Вы находитесь на странице: 1из 624

INFORMATION TO USERS

While the m ost advanced technology has been used to


photograph and reproduce this manuscript, the quality of
the reproduction is heavily dependent upon the quality of
the material submitted. For example:
• Manuscript pages may have indistinct print. In such
cases, the best available copy has been filmed.
• Manuscripts may not always be complete. In such
cases, a note w ill indicate that it is not possible to
obtain m issing pages.
© Copyrighted material may have been removed from
the manuscript. In such cases, a note w ill indicate the
deletion.
Oversize materials (e.g., maps, drawings, and charts) are
photographed by sectioning the original, beginning at the
upper left-hand corner and continuing from left to right in
equal sections with sm all overlaps. Each oversize page is
also film ed as one exposure and is availab le, for an
additional charge, as a standard 35mm slide or as a 17”x 23”
black and white photographic print.
Most photographs reproduce acceptably on p ositive
microfilm or microfiche but lack the clarity on xerographic
copies made from the microfilm. For an additional chaise,
35mm slides of 6”x 9” black and white photographic prints
are available for any photographs or illustrations that
cannot be reproduced satisfactorily by xerography.

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
8700854

A larcon-G uzm an, A dolfo

CYCLIC STRESS-STRAIN AND LIQUEFACTION CHARACTERISTICS OF


SANDS. (VOLUMES I AND II)

Purdue University Ph.D. 1986

University
Microfilms
International 300 N. Zeeb Road, Ann Arbor, Ml 48106

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
3$oi)
Vo I■^
V-lOuuua. '2h
G

CYCLIC STRESS-STRAIN AND LIQUEFACTION

CHARACTERISTICS OF SANDS

VOLUME I

A Thesis

Submitted to the Faculty

of

Purdue University

by

Adolfo Alarcon~-Guzman

In Partial Fulfillment of the

Requirements for the Degree

of

Doctor of Philosophy

August 1986

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
PURDUE UNIVERSITY

O Graduate School

This is to certify that the thesis prepared

g y _______________ Adolfo Alarcon-Guzman___________________________________

Entitled
Cyclic Stress-Strain and Liquefaction Characteristics of Sands

Complies with University regulations and meets the standards o f the Graduate School for
originality and quality

For the degree o f ___________ Doctor of Philosophy___________________________

Signed by the final examining committee:

h
c __________________ _ chair

J ffa., —

Approved by the head of school or department:

i ___________ IQ A f

O is
This thesis s c is not to be regarded as confidential

Major professor
Grad. School
Form No. 9
Revised 9-85

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
ii

ACKNOWLEDGMENTS

The author wishes to express his deepest gratitude to

Professor G. A. Leonards, the author's advisor, for his

guidance, encouragement and for the time he generously

shared with the author.

The author also wants to give his deepest thanks to

Professor J, L. Chameau, the author's co-advisor, for his

guidance, support and friendship during the past four

years.

^ The author is thankful to the other members of his

committee, Professors M. E. Harr and C. D. Sutton.

The financial support of the National Science Founda­

tion is gratefully acknowledged. The author would like to

thank Professor P. Drnevich for his contributions to the

design and fabrication of the new resonant

column/torsional shear apparatus. The author also wants

to express his gratitude to the National University of

Colombia for the leave of absence which permitted the

author to complete his doctoral program.

The author is deeply grateful to Kathie Roth, who

assisted with the typing, and to Margarita Marin de Gomez,

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
who assisted with the drafting of the figures. The author

would like to thank all his friends and colleagues at Pur­

due for their friendship and the stimulating environment

they provided.

Finally, the author wishes to thank his sisters and

brothers for their continuous support and affection.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
iv

o
TABLE OF CONTENTS

Page

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

LIST OF F I G U R E S ......................................... viii

LIST OF S Y M B O L S ......................................... xvii

A B S T R A C T ................................................ xxi

CHAPTER 1 - INTRODUCTION ............................ 1

C HAPTER 2 - APPARATUS AND TECHNIQUES FOR


CYCLIC SHEAR TESTING ..................... 20

2.1 The Resonant Column Test . . . . . . . . . . 26


2.2 The Triaxlal T e s t ........................... 37
2.3 The Simple Shear Test . . . . . . . . . . . 46
2.4 The Torsional Shear T e s t .................... 58
C 2.5 Prediction of Soil Liquefaction . . . . . . 66

CHAPTER 3 - CYCLIC UNDRAINED BEHAVIOR OF


SANDS: A FRAMEWORK OF UNDERSTANDING . . 75

3.1 Pore Pressure Buildup and Strain


Development . . . . . . . . . . . . . . . . 80
3.1.1 Typical Results of Cyclic
Triaxlal Tests . . . . . . . . . . . . 80
3.1.2 Pore Pressure Rise Mechanisms . . . . . 86
3.1.3 Liquefaction* Flc«r Deformation
and Strain Development . . . . . . . . 92
3.1.4 Threshold Shear Strain Concept . . . . 99
3.2 Evaluation of Liquefaction Resistance . . . 103
3.2.1 Cyclic Undrained Strength . . . . . . . 103
3.2.2 Cyclic Shear Strain Potential . . . . . Ill
3.2.3 Volume Decrease Potential and
Liquefaction Resistance . . . . . . . . 117
3.2.4 Effect of Initial Confining Stress . . 120

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
V

Page

3.3 Monotonic and Cyclic Strength


Relationship .................................. 121
3.3.1 Drained Behavior under
Monotonic Loading . . . . . . . . . . . 121
3.3.2 Undrained Behavior under
Monotonic Loading .............. 129
3.3.3 Effect of Sand Type on
Undrained Behavior . . . . . . . . . . 164
3.3.4 The State Diagram and Cyclic
Undrained Behavior . . ................. 174
3.4 Effect of Initial State of Stress and
In Situ Boundary Conditions . . . . . . . . 192
3.4.1 Level Ground Conditions ................ 194
3.4.2 Sloping Ground Conditions . . . . . . . 208

CHAPTER 4 - LABORATORY EQUIPMENT AND


TESTING PROGRAM ......................... 231

4.1 New Resonant Column/Torsional Shear


Test Apparatus . . . . . . . . . . . . . . . 231
4.1.1 Resonant Column System ................ 236
4.1.2 Torsional Shear System ................ 240
4.1.3 Data Acquisition S y s t e m .............. 249
4.2 Materials and Testing Procedures .......... 251
s 4.2.1 Drained Tests . . . . . . .............. 254
f 4.2.2 Undrained Tests . . . . . . . . . . . . 256
4.3 General Purposes and Scope of the
Investigations ................ 258

CHAPTER 5 - TEST RESULTS AND A N A L Y S E S ............. 264

5.1 Drained Cyclic Shear Tests . ................ 264


5.1.1 Effects of Specimen Geometry . . . . . 264 ~
'5.1.2 Effects of Relative Density . . . . . . 269
5.1.3 Effects of Mean Confining Stress . . . 272
5.1.4 Effects of Frequency ............ ~. . 287
5.1.5 Effects of Shear Strain Level . . . . . 288
5.1.6 Effects of Stress R a t i o ................ 290
5.1.7 Effects of Stress History . . . . . . . 298
5.1.8 Effects of Cyclic Prestraining . . . . 309
5.2 Undrained Cyclic Shear Tests . . . ........ 324
5.2.1 Results of Typical Stress and
Strain Controlled Tests ................ 324
5.2.2 Threshold Shear Strain . . . . . . . . 353
5.2.3 Monotonic and Cyclic Strength
Relationship under Torsional Loading . 369
5.2.4 Effects of Prior Cyclic
Stress/Strain History ............. 400

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
vi

f Page

CHAPTER 6 - SUMMARY AND C O N C L U S I O N S ................ 437

6.1 General Considerations . . . . . . . . . . . 437


6.2 New Apparatus for Cyclic Shear Testing . . . 442
6.3 Test R e s u l t s ...................... 443
6.3.1 Drained Conditions . . . . . . . . . . 444
6.3.2 Undrained Conditions 451
6.4 Commentary . . . . . . . . . . . . . . . . . 459

CHAPTER 7 - RECOMMENDATIONS FOR FURTHER RESEARCH . . 462

7.1 Improvement of Laboratory Equipment . . . . 462


7.2 Topics for Further Research . . . . . . . . 466

BIBLIOGRAPHY ........................................... 471

APPENDICES

Appendix A Data Reduction Procedures


for Resonant Column Tests ............ 492
Appendix B Data Reduction Procedures
for Torsional Shear Tests . . . . . . 535

V I T A .................................................... 590

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
vii

LIST OF TABLES

Table Page

4.1 Switch Settings for Peak R o t a t i o n .............. 245

4.2 Stress and Deformation Limits of New Apparatus 248

4.3 Properties of Ottawa 20-30 Sand .......... 248

5.1 Summary of Tests Performed . . . . . . . . . . 265

Appendix
Table

A .1 Calibration Factors for Hardin Oscillator . . . 518

B.l Relationship between Average Shear Stress


s- and Applied Torque for New Apparatus . . . . . 557

B.2 Results of Pressure Transducer Calibration • . 567

B.3 Results of Load Cell Calibration . . . . . . . 572

B.4 Results of Torque Transducer Calibration . . . 577

B.5 Characteristics of Transducers for


Measuring Applied Boundary Stresses . . . . . . 581

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
viii

LIST OF FIGURES

Figure Page

2.1 Laboratory Techniques for Measuring


Dynamic Soil Properties . . . . . . . . . . . 24

2.2 Common Resonant Column Boundary Conditions . . 28

2.3 Stress Components in Cyclic


Triaxlal Test Specimens. ..................... 38

2.4 Theoretical Stresses on the Boundaries of


a Square Simple Shear Specimen . . . . . . . . 49

2.5 Stress Components in Hollow Torsional


Shear Test Specimens . . . . . . . . . . . . . 61

2.6 Rotation of Principal Stresses in a Soil


C Element In Situ during an Earthquake . . . . . 68

2.7 Stress Paths for Isotropically


Consolidated Specimens . . . . . . . . . . . . 71

3.1 Definition of State of Liquefaction . . . . . 77

3.2 Typical Results of Cyclic Triaxlal Tests


on Isotropically Consolidated Specimens . . . 81

3.3 Typical Effective Stress Paths in


Cyclic Triaxlal Tests on
Isotropically Consolidated Specimens . . . . . 85

3.4 Changes in Fabric during Cyclic Loading . . . 89

3.5 Shear Strain Potential at Liquefaction . . . . 94

3.6 Rate of Pore Pressure Buildup


during Cyclic Loading . . . . . . . . . . . . 98

3.7 Concept of Threshold Shear Strain . . . . . . 100

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w ith o u t perm ission.
ix

f Figure Page

3.8 Cyclic Stress Ratio - Shear Strain (Single


Amplitude) Relationship for Different Number
of Cycles at a Constant Relative Density . . . 104

3.9 Relationship between Cyclic Stress Ratio


and Number of Cycles Causing Different
Level of Shear Strain ........... . . . . . . 106

3.10 Cyclic Stress Ratio - Shear Strain


Relationship for Different Relative
Densities at a Constant Number of Cycles . . . 108

3.11 Cyclic Stress Ratio - Relative Density


Relationship for Different Shear Strain
Levels at a Constant Number of Cycles . . . . 109

3.12 Cyclic Stress Ratio - Relative Density


Relationship and Cyclic Shear Strain
Potential from Large-Scale Simple Shear Tests 112

3.13 Tentative Relationship between Cyclic Stress


Ratio, (N.)fen Values, and Limiting
Strain for Natural Deposits of Clean Sand . . 114

3.14 Relationship between Volume Decrease


C Potential and Cyclic Stress Ratio Causing
Initial Liquefaction in 20 Cycles ............ 119

3.15 Effect of Confining Pressure on Cyclic Stress


Ratio Versus Number of Cycles Relationship . . 122

3.16 State Diagram for Sacramento River Sand . . . 125

3.17 Drained Properties as a Function of State


Parameter for Kogyuk 350 S a n d s ................ 127

3.18 Dndrained Load Controlled Tests on


Isotropically Consolidated Specimens
of Banding S a n d .................. 131

3.19 Undrained State Paths Corresponding


to Tests in Figure 3 . 1 8 ............. 132

3.20 Steady— State Diagrams for Banding Sand . . . . 136

3.21 Peak Undrained Shear Strength as a


Function of Relative Density ........... 143

c •

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
X

c Figure

3.22 Normalized Peak Undrained Shear Strength


as a Function of State Parameter . . . . . . .
Page

145

3.23 Axial Strain at Peak Deflator Stress


as a Function of State Parameter . . . . . . . 147

3.24 Angle of Shearing Resistance Mobilized


at Peak Deviator Stress as a Function
of Relative Density ....................... 149

3.25 Angle of Shearing Resistance Mobilized


at Peak Deviator Stress as a Function
of State P a r a m e t e r .............................. 150

3.26 Relationship between Mobilized and Pore


Pressure Af — Parameter at Peak
Devlator Stress .............................. 153

3.27 Angle of Shearing Resistance Mobilized


at Steady-State Condition as a Function
of Relative Density . . . . . . . . . . . . . 155

3.28 Index of Undralned Brittleness as a


Function of Relative Density . . . . . . . . . 156

c 3.29 Index of Undralned Brittleness as a


Function of State Parameter . . . . . . . . . 158

3.30 Undrained Behavior at Constant


Relative Density (Schematic) ......... . . . . 160

3.31 Undralned Behavior at Constant Consolidation


Confining Stress (Schematic) . . . . . . . . . 163

3.32 Undrained Behavior at Constant


State Parameter (Schematic) . . . . . . . . . 165

3.33 Angle of Shearing Resistance Mobilized at


Peak Deviator Stress for Two Different Sands . 167

3.34 Steady-State Diagrams for Two Different Sands 168

3.35 Undrained Steady-State Diagrams


for Two Different Sands . . . . . . . . . . . 170

3.36 Undrained Behavior under Cyclic Loading


of Magnitude Smaller than
Steady-State Strength (Schematic) . . . . . . 175

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
xi

Figure Page

3.37 Concept of Critical Cyclic Stress Ratio . . . 180

3.38 Undrained Behavior under Cyclic Loading of


Magnitude Larger than Steady— State Strength • 183

3.39 Angle of Shearing Resistance Mobilized at


Initiation of Strain Softening Response
during Cyclic Loading . . . . . . . . . . . . 185

3.40 Development of Limited Flow Deformation and


Liquefaction during Cyclic Loading (Schematic) 189

3.41 Typical Cyclic Torsional Shear Tests with no


Lateral Strain on ^ - C o n s o l i d a t e d Specimens • 196

3.42 Cyclic Torsional Shear Tests under


Different Boundary and Stress Conditions . . . 199

3.43 Potential Effective Stress Paths


during Cyclic Loading under Different
Boundary and Stress Conditions (Schematic) . . 201

3.44 Cyclic Undrained Strength from Plane


Strain Torsional Shear Tests ................ . 204

3.45 Cyclic Undralned Strength from


Torsional Shear Tests at Constant
Mean Confining Stress . . . . . . ............ 205

3.46 Effect of Overconsolidation on Plane


Strain Cyclic Undrained Strength . . . . . . . 207

3.47 States of Stress under Cyclic Loading . . . . 210

3.48 Torsional Shear Test on Anisotropically


Consolidated Specimen with ( t + t ) Smaller
than Steady-State Strength (Schematic) . . . . 216

3.49 Torsional Shear Test on Anisotropically


Consolidated Specimen with ( t + t _ ) Larger
than Steady-State Strength (Schematic) . . . . 219

3.50 Torsional Shear Tests on Anisotropically


Consolidated Specimen with Static
Shear Stress Larger than Steady— State
Strength (Schematic) . . . . . . . . . . . . . 221

3.51 Flow Failure Potential from Strain


Controlled Cyclic Torsional Shear Tests . . . 223

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
xii

, Figure Page

— ' 3.52 In Situ Effective Stress Path with Stress


Reversal during Cyclic Loading (Schematic) • . 228

3.53 Effect of Static Shear Stresses on the


Resistance to Strain Development under
Cyclic Triaxlal Loading Conditions ............ 230

4.1 New Resonant Column/Torsional


Shear Test Apparatus . . . . . . . . . . . . . 233

4.2 Electronic System for Resonant Column Tests . 239

4.3 Wiring Diagram for Hardin


Resonant Column Apparatus . . . . . . . . . . 241

4.4 Torsional Shear System ................ 247

4.5 Elements Used for Construction of


Hoilow Cylinder Specimens . . . . . . . . . . 252

5.1 Effect of Specimen Geometry on Shear


Modulus — Shear Strain Relationship . . . . . 267

5.2 Effect of Relative Density on Shear


f- Modulus — Shear Strain Relationship . . . . . 270

5.3 Effect of Relative Density on


Maximum Shear Modulus . . . . . . . . . . . . 271

5.4 Effect of Confinement on Shear


Modulus — Shear Strain Relationship . . . . . 273

5.5 Shear Modulus Versus Mean Confining Stress . . 274

5.6 Effect of Confinement on the Shape


of the Modulus Reduction Curve . . . . . . . . 276

5.7 Dimenslonless Shear Modulus —


Shear Strain Relationship . . . . . . . . . . 279

5.8 Prediction of Modulus Reduction Curve . . . . 281

5.9 Effect of Confinement on Damping


Ratio - Shear Strain Relationship . . . . . . 283

5.10 Damping Ratio Versus Shear Modulus Ratio . . . 285

5.11 Effect of Consolidation Stress


Ratio on Modulus Reduction Curve . . . . . . . 292

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
xiii

Figure Page

5.12 Effect of Stress Ratio on


Maximum Shear Modulus . . . . . . . . . . . . 293

5.13 Effect of Isotropic Prestressing on


Shear Modulus - Shear Strain Relationship . . 299

5.14 Effect of Axial Compression Prestressing on


Shear Modulus - Shear Strain Relationship . . 300

5.15 Effect of Axial Compression Prestressing


on Maximum Shear Modulus of Medium
Dense Sand Specimen ....................... 303

5.16 Effect of Confinement on Shear Modulus


of Medium Dense Specimen . ......... . . . . . 305

5.17 Effect of Torsional Prestraining on


Shear Modulus - Shear Strain Relationship . . 310

5.18 Effect of Torsional Prestressing on


Shear Modulus - Shear Strain Relationship . . 312

5.19 Effect of Strain History on


Shear Modulus - Shear Strain Relationship . . 316

5.20 Effect of Number of Cycles at Constant


Shear Strain Amplitude on Shear Modulus . . . 319

5.21 Undralned Torsional Shear Test on


Isotropically Consolidated Specimen (Dr~63Z) . 325

5.22 Undralned Torsional Shear Test on


Isotropically Consolidated Specimen (Dr*50%) . 331

5.23 Undrained Torsional Shear Tests on


Isotropically Consolidated Medium Loose
Sand Specimen (Dr~48%) . . . . . . . . . . . . 335

5.24 Shear Modulus - Effective Confining Stress


Relationship during Undrained Cyclic Loading . 339

5.25 Undrained Torsional Shear Test on


Isotropically Consolidated Loose
Sand Specimen (Dr-44Z) . . . . . . . . . . . . 341

5.26 Typical Results of a Strain Controlled


Torsional Shear Test on Isotropically
Consolidated Specimen . . . . . . . . . . . . 347

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
xiv

Figure Page

5.27 Results of Undralned Resonant Column Tests


on Isotropically Consolidated Specimen . . . . 355

5.28 Pore Pressure Ratio Versus


Shear Strain R e l a t i o n s h i p .................. . 359

5.29 Effect of Number of Cycles at Constant


Shear Stress Amplitude on Shear Modulus . . . 362

5.30 Torsional Shear Tests with Small Cyclic


Shear Stress Amplitude ( t9*11.2 k ? a ) ............ 363

5.31 Monotonic Torsional Shear Test


on Isotropically Consolidated Specimen . . . . 370

5.32 Monotonic Torsional Shear Test on


Specimen with an Initial Static Shear Stress . 375

5.33 Cyclic Torsional Shear Test with


Shear Stress Amplitude Larger than
Steady-State Strength . . . . . . . . . . . . 380

