemen­
tation of Hierarchical Single Surface and Models in Finite Element Procedures,"
International Journal o f Analytical and Numerical Methods in Geomechanics, Vol.
15, 1991, pp 649-680.

• Desai, C.S. and Siriwardane, H.J.(1984). Constitutive Laws for Engineering Mate­
rials, Prentice-Hall Inc., Angled Cliffs, New Jersey.

• Desai, C.S., Somasundaram, S. and Frantziskonis, G.(1986)."A hierarchical ap­


proach for constitutive modelling of geologic materials", International Journal fo r
Numerical and Analytical Methods in Geomechanics, Vol. 10, No. 3,225-257.

• Desai, C.S., Somasundaram, S., and Frantziskonis, G.(1986)."A hierarchical Ap­


proach for Constitutive Modeling of Geologic Materials," International Journal o f
Numerical and Analytical Methods in Geomechanics, Vol. 10, pp. 225-257.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
235

• Desai, C.S. and Toth, J.(1996). "Disturbed state constitutive modeling based on stress-
strain and nondestructive behavior", International Journal o f Soilds and Structures,
Vol. 33, No. 11,1619-1650.

• Desai, C.S. and Wathugala, G.W.(1987)."Hierarchical and unified models for solids
and discontinuities (joints/interfaces)", Short Course Notes, 2nd International Con­
ference on Constitutive Laws fo r Engineering Materials: Theory and Application,
Tucson, Arizona, pp. 31-124.

• Desai, C.S, Wathugala, G.W. and Matlock, H.(1900)."Constitutive Model for Cyclic
Behavior of Cohesive Soils: II Applications," Journal o f Geotechnical Engineering,
ASCE, Vol. ,pp.

• Desai, C.S., Wathugala, G.W., Sharma, S.K., and Woo, L.(1990).”Factors affecting
reliability of computer solution with hierarchical single surface models,” Interna­
tional Journal o f Computer Methods in Applied Mechanics and Engineering, Vol.
82, pp. 115-137.

• Desai, C.S. and Woo, L.(1993)."Damage model and implementation in nonlinear


dynamic problems", Computational Mechanics, Vol. 11, No. 2,189-206.

• Desai, C.S., Zaman, M.M., Lighter, G.J., and Siriwardene, H.J.(1984)."Thin-Layer


Element for Interfaces and Joints," International Journal o f Numerical and Analyti­
cal Methods in Geomechanics, Vol. 8 , pp. 19-43.

• Desai C.S.(1989)."Notes for advanced school for numerical methods in geomechan­


ics including constitutive modelling," Notes, July 10-14, 1989, International Center
for Mechanical Sciences, Udine, Italy.

• Duncan, J.M.(1994). "The role of advanced constitutive relations in practical appli­


cations," Proc. 13th International Conference on Soil Mechanics and Foundation
Engineering, New Delhi, India.

• Duncan, J.M. and Chang C.Y.(1970)."Nonlinear analysis of stress and strain in


soils," Soil Mechanics and Foundation Div., ASCE, Vol. 96, No. SM5, pp. 1629-
1653.

• Ehrlich, M. and Mitchell J.K.(1994)."Working stress design method for reinforced


soil walls," Journal o f Geotechnical Engineering, ASCE, Vol. 120, No. 4, pp. 625-
645.

• Eshelman, L.J. and Schaffer, J.D.(1993)."Real-Coded Genetic Algorithms and Interval-


Schemata," Foundations o f Genetic Algorithms II, D. Whitley ed„ Morgan Kauf­
man, San Mateo, California, pp. 187-202.

• Fannin, R.J., and Hermann, S.(1990)."Performance Data for a Sloped Reinforced


Soil Wall," Canadian Geotechnical Journal, 27(1990), pp. 676-686.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
236

• Farrag, K.A.(1990)."Interaction properties of geogrids in reinforced-soil walls, test­


ing and analysis", Ph.D. Dissertation, Louisiana State University, Louisiana.

• Farrag, K., Acar, Y.B. and Juran, I.(1993)."Pull-out resitance of geogrid reinforce­
ments", Geotextile and Geomembranes, Vol. 12, No. 2.

• Faruque, M.O., and Desai, C.S., (1985)."Implementation of a General Constitutive


Model for Geologic Materials," J. ofNumer. andAnalyt. Methods in Geomechanics,
Vol. 9, pp. 415-436.
• Fishman, K.L., Desai, C.S., and Sogge, R.L.(1993)."Field Behavior of Instrumented
Geogrid Soil Reinforced Wall," Journal o f Geotechnical Engineering, ASCE, Vol.
119, No. 8 , pp. 1293-1307.

• Frantziskonis, G.(1986)."Progressive damage and constitutive behavior of geomate­


rials including analysis and implementation", Ph.D. Dissertation, The University of
Arizona, Tucson, Arizona.

• Frantziskonis, G. and Desai, C.S.(1987)."Constitutive Model with Strain Softening,"


International Journal o f Solids and Structures, Vol. 23, No. 6 , pp. 733-750.

• Frantziskonis, G.N., Desai, C.S., and Somasundaram, S.(1986)."Constitutive Mod­


elling for Nonassociative Behavior," /. Eng. Mech., ASCE, Vol. 9, pp. 932-946.