5.34 Stress Conditions at Initiation of Strain


Softening Behavior under Cyclic Loading . . . 384

f 5.35 Number of Cycles to Initiation of


V Strain Softening Behavior . . . . . . . . . . 385

5.36 Cyclic Torsional Shear Test with


Non-Uniform Shear Stress Amplitude . . . . . . 391

5.37 Cyclic Torsional Shear Test with


Reversal Shortly after Initiation of
Strain Softening Behavior • ..................... 395

5.38 Effect of Number of Cycles on Secant Shear


Modulus under Undralned Cyclic Loading . . . . 402

5.39 Monotonic Torsional Shear Test


after Development of Initial
Liquefaction Due to Cyclic Loading . . . . . . 405

5.40 Cyclic Torsional Shear Tests on


Reconsolidated Specimen with Residual
Shear Strain ( - 3 . 8 Z ) ........................... 411

5.41 Cyclic Torsional Shear Test on


Virgin Specimen ( t^ ^ - 15.6 kPa) ........... .. 416

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
XV

Figure Page

5*42 Cyclic Torsional Shear Test on


Reconsolidated Specimens with
Residual Shear Strain (-1.0Z) 418

5 =43 Cyclic Torsional Shear Tests on


Virgin Specimen ( t^ - 17.5 kPa) ........... 422

5.44 Cyclic Torsional Shear Tests on


Reconsolidated Specimen with
Residual Shear Strain (+0.51) . . . 424

5.45 "Small Preshearlng" and "Large


Preshearing" Stress Paths under Cyclic
Loading of Magnitude Smaller than
Steady— State Strength . . . . . . . . . . . . 436

6.1 Development of Limited Flow Deformation


and Liquefaction under Cyclic Loading . . . . 440

Appendix
Figure

A.1 Torsional Vibrations of Elastic Rods . . . . 494

A.2 Resonant Column System (Schematic) . . . . . 498

A.3 Lissajous Figures from Two Sine


Waves 90° Out of Phase ........... 508

A.4 Methods for Determining Top


Platen System Rotational Inertia ............. 512

A.5 Steady-State Forced Vibrations of


Fixed Base Cylindrical Specimen . . . . . . . 525

A.6 Dimensionless Frequency Factor for Resonant


Column Apparatus with Base of Specimen Fixed . 530

A .7 Damping Amplification Coefficient


for Resonant Column Apparatus with
Base of Specimen Fixed . . . . . . 532

B.l Stress Components in Hollow Torsional


Shear Test Specimens . . . . . . . . . . . . 537

B.2 Radial Equilibrium in Torsional


Shear Test Specimens ............... 539

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
xv i

Appendix
( Figure Page

B.3 Potential Loading Schemes for


Cyclic Torsional Shear Tests ................ 542

B.4 Strain Components in Hollow


Torsional Shear Test Specimens . . . . . . . 547

B.5 Shear Stress Distributions in


Hollow Torsional Shear Tests Specimens . . . 560

B.6 Stress Components in Hollow Torsional


Shear Test Specimens Including End Effects . 562

B.7 Results of LVDT Calibration . . . . . . . . . 583

B.8 Definition of Equivalent Hysteretic Shear


Modulus and Damping Ratio . . . . . . . . . . 588

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
xv ii

LIST OF SYMBOLS

* Skempton's pore pressure parameter at failure

ADIO ■ Analog/Digital Input/Output

B - pore pressure parameter

CSR “ critical stress ratio

D * damping ratio

Dma x * limiting damping


r o ratio
Dr * relative density

£ m Young's modulus

e * void ratio

e » maximum void ratio


max
e .* minimum void ratio
mln
f ■ frequency

G = shear modulus

G = maximum shear modulus


max
IB * brittleness index

I.D. = inner diameter

= mass polar moment of inertia of active end

platen

K * principal stress ratio

Kg = consolidation stress ratio

Ko * coefficient of earth pressure at rest

(K = K in overconsolidated deposits
O UL O

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
MR » modular ratio

N « number of cycles

“ number of cycles liquefaction

(Ni)go “ standard penetration resistance at 60% of the

theoretical free fall energy

NC *• normally consolidated

NGI * Norwegian Geotecnnical Institute

OC = overconsolidated

OCR * overconsolidation ratio

O.D. - outer diameter

p" ■ /2* stress path parameter

PPS = pulses per second

q * (O j + O j )/2, stress path parameter

q » steady-state strength
ss
RAM 3* random access memory

SPT * standard penetration test

u *» pore pressure

Up = pore pressure at peak strength

ur = residual or terminal pore pressure

u^ «• pore pressure at triggering of strain

softening behavior

VD = volume decrease potential

VD = voltage divider

a = angle between direction and z axis

a ■ slope cf steady-state envelope


s
a <3 slope of critical stress ratio line

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
XIX

La cyclic deviator stress

L Ao
c
axial stress increment

La ar.ial stress decrement


e
At shear stress increment

axial strain

peak axial strain


P
Y shear strain amplitude

threshold shear strain

T -_/G reference strain

angle cf shear resistance

mobilized angle of shearing resistance at

steady-state condition

interparticle friction angle

mobilized angle of shearing resistance

c mob
* state parameter

confining pressure

radial stress

axial stress

we circumferential stress

consolidation deviator stress


Jdc
a. major principal stress

intermediate principal stress

minor principal stress

a* effective stress

o effective axial stress

effective confining stress

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
- effective horizontal stress

c o^ - mean effective confining stress

o *■ effective vertical stress


v
*
a » Initial effective vertical stress
vo
*
c*lc * consolidation major principal stress
*
o^c ■ consolidation minor principal stress

” effective confining stress at

steady-state condition

"a * average radial stress

o^g “ average circumferential stress

(O j -O j ) “ deviator stress

(Oj-o^Jp ■ peak deviator stress

(o^o^p « steady-state deviator stress

« shear stress

c t
t

s
* static shear stress

t * threshold shear stress

Tcy ■ cyclic shear stress

Tgg * steady-state strength

t - maximum shear strength of a soil


max
t . * horizontal shear stress
vh
g = shear stress component

t = shear stress component


zr
T^g = shear stress component

6 * angular rotation

6 = maximum amplitude of the twist angle


max r

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
xx i

ABSTRACT

Alarcon-Guzman, Adolfo, Ph.D., Purdue University, August


1986. Cyclic Stress-Strain and Liquefaction Characteris­
tics of Sands. Major Professors: G. A. Leonards and J.
L . Chameau

Liquefaction of saturated sand and silty sand depo­

sits has been recognized as a major cause of damage to

buildings and earth structures during earthquakes. How­

ever, in spite of many research studies during the past

two decades, there are still conflicting opinions on crit­

ical aspects of the phenomenon of soil liquefaction,

including the definition of the term "liquefaction." For

these reasons, a research program with an integrated

attack was initiated at Purdue University to include: a)

laboratory studies of the behavior of reconstituted cohe-

sionless soils under cyclic loading, b) a new technique

for obtaining superior undisturbed samples of sand below

the water table, and c) performing and interpreting a

variety of field measurements. This dissertation is con­

cerned with the first of these studies.

An important effort was devoted to delineate and

understand the pkysical factors controlling the response

of cohesionless soils to cyclic loading and consequently,

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
the factors involved in the different phenomena which have
c been described with the term "liquefaction" such as

steady-state flow, a condition of zero effective stress,

or cumulative residual strains due to cyclic loading. The

conditions leading to different modes of undrained failure

a re discussed in detail together with the direct relation­

ship that can be established between the monotonic and

cyclic stress-strain behaviors of sands.

Apparatus and techniques for cyclic shear testing are

discussed, with emphasis on the resonant column and tor­

sional shear tests. A new hybrid resonant

column/torsional shear apparatus was designed and built

(166) as a part of this research program. The new

c apparatus permits the determination of dynamic soil pro­

perties on a single solid or hollow cylinder specimen over

the entire range of shear strain amplitudes of engineering


—4
interest, i.e. from 10 % to 10%. Torsional shear tests

were performed with the new apparatus on reconstituted

specimens of Ottawa 20-30 sand. Within the design objec­

tives, the apparatus performed in a superior manner and

provided test results which illustrate the stress-strain

relationships under both drained and undrained conditions

for a wide range of shear strain amplitudes.

Particularly, it was found that: a) the maximum

shear modulus, Gm a x * relatively insensitive to impor­

tant factors affecting the undrained behavior of sands

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
such as stress ratio (and hence fabric), stress history,

(_ and cyclic prestraining. Consequently, a direct correla­

tion between an<* liquefaction potential is not possi­

ble; b) if staged testing includes the determination of

shear moduli by means of resonant column tests, which

imply the application of a considerable number of cycles,

it can lead to significant overestimations of shear


£
modulus for sands with no prior cyclic strain history even
_2
at shear strain amplitudes as low as 1,1 x 10 %; and c)

at a given relative density, the effective stress path

corresponding to an undrained monotonic loading tests

defines a "state boundary" or "collapse" surface marking

the initiation of strain softening behavior of loose sands

under cyclic loading at different shear stress amplitudes

smaller than the peak monotonic strength.

It is concluded that the concepts outlined in this

dissertation provide a framework for understanding the

stress-strain behavior of sands and offer clear guide­

lines for further research.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
1

c
CHAPTER 1

INTRODUCTION

Liquefaction of saturated sandy and silty soil depo­

sits has been recognized as a major cause of damage to

buildings and earth structures during earthquakes. Conse­

quently, assessment of liquefaction potential is of great

concern to engineers designing major structures in areas

of high seismic activity. However, in spite of many

research studies during the past two decades, there are

still conflicting opinions on critical aspects of the


c
phenomenon of soil liquefaction, including the definition

of the term "liquefaction."

On one side, the phenomenon of liquefaction is asso­

ciated with a condition of zero effective stress due to

progressive increases in pore water pressure resulting

from the tendency towards densification of the sand struc­

ture subject to cyclic loading (73,103,148). The condi­

tion of zero effective stress may be temporary during each

loading cycle but the associated shear strains may be so

large that the consequences can be catastrophic.

On the other hand, the term "liquefaction" has been

used to describe the strain softening behavior

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
2

characteristic of loose sands sheared under undrained con­

ditions (13,14). Ample experimental evidence has shown

that monotonic (or unidirectional) loading beyond the peak

point of the stress-strain curve (i.e., "failure") results

in a marked reduction in strength with increasing strain

until the stress stabilizes at an ultimate or residual

strength (9,14,55). The author attributes this behavior

to the fact that the structure of a loose sand is meta-

stable and hence small shear strains can produce rear­

rangement of grains and loss of contact points between

neighboring grains. Upon collapse of the structure, the

load is transferred from the soil skeleton to the water

and the pore pressure continues to increase in spite of

the reduction in applied shear stress. The shear strength


( of the sand is substantially reduced and the soil mass

undergoes large deformations in a very short period of

time. Consequently, a close analogy exists between the

"sensitive" behavior of saturated loose sands and that of

quick clays (9). After large strains, a residual fabric

is created which offers the minimum resistance to shearing

and the mass reaches a steady-state of deformation (131)

resembling the flow of heavy fluid (14). At this stage,

the shear strength of the sand is said to depend mainly on

the grain characteristics and the void ratio of the sand

(13,14.131,132).

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
3

Until recently, the two inodes of undrained deforma­

tion described above were studied independently without

attempting to establish a direct relationship between the

monotonic and cyclic undrained behaviors of sands. On the

contrary, some researchers even claimed that the two

phenomena were not related (13,15,17). In this connection

Casagrande (13) wrote that when a sand specimen is sub­

jected to cyclic loading the "peaks of pore pressures keep

increasing and finally rise momentarily to the confining

pressure, , every time the stresses cycle through the

hydrostatic state, or to use the term now generally used

in the literature for this phenomenon, the specimen

suffers 'liquefaction'. This response bears no relation­

( ship to actual liquefaction of sands in situ." The

independent consideration of the behavior of sands under

monotonic and cyclic loading and the indiscriminate use of

the term "liquefaction" have caused confusion regarding

the factors controlling the undrained behavior of

saturated sands. For example, experimental evidence was

presented showing either that the presence of initial

static shear stresses (anisotropic consolidation)

increases the cyclic undrained strength (104) or that the

monotonic undrained strength decreases significantly as

the consolidation shear stress increases (17).

The underlying cause of the problem is that the

mechanical behavior of granular materials is Influenced by

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
4

a large number of factors, which make It difficult to

establish general stress-strain relationships. There is

ample experimental evidence to show that relative density

and state of stress are not sufficient to characterize the

mechanical behavior of granular materials. Other factors,

such as fabric and its associated anisotropy, stress-

strain history, applied stress path (including rotation of

the principal stresses) and aging, among others, strongly

influences the in situ stress-deformation behavior of

granular soils (4,93,101,117,119,125). Consequently, all

these factors also affect the susceptibility of sands to

pore pressure development under undrained loading. More­

over, "undisturbed" sampling of cohesionless soils is

extremely difficult. With current sampling techniques it

is considered that effects such as stress-strain history

and aging may be largely removed (12,29,51). Therefore,

to date, most of the experimental work has been done on

reconstituted specimens. It is certainly recognized that

tests on reconstituted specimens may predict the in situ

performance poorly, because the effects of fabric, aging,

and stress-strain history cannot be accounted for

appropriately (93).

By far the most prevalent correlation extant today

between "liquefaction" potential and in situ measurements

has been with the standard penetration test (SPT), in

spite of the fact that the results from the test are

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
5

affected by a number of well-known yet uncontrolled vari­

ables: 1) the method of advancing and stabilizing the

hole, 2) hole diameter and depth, 3) type of drill-rod

and anvil used, 4) degree of interference with free fall

of the hammer, and, 5) operator's experience (145,153).

Accordingly, wide variations can be expected when blow

count data are collected from many different sources.

However, the main concern is with the fact that SPT is

inherently insensitive to important factors affecting

"liquefaction" potential including: sand fabric, pres­

tress, preshaking, and geologic aging. This is because

penetration resistance measures the shear resistance at

large strains, at which point the influence of fabric,

prestress, preshaking and aging have been largely des­

troyed (35,101). A field test capable of sensing the

effects of this factor must measure stress-deformation

relationships prior to failure.

For the reasons discussed above, a research program

with an integrated attack on the problem of predicting

reliably the "liquefaction" potential of saturated sand

deposits was initiated at Purdue University to include:

1. Laboratory studies of the behavior of cohesionless

soils under cyclic loading. It was considered neces­

sary to perform the most up-to-date cyclic shear

tests to assess the relationship between monotonic

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
and cyclic undrained behaviors and Che effects of

controlled differences in fabric, prestress and

cyclic prestraining on the stress-defornation charac­

teristics of sands under cyclic loading.

2. Development of a new technique for obtaining superior

undisturbed samples of sand below the water table

using an impregnation technique, and

3. Performing and interpreting field measurements that

make direct measurements of stress— deformation

characteristics, and hence can sense the effects of

all factors known to affect "liquefaction" potential

of saturated sands.

(
This dissertation is concerned with the first of

these studies.

An important effort was devoted to delineate and

understand the physical factors controlling the response

of cohesionless soil to cyclic loading and, consequently,

the factors involved in the different phenomena which have

been described with the term "liquefaction" such as

steady-state flow, a condition of zero effective stress,

or cumulative shear strains due to cyclic loading. This

exercise yielded an integrated framework of understanding

of the undrained behavior of sands under both monotonic

and cyclic loading, which is presented in Chapter 3. The


r
\

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
7

main features which characterize undrained behavior of


C sands under both cyclic and monotonic loadings are con­

sidered first independently, based on previous studies and

largely on the author's interpretation and understanding

of the factors controlling the shear strength of soils.

Particular attention is given to the fact that there

exists a region in the shear stress-normal stress space

beyond which the behavior of sands under cyclic loading

changes from contractive to dilative on loading. The lim­

its of this region seem to correspond to a particular

value of the principal stress ratio, which defines the so

called "phase transformation" lines (82) because reversal

in the direction of loading from stress states beyond such

( lines is accompanied by a significant increase in the pore

pressure and eventually by the development of a condition

of zero effective stress at the instant the stress cycles

pass through the hydrostatic state. The location of the

phase transformation lines in stress space depends on the

relative density and the type of sand. Once a state of

zero effective stress develops, a certain amount of shear

strain is required to mobilize a shear resistance in the

opposite direction, equal to a given applied shear stress,

which is a function of the relative density of the sand

(148,155). For very loose sands almost unlimited deforma­

tions will occur without mobilizing appreciable shear

resistance, whereas in the case of dense sands dilation

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
8

soon develops, Che p o r e pressure decreases and Che s a n d Is

able to withstand the applied shear stresses without

appreciable deformation.

On the other hand, it is shown that the undrained

behavior of sands under monotonic loading is characterized

by the existence of another region in the shear stress-

normal stress space beyond which a sand exhibits a marked

strain softening behavior, that is, it develops large pore

pressures which in turn cause a drastic reduction in

strength and the consequent development of large shear

strains. As mentioned earlier, this behavior is attri­

buted by the author to the "collapsiveness" of the sand

structure (2). The initiation of strain softening

V. behavior is observed to occur at a particular value of the

principal stress ratio or mobilized angle of shear resis-

tance, $ , , which is shown to depend at least on the type


moo
of sand, the relative density and the level of confining

stress. The locus of stress states triggering the strain

softening behavior defines a "critical stress ratio" (CSR)

line (186), or "collapse" surface (165). For certain ini­

tial states, the behavior of a sand may be limited strain

softening whereas as the relative density increases or the

confining stress decreases the behavior changes to

strongly dilative (strain hardening).

Under cyclic loading a sand specimen is subjected to

continuous reversals in the direction of shearing, that

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
9

is, experiences stages of loading, unloading and reload­

ing. Under these conditions, it is shown that depending

on the initial state of stress and the cyclic shear stress

amplitude, the resulting effective stress path may reach

the critical stress ratio (CSR) line during any loading

stage or move beyond the phase transformation lines, in

which case large pore pressures develop in the subsequent

unloading stages. In other words, under cyclic loading a

sand specimen may exhibit strain softening behavior or

reach a condition of zero effective stress. (Under cer­

tain conditions a combination of the two phenomena may

also develop.) Therefore, in the author's view, a basic

relationship must exist between the monotonic and cyclic

stress-strain behavior of sands. Accordingly, all the


c conditions leading to different modes of undrained failure

under cyclic loading are discussed in detail by using a

combination of state diagram and effective stress path

plots showing the location of the phase transformation and

critical stress ratio (CSR) lines. Results from previous

studies on the subject of "liquefaction" of sands are

sometimes presented to serve as an illustration. It is

also shown that the effects of the type of sand, that is

the mineral composition, size distribution, shape, and

roughness of the particles, on the susceptibility to

strain softening behavior (or collapse) under monotonic

loading and on the susceptibility to pore pressure buildup

due to sudden reversals in the direction of shearing, are

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
10

opposite in nature. For example, at the same relative

density, a sand with rounded grains is more susceptible to

flow failure than one with angular grains, whereas a sand

with angular grains may develop a condition of zero effec­

tive stress in a smaller number of cycles at a given shear

stress amplitude than a sand with rounded grains.