• Fwa, T.F., Tan, C.Y. and Chan, W.T.(1994)."Road-Maintenance Planning Using Ge­
netic Algorithms: II," Journal o f Transportation Engineering, v 120, n 5, pp. 710-
722.

• Galagoda, H.M.(1986). "Nonlinear Analysis of Soil Porous Media and Applica­


tion," Ph.D. Dissertation, University of Arizona, Tucson, USA.

• Ganendra, D. and Potts, D.M.(1994)."A comparison between explicit and implicit


stress point algorithms," Computer Methods and Advances in Geomechanics, Siri-
wardene and Zaman (editors), A.A. Balkema, Rotterdam, Netherlands, pp. 1975-
1980.

• Gens, A., Carol, E and Alonso, E.E.(1993)."An interface element formulation for
the analysis of soil-reinforcement interaction", Computers and Geotechnics, Vol. 7,
No. 1.

• Ghaboussi, J., Wilson, E.L. and Isenberg, J.(1973)."Finite element for rock joints
and interfaces," Journal o f Soil Mechanics and Foundation Div., ASCE, Vol. 99, No.
10, pp. 833-848.

• Goldberg, D.E.(1989). Genetic Algorithm in Search, Optimization and Machine


Learning, Addison-Wesley Publishing Company, Inc., Reading, Mass.

• Goldberg, D.E. and Samtani, M.P.(1986)."Engineering Optimization via Genetic Al­


gorithm", Proc. Ninth Conf on Electronic Computation, pp. 471-482.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
237

• Goodman, R.E., Taylor, R.L. and Brekke, T.L.(1968)."A model for the mechanics
of jointed rock," Journal o f Soil Mechanics and Foundation Div., ASCE, Vol. 94,
No. SM3, pp. 637-659.

• Gourc, J.P., Gotteland, P., and Villard, P.(1994)."Specific Problems Related to the
Design of Geosynthetics Earth Structures," Proc. 3rd. Int. Conf. Computer Methods
and Advances in Geomechanics, Siriwardene and Zaman (eds), pp. 1439-1354.

• Gourc, J.P., Ratel, A., and Delmas, P.(1986). "Design of Fabric Retaining Walls: the
Displacement Method," 3rd International Conference on Geotextiles and Geomem­
branes, Vienne.

• Hajela, P.(1990)."Genetic Search- An Approach to the Nonconvex Optimization


Problem", AIAA J., 28(7), pp. 1205-1210.

• Hajela, P. and Lee, E.(1995). "Genetic Algorithm in Truss Topological Optimiza­


tion," International Journal o f Solids and Structures, v 32, n 22, pp. 3341-3357.

• Hashmi, Q.S.E.(1986)."Nonassociative Plasticity Model for Cohesionless Materials


and Its Implementation in Soil-structure Interaction," Ph.D. Thesis, University of
Arizona, Tucson, Arizona, USA.

• HKS, Inc. (1996). ABAQUS/Standard User’s Manual: VolI-II, Version 5.6, Hibbit,
Karlsson and Sorensen, Inc., Pawtucket, Rhode Island.

• Ho, S.K. and Rowe, R.K.(1993)."Finite Element Analysis of Geosynthetics - Rein­


forced Soil Walls," Proceedings o f Geosynthetics ’93-Vancouver, Canada, pp. 203-
216.

• Ho, S.K. and Rowe, R.K.(1994)."Predicted Behavior of Two Centrifugal Model Soil
Walls," Journal o f Geotechnical Engineering, ASCE, Vol. 120, No. 10, pp. 1845-
1872.

• Ingold, T.S.(1993)."Laboratory pull-out testing of grid reinforcement in sand", Geotech­


nical Testing Journal, Vol. 6 , No. 3, 101-111.

• Jaber, M.B.(1989)."Behavior of Reinforced Soil Walls in Centrifuge Model Tests,"


Ph.D. Dissertation, University of California at Berkeley, California, USA.

• Jeremic, B. and Sture, S.(1995)."Implicit integration in geoplasticity," Proc. 10th


Engineering Mechanics Conference, ASCE, Boulder, Colorado, pp. 1099-1102.

• Jewell, R.A.(1987)."Reinforced Soil Wall Analysis and Behavior," Application o f


Polymeric Reinforcement in Soil Retaining Structures, NATO ASI Series E: Applied
Science, Vol. 147, Kluwer, Holand, pp. 365-408.

• Jewell, R.A.(1991)."Application of Revised Design Charts for Steep Reinforced


Slopes," Geotextiles and Geomembranes, 10(1991), pp. 203-233.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
238

• Jewell, R.A. and Milligan, G.W.E.(1989)."Deformation calculations for reinforced


soil walls,” Proc. 12th International Conference on Geotextiles, Vol. 4, Vienna,
Austria, pp. 1067-1072.

• Johnston, R.S. and Romstad, K.M.(1989)."Dilation and boundary effects in large


scale pull-out tests", Proc. XIII Int. Conf. Soil Mech. and Found. Engrg, Rio de
Janeiro, 1263-1266.

• Juran, I., and Farrag, K.(1994)."Strain Compatibility Simulation of Reinforced Soil


Structures," Proc. 13th. Int. Conf. Soil Mechanics and Foundation Engineering,
New Delhi, India.

• Juran, I., Ider, M.M., and Farrag, K.(1990)."Strain Compatibility Analysis for Geosyn­
thetics Reinforced Soil Walls," Journal o f Geotechnical Engineering, ASCE, Vol.
116, No. 2, pp. 312-329.

• Jwell, R.A. and Milligan, G.W.E.(1989)."Deformadon Calculations for Reinforced


Soil Walls," Proceedings o f 12th ICSMIE, Rio de Janeiro, Vol. 2, pp. 1257-1262.

• Karpurapu, R., and Bathurst, R.J.(1994)."Finite Element Analysis of Geosynthetic


Reinforced Soil Retaining Walls," 13th. Int. Conf. Soil Mechanics and Foundation
Engineering, New Delhi, India, pp. 1381-1384.

• Karpurapu, R. and Bathurst, R.J.(1995). "Behavior of geosynthetic reinforced soil


retaining walls using the finite element method", Computers and Geotechnics, Vol.
17,279-299.

• Katti, D.R.(1991)."Modelling including associated testing of cohesive soil using dis­


turbed state concept", Ph.D. Dissertation, The University of Arizona, Tucson, Ari­
zona.

• Katti, D.R. and Desai, C.S.( 1995)."Modeling and testing of cohesive soil using
disturbed-state concept", Journal o f Engineering Mechanics, Vol. 121, No. 5, 648-
658.

• Koemer, B.(1986)."Direct shear/pull-out tests on geogrids", Report No. 1, Dept of


Civil Engineering, Drexel University, Philadelphia, PA.

• Koumousis, V. K. and Georgiou, P. G.(1994)."Genetic Algorithms in Discrete Opti­


mization of Steel Truss Roofs," Journal o f Computing in Civil Engineering, v 8 , n
3, pp. 309-325.

• Lee, T.W.(1993)."Finite element analysis of reinforced soil structures," Ph.D. Dis­


sertation, The University of Texas at Arlington, Texas, USA.

• Leshchinsky, D. and Perry, E.B.(1987)."A Design Procedure for Geotextile - Rein­


forced Walls," Proceedings o f the Geosynthetics ’87 Conference, New Orleans, Vol.
1, pp. 95-107.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
239

• Li, J. and Kaliakin, V.N.(1994)." Application of improved zero-thickness interface


element to geosynthetically reinforced soil structures," Proc. 8th International Con­
ference on Computer Methods and Advances in Geomechanics , Siriwardene and
Zaman (Eds.), Morgantown, West Virginia, USA, June 22-28,1994, pp. 1367-1370.
• Ling, ELI. and Tatsuoka, F.(I992)."Nonlinear analysis of reinforced soil structures
by modified CANDE (M-CANDE)11, Geosynthetic-Reinforced Soil Retaining Walls,
ed. J.T.H. Wu, A.A. Balkema Publishers, Old Post Road, Brookfield, USA, pp.
279-291.
• Ling, H.I., Wu, J.T.H. and Tatsuoka, F.(1992)."Short-term strength and deformation
characteristics of geotextiles under typical operational conditions", Geotextiles and
Geomembranes, Vol. 10, No. 2.

• Ma, Y.(1992).”Constitutive modeling of joints and interfaces by using disturbed


state concept", Ph.D. Dissertation, The University of Arizona, Tucson, Arizona.

• Macari, E.J., Weihe, S. and Arduino, P.( 1997)."Implicit integration of elastoplastic


constitutive models for frictional materials with highly non-linear hardening func­
tions," Mechanics o f Cohesive-Frictional Materials, Vol. 2, pp. 1-29.
• Matsui, T. and San, K.C.(1989)."An elastoplastic joint element with its application
to reinforced slope cutting", Soils and Foundations, Vol. 29, No. 3,95-104.

• Matsui, T. and San, K.C.(1992)."Prediction of two test walls by elastoplastic finite


element analysis," Geosynthetic-Reinforced Soil Retaining Walls, ed. J.T.H. Wu,
A.A. Balkema, pp. 259-273.

• McGown, A., Andrews, K.Z. and Kabir, M.H.(1982)."Load-extension testing of


geotextiles confined in soil", Proc. Second International Conference on Geotextiles,
Las Vegas, 793-798.

• Murray, R.T.(1980)."Fabric Reinforced Earth Walls: Development of Design Equa­


tions," Ground Engineering, Vol 13, No. 7, pp. 29-36.

• Navayograjah, N., Desai, C.S., and ICiousis, P.D.(l992)."Hierarchical Single Sur­


face Model for Static and Cyclic Behavior of Interfaces," Journal o f Engineering
Mechanics, ASCE, 118(5), pp. 990-1011.

• Nayak, G.C. and Zienkiewicz, O.C.(1972)."Elasto-plastic Stress Analysis: A Gen­


eralization for Various Constitutive Laws Including Strain Softening," Int. J. Numer.
Methods Eng., Vol. 5, pp. 113-135.
• Ortiz, M. and Simo, J.C.(1986). "An Analysis of a New Class of Integration Algo­
rithms for Elasto-plastic Constitutive Relations," I. J. o f Numer. Methods in Engi­
neering, Vol. 23, pp. 353-366.
• Owen, D.R.J. and Hinton, E.(1980).Finite Elements in Plasticity: Theory and Prac­
tice, Pineridge Press, West Cross Lane, Swansea, U.K.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
240

• Potts, D.M., and Gens, A.(1985)." A Critical Assessment of Methods of Correction


for Drift from the Yield Surface in Elasto-plastic Finite Element Analysis," Int. J.
Num. Analyt. Meth. Geomechanics, Vol. 6 , pp. 149-159.