The discussion in Chapter 3 also shows that the way

in which the effective stress path approaches the strength

envelope determines the failure modes and the resulting

deformations. Starting from an initial anisotropic state

of stress, which is typical of most in situ conditions,

the failure line can be approached following different

stress paths which depend not only on the nature and mag­

nitude of the applied loads but also on the strain boun­

dary conditions. These stress paths may reflect changes

in the initial state of stress, including rotation of the

planes of principal stress (reversal of shear strain) as a

result of shaking during an earthquake. For example, it

is shown that under plane strain conditions, such as in

sand deposits under level ground, the effective stress

path moves down and towards the left, reflecting continu­

ous reductions in the static shear stress. The effective

stress path eventually reaches the origin of the stress

space (zero effective stress), irrespective of the initial

state of stress. It is also shown that in this case, the

effective stress path cannot reach the critical stress

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
11

t ratio line and therefore, strain softening behavior is not


V ...
a possibility under plane strain conditions. On the other

hand, it is shown that if lateral strains are not con­

strained, such as in soil elements close to sloping

ground, static shear stresses remain throughout the cyclic

loading due to the need to maintain equilibrium imposed by

geometry. Accordingly, in this case, the effective stress

path moves essentially horizontally towards the left

approaching eventually the critical stress ratio (CSR)

line, provided the total shear stress (static + cyclic) is

larger than the steady-state or residual strength of the

sand. Hence, strain softening behavior is possible under

these conditions. It is noted that as the static shear

^ stress increases (anisotropic consolidation), the state of

stress gets closer to the critical stress ratio (CSR) line

and hence, the susceptibility to strain softening behavior

under cyclic loading increases. On the other hand, as the

static shear stress increases the possibility of a com­

plete stress reversal (hydrostatic state of stress)

decreases. A hydrostatic state of stress is a prere­

quisite for the development of a zero effective stress

condition. Consequently, the effects of initial static

shear stresses on the susceptibility to strain softening

behavior under monotonic loading and on the susceptibility

to the development of a zero effective stress condition

due to sudden reversals in the direction of shearing, are

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
12

opposite. This explains the conflicting conclusions dis­


L, cussed earlier.

In summary, it is shown in Chapter 3 that the problem

of undrained stability of a contractive saturated sand

during earthquake loadings can be studied by considering

the way in which the in situ effective stress path

approaches the failure envelope. It is this path that

determines the failure modes and the resulting deforma­

tions. An agreement on the definition of the term

"liquefaction" is not really needed. It is considered

that a combination of failure mechanisms, corresponding to

either strain softening behavior or the development of

zero effective stress conditions or both may contribute to

the failure of a complex structure, such as an earth dam.

From the practical point of view, it is immaterial which

mechanism is called "liquefaction" since both may cause

intolerable deformations leading to the loss of the struc­

ture. However, the author will use the term "flow

failure" as indicative of strain softening behavior and

will reserve the term "liquefaction" for describing zero

effective stress or "quick" conditions, in accord with the

etymology of these words. It should be noted that deter­

mination of the way in which the state of stress actually

changes within an earth structure in the course of earth­

quake loading is a goal that still remains to be achieved.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
13

G Apparatus and techniques for cyclic shear tests were

examined in detail and the advantages and limitations of

each technique are summarized in Chapter 2. Emphasis is

placed on the factors discussed below. A variety of

laboratory testing techniques have been developed over the

years to evaluate the stress-strain behavior of soils

under cyclic loading but the capabilities of each device

are in general limited to a specific strain range. For

example> resonant column tests apparatus have been build

to determine dynamic soil properties at low strains rang-


-4 -2
ing from 10 % to 10 %. These strain levels may be typi­

cal of geophysical measurements or of vibrations due to

traffic loadings. On the other hand, cyclic triaxial,

( simple shear or torsional shear equipment is used to simu­

late the larger strain levels associated with earthquake

loading conditions.

In practice, the use of two separate pieces of test­

ing equipment implies that the dynamic properties of the

tested material, over the range of shear strains of

interest, are determined on different specimens, which are

unlikely to have identical characteristics. Moreover,

very often there is no overlap between the ranges of shear

strains for which the results are available from each

test. This is a significant limitation when investigating

either the reduction of shear modulus as a function of

shear strain amplitude or the existence of a critical

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
14

shear strain at which the fabric of the sand eventually

yields during a given loading sequence (1).

Although the cyclic triaxial test has been used

extensively during the past two decades to evaluate cyclic

undrained strength of saturated sands, its limitations to

simulate earthquake loading conditions have long been

recognized (13,14,145,192). The stresses induced in the

ground by an earthquake are considered to be due primarily

to the upward propagation of shear waves from the underly­

ing rock formation. Consequently, during an earthquake,

elements in the soil deposit are subjected to a series of

horizontal shear stress applications. Accordingly, simple

shear and torsional shear tests, in which horizontal shear

stresses can be applied on the upper and lower face of a

specimen are now recognized as having superior capabili­

ties for determining soil behavior under earthquake load­

ing, including "liquefaction" phenomena

(30,69,75,140,173.192). The previous considerations led

to the design and development of a new hybrid torsional

shear apparatus which combines the features of both the

resonant column and torsional shear devices. The main part

of the apparatus was built by Soil Dynamics Instruments

(166) under the supervision of Professor V. P. Drnevich,

based on ideas and specifications provided by the author's

advisors, Professors G. A. Leonards and J.L. Chameau. The

new apparatus permits the determination of dynamic soil

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
15

properties on a single solid or hollow cylinder specimen

over the entire range of shear strain amplitudes of


-4
engineering interest, i.e. from 10 % to 10%, as well as a

better simulation of the stress paths associated with

earthquake loadings. A detailed description of the new

apparatus is presented in Chapter 4.

Torsional shear tests were performed with the new

resonant column/torsional shear apparatus on reconstituted

specimens of Ottawa 20-30 sand. Specimen preparation pro­

cedures and the general purposes and scope of the testing

program are described in Chapter 4. Preliminary tests

were performed under both drained and undrained conditions

to evaluate the whole range of capabilities of the new

apparatus and to compare these results with those avail­

able in the literature, which had been obtained using

other devices and similar (or different) types of sand.

After these tests, a comprehensive testing program was

initiated in accord with the main goals of the investiga­

tion as described earlier. In addition to obtaining addi­

tional support for the author's views on the relationship

between monotonic and cyclic undrained behavior of sands,

it was considered that some aspects of the drained and

undrained behaviors of sands were still subject to con­

flicting opinions, thereby requiring further investiga­

tion, as discussed below.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
16

1. Some researchers consider that a relationship

between shear wave velocities and liquefaction potential

of sands can be developed (29,32,33,89,147)• It is con­

sidered that shear wave velocities can be readily deter­

mined by means of geophysical tests in situ and that all

the factors known to affect the maximum shear modulus of

dry sands, which relates to the shear wave velocity, will

also affect the stress-strain behavior and liquefaction

potential of the materials when subjected to undrained

loading in a saturated condition. It is also claimed that

elastic wave velocities respond to fabric differences

(29,36,38,39) and therefore should reflect the effect of

factors such as stress-strain history and aging on the

r strength-deformation characteristics of sands. Conse—

quently, it was thought appropriate to determine to which

extent shear modulus values determined by means of

resonant column tests at strain levels corresponding to

those produced in geophysical tests for measuring in situ

shear velocities, can reflect the effect of factors such

as stress ratio (anisotropic consolidation), prestressing

and cyclic prestraining which in addition to relative den­

sity and confining stress are known to influence the sus­

ceptibility of sands to pore pressure development under

undrained loading.

2. Experimental evidence has been presented showing

the existence of a threshold shear strain for sands on the


r

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
17

order of 10~^% (24,32,33,43), below which there is in

principle neither densification of dry sands nor pore

pressure buildup in saturated sands. It is claimed that

the threshold shear strain is a constant characteristic of

each type of sand for a wide range of particle arrange­

ments, relative densities and confining pressures. These

findings were considered to conflict with the results of a

significant amount of research, which has shown that the

mechanical behavior of granular materials is affected sig­

nificantly by factors such as fabric and its associated

anisotropy, stress-strain history, stress path, etc.,

among others. It was thought that the initiation of pore

pressure buildup could be associated with the shear strain

marking the initiation of degradation in the shear

modulus, that is, the shear strain at which the behavior

of sands changes from linear elastic to nonlinear and Ine­

lastic. Consequently, particular attention was given to

the degradation of shear modulus from the G value as a


0 max
function of strain level under both drained and undrained

conditions.

3. Most investigations on the undrained steady-state

strength of sands have been based on axial compression

triaxial tests, and hence they have been limited to a par­

ticular stress path. However, there is ample experimental

evidence to show that the fabric developed under condi­

tions prevalent during deposition of most natural deposits

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
18

is highly anisotropic, which causes changes in the

undrained deformation-strength characteristics with the

direction of loading (55,118,119,120,169), although others

claim that the initial fabric has no effect on the large

strain deformation characteristics of sand under undrained

conditions (13,14,18,131,132). A few recent studies,

which also have attempted to establish a relationship

between the monotonic and cyclic undrained behavior of

sands (34,121,165,184,185,186) still contain conflicting

opinions on factors such as the effect of consolidation

shear stresses (anisotropic consolidation) on the subse­

quent behavior under undrained shearing, and on the state

conditions marking the initiation of strain softening

behavior and hence the relationship between monotonic and

cyclic loading of sands. Some of these aspects were also

investigated in the new torsional shear apparatus, includ­

ing the effects of prior strain history, and hence of

fabric, on the subsequent undrained behavior. It should

be noted that the orientation of the principal stresses in

torsional shear tests specimens differs from that in

triaxial test specimen. Therefore, these effects were

investigated along different stress paths.

The results of the tests performed are summarized and

discussed in Chapter 5. These results illustrate the

behavior of Ottawa 20-30 sand under both drained and

undrained conditions for a wide range of shear strain

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
19

amplitudes. A summary of the main conclusions and find­

ings of this investigation is presented in Chapter 6.

Finally, suggestions for further research are made in

Chapter 7.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
20

CHAPTER 2

APPARATUS AND TECHNIQUES FOR CYCLIC SHEAR TESTING

Analytical methods for predicting ground motions due

to dynamic loading require an appropriate knowledge of the

stress-strain and energy dissipating properties of soil

deposits. The problems of interest cover a wide range of

situations. At one end, very small amplitudes of motion

are expected when considering vibrations due to traffic

loads. On the other hand, blast and seismic loadings often

produce shear strains in the soil deposit which are high

enough to bring about shear failures. In the latter

category, the phenomenon of soil liquefaction and its

catastrophic consequences are of great concern to

engineers evaluating the stability of natural or man-made

deposits of saturated sands in areas of high seismic

activity.

The need to evaluate stress-strain characteristics

over a wide range of shear strain amplitude has stimulated

the development of a variety of laboratory and field

measuring techniques (136,140,192). Ideally, the purpose

of any laboratory testing technique is to study the

behavior of an element of soil under conditions similar to

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
21

those encountered in the field. In a laboratory test, the


c
specimen is assumed to represent a single point in a soil

deposit. The validity of this assumption depends on the

uniformity of the distributions of stress and strain

within the specimens, which in turn depends on the confi­

guration of the specimen and the control and measurement

of stress and strain on the boundaries (139,140). Dis­

cussing these effects Woods (192) wrote: "If these factors

are not considered, there is little hope of obtaining

agreement between field results where nature applies the

boundary conditions and laboratory results where engineers

must try to reproduce or otherwise to account for boundary

c ond iti ons ."

Although considerable improvement has been achieved

in testing equipment, in the interpretation of boundary

and membrane effects, and in the understanding of the

capabilities and limitations of each technique, the com­

parison between predictions based on laboratory results

and field performance might still be disappointing (192).

This can be attributed firstly to the large number of fac­

tors that control the mechanical behavior of granular

materials, which make it difficult to establish general

stress-strain relationships. It is widely accepted now

that relative density and .state of stress are not suffi­

cient to characterize the mechanical behavior of granular

materials. Other factors, such as fabric and its

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
22

associated anisotropy, stress-strain history, aging, and

the stress path, among others, must be accounted for

(4,93,101,117,119,125). The problem is aggravated by the

fact that "undisturbed" sampling of cohesionless soils is

extremely difficult. With current sampling techniques it

is considered that effects such as stress-strain history

and aging may be largely removed (12,29,51). Conse­

quently, to date, most of the experimental work has been

done on reconstituted specimens. It is certainly recog­

nized that tests on reconstituted specimens may predict

the performance of the soil in situ poorly, because the

effects of fabric, aging and stress-strain history cannot

be considered appropriately (93).

c A number of testing devices have been developed to

deal with the wide range of strain levels that a soil may

be subjected to under different conditions ranging from

geophysical measurements to earthquake loading. High fre­

quency apparatus, such as the resonant column test, have

been used to determine dynamic soil properties at low


/ *)
strains, ranging from 10 % to 10 %. On the other end of
-2
the spectrum, for strains ranging from 10 % to 1%, low

frequency devices, such as cyclic triaxial, simple shear,

and torsional shear equipment, have been used to simulate

earthquake loading conditions. Very few devices are capa­

ble of providing dynamic soil properties for strain span­

ning over these two strain ranges. A significant

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
23

improvement was accomplished by Kokusho (89), who developed

a cyclic triaxial apparatus for measuring modulus and damp­

ing values of soils over a range of shear strains on the

order of 10"4% to 10- 1 %.

A summary of the potential shear strain ranges over

which each device can be used and the relative capabili­

ties of each technique is presented in Figure 2.1. These

techniques are very often classified into two groups:

namely, dynamic properties tests and cyclic strength tests

(177). Dynamic properties tests are supposed to provide

soil properties (modulus and damping values) from which

the response of an embankment or a natural soil deposit to

a given dynamic loading time history can be calculated.

This analysis yields the time history of shear stresses

induced within the soil mass. The cyclic strength tests

are used to evaluate the ability of the soil to resist the

induced shear stresses. In the author's opinion and in

the light of Figure 2.1, the difference between dynamic

properties tests and cyclic strength tests is rather

artificial, since the variation of the so called dynamic

properties with strain amplitude simply reflects the gra­

dual mobilization of the shearing resistance of the soil.

Consequently, uncoupling between the induced shear

stresses and the shear strength of the soil in the deposit

under consideration is not desirable.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
S h e a r i n g strain amplitude (%)

1 0 -4 1 0 -3 1 0 ~2 10 r1

Resonant Column (Solid samples)


Resonant Column (Hollow samples)
Torsional Sheaf (Hollow samples)
Pulse methods Cyclic Triaxial
Cyclic simple shear
Shake Table

Typical Motion Characteristics


"wny
machine foundation
5SBW7S?
earthquake
sssi”
explosion
10~4 to-3 10—2 ■jo-1

a) S h e a r Strain R a n g e s

Shear Young's Material Cyclic Stress Attenuation


Modulus Modulus Damping Behavior
Resonant
Column X X X

with
adaptation X

Ultrasonic
Pulse X X X

Cyclic
Triaxial X X X

Cyclic
Simple X X X
Shear

Cyclic
Torsional X X X
Shear
Shake
Table X X

b) Property Capabilities
Figure 2.1 Laboratory Techniques for Measuring Dynami
Soil Properties (After Ref. 192).

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
25

The stresses induced in the ground by an earthquake

are considered to be due primarily to the upward propaga­

tion of shear waves from the underling rock formation,

although other forms of wave motion are also likely to

occur. Consequently, during an earthquake, elements in a

soil deposit are subjected to a series of cyclic shear

stress applications, which in general involves changes not

only in magnitude but also in direction. Accordingly,

simple shear and torsional shear tests, in which horizon­

tal shear stresses can be applied on the upper or lower

face of a specimen are now recognized as having superior

capabilities for determining soil behavior under earth­

quake loading, including liquefaction phenomena

(30,69,75,140,173,192).
c
This chapter examines some of these laboratory test­

ing techniques with emphases on the resonant column and

torsional shear tests. A new testing device, which is

considered to improve techniques for measuring the

response of granular soils (and of other soils) to a wide

range of cyclic loading conditions, was designed and built

(166) as a part of the research program already outlined

to assess the liquefaction potential of saturated sands

and silts. It combines the resonant column and torsional

shear tests into one apparatus, which permits the determi­

nati on of dynamic soil properties over the whole range of

strains of engineering interest, along with a better

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
26

simulation of the stress paths associated with earthquake


L,
loadings. The description of this new equipment is

presented in Chapter 4.

2.1 The Resonant Column Test

The principle of this test is to determine the

dynamic properties of sands (and of other materials) from

observation of the steady-state response of a cylindrical

specimen subjected to forced vibrations. In general, the

frequency of the applied vibration is adjusted until reso­

nance develops. Consequently, the frequency and dynamic

amplification factor at resonance can be correlated with

the stiffness and energy dissipating characteristics of

c the soil specimen.

It is clear that in order to excite and measure the

vibration of a system (soil specimen), it must be altered

by the attachment of the vibration excitation device and

the required transducers (56,62). Hence, the observed

vibration is not the response of the soil specimen alone

but of the system composed of the specimen and its

attached apparatus including end platens, O-rings, etc.

Consequently, the frequency response cannot be interpreted

solely in terms of specimen properties but the contribu­

tion of all other elements in the system should be

accounted for (see Appendix A).

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
27

Either compression or torsional vibrations can be


V..
applied to the soil spec-imen. Accordingly, either Young's

modulus or shear modulus (and any corresponding damping)

can be determined from the observed response. However,

torsional vibration where shear waves are generated is by

far the most common (39). For small amplitude vibrations

of generally high frequency, the response of sand speci­

mens is essentially elastic. Hence, the theory of propa­

gation of waves (shear or compression) in elastic bars is

used to interpret the resonant column test results

(28,133,156). However, as the amplitude of vibration

increases, the stress— strain characteristics of sands

become nonlinear and inelastic. Accordingly, the theory

of elastic wave propagation alone is not sufficient to


c
model the stress-strain response; in this case, the energy

dissipating properties of the soil should be accounted for

(39,40,56,62).

A variety of resonant column test apparatus have been

developed using different boundary conditions to constrain

the specimen (39,56). Most of these apparatus can be

represented closely by the models shown in Figure 2.2.

Each configuration requires a slightly different driving

equipment and methods of data interpretation. The most

widely used is the fixed-free apparatus, in which the

vibration is applied, and the response is measured at the

top of a specimen (active-end platen), that is rigidly

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
Reproduced with permission of the copyright owner. Further reproduction r

MASS RIGIDLY CONNECTED TO PLATEN

PASSIVE END PLATEN

SOIL SPECIMEN
c)
ACTIVE END PLATEN
^-TORSIONAL SPRING

TORSIONAL DASHPOT
prohibited without perm ission.

PORTION OF VIBRATION EXCITATION DEVICE


RIGIDLY CONNECTED TO PLATEN
— LONGITUDINAL DASHPOT

LONGITUDINAL SPRING

Figure 2.2 Common Re s o na n t Column Boundary C o n d i t i o n s . a) F i x e d - F r e e ,


b) Fi xe d S p r i n g Top, and c ) F r e e - F r e e ( A f t e r Ref. 39 ).

S3
00
29

fixed at its base (passive— end platen). If a sinusoidal

driving torque is applied at the top of the specimen, the

distribution of angular rotation 6, along the fixed base

specimen is a 1/4 sine wave but adding a polar mass moment

of inertia at the top (Fig. 2.2a and 2.2b) the distri­

bution of 6 along the specimen becomes nearly linear with

a node at the fixed end of the specimen (192). Conse­

quently, most fixed-free models take advantage of end-mass

effects to obtain uniform strain distribution throughout

the length of the specimen. The top mass may also have

springs and dashpots attached to it (Fig. 2.2b). The pur­

pose of the springs and dashpots is to account for

apparatus damping in all types of apparatus and to account

for springs which may be part of the excitation device.

These springs allow for application of anisotropic ambient

stresses or for supporting the top platen. However,

springs limit the strain amplitude capabilities of an

apparatus (39). If a single spring is used to restrain

the top of the specimen, the excitation device and the

active-end platen (without specimen in place) form a one-

degree-of-freedom system. This feature makes it con­

venient for data reduction and apparatus calibration pur­

poses. A new device of this type is described in detail

in Chapter 4 and in Appendix A.