• Runesson, K. and Samuelsson, A.(1985)."Aspects on numerical technique in small


deformation plasticity," NUMETA 85 Numerical Methods in Engineering, Theory
and Applications, G.N.P.J. Middleton (Ed.), A.A. Balkema, pp. 337-347.

• San, K.C., Leshchinsky, D., and Matsu, T.(1994)."Geosynthetic Reinforced Slopes:


Limit Equilibrium and Finite Element Analyses," Soils and Foundations, Vol. 34,
No. 2, pp. 79-85.

• Sawicki, A., and Lesniewska, D.(1987)."Failure Modes and Bearing Capacity of


Reinforced Soil Retaining Walls," Geotextile and Geomembranes, 5(1987), pp. 29-
44.

• Sawicki, A., and Lesniewska, D.(1988). "Limit Analysis of Reinforced Slopes," Geo­
textiles and Geomembranes, 7(1988), pp. 203-220.

• Sawicki, A., and Lesniewska, D.(1991).”Stability of Fabric Reinforced Cohesive


Soil Slopes," Geotextile and Geomembranes, 10(1991), pp. 125-146.

• Schlosser, F.(1990)."Mechanically Stabilized Earth Retaining Structures in Europe,"


Design and Performance o f Earth Retaining Stuctures, ASCE Geotecthnical Special
Publication No. 25, pp. 347-378, Labbe, P.C., Hansen, L.A., Eds.

• Schmertmann, G.R., Chouery-Curtis, V.E., Johnson, R.D., and Bonaparte, R.(1987).


"Design Charts for Geogrid-reinforced Soil Slopes," Proceedings o f Geosynthetics
’87 Conference, New Orleans, pp. 108-120.

• Schneider, H.R., and Holtz, R.D.(1986)."Design of Slopes Reinforced with Geotex­


tiles and Geogrids," Geotextiles and Geomembranes, 3(1986), pp. 29-51.

• Schofield, A.N. and Wroth, C.P.(1968). Critical State Soil Mechanics, McGraw-Hill
Book Company.

• Sekiguchi, H. and Ohta, H.(1977),"Induced anisotropy and time dependency in


clays," Proc. 9th International Conference on Soil Mechanics and Foundation En­
gineering, Tokyo, Japan.

• Sharma, K.G. and Desai, C.S.(1992)."Analysis and Implementation of Thin-layer


Element for Interfaces and Joints," Journal o f Engineering Mechanics, ASCE, Vol.
118, No. 12, pp. 2442-2462.

• Sharma, K.G., Rao, G.V., and Raju, G.V.S.S.(1994)."Elasto-plastic Analysis of a


Reinforced Soil Wall by FEM," Proc. 3rd. Int. Conf. Computer Methods and
Advances in Geomechanics, Siriwardene and Zaman (eds), pp. 1415-1420.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
241

• Siel, B.D., Wu, J.T.H. and Chou, N.N.S.(1987)."In-soil stress-strain behavior of geo­
textile", Proc. Geosynthetics ’87 Conference, Feb 24-25, New Orleans, LA.

• Simac, M.R., Christopher, B.R., and Bonczkiewicz, C.(1990).”Instrumented Field


Performance of a 6 m Geogrid Soil Wall," Proceedings o f 4th International Con­
ference on Geotextiles, Geomembranes and Related Products, Vol. 1, The Hague,
Netherlands, A.A. Balkema, pp. 53-59.

• Simo, J.C. and Taylor, RJL.(1985)."Consistent tangent operators for rate indepen­
dent elastoplasticity," Computer Methods in Applied Mechanics and Engineering,
Vol. 48, pp. 101-118.

• Simpson, A.R. and Priest, S.D.(1993).”The Application of Genetic Algorithms to


Optimization Problems in Geotechnics," Computers and Geotechnics, v 15, n l,pp.
1-19.

• Sloan, S.W.( 1987)."Substepping Schemes for the Numerical Integration of Elasto­


plastic Stress-Strain Relation," Int. J. Numer. Methods Eng., Vol. 24, pp. 893-911.

• Stewart, J.E., Williamson, R., and Mohney, J.(1977)."Guideline for Use of Fabrics
in Construction and Maintenance of Low-volume Roads," National Technical Infor­
mation Service, Revised (1983).

• Syswerda, G.(1989)."Uniform Crossover in Genetic Algorithms", Proc. Third Int.


Conf. on Genetic Algorithms, George Mason University, June 4-7, Morgan Kauf-
mann Publishers, Inc., pp. 2-9.

• Tavassoli, M., and Bakeer, R.M.(1994)."Finite Element Study of Geotextile Em­


bankments," Proc. 13th Int. Conf. o f Soil Mechanics and Foundation Engineering,
New Delhi, India, pp. 1385-1388.

• TC(1986)."Guideline for the Design of Tensar Geogrid Reinforced Soil Retaining


Walls," Tensar Technical Note, TTN:RW1, The Tensar Corporation.

• Tesar, A. and Drzik, M.(1995)."Genetic Algorithm for Dynamic Tuning of Struc­


tures," Computers and Structures, v 57, n 2, pp. 287-295.

• Wathugala, G.W.(1990)."Finite element dynamic analysis of nonlinear porous me­


dia with applications to piles in saturated clays", PfuD. Dissertation, University of
Arizona, Tucson, Arizona.

• Wathugala, G.W. and Desai, C.S.(1987)."Constitutive Model for Soils with Strain
Softening and Shear Dilation," Internal Reserach Report, Dept, of Civil Engineering
and Engineering Mechanics, Unversity of Arizona, Tucson, Arizona.

• Wathugala, G.W. and Desai, C.S.(1989)."Damage based constitutive model for soils",
Proc. 12th Canadian Congress o f Appl. Mech., Ottawa, Canada, 530-531.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
242

• Wathugala, G.W., and Desai, C.S.(1991a)."Hierarchical Single Surface Model for


Anisotropic Hardening Cohesive Soils," Computer Methods and Advances in Ge­
omechanics: Proceedings o f the 7th Int. Conf. o f the International Association fo r
Computer Methods and Advances in Geomechanics, Eds. J. Beer, J.R. Booker and
J.P. Carter, May 6-10, Caims, Australia, A A Balkema, Rotterdam, Netherlands, pp
1249-1254.

• Wathugala, G.W., and Desai, C.S.(1991b)."Hierarchical Single Surface Model for


Isotropic Hardening Cohesive Soils,” Constitutive Laws fo r Engineering Materials:
Recent Advances and Industrial and infrastructure Applications, Desai et al (Edi­
tors), ASME Press, New York, pp 231-234.

• Wathugala, G.W. and Desai, C.S.,(1993)."Constitutive model for cyclic behavior of


clay I: Theory," Journal o f Geotechnical EngineeringASCE, Vol. 119, No. 4,pp.
714-729.

• Wathugala, G.W. and Pal, S.(1996)."Improving robustness of algorithms to imple­


ment HiSS models in FEM," Proc. 11th Engineering Mechanics Conference, ASCE,
Fort Lauderdale, Florida, USA, May 19-22,1996, pp. 144-147.

• Wilson E.L.(1977)."Finite elements for foundations, joints and fluids," Chap. 10 in


Gudehus (Ed.), Finite Elements in Geomechanics, Wiley, Chichester.

• Woods, R.I. and Jewell, R.R.(1990)."A computer Design Method for Reinforced
Soil Structures," Geotextiles and Geomembranes, 9(1990), pp. 233-259.

• Wright, A.H.(1991)."Genetic Algorithms for Real Parameter Optimization," Foun­


dations o f Genetic Algorithms, G. J. E. Rawlins ed., Morgan Kaufman, San Mateo,
California, pp. 205-218.

• Wu, J.T.H.(1991)."Measuring inherent load-extension properties of geotextiles for


design of reinforced structures", Geotechnical Testing Journal, Vol. 14, No. 2, 157-
165.

• Wu, J.T.H.(1992)."Predicting Performance of the Denver Walls: General Report,"


Geosynthetic-Reinforced Soil Retaining Walls, A.A. Balkema, pp. 3-20.

• Wu, J.T.H.,(1992) (Editor). Geosynthetics-Reinforced Soil Retaining Walls, A.A.


Balkema Publishers, Old Post Road, Brookfield, VT 05036, USA.

• Wu, S. and Chow, P.(1995)."Steady-State Genetic Algorithms for Discrete Opti­


mization of Trusses," Computers and Structures, v 56, n 6 , pp. 979-991.

• Yeo, K.C. and Yogarajah, I.(1992)."Prediction analysis of two geosynthetic rein­


forced retaining walls using finite element method", Geosynthetics-Reinforced Soil
Retaining Walls, ed. J.T.H. Wu, A.A. Balkema Publishers, Old Post Road, Brook­
field, USA, pp. 193-203.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
243

• Zienkiewicz, O.C.(1970)."Analysis of nonlinear problems with particular reference


to jointed rock systems," Proc. 2nd International Conference o f Society o f Rock
Mechanics, Belgrade, Vol. 3, pp. 501-509.

• Zienkiewicz, O.C.(1986). The Finite Element Method, McGraw-Hill Book Com­


pany (UK).

• Zomberg, J.G. and Mitchell, J.K.(1994)."Finite Element Prediction of the Perfor­


mance of an Instrumented Geotextile-reinforced Wall," Proc. 3rd. Int. Conf. Com­
puter Methods and Advances in Geomechanics, Siriwardane and Zaman (eds), pp.
1433-1438.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
APPENDIX A
INCORPORATION OF HiSS <yx MODEL INTO ABAQUS
The H iS S 61 model has been incorporated into A B A Q U S through an user defined

subroutine U M A T. The information concerning element number, gauss point number,

stresses at the beginning of the increment, total strain , strain increment, step no., and
iteration no. is passed to the subroutine, and the subroutine calculates stresses for the

given strain increment, and updates all the state dependent variables, such as trajectory

of volumetric, deviatoric and total plastic strains; accumulated plastic strains; and a sta­
tus value corrsponding to unloading, reloading and virgin loading. Developing U M A T

has been very involved, and most of the research effort was consumed by its coding and
implementation.