In the spring-base model (Fig. 2.2c) the motion of

the system is excited at the base of the specimen and the

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
30

upper end is free except for the relatively rigid end pla­

ten system. When the base spring is stiff compared to the

specimen, the mode of vibration at the lowest resonance is

wi th a node near the base of the specimen (56). On the

other hand, if the stiffness of the base spring is much

smaller than the stiffness of the specimen, and for rigid

masses at the ends which are approximately equal, the

lowest mode of vibration of the specimen at resonance (as

measured by a pickup at the top) is with a node at the

center and maximum amplitude at the ends (56,192). Sys­

tems where neither end of the specimen is fixed are also

called "free-free" type of apparatus, which are advanta­

geous for testing large size specimens which are unusually

stiff such as shales, rock, concrete, asphalt, cemented

soils, etc. (39,64). The rotation distribution for tor­

sional vibrations will range from a 1/2 sine wave to

nearly linear depending on the size of the end masses

(192).

Although the response of the spring base model is

markedly changed by changing the base properties (bottom

mass and spring constant), the resonance ratio, defined by

the displacement amplitude at the top of the specimen

divided by the displacement amplitude at the base of the

specimen, is independent of these properties (56). Conse­

quently, when testing a specimen the resonant frequency

and amplification factor can be found directly from the

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
31

frequency response curve by recording the amplitude ratio

of displacement (longitudinal or angular) at both ends of

the specimen, plotted as a function of excitation fre­

quency. The driving force does not have to be known,

because it is divided out in the resonance ratio. There

is yet another advantageous feature of spring— base models.

The acceleration (or displacement) output at the base or

at the top of the specimen can serve as feedback for the

control of the vibrator to get a constant pre-selectable

input amplitude at any end of the specimen (64). In this

way, for measurements at low strain amplitudes, feedback

control can be used to ensure that the strain developed in

the specimen is limited.

c For measurements of shear modulus, G, torsional

vibrations are applied at the active end of the specimen.

Consequently, the direction of wave propagation is verti­

cal and the direction of particle motion or dynamic shear­

ing is tangential (198). Accordingly, the stress condi­

tion in the torsional resonant column test is axially sym­

metric and the associated stress and strain distributions

are analogous to the ones in cylindrical specimens sub­

jected to static torsional loading. These corresponding

stress and strain distributions are reviewed in detail in

the section on torsional shear tests together with the

advantages and disadvantages of testing hollow cylinder

s pe ci me ns .

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
32

f In principle, torsional resonant column specimens


^ J r ' ' '
could be subjected to end restraint effects (also dis­

cussed in section 2,4) since the end platens cannot be

lubricated in order to be able to transmit the desired

torque to the specimen. However, a study by Yu and

Richart (198) to evaluate the effect of end restraint on

the maximum shear modulus of sands, Gmax, showed that

these effects are negligible at small shear strain ampli-


—4
tudes ( y < 6x10 %). In their investigation, a specially

made top cap and base pedestal consisting of radial teeth

and greased membranes were used. The end restraint was

reduced because the total area of the teeth was only 36%

of the area of the end platen. However, no difference

c between the shear modulus values with the reduced end res­

traint and the usual end restraint was observed. In Yu

and Richart^s opinion, results from resonant column tests

under both triaxial compression and extension conditions

also confirmed that the effect of end restraint on maximum

shear modulus of sands is not very large. It was expected

that the effects of end restraint using standard platens

on the tangential stress distribution under the triaxial

compression and extension conditions were opposite. How­

ever, the difference between resonant column tests for two

different sands under both conditions was not significant.

The effects of end restraint may become significant as the

shear strain amplitude increases in a similar fashion as

/' discussed later on for torsional shear test specimens.


i

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
33

Opposite to end restraint effects, there is the prob­


L
lem of coupling between the specimen and end platen for

transmitting the torque. The effect of slipping is to

give lower shear moduli and higher damping than actually

exist at a given shear strain amplitude (37). End platens

with rough surfaces are more appropriate than teethed pla­

tens to ensure a uniform distribution of stresses in the

circumferential direction. However, the torque transmis­

sion capacity of end platens without teeth is always lim­

ited and may affect torsional resonant column test

results. Drnevich (37,39) developed a criterion based on

the mobilization of a minimum coefficient of friction of

0.2. Therefore, the maximum shear strain amplitude, Y,

that should be applied to a specimen with shear modulus,


(
G, is limited according with the expression

YG/o < 0.2 (2.1)


a

where a is the effective axial stress applied by the pla-


di
ten to the specimen. Drnevich (39) considers that the use

of porous disk of stone, metal or other material provide

sufficient roughness for most soils. However, these

porous stones should be rigidly fastened to the platens.

Resonant column tests can give very accurate shear

modulus results for shear strain amplitudes as low as


-4
10 %. The upper limit of shear strain amplitude can be

as high as 10~*% or even 1%, as far as proper

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
34

consideration is given to the nonlinear and inelastic

stress-strain characteristics of the soil at large shear

strain amplitudes (37). However, most typical results

have been reported for shear strain amplitudes ranging

from 1(T*% to 10~^% (22,88). In spite of different confi­

gurations, all devices used for resonant column tests pro­

duce relatively consistent results, when the hysteresis

loop model with equivalent shear modulus and damping ratio

is assumed for the stress-strain relationship of the

specimens (164,192). The response of sand specimens to

forced vibrations has been found to be similar to the

theoretical response of an equivalent uniform linearly

viscoelastic (Kelvin-Voigt model) specimen of the same

c mass "density and dimensions of the soil specimen, even

though the damping of sands is not of a viscous nature

(56,62). This comparison is illustrated in detail in

Appendix A. Nevertheless, it should be noted that the

Kelvin-Voigt model is inadequate to describe more compli­

cated stress-strain behavior such as under undrained load­

ing where simplification of the behavior in terms of an -

equivalent secant shear modulus and hysteretic damping

ratio is not appropriate. Results from resonant column

tests have been found to be also consistent with results

from other cyclic tests, especially cyclic torsional shear

tests (8 8 ).

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
35

In the past, one of the main advantages of the

resonant column test was considered to be the low shear

strain levels that can be applied to a specimen

corresponding to those produced in geophysical tests in

situ (136,177). Conceptually, it has been considered that

elastic wave velocities should respond to minor changes in

soil structure or fabric (29). This feature supposedly

makes the resonant column test especially powerful for

investigating the effects of specimen disturbance, remold­

ing, etc., in cases where "undisturbed" sampling is

attempted (38,177). When working with reconstituted speci­

mens, the measurement of elastic wave velocities by means

of resonant column tests has been recommended to investi­

gate the effects of factors such as anisotropic states of


c. stress, stress-strain history and aging on the stress-

deformation and damping characteristics of sands

(36,38,39). These effects are reflected in the sand

structure since they affect the predominant grain orienta­

tion or the conditions at the grain contacts. It has also

been considered that the measurement of maximum shear

modulus, 6 , by resonant column tests also allows one to


* max J
judge if laboratory specimens have been replicated, or to

examine to what extent they have been reconstituted to an

"equivalent fabric" which exhibits a shear modulus in

agreement with the values determined from in situ seismic

techniques (29). Nevertheless, tests results are

presented in Chapter 5, which show that G is not very


r max

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
36

r sensitive to drastic changes in fabric resulting from high

consolidation stress ratios or prior strain histories

involving large strains. Moreover, in the author's opin­

ion some judgement should be exercised when evaluating

stress-history effects by means of resonant column tests

since the measurement of G in the resonant column test

implies the application of a considerable number of

cycles. It is not clear yet at which shear strain level

the stress-strain characteristics of sands change as a

function of the number of cycles. Consequently, at rela­

tively large shear strain levels, resonant column test may

override any prior stress-strain history effects.

As the resonant column test equipment can be designed

for independent measurement of the shear and Young's

moduli, values of Poisson's ratio can be calculated, or

inherent anisotropy evaluated to some extent, from the

test data. As summarized by Wood (192), modifications to

the resonant column test would include random vibrations

in addition to sinusoidal vibrations and the measurement

of attenuation as a function of frequency. This may be

very important to study further the effect of frequency on

shear modulus and damping of sands and of other soils.

The latest developments in resonant column testing have

been recently reviewed by Drnevich (39), together with a

comprehensive discussion on the effects of "fixity" condi­

tions in fixed-free apparatus on the measured modulus and

C damping values.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
37

^ 2.2 The Triaxial Test

The cyclic triaxial test has been used since the mid­

dle 1960's (103>148) and in spite of its limitations as

discussed later on, it is still the most widely used

laboratory method of determining dynamic properties and of

evaluating liquefaction characteristics of cohesionless

soils. In common practice, cylindrical specimens consoli­

dated either isotropically or anisotropically are sub­

jected to a cyclic variation in the axial stress, which is

intended to simulate the cyclic shear stresses experienced

by an element of soil in the field during an earthquake.

The validity of this simulation is discussed later in this

chapter and also in Chapter 3.

c The resulting stress conditions on soil specimens in

the triaxial test are illustrated in Figure 2.3. It is

evident that the applied boundary stresses, that is, the

radial and- axial stresses, correspond to principal

stresses. Under isotropic consolidation, the major prin­

cipal stress, o^, acts alternately in the vertical and

horizontal directions, depending on whether the compres­

sion or the extension portion of the loading cycle is

applied. It is to be noted that to prevent the top platen

from being lifted from the specimen the maximum reduction

in the axial stress cannot exceed the effective consolida­

tion pressure. As the compressive portion of the loading

cycle, Ao is applied, the maximum shear stress, + t ,


cy

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
38

c.

ct^ A ct <t 'c-A<t


i
o-U ©
T
T,q Total stress path
A ct

- A ct

a) lsotropically consolidated
rV 1C

3c

T,q T Total stress path


+A<r
<D
max

/ consolidation
(0*a)min state

-AC ACT

b) Anisotropically consolidated
Figure 2.3 Stress Components in Cyclic Triaxial Tests
Spec ime ns.

R eproduced with
perm ission o „ P e copyrigm owner. Pu rth er reproduction p r o v e d w ith o u t perm ission.
39

which develops on a 45° plane in the specimen, is equal to


L
Ac^/2, On the other hand, during application of the

extension portion of the loading cycle, Ao^ the shear

stress on any plane reverses and the maximum shear stress

then equals AoE /2. Since Aoc and Ao^, are usually equal in

magnitude (Aa in Fig, 2,3), a symmetrically reversing

shear stress condition is' obtained.

Under anisotropic consolidation, the orientation of

the major principal stress depends on the relative magni­

tude of the cyclic deviator stress, Ao, with respect to

the consolidation deviator stress, a, ( o , = o, -c_ ). If


a c ac 1c ic
Aa is smaller than the major principal sisress acts in

the vertical direction throughout the loading sequence.

C However, in cases where A a is larger than o<jc » the net

axial stress acting on the specimen will become less than

the radial or lateral stress during a portion of the load­

ing cycle. Accordingly, in this case the directions of

the principal stresses are temporarily reversed. It is

clear that under anisotropic consolidation, the variation

in the shear stress on any plane is non-symmetric

throughout the cyclic loading.

From either elasticity or plasticity theories, the

circumferential stress, which corresponds to the inter­

mediate principal stress, , is found to be equal to the

confining pressure, provided there are no end restraining

effects (198). Accordingly, during the compression

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
40

portion of the loading cycle the intermediate principal

stress is equal to the minor principal stress, • On the

other hand, when the extension portion of each cycle is

applied the intermediate principal stress becomes equal to

. This is an unfavorable feature of the cyclic triaxial

test, which produces undesirable effects especially under

undrained loading. It is well established that the pore

pressures generated during axial compression and axial

extension under the same deviator stress are different,

even in the case of a soil with no inherent anisotropy.

This is attributed to the influence of the intermediate

principal stress, since the octahedral shear stresses are

different during the corresponding compression and exten­

C sion portions of each cycle, which results in different

pore water pressures.

In common with other testing apparatus, end restraint

effects develop in cyclic triaxial test specimens

(107,140). These end effects are associated with radial

forces due to frictional restraint and stiffness of the

rigid platens used at the top and the bottom of the speci­

men. The radial frictional forces are always present if

the specimen tends to contract or expand due to volume

changes or when there is a change in length at constant

volume during the triaxial test. The most noticeable

consequences of those unwanted end effects are the

development of non-uniform stress distributions and

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
41

bulging of Che specimen aC large axial strains. In addi­


L
tion Co radial shear scresses (and hence complementary

shear scresses on circumferential surfaces), end restraint

effects produce changes in the distribution of the axial,

radial and circumferential normal stresses, which depend

on the constitutive law of the soil. Under these condi­

tions, even the radial stress may not be uniformly distri­

buted across the triaxial cylinder. The influence of end

restraint is more significant for brittle materials

because of the potential occurrence of progressive

f al lur e.

The most undesirable result of non-uniformities of

strains in a triaxial specimen is that the pore water


f
V pressure migrates within the specimen as different strains

induce different volume change tendencies, producing a

significant redistribution of water content or void ratio

throughout the specimen which increases with the shear

strain amplitude (14,15,53). Sands which are medium dense

or dense exhibit a tendency to contract under small shear

strains and to increase in volume substantially under

large shear strains. Consequently, if during an undrained

test a portion of the specimen is subjected to small shear

strains while another is developing large shear strains,

this latter part of the specimen will become looser at the

expense of the compaction of the other part (14).

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
42

Under cyclic loading, redistribution of water content


u
is further increased by the switching of loading from the

compression to the extension side. As the cyclic strain

amplitude increases, it has been observed that the axial

strains develop manly in a narrow zone at the top of the

specimen in the form of bulging during compression and

necking during the extension portion of the cycle. In

some cases, it is considered that even a water layer can

be observed at the top of the specimen when it is cycled

through the hydrostatic state of stress (14). Based on

these observations, Castro (14) and Casagrande (13) have

concluded that there is a "pumping action" of the vertical

cyclic forces which seems to draw water to the top of the

^ specimen in the instance of cycling through the hydros­

tatic state of stress. This action combined with the

effects of non-uniform distribution of stresses on the

boundaries causes redistribution of water content and

softening of the top of the specimen. It is clear that If

there is a significant redistribution of water content,

both the recorded pore pressure and the axial strains dur­

ing cyclic triaxial tests cannot be attributed to the

behavior of a uniformly strained specimen but in great

proportion to the looser layer which forms at the top of

the specimen. Accordingly, the results of such tests

would greatly underestimate the ability of sands to with­

stand cyclic loading, especially in case of dense sands.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
43

( Non-uniform distribution of strains and the resulting

redistribution of water content is a limitation that to

some extent is common to all cyclic shear tests. However,

the reasons for the "pumping” mechanism are not clear yet.

As discussed in Chapter 3, high pore water pressures may

develop in cyclic shear tests at the moment the deviator

stress is zero and in some cases they may equal the

applied confining pressure. Accordingly, at this stage

the effective stress acting on the specimen is zero. In a

sense this is only true at the top of the specimen. At

the bottom of the specimen the sand is subjected to the

weight of the overlying sand. Therefore, when the axial

load is reapplied, it is reasonable to expect strain con-

^ centration to take place at the top of the specimen. The

same effects, although to a limited extent, have been

recognized in other cyclic shear tests (174). However, if

the "pumping" effect and hence the development of a looser

layer at the top were due to the effect of the weight of

the specimen under low levels of effective confining

stresses, they would be likely to develop in other cyclic

test specimens. However, in torsional shear test and sim­

ple shear test specimens, the greatest loosening of the

sand has been found to develop in the middle plane and not

at the top of the specimen (13). On the contrary, as

reported by Casagranae (13), these specimens were found to

develop compacted zones adjacent to the cap and the base.

Consequently, it is very likely that the concentration of

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
44

strains at the top of the specimen is peculiar to the

cyclic triaxial test, associated mainly with necking of

the specimen under axial extension. Of course, this con­

dition is not characteristic of field conditions, and

hence, cyclic triaxial tests will exaggerate the cyclic

deformations that might develop in the field.

Lubricated end platens can be used to reduce end res­

traint effects and hence to improve the uniformity of

strains in triaxial test specimens (140). Lubrication is

readily provided by means of a combination of rubber

sheets and lubricating silicone grease between them, which

are placed between the specimen and the surface of the

platens. Sometimes a short porous dowel is used at the

center of the platens to avoid side-slipping of the speci­

men. This arrangement permits more uniform strain condi­

tions to develop throughout the length of the specimen.

Consequently, the assumption of equal radial and tangen­

tial strains may be valid for lubricated platens but

largely inadmissible for specimens tested with rough pla­

tens. In the latter case, quasi— rigid zones form at both

ends of the specimen producing non-uniform strain condi­

tions throughout the specimens (140). Since lubrication

results on a much more uniform strain pattern, it will be

reflected in more uniform volume changes and crushing of

particles under drained shearing or more uniform pore

water pressures in undrained tests. In general, a

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
i significant increase in strength may be observed in tests

with lubricated ends as compared with tests using regular

ends, due perhaps to the fact that the uniform distribution

of strains reduced the possibility of a single failure

plane (140). However, it should be noted that lubricated

ends reduce but they do not eliminate the problem of neck­

ing under axial extension.

The cyclic triaxial test can be used to determine

dynamic soil properties (Young's modulus and damping

ratio), or cyclic undrained strength. In the first case,

the Young's modulus, E, and the axial strain, e3> , are usu-

ally converted to values of the shear modulus, G, and the

single amplitude shear strain, y, by using the elasticity

C theory relationships (89,128). However, shear strain


_2
measurements below about 10 Z are in general difficult to

achieve with cyclic triaxial test apparatus. The results

of cyclic triaxial tests for determining undrained

strength are usually presented in the form of time his­

tories of the pore water pressure and the axial strain.

The cyclic undrained strength is then expressed as the

relationship between the cyclic stress ratio and the

number of cycles required to induce liquefaction or a cer­

tain level of axial strain. The cyclic stress ratio is

defined as the amplitude of the cyclic deviator stress,

Ao, divided by twice the initial confining pressure. In

spite of all the problems discussed above, Seed (145)

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
46

insists that carefully conducted cyclic triaxial tests can

provided "valid" data up to the development of pore pres­

sure ratios of 100% and strains on the order of about 5%

for dense specimens or 20% for loose specimens. This view

is not shared by other investigators (13,14,15).

2.3 The Simple Shear Test

Among the available cyclic strength tests, the simple

shear test is considered to simulate most closely the

stress conditions to which soil elements in a horizontal

deposit under level ground are subjected during an earth­

quake. Under level ground conditions, which can be

approximately modeled by a half-space, lateral expansion

c or contraction of the soil elements in the horizontal

direction cannot occur. Consequently, under the

earthquake-induced shear stresses, the soil elements

undergo shear deformations but without change in the cross

section perpendicular to the direction of shear wave pro­

pagation. In the simple shear test, a cyclic horizontal

shear stress, Tv k» applied'at the top or bottom of a

specimen which is constrained laterally, after consolidat-

ing it under an initial vertical stress, o^.

Two main types of simple shear apparatus have been

developed which use slightly different conditions to con­

strain the specimen. They are commonly identified as the

Roscoe device and the Norwegian Geotechnical Institute

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
47

(NGI) device, respectively (140,177). The Roscoe-type

apparatus uses a square specimen placed within a box with

hinged edges, whereas the NGI device uses a cylindrical

specimen which is restricted from deforming laterally by

thin wires embedded in a rubber membrane. It is evident

that in both configurations K conditions are intended to


o
develop during consolidation under the static vertical
*
stress, <?v , whereas the specimen is able to undergo shear

deformations without change in the cross section upon

applying the cyclic horizontal shear stress,

Nevertheless, it should be noted that a slight lateral

expansion of the membrane in the NGI— type apparatus could

result in the specimen attaining an active stress state.

In general, neither apparatus permits control of the

lateral stress, o^.

Under the idealized earthquake loading conditions,

soil elements are considered to be subjected to reversing

shear stresses on both horizontal and vertical planes.