There are forty quantities which have to be defined in the input file. These quantities
consist of the material parameters, different flags, and different tolerance values. All these

quantities are defined, and their order in the input file is shown in Table A. 1.

The initial state dependent vaiables are defined through a user defined subroutine

S D V I N I . This subroutine calculates and returns all the initial values of the state de­
pendent variables for each gauss point depending on the initial state of stresses. Unfortu­
nately, A B A Q U S does not pass material parameters and initial stresses to this subroutine.
This information is provided by an input file called insdv.in. The quantities required to be
defined and their order in insdv.in are shown in Table A.2.

244

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
245

Table A.1: Parameters to be defined in the input file for the HiSS Si model

First Line:

EMAXVL Maximum strain increment for virgin loading.

EMAXRL Maximum strain increment for reloading.


EMAXUL Maximum strain increment for unloading.

SIGFACQ Factor used to calculate nP in the drift correction algorithm.


SIGFACF Factor used to calculate nF in the drift correction algorithm.

CEPFAC Factor used to calculate C'ep matrix in the CEP iteration

algorithm.

TOLF Tolerance for F.

TOLALPHA Tolerance for a.

Second Line:

M1TF1NDF Maximum iterations to find F.

MTTCEP Maximum iterations to find matrix.

M3TDRIFT Maximum iterations for drift correction.

METHOD Method of Integration (3 for E P — PC).

JHARD Type of hardening function (1 for the Si model).


JUL Type of interpolation function for unloading.
JRL Type of interpolation function for reloading.

JASO Type of flow rule (0 for associative and 1 for

nonassociative <5i model).

Third Line:

PA Atmospheric pressure.

XJ1SHBFT Ji intercept of the ultimate line.

continued on next page

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
246

ELASCIND (E/Pa)(l-i/ )/(l+r/ )(l-2i/ )


ELASC2ND (E/Pa)/(l+t/ )(l-2i/)

ELASC3ND (E/Pa)/2(l+i/)

ELASD1ND Pa/E

ELASD2ND -v Pa/E

ELASD3ND 2(l+i/ )Pa/E

Fourth Line:

GAMMA HiSS model parameter 7 .

BETA HiSS model parameter /?.

EM HiSS model parameter m.

EN HiSS model parameter n.


AKAPPA HiSS model parameter k .
AQ Anisotropy parameter A q (zero for the model).

HARDPM(l) Hardening parameter hi.


HARDPM(2) Hardening parameter h2.

Fifth Line:

HARDPM(3) Hardening parameter h3 (zero for the £1 model).


HARDPM(4) Hardening parameter h4 (zero for the 61 model).

RLPM(l) Reloading parameter ri (zero for elastic reloading).


RLPM(2) Reloading parameter r 2 (zero for elastic reloading).

RLPM(3) Reloading parameter r 3 (zero for elastic reloading).


ULPM(l) Unloading parameter ui (zero for elastic unloading).
ULPM(2) Unloading parameter u2 (zero for elastic unloading).

ALPHAQO Initial size of the potential surface Q, ctQa.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
247

Table A.2: Format of the input file insdv.in required for the subroutine SDVINI

First Line:

UNITWTS Unit weight of the material.


AKO Coefficient of earth pressure at rest, K 0.

Second Line:

YSURF Y-coordinate of the surface.

SSTRESS Surcharge pressure on the surface.

Third Line:

PA Atmospheric pressure.

Fourth Line:

GAMMA HiSS model parameter 7 .

BETA HiSS model parameter /?.

EM HiSS model parameter m.


EN HiSS model parameter n.

Fifth Line:
JHARD Type of hardening function (1 for the £1 model).
HARDPM(4) Hardening parameters (hi, h$ and h4).

Sixth Line:

JASSO Type of flow rule.

AKAPPA Nonassociative parameter «.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
APPENDIX B
DENVER WALL

248

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
249

300.0

200.0

JZ
an
'53
X
100.0

• Measured (103 kPa)


■ Measured (186 kPa)
~ FEM Prediction (103 kPa.)
~ FEM Prediction (186 kPa)

0.0
0.0 2.0 4.0 6.0 8.0 10.0 12.0 14.0
W all M ovem ent, cm

Figure B.l: Facing movement of Denver Wall (von Mises model for reinforcements, no
interfaces)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
250

0.0
cm

-4.0
Vertical Displacem ent,

- 12.0 • Measured (103 kPa)


■ Measured (186 kPa)
~ FEM Prediction (103 kPa)
- FEM Prediction (186 kPa)

-16.0
0.0 50.0 100.0 150.0 200.0 250.0
H orizontal D istance fro m the F acing, cm

Figure B.2: Top surface displacement profile of Denver Wall (von Mises model for rein­
forcements, no interfaces)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
251

15.0

Measured ( EOC)
Measured (103 kPa)
10.0 Measured (186 kPa)
FEM Prediction (EOC)
FEM Prediction (103 kPa)
FEM Prediction (186 kPa)

0.0

-5.0
0.0 50.0 100.0 150.0 200.0
H o rizo n tal D istan ce fro m th e F acing, cm

Figure B.3: Axial strain in reinforcement: Height = 0.15H (von Mises model for rein­
forcements, no interfaces)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
252

15.0

Measured ( EOC)
Measured (103 kPa)
Measured (186 kPa)
FEM Prediction {EOC)
FEM Prediction (103 kPa)
* 1° '° FEM Prediction (186 kPa)
‘I
a
C/3
si
‘S
< 5.0

0.0
0.0 50.0 100.0 150.0 200.0
Horizontal Distance from the Facing, cm

Figure B.4: Axial strain in reinforcement: Height = 0.52H (von Mises model for rein­
forcements, no interfaces)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
253

8.0

Measured {EOC)
■* Measured (103 kPa)
6.0 -*■ Measured (186 kPa)
“ FEM Prediction ( EOC)
■ FEM Prediction (103 kPn) .
a “ FEM Prediction (186 kPa)
•a
£ 3 4 0

•a
X
<

2.0

0.0
0.0 50.0 100.0 150.0 200.0
Horizontal Distance from the Facing, cm

Figure B.5: Axial strain in reinforcement: Height = O.8 8 H (von Mises model for rein­
forcements, no interfaces)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
254

300.0

200.0

100.0
• Measured (103 kPa)
■ Measured (186 kPa)
” FEM Prediction (103 kPa)
" FEM Prediction (186 kPa)

0.0
0.0 2.0 4.0 6.0 8.0 10.0 12.0 14.0
Wall Movement, cm

Figure B.6: Facing movement of Denver Wall (Hyperbolic model for reinforcements, no
interfaces)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
255

0.0
Vertical Displacement, cm

-4.0

- 12.0 • Measured (103 kPa)


■ Measured (186 kPn)
“ FEM Prediction (103 kPa)
■ FEM Prediction (186 kPa)

-16.0
0.0 50.0 100.0 150.0 200.0 250.0
Horizontal Distance from the Facing, cm

Figure B.7: Top surface displacement profile of Denver Wall (Hyperbolic model for rein­
forcements, no interfaces)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
256

15.0

Measured ( EOC)
Measured (103 kPa)
10.0 - — * Measured (186 kPa)
FEM Prediction (EOC)
1 FEM Prediction (103 fcPa)
c FEM Prediction (186 A:Pa)
’3
a 5.0 -
CO
■a
<

0.0

-5.0
0.0 50.0 100.0 150.0 200.0
Horizontal Distance from the Facing, cm

Figure B.8: Axial strain in reinforcement: Height = 0.15H (Hyperbolic model for rein­
forcements, no interfaces)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
257

15.0

Measured (EOCT)
Measured (103 kPa)
Measured (186 kPa)
FEM Prediction ( E O C)
10.0
FEM Prediction (103 kPa)
FEM Prediction (186 kPa)

5.0

0.0
0.0 50.0 100.0 150.0 200.0
Horizontal Distance from the Facing, cm

Figure B.9: Axial strain in reinforcement: Height = 0.52H (Hyperbolic model for rein­
forcements, no interfaces)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
258

8.0

Measured (EOC)
Measured (103 kPa)
6.0 Measured (186 kPa)
FEM Prediction {EOC)
FEM Prediction (103 kPa)
FEM Prediction (186 kPa)

2.0

0.0
0.0 50.0 100.0 150.0 200.0
Horizontal Distance from the Facing, cm

Figure B.10: Axial strain in reinforcement: Height = 0.88H (Hyperbolic model for rein­
forcements, no interfaces)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
259

300.0

200.0

sz
OD
‘53
DC
100.0

• Measured (103 kPa)


■ Measured (186 kPa)
- FEM Prediction (103 kPa )
* FEM Prediction (186 kPa)

0.0
0.0 2.0 4.0 6.0 8.0 10.0 12.0 14.0
W all M ovem ent, cm

Figure B.L1: Facing movement of Denver Wall (Hyperbolic model for reinforcements,
Goodman element for interfaces)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
260

0.0
Vertical Displacement, cm

-4.0

• Measured (103 kPa)


■ Measured (186 kPa)
- FEM Prediction (103 kPa)
" FEM Prediction (186 kPn)
-16.0
0.0 50.0 100.0 150.0 200.0 250.0
Horizontal Distance from the Facing, cm

Figure B.12: Top surface displacement profile of Denver Wall (Hyperbolic model for re­
inforcements, Goodman element for interfaces)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
261

15.0

Measured (EOC)
Measured (103 kPa)
10.0 Measured (186 kPa )
FEM Prediction ( EOC)
FEM Prediction (103 kPa)
FEM Prediction (186 kPa)

x
<

0.0

-5.0
0.0 50.0 100.0 150.0 200.0
Horizontal Distance from the Facing, cm

Figure B.13: Axial strain in reinforcement: Height = 0.15H (Hyperbolic model for rein­
forcements, Goodman element for interfaces)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
262

15.0

- ' • Measured ( EOC)


■ Measured (103 kPn)
* Measured (186 kPa)
FEM Prediction (EOC)
10.0
FEM Prediction (103 kPa)
c FEM Prediction (186 kPn)
•a \
£3
00
’S
<
5.0