However, it is evident that complementary shear stresses

cannot develop on the vertical boundaries of the NGI-type

specimen, whereas in the Roscoe-type apparatus, the hinged

rigid surfaces are supposed to provide the necessary boun­

dary stresses. Accordingly, in either case there is an

inherent difficulty to apply the desired shear stresses to

the specimen. This limitation leads in the first place to

a non-uniform distribution of shear stresses on the

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
48

c surfaces of the specimen, which in turn affect the distri­

bution of normal vertical and horizontal stresses in order

to maintain equilibrium (140). In this connection, it is

easy to infer that especially in the case of the NGI-type

device the resultant of the vertical force on the upper

and lower faces of the specimen must form a couple in

order to balance the moment produced by the shearing

forces and hence the vertical normal stresses are unlikely

to be uniformly distributed.

Comprehensive series of studies, both theoretical and

experimental, have been performed to study the distribu­

tion of stresses in simple shear test specimens, both

square and circular ones. These studies have been

c reviewed in detail by Saada et al. (140).

have shown that in order to satisfy the assumed boundary


Such studies

conditions (no lateral deformation), the stress distribu­

tions on the faces of the specimen should be as illus­

trated in Figure 2.4. These results confirm the non­

uniformities and high concentration of stresses. However,

it should be noted that to maintain equilibrium and to

satisfy the plane strain condition the solutions involve

the existence of tensile and compressive stresses at the

end platens and on the sides of the specimen. It is phy­

sically impossible that these boundary stresses could be

applied by the hinged sides of the Roscoe device or any

reinforced membrane. Consequently, it is difficult to

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
1.2

0.8

X= 0.2
-------------X= 0.5
0.4
RATIO
STRESS

0.4

0 .8f-

1.2b
0.25 0.50 0.75

X/ 1

t = average shear stress

Figure 2.4 Theoretical Stresses on the Boundaries of a


Square Simple Shear Specimen (After Ref. 140)

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
50

justify the assumption of perfect plane strain since it

requires boundary conditions that no device can physically

apply. However, it is important to note that these solu­

tions suggest that stress non-uniformities are less severe

in the square specimen than in the circular one. In the

latter case, even significant shear stresses in the direc­

tion normal to that in which the shearing force is applied

must develop, which further precludes the development of a

plane strain condition.

On the other hand, both shear strains and volumetric

strains in soils may result from either shear or normal

stresses and vice versa. Therefore, the volumetric

strains resulting from non-uniform shear stresses will

also be non-uniform; and since the top platen through

which the vertical stress is applied is rigid, an addi­

tional non-uniform and unknown redistribution of vertical

normal stresses will take place in response to this ten­

dency for volume changes (140,192). Since the thickness

of the specimen is generally less than 1/3 of the size of

the loading plate, significant end restraint effects may

develop even without an applied shear stress (138,140).

Furthermore, additional changes in the normal stresses may

occur as shearing progresses, especially associated with

the fact that the horizontal direction is a zero-extension

direction (no lateral deformation) and additional couples

are applied by the normal vertical stresses because of

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
increasing eccentricity. Consequently, even though the
' W
specimen is subjected to a "simple shear" state of strain,

the resulting state of stress throughout the specimen is

not simple but very complex. Nevertheless, measurements

of strain distributions show relatively high uniformity

(113). Therefore, a "uniform" state of strain does not

necessarily imply a uniform state of stress (140). Nei­

ther the vertical normal stress nor the applied shear

stresses can be uniformly distributed. Obviously, the

non-uniformities increase with increasing shear deforma­

tion. These features make it very difficult to make an

appropriate interpretation of simple shear test results.

According to Saada et al. (140), simple shear tests cannot

^ yield either reliable stress— strain relations or absolute

failure values.

In spite of the limitations just discussed, the

author finds some positive features in the simple shear

test. First of all, it is easy to develop K q conditions

prior to the application of the cyclic shear stress.

Therefore, it permits to study the effects of prior stress

history under plane strain conditions on the cyclic

behavior of sands, and hence to establish a relationship

between the overconsolidation ratio, OCR, and liquefaction

resistance (6,145). When working with dry specimens, it

provides volume change measurements to a degree of preci­

sion which is rather difficult to obtain with any other

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
52

cyclic shear test apparatus. These measurements can be

used to predict settlements due to earthquake loading in

deposits under level ground. Constant volume simple shear

tests have been performed by locking the loading head at a

fixed vertical position. The tendency of a dry sand

specimen to compact during shaking causes a progressive

reduction in the vertical pressure on the loading head.

The reduction in the average vertical stress is assumed to

be equivalent to the increase in pore water pressure in a

corresponding undrained test (46,140). However, in view

of the foregoing discussion it is difficult to ensure that

the pore water pressure, if uniform, is equal to the

reduction in the average vertical stress. Moreover,

results from simple shear tests performed by Saada et a l .

(139) clearly indicate that the reduction in vertical

stress in a constant volume test does not equal the

increase in pore pressure at the same shear strain in a

corresponding undrained test.

The deformations imposed on a simple shear specimen

simulates the deformations thought to occur for example at

the bottom of a circular rupture surface. Accordingly,

the ability to produce a horizontal rupture layer in a

cylindrical or square specimen makes it an excellent tool

for evaluating the influence of anisotropy on the shear

strength of soils (5). In principle, the test could be

used to investigate the effect of rotation of the planes

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
53

of principal stress. However, studies have concluded that


( ..J
the fixed no-extension condition in the horizontal direc­

tion severely limits the actual rotation of the direction

of the major principal stress within the specimen during

straining (5,111). To overcome this limitation, Arthur et

al. (4,5) developed a directional shear cell, in which a

cube is subjected to normal and shearing stresses on four

of its faces while the others are not allowed to deform.

This device has some common features with torsional shear

test apparatus which are discussed later on.

Simple shear tests can be used to determine

equivalent dynamic properties (shear modulus and damping),

a case in which the tests are usually run on dry speci­

( mens, and to evaluate cyclic undrained strength, which is

expressed in terms of the ratio between the peak alternat­

ing shear stress, t and the initial vertical stress,


^ *»
o . The plot of stress ratio, x , versus the number
vo vh vo’
of cycles to induce liquefaction or a given level of shear

strain, constitutes the liquefaction resistance curve.

However, the non-uniform stress and strain conditions may

produce significant redistribution of water content, thus

affecting the development of pore water pressures.

Accordingly, it is generally considered that the simple

shear test may give an underestimate of the cyclic

undrained strength of sands, especially in the case of

dense sands (15,173).

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
54

In the early studies, it was found that the resis­

tance to liquefaction under simple shear conditions was

smaller than the resistance under analogous conditions in

cyclic triaxial test apparatus (129). This discrepancy

was attributed in the first place to the complex non-

uniform distribution of stresses and strains within the

simple shear test specimen. Among other reasons given for

explaining the difference between the cyclic triaxial and

simple shear strengths were that the orientation and rota­

tion of the principal stresses, the relative magnitude of

and the lateral constraint in the two specimens were

different (75). However, Finn et a l . (50) pointed out

that if the cyclic stress ratio in the simple shear test

is defined as the ratio of the peak cyclic shear stress,


(
Tv k> to the initial mean confining stress, o^, approxi­

mately equal liquefaction resistance curves are defined

from both tests. The initial mean confining stress, o ,


ni
is related to the vertical stress by the expression

1+2 K
a = =-- 2. a ( 2 .2 )
m 3 v

where Kq is the coefficient of earth pressure at rest.

Therefore, if

<2.3)
o 2o
m c

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
55

It follows that

x . 1+2 K .
5— - (2.4)
2 »c

Accordingly, it was concluded that the cyclic triax­


ial undrained strength had to be corrected in order to

obtain the corresponding plane strain strength, which from

the simple shear test is obtained in terms of the ratio,

t ,/a , where., a would correspond to the effective


vh vo vo
overburden pressure at a given depth within a sand deposit

under level ground. Good agreement in terms of shear

modulus and damping ratio values was also found by Park et

al. (128), when the values obtained from the two tests
*
were compared at the same mean confining stress, a .
m

Nevertheless, in the author's opinion, the difference

in results from cyclic triaxial tests and simple shear

tests cannot be just a matter of definition of the cyclic

stress ratio as discussed in the last section of this

chapter and in Chapter 3. On the other hand, it has been

shown conclusively that in a short cylinder specimen such

as the NGI-type specimen, the end restraint effects may be

significant and therefore the radial and circumferential

stresses are likely to be different (138,139). The end

effects are attributed to the roughness of the platens and

to expansion of the membrane when the specimen is sub­

jected to axial load. Consequently, if the results from

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
56

( triaxial and simple shear tests agree by normalizing with


^— '' *
respect to o , it is perhaps just as matter of coincidence
m
since the actual stresses acting inside the specimen may

be far from the values implied by E q . 2.2. In this con­

nection, it is worth mentioning that in a comprehensive

study by Tatsuaka et al. (175) by means of cyclic triaxial

and torsional shear tests, it was found that the relation­

ship between the cyclic triaxial strength and the plane

strain strength is not only a function of the coefficient

of earth pressure at rest, Kq , but also of at-least the

type of sand and the method of specimen preparation.

In an effort to eliminate the boundary effects asso­

ciated with tests on small-scale simple shear test speci­

es mens of sands, test equipment utilizing a shaking table as

a means of inducing cyclic shear stresses on large-scale

specimens under simple shear conditions have been

developed (30,1J51). By using specimens with a large

length/height ratio it is possible to minimize the effects

of lack of complementary stresses at the ends of the

specimen. The boundary effects are not completely elim­

inated and therefore the results of actual tests are

influenced by the dimension of the specimen and by the

stiffness of the specimen container in case the specimen

is in contact with the walls of the container. The boun­

dary conditions should be clearly and properly defined if

meaningful data are to be obtained from shaking table stu­

dies.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
57

^ When working with specimens with large surface areas,

a major difficulty arises because of penetration of the

membrane between sand grains (30,192). This is the case

with large-scale specimens with a large length/height

ratio. A small penetration may permit significant volume

changes to develop in response to pore water pressure

changes, thereby altering the behavior of saturated speci­

mens. Consequently, consideration should be given to

these effects and corrections are required in order to

obtain accurate results.

The large-scale shaking table tests performed by De

Alba et al. (30) were considered to provide more accurate

data, on the development of shear strains in denser speci­

es mens after initial liquefaction and on the behavior of

dense sands at relatively high stress ratios, than previ­

ously obtained from cyclic triaxial tests. Contrary to

what could be expected, De Alba et al. (30) found a very

good agreement between the results obtained using the

shaking table and those obtained in simple shear tests on

small specimens. Based on these results, it has been con­

cluded that the errors due to stress concentrations in the

small-scale tests may not be as large as has often been

claimed or they are counterbalanced by some other feature

of the tests (30,145,192). In view of these findings, it

becomes very difficult to question the value of the test

to simulate simple shear field conditions in spite of the

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
58

internal complexities and uncertainties previously dis­

cussed •

2.4 The Torsional Shear Test

Among all the cyclic strength tests, the torsional

shear test on solid and hollow cylinder specimens allows

for reproducing the most general stress conditions. A

torque is reversed and cycled to simulate earthquake load­

ings and the resulting rotation of the planes of principal

stress. The torque develops shear stresses on horizontal

planes and complementary shear stresses also develop

automatically on vertical radial planes. By applying this

torque along with the confining pressure and the axial

load, nearly any desired combination of ambient principal

stresses or stress path may be produced in the specimen

(52,65,97,140,169,181,182), particularly if different

internal and external pressures are applied to a hollow

specimen (65,66). Another major advantage of the tor­

sional shear test is that the circular shape and the cir­

cumferential direction of shearing ensures under torsion a

condition of pure shear and makes it possible to duplicate

in the laboratory the large strains which occur in situ,

for example during liquefaction failures (14,30,59).

On a solid cylindrical column of soil subjected to

torsional loading, the shear strains range from zero at

the center of the specimen to a maximum at the outside

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
59

c radius; Che resulting variation in shear stresses will

depend upon the stress-strain properties of the tested

specimen. This was pointed out early as a major disadvan­

tage of this type of test (140,192). However, it has been

shown that the effects are not always significant

(38,75,88), especially in the case of drained tests. The

relationship between the average shear stress and the

applied torque does not differ appreciably (about 11% at

most) , whether a linear variation of shear stresses (elas­

tic model) or a uniform shear stress distribution (rigid

plastic model) is assumed. The most correct relation

between torque and shear stress is non-linear and shear

strain dependent, lying between the two extremes.

c The non-uniformities produced by the applied torque

can be reduced if a hollow cylinder is used. For

liquefaction studies of sands, the use of hollow specimens

is advantageous because pore pressure gradients are

reduced (59). The use of hollow specimens poses, however,

some difficulties regarding specimen preparation, espe­

cially for testing "undisturbed" cohesionless soils

(140,192). Moreover, when hollow cylinders are used, mem­

brane penetration effects are much more severe since the

ratio of the specimen surface in contact with the membrane

to the volume of the specimen is about five times that of

a solid cylinder specimen (39).

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
60

Figure 2.5 shows Che stress components in hollow tor­

sional shear test specimens. It is evident that the

stress condition is axially symmetric and that six

independent stresses must be specified to completely

define the state of stress. The axial stress, a , the


z
radial stress, a , and the shear stress, t fi, correspond
IT ZO
to the applied boundary stresses. Since the confining

pressure, , acts through flexible membranes, there are

no shear stresses on vertical circumferential surfaces.

Accordingly, t^ z and Tr g are equal to zero and hence the

radial stress, o ^ , is a principal stress. Consequently, in

addition to the distribution of the shear stress, t q, the


Z0
only unknown is the circumferential stress, Og, which is

c the normal stress in the direction of particle shear

deformation. This stress can be determined by considering

force equilibrium along the radial direction. The magni­

tude and direction of the major and minor principal

stresses, which act on the circumferential plane, are

determined from the magnitude of shear, axial and circum­

ferential stresses. The orientation of and can be

maintained or rotated, if desired, during the test. The

radial stress corresponds to the intermediate principal

stress, w2 • A comprehensive discussion on stress and

strain components and the corresponding distributions

across the cylinder wall is presented in Appendix B.

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
61

K.J

a\ *■ r

Tze

Oz
Tzi

Tez

<Tr

a) A pplied B o u n d a r y b) S t r e s s e s on an
S tresses
C E le m e n t in the Wall

Q - a n g le b e tw e e n

T O i d ire c tio n
Pole
an d z a x is
Z0

OS

ez

c) M o h r's C ircle D ia g r a m

Figure 2.5 Stress Components in Hollow Torsional Shear


Test Specimens.

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
62

Assuming linearly elastic isotropic properties, and

for the case of equal internal and external pressure, the

radial and circumferential stresses turn out to be equal

and uniformly distributed across the wall thickness of the

test specimen, irrespective of the applied shear stress,

provided there are no end effects (140). However, the

non-uniform distribution of shear stresses ( t q) causes


Z O
different tendencies for volumetric strains across the

wall of the specimen. Consequently, at larger strains,

non-uniform distribution of axial, radial and circumferen­

tial stresses may also be induced. Even on an isotropic

material, the radial stress need not be constant and equal

to the circumferential stress when a torque is applied.

Only the average values of these stresses, "or and "5q, are

equal. The actual distributions of these stresses depend

on the specimen geometry and the stress-strain properties

of the soil. For anisotropic materials, stress non­

uniformities might arise, even without an applied torque

(6 6 ). By adopting a specimen geometry with large outer

radius and thin wall, the shear stress gradient can be

substantially reduced and with it the non-uniformities in

the distribution of normal stresses across the cylinder

wall.

If the internal and external pressures are different,

the radial and circumferential stresses are not equal in

any case. This testing procedure can be used to control

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
63

Independently the effects of changes in the magnitude and


c
orientation of the major and minor principal stresses and

relative changes in the value of the intermediate princi­

pal stress, but at the expense of additional non­

uniformities in the stress distributions (65,66).

In the torsional shear test the distribution of shear

strains is approximately the same at all the cross sec­

tions along the height of the specimen. Minor deviations

from these "uniform" patterns have been attributed to the

effect of the weight of the specimen under low levels of

confining stresses (174). Consequently, the specimen

deformation is not concentrated on one narrow band except,

possibly, when failure develops. Actual specimens often

c show a number of shear planes or a continuous spiral shear

zone after the peak shear stress has been reached (52,66).

This is certainly a major advantage for investigating the

post-peak behavior. For example, drained torsional shear

tests on dense sand specimens may result in strain soften­

ing behavior starting at much larger strains than in

corresponding triaxial tests, probably because a large

portion of the specimen in the torsional shear test

approaches failure simultaneously, in comparison with a

failure along a single plane in the triaxial test.

However, it is not to be concluded that a hollow

specimen can be considered completely free from stress

concentration problems, as has been suggested elsewhere

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
64

(173). Because of the tendency of the sand to dilate when


c
shear stresses are applied* it is likely that radial shear

stresses are induced at both ends of the specimen due to

the end restraints and stiffness of the platens. Comple­

m e n ta ry shear stresses on circumferential surfaces also

develop and the intermediate principal stress rotates out

of the horizontal direction. These shear stresses produce

changes in the circumferential stress and induce bending

moments which affect the distribution of vertical normal

stresses (52*65*97,140). The magnitude of stress changes

due to end restraints depends on the stress-strain proper­

ties and geometry of the tested specimen* and on the com­

bination of applied loads. However, as the induced shear

stresses decrease with distance from the ends of the


(
specimen (St. Venant's principle), the central part of a

properly proportioned hollow specimens considered to be

relatively free of such effects (65,97,140,192). There­

fore, it would be more meaningful for interpreting the

test results to measure the deformations by sensors

located within the central part of the specimen (65,168).

The torsional shear test has been shown to be a valu­

able tool for investigating the behavior of soil specimens

under a wide range of loading conditions. A variety of

stress conditions can be reproduced on torsional shear

test specimens with no restraint on their strain response

other than those resulting from specimen shape and end

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
65

restraints, and both can be reduced to acceptable limits

with a properly selected specimen geometry (66).

Undrained plane strain conditions can be developed by

preventing any change in the volume of the cell water or

by fixing the length of the specimen. In the first case,

the cross-sectional area of the vertical loading piston

should be the same as that of the hollow cylinder specimen

(75,81,175). Torsional shear test apparatus can be used

in fundamental investigations of anisotropy and principal

stress rotations in sands and clays. It is well esta­

blished that to study anisotropy, it is better to keep

specimens along the axes of symmetry and incline the prin­

cipal stresses in order to obtain meaningful strain meas­

urements (139,140). Continuous rotation of principal

stress directions appears to have an important role in

properly simulating the stress paths which are follow,

e.g. by elements beneath ocean structures (83,181,182).

The particular advantages of the torsional shear test for

evaluating liquefaction potential are discussed in the

next section. Stress non-uniformities and end restraints

obviously become larger as failure is approached and their

influence must be considered in interpreting the test

results. For example, uncertainty in the value of the

mobilized friction angle, , can arise when failure is

approached and the specimen shows a tendency to bulge.

Under these conditions, the circumferential stress cannot

be assumed to be equal to the confining pressure because

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
66

/
of Che development of radial shear stresses due to end
L.
restraint effects (see discussion In Appendix B ) .

2.5 Prediction of Soil Liquefaction

Limitations of the cyclic triaxial test and differ­

ences between predictions of liquefaction potential

obtained from it versus those obtained from cyclic simple

shear or torsional shear tests have long been recognized

(13,14,145,192). However, several additional effects have

not been fully or explicitly recognized when comparing

results of liquefaction tests from cyclic triaxial and

torsional shear tests.

The stress-strain characteristics of sands change

with the direction of loading reflecting the effect of

inherent anisotropy developed during deposition or as a

result of the stress-strain history of the deposit

(4,118,120,125,127,16-9). For example, ample experimental

evidence has shown that specimens prepared by the method

of pluviation through water or air are more compressible

laterally than vertically (78,119,125). In the triaxial

test specimen, the principal stresses are respectively

perpendicular and parallel to the direction of the assumed

bedding planes. On the contrary, in the torsional shear

specimen, the axis of the major principal stress changes

its orientation from +a t»/ -a from the vertical (Fig.

2.5c), in the course of reversing the shear stress from

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
67

clockwise to counterclockwise (181). Therefore, it is

important to recognize that in the torsional shear test,

the orientation of the major principal stress relative to

the planes of anisotropy is similar in each half of the

loading cycle. ' From this point of view the response of

the specimen is expected ‘to be the same whether it is

rotated right or left. This is not the case in the triax­

ial test, in which the cyclic load alternately induces

axial compression and extension in the specimen. These

two phases of each cycle produce different results in the

great majority of cyclic triaxial tests. Accordingly,

hysteresis loops are not symmetric in strain-controlled

tests whereas larger deformations are observed in exten­

sion than in compression in stress controlled tests, which


c
increase the tendency to neck.

Another aspect which is more conveniently simulated

by the torsional shear test than by the triaxial test is

the magnitude of the rotation of the principal stresses.

Figure 2.6 presents the relationship between the initial

(K ) state of stress, the applied shear stress ( t , ), and


o vn
the magnitude of the rotation of the principal stresses

for a soil element in situ assuming vertically propagating

shear waves during an earthquake. It is evident that the

magnitude of such rotation decreases as the existing in

situ shear stress increases, especially if the applied

shear stress ratio is not very large. This is of major

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
68

V. /

ov
Pole

J vh

cr

wo

50"
Ko-1.0
OF ROTATION, Gt (degress)

0.90
40"

0.75

0.5

20 "
ANGLE

10- -

0.1 0.2 0.3 0.4 0.5

APPLIED SHEAR STRESS RATIO, Tvh/<7$0

Figure 2.6 Rotation of Principal Stresses in a Soil


Element In Situ during an Earthquake.

R eproduced w ith perm ission o f the copyright owner. F urth er reproduction prohibited w itho ut perm ission.
69

significance when investigating the effect of inherent

anisotropy or of changes in fabric due to the rotation of

principal stresses. It is also relevant for evaluating the

effects of the initial state of stress and cyclic stress

histories on liquefaction potential. In the cyclic triax­

ial test, changes in major principal stress directions are

restricted to sudden jumps of 90°, whereas in the tor­

sional shear test the planes of principal stress gradually

rotate as the applied shear stress increases (Fig. 2.5c),

resembling the rotation that occurs in situ as depicted in

Figure 2.6.

In the cyclic triaxial test, both the maximum conso­

lidation shear stress, if any, and the maximum cyclic

shear stress are applied on the same 45° planes, whereas

in the field they usually act on different planes. It is

evident that the latter is the case in the torsional shear

test. Moreover, in the cyclic triaxial test, the maximum

shear stress is reduced below the initial consolidation

value periodically during each cycle, while in most field

situations static shear stresses are sustained due to the

need of maintaining equilibrium of shear stresses imposed

by geometry. Therefore, the shear stresses must usually

remain larger than the static pre-earthquake maximum shear

stress. In the torsional shear test, the minimum value of

the maximum shear stress is equal to ( - or )/2, which

corresponds to the consolidation shear stress. The

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
70

effects are further discussed in Chapter 3 in connection

with the effect of initial states of stress and boundary

conditions on the liquefaction resistance of sands.

A final factor to be discussed is the effect of

changes in the total stress as related to the pore pres­

sure increment during each cycle of loading. The changes

in pore pressure during undrained shear in monotonic load­

ing are related to the changes in the total principal

stresses and can be expressed in terms of Skempton's or

Henkel's pore pressure parameters. Figure 2.7 compares

the stress paths during the loading stages of cyclic

triaxial and torsional shear tests on isotropically conso­

lidated specimens. In the triaxial test, pore pressure

c build-up is a consequence of both the applied shear

stresses and changes in the mean normal stress, whereas in

the torsional shear test the mean normal stress remains

constant through the cycle and therefore pore water pres­

sures are caused only by the applied shear stresses. In

order to maintain constant the main principal stress act­

ing on a triaxial test specimen, it would require to

increase the axial stress by Aa, and a simultaneous reduc­

tion in the cell pressure by an amount Ao/2, which is a

difficult procedure using conventional triaxial apparatus.

On the other hand, it is argued that reducing or increas­

ing the cell pressure by a given amount, would simply

change the pore water pressure in the same amount, without

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
Reproduced with permission of the copyright owner. Further reproduction r
v ,

Compression Counter-Clockwise rotation

AU
—i Pole
ESP
ESP —

cr.P cr.p
Pole

Oitf. In mean
effective stress

Clockwise rotation
T,q T,q
Extension
prohibited without perm ission.

cr.p cr.P
Pole ESP

ESP
i-- 1
AU Pole
Note that pc =' pE

a) Cyclic Triaxial Test b) Torsional Shear Test

Figure 2.7 Stress Paths for Isotropically Consolidated Specimens.


72

causing any change in effective stresses and therefore,


c with no effect on the deformation of the specimen

(103,104,148). Because of these reasons, and of course

for convenience, the test is normally performed by main­

taining the cell pressure at a constant value and cycling

the axial stress.

An increment in the axial stress of Ao implies a

change in the mean principal stress and in the pore water

pressure equal to Ao/3, and hence no change in the mean

effective principal stress. Nevertheless, it should be

noted that there are changes in the individual values of

o ^ , o^, and o^. It is not clear yet whether the shearing

strength of soils is controlled by the mean effective

c principal stress or by the relative contribution of each

stress component to the effective normal stress on the

potential failure planes. The author considers that as

shear strains increase, soil properties such as dilation

could be more a function of the major and minor principal

stress. Yu et al. (198) have reported that even the max­

imum shear modulus, G , determined at very small shear


* max
strains by means of resonant column tests seems to be

relatively independent of the intermediate principal

stress. Accordingly, in the author's opinion, a change on

the mean confining stress acting on a triaxial test speci­

men may contribute to changes in the relevant stress com­

ponents, therefore influencing the deformation response of

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
73

the specimen. Additionally, as discussed earlier, the


c .
pore water pressures caused by the applied deviator

stresses are different during axial compression and axial

extension.

In Figure 2.7 the effective normal stresses are

represented in terms of the stress path parameter


0> + *
(p" « ( C j + o ^ ) ^ ) . It is clear that the change in p will

be different during the two loading stages (compression

and extension) in the triaxial test, even for the case of

a specimen with no inherent anisotropy. A symmetrical

effective stress path is expected for the two loading

stages in the torsional shear test. Hence, the simplified

procedure recommended by Seed and Lee (148) , which util­

c izes the results of cyclic triaxial tests in practice, is

conceptually inadequate, in the sense that the relative

contributions of the normal and shear stresses to the

change in pore pressure cannot be independently deter­

mined. Consequently, the pore pressure build-up during

cyclic triaxial tests is not representative of earthquake

loadings.

In summary, the features discussed above strongly

favor the use of cyclic torsional shear tests in lieu of

triaxial tests to study the liquefaction potential of

saturated sands. The prevalent belief that the difference

in results from cyclic triaxial and simple shear tests is

only a matter of definition of the cyclic stress ratio or

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
74

the nature and degree of non-uniform distribution of


o strains in the specimen (13,30,145), is no longer accept­

able.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
C HAP TER 3

CYCLIC UNDRAINED BEHAVIOR OF SANDS:


A FRAMEWORK OF UNDERSTANDING

There is ample experimental evidence which shows that

the application of small amplitude cyclic shear stresses

to sands under drained conditions results in a progressive

decrease in volume, even in the case of a dense sand with

a void ratio below that which might cause dilation during

unidirectional, monotonic loading at a given confining

pressure. Consequently, when a saturated sand deposit is

subjected to propagating shear waves during an earthquake,

the sand structure tends to decrease in volume. However,

since the duration of the cyclic stress applications is in

general short as compared to the time required for

drainage of water to occur, then the tendency towards den-

sification of the sand structure during each successive

application of shear stress results in a progressive

increase in pore water pressure, which causes a continued

reduction in the effective stress and a corresponding drop

in the shear strength of the sand. A substantial reduc­

tion in shear strength may lead to shear failures, which

sometime are of catastrophic consequences.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
76

If in the course of cyclic loading the pore pressure


o builds up to a value equal to the initially existing con­

fining stress, no effective stress or intergranular stress

is acting on the sand, which having lost its strength is

considered to have "liquefied," although this may be only

a temporary state. At this state, individual particles

are postulated to exist as if they were in a state of

suspension in water (74), as illustrated in Figure 3.1.

(This is a useful concept but its physical existence has

not been proved.) Accordingly, such a state will be called

"liquefaction" throughout this dissertation. The catas­

trophic effects of liquefaction on overlying structures

are obvious. As liquefaction develops, drainage of water

towards more permeable boundaries of the deposit or


c towards areas of smaller pore water pressure is initiated

together with rearrangement of individual particles of

sand into a denser packing (Fig. 3.1c). The length of

time for which liquefaction state continues to exist

depends upon the drainage conditions and also on the dura­

tion of cyclic stress applications following the onset of-

liquefaction.

The accumulation of pore water pressure prior to the

development of liquefaction is controlled by the compres­

sibility characteristics of the sand, primarily within the

range of small shear strains. As discussed earlier, it

has been shown that relative density and the initial state

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
Reproduced with permission of the copyright owner. Further reproduction »

Water Increase in pore Surface


surface water pressure settlement

\ ^7
06o&rodo

oo£bo o®
°ooxooxo
prohibited without perm ission.

Effective
stress

Total
stress

(a) Deposit prior (b) State of suspension (c) Deposit after


to liquefaction during liquefaction liquefaction

Figure 3.1 Definition of State of Liquefaction (Aft er Ref. 74).


78

of stress are not sufficient to characterize the mechani­

cal behavior of granular materials. Other factors, such

as composition, size distribution and shape of the parti­

cles, fabric and its associated anisotropy, stress-strain

history, stress path, and aging since the time of deposi­

tion, among others, strongly influence the stress—

deformation behavior of cohesionless soils

(4.93.101.117.119.125). Consequently, all these factors

also affect the pore pressure buildup and the consequent

strain development in sandy and silty soil deposits during

an earthquake, as discussed in this Chapter.

The importance of initial fabric on the stress-strain

response of sands has long been recognized

(94.101.123.125). The mechanism of deposition of the sand


c
deposit has a significant effect on the initial fabric and

hence, on the deformation characteristics. Minor differ­

ences in particle packing, resulting from different

preparation methods, significantly affect the stress-

strain behavior of reconstituted specimens at the same

relative density (94,101,123,175). Similar effects will

occur in situ because of variations in the deposition pro­

cess. It has also been recognized that the fabric

developed under conditions prevalent during the deposition

of most natural sand deposits is anisotropic. This

inherent anisotropic fabric causes changes in

deformation-strength characteristics with the direction of

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
79

loading (118,120,125,169). Moreover, the stress path


o itself can induce an anisotropic fabric in sands in the

course of shear loading (4,126,178). This induced aniso­

tropy may be associated with rotation of the planes of

principal stress and changes in the state of stress

required to produce a given orientation of potential

failure planes. Accordingly, the mechanical response of

sands cannot be independent of inherent anisotropy and of

the nature of the stress path with respect to the orienta­

tion of the principal axes of anisotropy.

Prior stress-strain history is one of the most signi­

ficant factors affecting the response of cohesionless

soils. These effects are partly reflected in the soil

^ : fabric as they change the predominant spacial distribution

and orientation of the grains or grain-to-grain contact

conditions (29,93,101,145). The potential for plastic

deformation may be greatly reduced by prestressing. How­

ever, the actual effects of past loading histories depend

on the particular stress path followed during prestressing

and on the level of the associated shear strains (101).

Consequently, prestressing might lead to stiffening of the

sand fabric in a preferred direction. The new fabric is

better able to resist shear stresses acting in this direc­

tion is such a way that below some threshold shear stress

or strain virtually no further plastic strain will occur

(93,101). However, for other loading directions, the

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
80

fabric might be even softer than the one existing prior to

prestressing (93). The stress path dependent nature of

prestressing is of particular importance when evaluating

its effects on the response of saturated sands to earth­

quake loading. In this case, the planes of principal

stress continually rotate and, thus, the relative impor­

tance of prestressing along a particular stress path may

also change.

3.1 Pore Pressure Buildup and Strain Development

Since the pioneer work by Seed and Lee (148), the

cyclic stress conditions leading to liquefaction of

saturated sand have been Investigated extensively by means

of tests on isotropically consolidated reconstituted

specimens using various types of laboratory techniques.

However, until recently, the cyclic triaxial tests has

been the most widely used among all- testing techniques.

In spite of its recognized limitations (1,14,145,192),

results from cyclic triaxial tests have provided a great

deal of valuable information for the proper understanding

of the behavior of saturated sands under cyclic loading.

3.1.1 Typical Results of Cyclic Triaxial Tests

Figure 3.2 shows typical results from undrained

cyclic triaxial tests on reconstituted specimens of a sand

at two different relative densities, which correspond to

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
81

o
e
o 30r
CO
€ at ACT- 3 9 kPa
© 20
Q.
—' F
w o 10
c O
co
I— 0I
CO
m c 10 Sacram ento River sand
o
< CO 0- * 0.87, 0 . - 3 8 %
t: 20- O g -1 0 0 kPa
CD
OS' - ± 3 9 kPa t at act- - 3 9 kPa
X
—1T“

LU
CO
o

N u m b e r of cycles

co
CL
+150

c -X

© +100
Initial effective
confining pressure
&U . at a© = 0

3 aU at
CO a c t = + 3 9 kPa
CO
©
a ll at
w +50 a0*= - 3 9 kPai
a>
co
5
©
o
Q.

©
O)
C
CO 10 20
.C
O N u m b e r of cycles

a) L o o s e s a n d
Figure 3.2 Typical Results of Cyclic Triaxial Tests
on Isotropically Consolidated Specimens
(After R e f . 148).

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
82

Jl5
CO
2 g10
€ at
w
c ao*» +70 kPa
CO
w
to I I- - - - - - - - -

. Sacramento River s a n d .
to § 5
"x CO e -“ 0.7 1, Dr - 78% e at
< =10 . < ^ - 1 0 0 kPa ______ act- - 7 0 kPa
Q> *& --± 7 0 k P a ,
"><15
UJ
10 20 40 100
N u m b e r of cycles

CO
a. -150
.X
Initial effective
at act- o
C <3
©
3
+100
confining pressure _
I ✓
a u

CO
CO
CD AO. at A G V +7Q k
— ^ — at AO"- - 7 0 k
a>
co
5
C
iD
_
O
Q.

a> 10 20 100
O)
c N u m b e r of cycles
co
O b) D e n s e s a n d

Figure 3.2, continued.

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
83

contractive (loose) and dilative (dense) states, respec­


o tively. Relative density is of course a first order

parameter, which is indicative of volume change tendencies

due to shearing. It is clear that the greater the ten­

dency towards volume contraction of the sand skeleton, the

faster the pore pressure buildup and hence, the higher

potential of liquefaction under a given series of shear

stress applications. For example, it is noted that

approximately ten cycles caused "liquefaction" in the two

cases in Figure 3.2. However, for the test on loose sand

the deviator stress was 39 kPa, whereas for the dense sand

Ao was 70 kPa.

In cyclic triaxial tests on isotropically consoli­

c; dated specimens of saturated loose sands (Fig. 3.2a) the

following events are generally observed (15,148,155):

during the first cycles of loading (cycling of deviator

stress) the specimen shows no noticeable deformation,

although the pore water pressure builds up gradually.

However, after a number of cycles, the pore water pressure

suddenly increases to a value equal to the externally

applied confining pressure. At this stage, the specimen

develops large deformations, which increase in amplitude

under subsequent loading cycles.

On the other hand, as the relative density of the

sand increases, the axial strains accumulate at a much

slower rate with increasing number of cycles (Fig. 3.2b),

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
84

even after pore water pressure becomes equal to (15).


G Following the onset of liquefaction, denser specimens

develop smaller deformations, which accumulate at a

decreasing rate reaching eventually a limiting value,

irrespective of the number of cycles of a given stress

amplitude, as will be illustrated later on.

It is important to note that in the course of these

cyclic triaxial tests, the pore water pressure may become

equal to the confining pressure, irrespective of the rela­

tive density of the sand. As illustrated in Figure 3.2,

this zero effective stress condition always occurs at the

moment the deviator stress is zero, that is, when the

specimen is subjected to an isotropic state of stress.

However, the pore pressure drops substantially when either


C
the axial compression or axial extension load is applied.

The deformations induced in the specimen each time a zero

effective stress condition develops are dependent on the

relative density of the specimen (155). Almost unlimited

deformations may develop in loose specimens, whereas the

corresponding deformations of dense specimens may be of

limited magnitude.

The behavior just described and the relationship

between pore water pressure buildup and strain development

during cyclic loading can be visualized more clearly if an

effective stress path plot is used to present the results,

as illustrated in Figure 3.3. This approach will be

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
85

cm + 0.6 a)Loose
O ttawa sand
E ASTM C - 1 09 Dr = 3 3 .8 %
■g +0.4
Kc-1
o>
7 c y / ^ c = 0 .0 9 5
■“ + 0.2
cm"
0.0 Flow def.

g"-0.2
-0 .4
0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0 2.2 2.4
(0 ^+0 ^)/2,k g f / c m 2
. cj +0.6
E
2 +0.4
o>
( ■*„ +0.2

0.0

§" - 0.2 b) Dense


Dr = 71.3%
-0 .4
i 0.251
- 0.6
0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0 2.2 2.4
{(7^+0 ^)/2, k g f / c m 2

Figure 3.3 Typical Effective Stress Paths in Cyclic


Triaxial Tests on Isotropically Consolidated
Specimens (After Ref. 184).

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
86

followed throughout this dissertation.


o
3.1.2 Pore Pressure Rise Mechanisms

Loose sand specimens are contractive and therefore

develop positive pore pressure under undrained shearing.

Consequently, as the pore pressure increases the effective

stress path moves gradually towards the left with increas­

ing number of cycles (Fig. 3.3a) . The magnitude of the

pore water pressure buildup for any given stress cycle

depends on the potential volumetric strain associated with

slippage between sand grains. Consequently, the rate of

pore pressure buildup is expected to be inversely propor­

tional to the initial stiffness of the sand specimen,

c which is determined by a combination of factors such as

composition, size distribution and shape of the particles,

initial fabric, relative density, state of stress,

stress-strain history, etc. It has been shown con­

clusively that, at the same relative density, sand speci­

mens with different fabric, that is, different grain and

interparticle contact orientations, have different

stress-strain and volume change-strain characteristics

(94,119,123,125,175). Stress-strain histories may cause

changes not only in relative density but they may also

change the predominant spacial distribution and orienta­

tion of the grain-to-grain contact conditions, hence

affecting the stiffness of the sand and the potential

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
87

Cowards densification under cyclic shearing

(29,53,101,145).

It is very unlikely that the fabric of in situ loose

sands is as "uniform" as the fabric of laboratory speci­

mens reconstituted by a particular preparation procedure.

On the contrary, it may be characterized by the occurrence

of pockets or groups of particles forming "holes" or

bigger voids which might be stable under static shear

stresses but they would be rather unstable if the sand

skeleton is subjected to a series of shear stress applica­

tions (196). Small shear strains may suffice to cause the

collapse of some of these unstable groups, which leads to

an overall increase in pore water pressure under undrained

conditions. Additional cycles of shear straining would

lead to additional collapses and a continued increase in

pore water pressure. There may be some amount of particle

crushing during the cycling process, especially breakage

of sharp contact points.