0.0
0.0 50.0 100.0 150.0 200.0
Horizontal Distance from the Facing, cm

Figure B. 14: Axial strain in reinforcement: Height = 0.52H (Hyperbolic model for rein­
forcements, Goodman element for interfaces)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
263

8.0

"• Measured {EOC)


'■ Measured (103 kPn)
6.0 Measured (186 kPa)
- FEM Prediction ( EOC)
” FEM Prediction (103 kPa)
ef ~ FEM Prediction (186 kPa)
•a
b 40
CO
13
<

2.0

0.0
0.0 50.0 100.0 150.0 200.0
Horizontal Distance from the Facing, cm

Figure B.15: Axial strain in reinforcement: Height = 0.88H (Hyperbolic model for rein­
forcements, Goodman element for interfaces)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
APPENDIX C
RMC WALL

264

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
265

3.0

2.0

Measured)
FEM Prediction

0.0
0.0 20.0 40.0 60.0 80.0 100.0
Wall Movement, m m

Figure C.l: Facing movement of R M C Wall (E = .24 x 10-7 kPa facing stiffness)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
266

100.0

Measured)
FEM Prediction
80.0
Displacement, m m

60.0

20.0

0.0
0.0 20.0 40.0 60.0 80.0 100.0
Surcharge Pressure, kPa

Figure C.2: Facing movement at mid-height with surcharge (E = .24 x 10-7 kPa facing
stiffness)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
267

4.0 1 " ------- I " ■ “»-------

.
-------- Measured
-------- FEM Prediction (average)
3.0 -------- FEM Prediction

4CC 2.0 \
\
a \
00
\
I
\
1.0 - \
\
— \ ----------------
'
r ’ \”
V
^_
0.0 1
0.0 1.0 2.0 3.0
Distance from the facing, m

Figure C.3: Axial strain in geogrid, layer 1 (E = .24 x 10"7 k P a facing stiffness)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
268

6.0 r »■ i ■

------- Measured
------- FEM Prediction (average)
------- FEM Prediction
4.0 -
Strain, %

—\ x _.
\
\
2.0 \
\
\
r \
\
\
\
s

0.0 -- 1------ i . f— _
0. 0 1.0 2.0 3. 0
Distance from the facing, m

Figure C.4: Axial strain in geogrid, layer 2 (E = .24 x 10“ 7 k Pa facing stiffness)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
269

8.0

Measured
FEM Prediction (average)
6.0 FEM Prediction
-\-
\
\
$
4C3 4.0 ux..
b
CO

i\
• \
2.0 ----- T

0.0
0.0 1.0 2.0 3.0
Distance from the facing, m

Figure C.5: Axial strain in geogrid, layer 3 (E = .24 x 1Q“7 k Pa facing stiffness)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
270

20.0

Measured
FEM Prediction (average)
15.0 1 FEM Prediction
I
\
I
I
\
Strain,

10.0 1- \
\
\

Ut" i
5.0 \ '
------ V
V

0.0
0.0 1.0 2.0 3.0
D istance from the facing, m

Figure C.6: Axial strain in geogrid, layer 4 (E = .24 x 10-7 k P a facing stiffness)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
271

50.0

Horizontal (measured)
Horizontal (FEM)
40.0
Vertical (measured)
Vertical (FEM)
Toe Load, kN/m

30.0

20.0

10.0

0.0
0.0 20.0 40.0 60.0 80.0
Surcharge Pressure, kPa

Figure C.7: Horizontal and vertical loads at toe (E = .24 x 10" 7 k P a facing stiffness)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
VITA
Mr. Surajit Pal was bom in Calcutta, India, in 1965. He received a bachelor of civil
engineering degree from Jadavpur University, Calcutta, India, in 1987. He was awarded

the Jadavpur University Gold Medal for standing first in the Bachelor of Civil Engineering
Examination, 1987. After his graduation, he worked for four years as a senior design

engineer in a consulting engineering firm located in Calcutta, India, and was involved in
the analysis and design of the various structures of a nuclear power plant built in India. In

1991, he joined the State University of New York at Buffalo to pursue a master of science

degree in Civil Engineering. After completion of his master’s degree, he came to Louisiana

State University for doctoral study in Civil Engineering. He is also pursuing a master’s

degree in Systems Science at L.S.U. He has already published five technical papers. Mr.
Pal was married in January, 1997, and his wife, Gargi, now in Calcutta, is planning to join
him soon.

272

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
DOCTORAL EXAMINATION AND DISSERTATION REPORT

Candidates Surajit Pal

Major Fields Civil Engineering

*i^e Dissertations Numerical Simulation of Geosynthetic Reinforced Earth


Structures Using Finite Element Method

Appr oveds

(£»• W-
Major Professor and Chairman

Gradu'ate School

EXAMINING COMMITTEE:

Date off R»amination:

May 1 . 1QQ7

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
IMAGE EVALUATION
TEST TARGET (Q A -3 )
A

4
k
a

150m m

6"

V - i.
A P P L IE D A INA4 G E . Inc
1653 East Main Street
- = ~■ Rochester. NY 14609 USA
A
►'♦v - = ~— Phone: 716/482-0300
- ^ = r -= = Fax: 716/288-5989

i»y. / / 0 1993, Applied Image, Inc.. All Rights Reserved

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

Das könnte Ihnen auch gefallen