In spite of the progressive increase in pore water

pressure, it is a well known fact that the contractive

tendencies of loose sand specimens are smaller at lower

confining pressures (67,105,149). Therefore, once the

effective confining pressure becomes low enough, the

response of the specimen may change from contractive to

dilative under shearing in either direction. The state

conditions at which the behavior changes from contractive

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
88

Co dilative are discussed in detail in a later section of


o
this chapter. It is clear that during this stage of load­

ing, the initial particle arrangement is drastically

altered. Accordingly, the pore pressure response rather

reflects the changes in fabric or particle packing induced

by each cycle of shearing (196). In other words, the ini­

tial fabric may control the pore pressure response in the

first cycles of loading (Fig, 3.4a). but a new fabric is

gradually created with increasing number of cycles, which

changes with the magnitude of the shear strains that

develop in each cycle. These changes in fabric are evident

once dilation develops (Fig. 3.4b).

The transition from contractive to dilative behavior

c marks a drastic change in the mechanism of pore water

pressure generation. The dilatancy tendencies are caused

by sand particles attempting to roll or slide up and over

one another, which under constant volume causes a reduc­

tion in the pore water pressure. The rolling and sliding

might even create some locally bigger voids near the

dilating particles (196). Upon reversal of stress, most

of these unstable holes would readily collapse producing a

corresponding high increases in pore water pressure (Fig.

3.4b). This may be associated with the physical fact that

once unloading begins some of the grains of the dilated

skeleton may actually start a downward motion under the

action of the normal forces, which leads to an increase in

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
89

Small strain
c y
S tatically stable
collapses hole
large hole

L) C h a n g e s in initial fabric at small strains

Upwards motion Downwards motion


(.

b) C h a n g e s in fabric d u e to dilatancy at large strains

Figure 3.4 Changes in Fabric during Cyclic Loading


(Adapted from Ref. 196).

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
90

c pore water pressure (124).

the direction of shearing is reversed,


It is also very likely that as

the number of con­

tact points between neighboring grains is drastically

reduced. As a result of these mechanisms, the effective

stress path moves towards the zero effective stress state

during the unloading stage, that is, as the applied shear

stress is reduced to zero (Fig. 3.3b).

Once the effective stress becomes zero, the shear

strength of a cohesionless soil is theoretically zero.

However, the condition of 100% pore pressure buildup

develops at the instant the specimen is subjected to an

hydrostatic state of stress, and therefore it does not

have shear stresses to resist. Nevertheless, the impor­

c tant consequence of the momentary occurrence of liquefac­

tion is the deformations which develop upon application of

a small shear stress in the opposite direction of loading.

For very loose sand almost unlimited deformations develop

under essentially zero shear stresses. Accordingly, the

specimen is said to have liquefied, since it exhibits no

resistance to deformation over a wide shear strain range

(74,102,148).

Nevertheless, due to the process of continuous defor­

mation it is very unlikely that the sand specimen can

remain indefinitely in a "liquefied" state, that is, with

a minimum number of contacts between neighboring grains.

On the contrary, after some straining, new contact points

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
91

might develop allowing dilation of the sand skeleton since

the effective confining pressure is extremely low. Conse­

quently, the process of continuous deformation may be

arrested after a finite strain by dilatancy caused pore

water pressure reduction and the system reverts to a

"solid" state (155,196). It is clear that the strain

range within which the specimen remains in a "liquefied"

state depends on the relative density of the sand. If

cyclic loading continues, the specimen undergoes repeated

cycles of liquefaction, finite continuous deformation and

stiffening or solidification due to dilation, with the

stress path climbing upwards and downwards along the

failure line during the loading stages and approaching the

c origin during the unloading phases

resistance envelope has been found to correspond to the


(Fig. 3.3b). The

failure line determined after large strains in undrained

tests under monotonic loading (82,98,141,157). It is

important to note that the condition of 100% pore pressure

buildup (Fig. 3.2) corresponds to the moment the stress

path passes through the hydrostatic state (Fig. 3.3),

since the failure lines for a cohesionless material extend

from the origin. Consequently, liquefaction only occurs

at the instant the major and minor principal stresses

become equal (o,=o,).


1 J

As the relative density increases, the sand specimen

behaves strongly dilative even at relatively high levels

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
92

of confining pressure (Fig. 3.3b). Consequently, the pore

water pressure accumulation during cyclic loading results

from the alternating cycles of dilation upon shearing in

either direction (rolling and sliding of grains on top of

each other) and contraction upon reversal (loss of con­

tacts and downwards motion ) (Fig. 3.4b), which leads to a

progressive increase in pore water pressure. It is well

known that even dense sands are slightly contractive when

subjected to small shear strain amplitudes. The contrac­

tive behavior is enhanced by the physical fact a reversal

in the direction of shearing occurs. Accordingly, each

time the specimen is close to a hydrostatic state of

stress, it develops a contractive response over a small

range of deviator stress and a pore water pressure incre­


c ment is induced (13); eventually a zero effective stress

condition develops.

3.1.3 Liquefaction, Flow Deformation and Strain Development

The development of a zero effective stress condition

for the first time is usually called "initial liquefac­

tion" (25,74,145), which in very loose sand is associated

with the development of very large deformations. However,

in dense specimens, minor shear strains may be sufficient

for "rebuilding" most of the grain contacts. Conse­

quently, dilation develops and the corresponding reduction

in pore water pressure allows the specimen to resist the

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
93

applied stresses without a significant deformation,

irrespective of the number of cycles after "initial

liquefaction." Accordingly, "liquefaction" of dense sands

is only a momentary state which does not imply any insta­

bility of the specimen. The associated deformations

decrease as the relative density increases, becoming

insignificant at very high Dr, as illustrated schemati­

cally in Figure 3.5. This curve conceptually represents

the strain potential following the onset of liquefaction.

It should be noted that the limiting shear strains are

likely to be a function of the type of sand and of the

level of confining stresses.

Some researchers argue that the development of a

momentary zero effective stress condition in a specimen of


c
dense sand, which implies a finite strain potential,

should be termed "cyclic mobility" (13,15,17,132) to dis­

tinguish it from the behavior of loose specimens. In the

author's view, the structure of loose sands is metastable,

that is, susceptible to collapse. In a collapsive skele­

ton, small shear strains may be sufficient to produce a

sudden rearrangement of grains and loss of contact points

between neighboring grains, thereby imparting the brittle

behavior characteristic of loose sands (9,14,55). Upon

collapse of the structure, the load is suddenly

transferred from the sand skeleton to the water, resulting

in a sharp increase in pore water pressure. As a result

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
0 20 40 60 80 100

R E L A T I V E D E N S I T Y (%)

Figure 3 .5 Shear Strain Potential at Liquefaction.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
95

, of the two mechanisms, namely, loss of contacts and

increases of pore water pressure, the shear strength is

reduced substantially and the specimen undergoes large

deformations in a very short period of time. Reduction in

the rate of loading does not avoid the phenomenon (110).

Due to the process of deformation, the sand grains become

oriented and reach a statistically steady-state condition

and after all particle breakage, if any, is completed, the

shear stress needed to continue deformation eventually

reaches a very low but nevertheless constant value (131).

This type of deformation will be termed "flow deformation"

or "strain softening behavior" throughout this disserta­

tion.

It has been shown that the "sensitive" response of

loose sands and the associated strength loss due to sudden

breakdown of the structure, can develop at any loading

stage (increase of shear stress) in the course of cyclic

loading as a result of pore water pressure buildup and

consequent strain development (14,34,184,186). Conse­

quently, during cyclic loading of saturated loose sands,

the level of deformation after a sufficient number of

cyclic stress applications could be due to "flow deforma­

tion" or to a combination of limited flow deformation fol­

lowed by liquefaction. In other words, in the author's

view, the structure of a loose sand may collapse (sudden

rearrangement of grains) either because the applied shear

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
stresses are too high or as a result of a sudden reversal

in the direction of shearing under low levels of contact

pressures. Under undrained conditions, both types of

"collapse" result in a significant increase in pore pres­

sure and a consequent reduction in strength. The first

collapse mechanism leads to flow deformation whereas the

second one results in liquefaction, as defined herein. It

will be shown later that the accumulation of strains due

to flow deformation occurs always during a loading stage

before the development of momentary states of zero effec­

tive stress. A stage of limited flow deformation is

observed in Figure 3.3a, where the stress path quickly

approaches the failure line. At this stage a reduction in

strength is clearly observed. It is important to note

that strains due to limited flow deformation may be as

significant as the strains which develop upon occurrence

of limited liquefaction. The conditions leading to strain

softening behavior of loose sand and the factors control­

ling the steady-state strength of sands are discussed in

the next section together with its relevance in under­

standing the response of saturated sands under cyclic

lo ad ing .

Experimental evidence indicates that for loose sands,

the relationship between the normalized pore pressure,

u/o , and the cycle ratio, N/NT , in which N_ is the number


C L u
of cycles required to cause liquefaction, will correspond

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
97

approximately to the curve shown in Figure 3.6 (147).

Three stages in the process of pore water pressure genera­

tion are clearly evident. The pore pressure generated

during the first cycles is significant but the rate of

pore pressure buildup per cycle becomes smaller as loading

proceeds. However, the rate of pore pressure buildup

increases again during the last cycles prior to the

development of liquefaction. It is likely that the pore

pressure buildup during the initial cycles reflects the

collapse of unstable particle groups of the Initial

fabric. The elimination of local instabilities at the

contact points produces a more stable structure, which

results in a reduction in the rate of pore pressure

c buildup. However,

deformation develops;
this rate increases again once flow

this rate is further increased when

the specimen becomes dilative upon loading and contractive

upon unloading, which precedes the occurrence of liquefac­

tion. This behavior is further illustrated with the test

results presented in Chapter 5. Also shown in Figure 3.6

is the normalized pore pressure curve typical of very

dense specimens. It is seen that in this case, the rate

of pore pressure buildup is relatively uniform with

increasing number of cycles until liquefaction develops.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
98

I Initial t e n d e n c y to contract
II M o r e stable fabric d u e to prior cycles
ill F l o w deformation a n d contraction u p o n u n lo a d i n g
j. |--- +--------------- II T ---- HI---- 1

1.0

0.8--
Denser

0.6-
RATIO,
PRESSURE

0 .4 -
Looser

0.2 -
PORE

0.2 0.4 0.6 0.8 1.0

C Y C L E RATIO, N / N L

Figure 3.6 Rate of Pore Pressure Buildup during Cyclic


Loading.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
99

3.1.4 Threshold Shear Strain Concept


c
There is yet another aspect related to pore pressure

buildup and strain development during cyclic loading.

There is experimental evidence which indiated the

existence of a threshold shear strain for sands

(24,32,33,43). If cyclic strains below this value are

induced in the sand deposit, it is considered that there

is neither densification of dry sands nor pore pressure

buildup in saturated, sands. Investigations have been

carried out to study the effects of initial state of

stress and of relative density on the threshold shear

strain, y , by means of strain controlled cycled triaxial

tests (43). In these tests, the threshold shear strain

c was determined by subjecting sand specimens to short

sequences of sinusoidal axial strain cycles at increasing

amplitude until a residual pore water pressure gerater

than zero was measured. The results of these tests seem

to suggest that the threshold shear strain is a constant

characteristic of each type of sand for a wide range of

relative densities, as illustrated in Figure 3.7 for Mon­

terey No. 0 sand. The value of the threshold shear strain

for NC sands has been found to be of the order of 0.8 to


—2
1.5 x 10 %., with a trend for y to decrease slightly as

the grain size decreases (43). For two different sand

types, Monterey No. 0 ar»d banding sand, which is finer

than Monterey No. 0, the value of y£ was unaffected by the

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
Reproduced with permission of the copyright owner. Further reproduction r

0.06
Monterey No. O Sand
Cyclic Triaxial Tests
0.05 ^ ' - 2 0 0 0 psf
n - 10 cycles

0.04
o O D r- 45%
H- □ D r - 60%
< A D r - 80%
0C 0.03
Ui
DC
Z>
</) 0.02
<n Threshold shear strain
Hi
oc
prohibited without perm ission.

Ql
ui
0.01
DC
o
£L
ooo
»-3

C Y C L I C S H E A R S T RA I N , / (%)

Figure 3.7 Concept of Threshold Shear Strain (After Ref. 33).

100
c consolidation stress ratio, K£ *

range from 1 to 2 (43).


w it ^ Kc in the

In tests on overconsolidated

specimens of Monterey No. 0 sand, at a relative density of

60%, the threshold shear strain was observed to increase

with the overconsolidation ratio, OCR, from


_2
Y = 1.1 x 10 % for NC specimens to a value close to 3 x

10_2 % for an OCR of 8 (43).

Consequently, it has been suggested that for some

practical applications, an engineering value for Yt of

about 0.001% to 0.015% may be used as a threshold shear

strain level marking the beginning of significant pore

water pressure increases (25,32,43). This value has been

found to correspond to the estimated average shear strain

c induced in sand deposits by earthquakes of moderate magni­

tude, which have been observed to cause negligible pore

water pressure buildup (147). Dobry et a l . (32) have pro­

posed a method for determining the threshold ground

acceleration, at , for a sand layer in situ. at

corresponds to the ground acceleration which just gen­

erates an excess pore water pressure in the layer. The

method is based on the threshold shear strain concept and

on the field measurement of the shear modulus of sands at

small strains, Gm a x » using geophysical techniques. If the

peak acceleration is greater than a£ , the actual

occurrence of liquefaction and its effects on overlying

structures will depend on the factors discussed above such

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
102

as duration of the earthquake, relative density, etc., and

the drainage boundaries of the sand layer.

An alternative approach for evaluating threshold

shear strains is by observing the shape of the effective

stress path from tests that include large amplitude shear

strain. In this connection, Towhata and Ishihara (181)

observed that in cyclic tests involving just triaxial

compression cycles, the stress path corresponding to the

initial stage of the first loading was almost parallel in

direction to the stress path observed at the later stage

of cyclic loading, where it was producing just recoverable

deformation. Accordingly, they concluded that below a

certain threshold shear stress level the pore pressure is


r recoverable on unloading and that a truly residual pore

pressure can only be produced when the applied shear

stress exceeds this limit. The concept of a critical

shear stress, below which the pore pressure does not build

up, had been previously developed by Shibata et a l . (157)

and Yoshimi and Oh-Oka (193). It should be noted that

while, the threshold shear stress may vary over a wide

range depending on factors such as type of sand, density,

confining stress, etc., the corresponding threshold shear

strain appears to vary over a relatively small range,

irrespective of density, fabric or confining pressure

(24) .

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
103

3.2 Evaluation of Liquefaction Resistance

In addition to the potential tendency towards densif-

ication, the occurrence of liquefaction will depend on the

amplitude and number of cycles in the series of stress

applications. Conse que nt ly, laboratory measured liquefac­

tion resistance, or cyclic undrained strength, is usually

defined as the combination of the cyclic shear stress and

the corresponding number of cycles causing initial

liquefaction or a given level of shear strain, as dis­

cussed below.

3.2.1 Cyclic Undrained Strength

c It is common practice to subject "identical" speci­

mens of a given sand with the same initial state (relative

density and confining pressure to cyclic loading at dif­

ferent levels of shear stress amplitude. From the time

histories of shear stress, shear strain and pore water

pressure, it is possible to plot for each specimen the

applied shear stress ratio (ia/2oc in the triaxial test,

or t / c>c in the torsional shear test) versus the single or

double amplitude shear strain with increasing number of

cycles. By this procedure a plot like the one shown in

Figure 3.8 can be obtained, in which the data points

corresponding to the same number of cycles have been con­

nected with a solid line. These curves constitute a kind

of stress versus peak strain curves under cyclic loading

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
Reproduced with permission of the copyright owner. Further reproduction

1.2

Number of Dr - 77%
cycles- 5
1.0

7O 10,
P 0.8
O 20
h-
<
30
cc 0.6
<o
CO
UJ
cc
H 0.4
CO Fuji river sand
o Torsional test
prohibited without perm ission.

CTe- 98 kPa
O
>- 0.2
o

0 1 2 3 4 5 6 7 8 9 10

S H E A R S T RAIN, / (%)

F i g ur e 3.8 C y c l i c Stress Ra ti o - Shear S tr ai n ( Si ngle A m pli tu de )


R e l a t i o n s h i p for D i f f e r e n t N u m b e r of Cyc l e s at a C on s t a n t
Relative Density (After Ref. 74).
105

o conditions. It is interesting to observe that for each

curve there is an initial convex part in the range of

small strains, which is followed by a concave portion with

increasing shear strain amplitude (74). This change in

curvature seems to suggest that at this level of relative

density (77%) the curves for Fuji river sand would tend to

reach a limiting shear strain level of increasing magni­

tude for each N value, irrespective of the applied shear

stress amplitude below that which might cause cavitation

in the pore water.

From the shear stress ratio versus peak shear strain

amplitude plots, the relationship between the cyclic shear

stress ratio and the number of cycles causing initial

liquefaction or a predetermined level of single or double


(
shear strain amplitude can be readily obtained. A typical

plot of this kind is presented in Figure 3.9. This type

of plot has been consistently used to represent the

liquefaction or cyclic undrained strength characteristics

of a particular sand at a given relative density.

In analyzing any given situation, the number of sig­

nificant stress cycles in the design earthquake or other

cyclic disturbance to which the sand deposit might be sub­

jected can be estimated. For strong earthquakes the sig­

nificant number of cycles is likely to lie in the range of

about 10 to 30 cycles (147). Consequently, it is common

practice to define the resistance to liquefaction for a

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
106

5%

0.3

10%
5%
Dr- 90%

25
>*
o 20 %
0.2f
Dr- 82%

X
c CO
Ui

0.1
LU
CC

Large Scale Simple Shear Tests


Monterey N0 0 Sand j
O ’c - 5 5 kPa

100

NUMBER OF CYCLES

Figure 3.9 Relationship between Cyclic Stress Ratio and


Number of Cycles Causing Different Level of
Shear Strain (After Ref. 30).

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
107

particular number of stress applications. Accordingly,

the cyclic shear stress ratio versus peak shear strain

amplitude relationship for a fixed number of cycles, can

be compared at different relative densities on the same

plot, as illustrated in Figure 3.10. It is again observed

that for high relative densities and for both types of

sand the curves in Figure 3.10 have a convex part in the

range of small strains, which is followed by a concave

portion which tends to reach a limiting shear strain

level. Figure 3.10 also compares stress— strain curves

from monotonic tests for Toyoura sand with corresponding

shear stress versus peak strain curves under cyclic load­

ing conditions. It is interesting to observe that at a

c relative density of about 84%, the shear strain induced by

a stress ratio of 0.6 increases from 1.5% in the monotonic

test to 7.5 under cyclic loading. The large increase in

deformations due to the development of excess pore pres­

sures in the course of cyclic stress applications is

clearly evident.

An alternative method of representing the liquefac­

tion resistance of sands for a given number of cycles is

by plotting the values of cyclic stress ratio causing ini­

tial liquefaction, or different levels of shear strain,

versus the relative density, as illustrated in Figure

3.11. When the cyclic undrained strength is presented in

this way the following features are generally observed.

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
108

o 1.0
/Torsional Shear Test
0.9 (a) • Toyoura Sand
CO A ir-P lu viated
L? £ 0.8 -I- & c “ 9 8 kPa
fe ta tic Undrained Test ( D r - 84 % )

Dr-9 0 %
a w 0.6 x 85 % P

Cyclic
CO L ° 0.4
0.3
o o
o < 0.2, ___
S ta tic Undrained Test D r - 4 0 % / ■

7.5
S I N G L E A M P L I T U D E S H E A R STRAIN, (FOR C Y C L I C TEST)

c S H E A R S T R A I N , Y(%) ( F O R S T A T I C T E S T )
Strain Range At
>. D r - 88% Observed Initial Number of
Liquefaction cy c le s - 10 i
O
H—
<
cc
CO
CO
UJ
tr
i— Fuji river sand
CO
Torsional test
o 47% ^ - 9 8 kPa
—I
o
> I
o
0 1 2 3 4 5 6 7 8 9 10
S I N G L E A M P L I T U D E S H E A R S T R A I N , Y (%)
Figure 3.10 Cyclic Stress Ratio - Shear Strain Relation­
ship for Different Relative Densities at a
Constant Number of Cycles (After Refs. 74,
173) .

R eproduced w ith perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
109

o 1
CYCLIC TORSIONAL
0 .7 ' SHEAR TEST
TOYOURA SAND
i

(a)
“ I--------------

0.6 Min r u w v
* c - 9 8 k Pa
b° 0 .5 - NC- 20
/ /
DA SHEAR S T R A IN -1 5 %
- J.4
7 .5 %

< 0 .3 -
0C
CO
CO v 1
UJ 0.2 \3 %
tr 1
h
CO 0.1 ! Dr critical -
(b)
0.0 I i i :/ ,
30 40 50 60 70 80 90

R E LA TIV E D E N S IT Y , Dr (% )

1.4 Number of
c y c le s - 20 (b)

r
v 1.2
>.

O
I- SA SHEAR S T R A IN - 6% 5 % 3%
<
cc 0.8
CO
CO
UJ F u ji-riv e r sand
0.6 - Torsional test 1 -
CO
0 c - 9 8 kPa
o
_1 0 .4
o
>
o
0.2 -

Dr critical

0 50 100
R E LA TIV E D E N S IT Y , Dr (% )
Figure 3.11 Cyclic Stress Ratio - Relative Density
Relationship for Different Shear Strain
Levels at a Constant Number of Cycles
(After Refs. 74,173).

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
The cyclic undrained strength increases at an increasing

rate as the relative density increases, irrespective of

the failure criterion, e.g. initial liquefaction or strain

level. However, there seems to be a particular value of

relative density, termed "critical relative density,"

below which only a small increase is noted in cyclic

resistance with increase in Dr (173). On the other hand,

if the relative density is greater than the critical, the

dynamic resistance builds up at a much faster rate with

increasing relative density. Consequently, it cannot be

said that the cyclic undrained of sands is proportional to

relative density (173). It has been also observed that

below the critical Dr, a small increase in shear stress

ratio above that required to cause initial liquefaction

results in the development of large strain amplitudes. On

the contrary, at relative densities above the critical,

initial liquefaction occurs at a small strain amplitude.

A large increase in the cyclic stress, above that causing

initial liquefaction, is required to produce large strain

amplitudes (30,184).

Vaid et a l . (184) have suggested that below the Dr

critical, sand are susceptible to flow deformation, which

might be responsible for the sudden development of large

strains with a minor increase in stress ratio (see Fig.

3.10) and, consequently, only a small increase in cyclic

strength is noted with increase in Dr. On the other hand,

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
111

at relative densities greater than the critical, flow

deformation does not develop. Accordingly, it is likely

that the large resistance of dense specimens against

cyclic undrained loading at larger shear strain amplitudes

is due to their stronger dilative tendencies (173,184).

The critical relative density, as defined, is a function

of at least the type of sand, the confining stress level,

the number of cycles and the stress-strain boundary condi­

tions. An important inference from the concept of criti­

cal relative density is that the beneficial effects of

densification in increasing the shear stress required to

induce large strain amplitudes only develops above the

critical relative density.

c 3.2.2 Cyclic Shear Strain Potential

It is noted that once the critical value of Dr is

exceeded, all the curves in Figure 3.11 are very steep

suggesting that a limiting shear strain will develop in a

given number of cycles, irrespective of the level of

cyclic stress ratio that is applied. Consequently, a

curve of limiting shear strain versus relative density

might be established, as presented by De Alba et a l . (30).

The form of this relationship, which herein will be termed

"cyclic strain potential," is presented in Figure 3.12.

The data were obtained from large-scale simple shear tests

on freshly deposited sand in the laboratory. It is seen

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
112

40
o Monterey No 0 Sand
O q - 5 5 kPa
Nc-1.0

p <10 -

20 40 60 100
0.6
15% 5%

0.5

TO
0.4

I—

c 0.3

I—
Monterey No o sand
O 0.2
C7 ‘c - 55 kPa
Nc= 1 0

Initial Liquefaction
0.1

0 20 40 60 80 1 00

RELATIVE DENSITY- %

Figure 3.12 Cyclic Stress Ratio - Relative Density Rela­


tionship and Cyclic Shear Strain Potential
from Large-Scale Simple Shear Tests
(After Ref. 30).

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
113

that for this particular sand at relative densities less

than about 45%, the application of cyclic stress ratios

sufficiently high to cause initial liquefaction also

causes extremely high shear strains in the sand. However,

for relative densities greater than 45%, the development

of initial liquefaction results in only a limited amount

of shear strain, the corresponding shear strain potential

decreasing with increasing relative density. Figures 3.12

confirms the general trend of the shear strain potential

postulated in Figure 3.5. As explained earlier, dense

sands only remain momentarily in a liquefied state, which

in general does not imply instability.

Based on data such as those shown in Figure 3.12,

c Seed (145) proposed a tentative range of limiting shear

strain for natural deposits as a function of normalized

standard penetration resistance C1^ ) ^ , which is shown in

the upper part of Figure 3.13. ^n i^6o iS de f±ned as the

standard penetration resistance determined in SPT tests

where the driving energy in the drill rods is 60% of the

theoretical free-fall energy (153). The limiting strains

measured in tests on undisturbed sand specimens obtained

by freezing techniques have been related to the in situ

^N 1^60 v a ^ues ky Tokimatsu et a l . (180) and the results

are superimposed on the original estimated range in Figure

3.13. According to Seed et al. (153), the generally good

agreement suggests that the indicated values of limited

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
114

w 2
c UJ 30-
< O
uj oc Estimated range (1 384)
X W __
co o. 20-
CD I Proposed Based on test data by
36 by Seed (1979)
I 1» Tokimatsu and Yoshimt
oc (1984)
. F*
03 OL
— High damage potential-jintermediate!—No significant d a m ag e-
0.6
Liquefaction with
~ 20% * 1 0 % =3%

No Liquefaction

(N.,) 6 0

Figure 3.13 Tentative Relationship between Cyclic Stress


Ratio, Values, and Limiting Strain
for Natural Deposits of Clean Sand (After
Ref. 153).

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
115

shear strains are of the right order of magnitude.

In summary, for a given cyclic loading condition,

medium dense to dense sand specimens are able to withstand

any number of stress repetitions without exceeding a lim­

iting maximum strain. This is the significance of the

cyclic strain potential presented in Figures 3.5, 3.12,

and 3.13. Sands at high relative densities would have

very low strain potential under cyclic loading conditions

at moderate confining pressures, even if they develop 100%

pore pressure buildup during unloading stages. It is

important to note that the results in Figures 3.12 and

3.13 do not include the effects of initial sustained shear

stresses. Accordingly, their applicability is restricted

to cases in which these effects can be considered negligi­

ble, as discussed in a later section. The effects of con­

fining pressure on the limiting shear strain are yet to be

determined.

The shear strains resulting from cyclic loading may

be beyond what is considered to be acceptable for a struc­

ture. Consequently, a failure criterion should be esta­

blished for any given situation. For relative loose sands

a single definition of failure will usually suffice

because initial liquefaction and large deformations

develop almost simultaneously. For denser sands, however,

the relationship between the magnitude of the cyclic

stress ratio and the number of cycles required to cause

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
"failure," that is a predetermined level of strain, will

vary considerably depending on the failure criterion

adopted. Ishihara (74) has recently called attention to

the fact that the magnitude of the peak shear strain at

which 100% pore pressure buildup takes place may lie

within a relatively narrow range from 2.5 to 3.5% (in

terms of single amplitude), as indicated in Figure 3.10b.

Consequently, Ishihara recommends the use of the develop­

ment of a peak shear strain of about 3%, which corresponds

to the development of initial liquefaction regardless of

the Dr, as a liquefaction criterion to consistently define

cyclic strength characteristics of sands at any density

from loose to dense states. In the case of loose sands,

intolerable large deformations are produced immediately

after the inducement of 3% cyclic shear strain accompany­

ing the onset of liquefaction. In the author's opinion,

this criterion may be valid only if the relative density

is above the critical value or when the cyclic shear

stress amplitude is smaller than the steady-state

strength. As discussed later on, a process of flow defor­

mati on in loose sand can be initiated at much smaller

shear strains.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
, 3.2.3 Volume Decrease Potential and
C Liquefaction Resistance

It is well known that different sands at identical

relative densities may not have the same drained or

undrained stress-deformation characteristics

(94,118,123,125). Accordingly, it is recognized that

relative density is neither a unique index to represent

the effect of density of different kinds of sand on the

cyclic undrained strength, especially when the mineral

composition, size distribution and shape of the particles

are sharply different. As discussed, the potential for

liquefaction is associated with the tendency towards den-

sification under a series of stress applications in

f undrained conditions. Relative density indicates how the


V
current packing of a sand, in terms of void ratio, com­

pares with its loosest and densest states, but it does not

give any indication of the actual volume changes between

the limiting states. Consequently, it has been suggested

(87) that liquefaction potential of different kinds of

sands should be more closely related to a parameter indi­

cative of actual volume decrease potential, which under

cyclic loading conditions is given by

VD " e - *«!« ( 3 -U

where, VD , the volume decrease potential, is the differ­

ence between the in situ void ratio, e, and the void ratio

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
118

corresponding to the densest state, em in * Thus,


c
(3.2)

The difference in VD , and hence in liquefaction

potential at the same Dr, arises from the differences in

composition, size distribution and shape of the particles

of different sands, as evaluated by the term

^emax ~ emin^* m a 7 tend to decrease as the average


sand grain increases (87).

Comprehensive studies by Ishihara et a l . (87) by

means of cyclic triaxial tests, have shown that there is

indeed a good correlation between the stress ratio causing

initial liquefaction and , as illustrated in Figure

3.14, in spite of the wide range in size distribution,

shape of the particles, origin and relative density of the

materials used in the investigation. Consequently, the

volume decrease potential, as defined, seems to be a more

convenient parameter for comparing liquefaction potential

of different kinds of sands at the same o . This is sup-


c
ported by the physical fact that during cyclic loading the

volume of the sand skeleton will tend to decrease from its

present state of packing to the densest state. Accord­

ingly, the higher Vp , the more likely the sand is to

liquefy under any given loading conditions. Unfor­

tunately, the use of relative density is so extensive

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
119

o 100

natural
80 11
sieved
cc :soil No.
Ui
2 60 - in figure
u_ below
»-
uj 40 -
O
cc
UJ
o. 20 -

0.01 0.1 1.0


P A R T I C L E SIZE (mm)

-- 1---r~ ~\
1-- 1---1 r "i r
0.6 Number of cycles
o Glass beads (D
: 20, D r - 4 0 - 7 6 % • Fuji river sand <2>
( 0.5
c r ^ - lO O kPa
a Toyoura sand
a Niigata sand
®
0
<=>Chiba sand (g)
0.4 ■ Shirasu (g)
v Shiriuchi sand (2)

0.3
b
CM

S ’ ° '2

0.1
Cyclic Triaxial Tests

0.1 0.2 0.3 0.4


V O L U M E D E C R E A S E P O T E N T I A L , e - e mjn

Figure 3.14 Relationship between Volume Decrease Potential


and Cyclic Stress Ratio Causing Initial
Liquefaction in 20 Cycles (After Ref. 87).

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
120

/' ■ among researchers and practitioners that its replacement,


V_/j
even for a more convenient parameter, seems to be unlikely

in the very near future.

Nevertheless, it should be noted that may be indi­

cative of the tendency to contract and hence, of the

liquefaction potential because of continuous reversals in

the direction of shearing,- but it may not be indicative of

the tendency to flow deformation under monotonic loading.

First of all, the volume decrease potential during mono­

tonic loading is determined by the critical void ratio

line as discussed in the next section. On the other hand,

a sand with rounded grains, which is highly susceptible to

collapse under increasing shear stresses, is likely to

c have a smaller VD than an angular sand at the same Dr.

summary, according to Figure 3.14, a sand with a low


In

has a higher liquefaction resistance. However, that sand

may be highly susceptible to flow deformation or strain

softening behavior on loading, as discussed later on.

3.2.4 Effect of Initial Confining Stress

There is yet another factor which itself strongly

influences the cyclic undrained strength of sands; namely,

the magnitude of the initial confining pressure. The mag­

nitude of the shear stress level or the number of cycles

(or both) needed to cause liquefaction increases with

increasing confining pressure. This can be associated in

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
121

the first place with the fact that the higher the confin­

ing pressure, the larger the distance the effective stress

path (Fig. 3.3) needs to traverse in order to approach the

failure line and finally the zero effective stress condi­

tion. Moreover, for a given volume decrease potential the

forces resisting slippage at the grain contacts increase

with the confining pressure. However, the cyclic shear

stress which produces liquefaction in a given number of

cycles in general increases with increasing confining

pressure at a rate that is less than linear (17,103).

This behavior may be attributed to the increased contrac­

tiveness of sands at higher confining pressures. Accord­

ingly, although the cyclic shear stress level increases

with confining pressure, the cyclic stress ratio decreases

with increasing confining pressure ( 17 ,186,19A) , as illus­

trated in Figure 3.15. Thus, in selecting cyclic shear

strengths, due cognizance must be taken of the in situ

confining stresses. The level of confining pressure is

likely to affect properties such as the critical relative

density and the limiting shear strains.

3.3 Monotonic and Cyclic Strength Relationship

3.3.1 Drained Behavior under Monotonic Loading

The macroscopic deformation of a specimen of cohe-

sionless soil under drained shear is the result of

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
Reproduced with permission of the copyright owner. Further reproduction
c

Cyclic triaxlal tests

Tailings sand
0.4
Dr“ 70%

Axial strain, G fl= 2.5%

o
0.3
CM

>*
C q =2.0 kg/cm 1

0.2
prohibited without perm ission.

16.0 kg/cm2

0.1

± .L
0 10 20 50 100

N U M B E R O F CYCLES, N

12 2
Figure 3.15 E f f e c t o f C o n f i n i n g P r e s s u r e on C y c l i c S t r e s s R a t i o Ve r s us
Number o f C y c l e s R e l a t i o n s h i p ( D a t a from R e f . 1 8 6 ) .
123

microscopic motion (rolling and sliding) of the grains

relative to each other at the contact points, producing

rearrangement of grains and corresponding changes in volume.

There may be some amount of particle crushing during the

process of deformation, especially breakage of sharp contact

points at higher confining pressures. Accordingly, during

drained monotonic shear tests, loose sands decrease in

volume, whereas dense sands increase in volume, that is,

they dilate. Nevertheless, due to the process of continu­

ous deformation, the sand particles in either case eventu­

ally reach a statistically steady-state condition, at

which the volume and the shear stress needed to continue

deformation remain constant. The void ratio at this state

is termed "critical void ratio." It is clear that after

large strains, the original structure of sand is drasti­

cally altered. A "statistical fabric" is created, which

depends mainly on the mineral composition, size distribu­

tion and shape of the particles, and on the level of

effective normal stress. The steady-state fabric exhibits

a constant resistance to continuous deformation.

The condition at which a sand specimen is continu­

ously deforming at constant volume, constant effective

normal stress, constant shear stress and constant velocity

has also been termed "critical state" (137,144) or

"steady-state of deformation" (131). However, in the

author's opinion, it is not yet known whether the rate of

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
124

deformation at the critical state is really constant

(steady). Since dilation tendencies are smaller at higher

confining pressures, the critical void ratio for a sand

decreases as the confining pressure increases. Moreover,

for each type of sand, the relationship between void ratio

and confining pressure at the critical state is unique,

irrespective of the initial density of the specimen. Con­

sequently, under drained shearing a loose sand contracts

and eventually approaches a steady-state volume. At the

same confining pressure a dense specimen of the same sand

dilates until it also reaches the same steady-state condi­

tion (in terms of void ratio and shear resistance) as the

loose sand. The relationship between void ratio and con­

fining stress at the critical state is represented by the

steady-state or critical void ratio line.

Consequently, the drained behavior of a sand is con­

trolled in the first place by its initial state (in terms

of void ratio and level of confining stress) with respect

to the critical void ratio line. The void ratio versus

confining stress space is usually termed the state

diagram, in which the critical void ratio line separates

contractive from dilative states, as shown in Figure 3.16

for Sacramento River Sand (105). If a sand specimen with

initial state at point A is sheared under drained condi­

tions, the void ratio would decrease at constant effective

confining stress until the critical void ratio line is

reached at point A".

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
125

o
1.0
3ontractiv« ■

0.9
Critical vo d ratio iinc

0.8
Con so idation stsie
Q>
RATIO,

0.7
VOID

0.6

0.5
Dilative

0.4
0 5 10 15 20 25 30

EFFECTIVE CONFINING PRESSURE, (Kg/cm2 )

Figure 3.16 State Diagram for Sacramento River Sand


(After Ref. 105).

R eproduced w ith perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
126

Test results have indicated that the drained behavior

of specimens with different initial states may be similar

if the initial conditions are such that there is an equal

proximity to the steady-state line (8,165). Consequently,

a state parameter for sands has been proposed (8), which

is defined as the void ratio difference between the ini­

tial state and the steady-state condition at the same con­

fining stress. For example, Figure 3.17 shows the drained

behavior of Kogyuk sand (8), corresponding to a wide range


*
of combinations of e and , mainly in the dilative side,

as a function of the state parameter, ij/, which in this

case was defined with respect to the steady-state line

obtained from undrained tests (8). The peak angle of

shearing resistance and dilation rate at the peak stress

c ratio for Kogyuk sand, are reasonably well-defined func­

tions of the state parameter. In general, for specimens

with negative 'i' (initial dilative state), a clear peak is

observed in the stress-strain curve, which becomes less

marked with decreasing negative 4', until there is gen­

erally no peak for specimens with positive ij> (initial con­

tractive st ate ). The volumetric strain behavior may be

similarly dependent on state parameter. Strong dilation

will be apparent for high negative states with little or

negative dilation observed once the state parameter

becomes positive. The apparent dependence of sand charac­

teristics on the state parameter means that the mechanical

properties of sands are influenced by actual volume

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
127

o a>
©
42 Fine content (%)
"O
40-
UJ
o 38-
U_ Z
o g
UJ CO 36r ■ ■.
m 05
0 UJ
z cc 34
< CD
Q Z
32 ■
| <
< UJ
£
Q x
co
301-
28
-0.15 -0.10 -0 .0 5 0 0.05
STATE P A R A M E T E R ^
a) P E A K A N G L E O F S H E A R I N G R E S I S T A N C E

c ■cl.«
)
0.8r
Fine content (%)

5 0.6
tif
•o
UJ
)-
< 0 .4 r
tr
z
Q
h- 0.2
<
—i
o

- 0.10 -0.05
STATE P A R A M E T E R
b) D I LA T I O N R A T E A T P E A K S T R E S S R A T I O

Figure 3.17 Drained Properties as a Function of State


Parameter for Kogyuk 350 Sands (After Ref. 8)

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
128

changes during shearing and not by the absolute magnitude

of either void ratio or confining stress. It is clear

that under high confining pressure, dilation will begin at

a much lower void ratio. However, it is the magnitude of

dilation which seems to determine strength, not the void

ratio at which dilation occurs (8). Thus, the state

parameter reflects the volume change potential upon shear­

ing under the applied confining pressure.

The state parameter approach still has some limita­

tions. First of all, it does not include the effect of

interparticle fabric of the sand, which might be the cause

of some of the scatter in Figure 3.17. It is well known

that a given cohesionless soil may have different fabrics

at the same void ratio or Dr. The initial fabric might

have a considerable influence on the behavior of sands,

especially under undrained conditions. Consequently, an

additional fundamental parameter of state is necessary to

describe fabric of sands (8). This is a goal yet to be

achieved. Secondly, the state parameter, as defined, does

not account for the effect of initial static shear stress

on the subsequent drained and undrained response, as will

be illustrated later herein. Nevertheless, it is impor­

tant to emphasize that the critical void ratio line pro­

vides a reference state which apparently is characterized

by a unique fabric or particle arrangement. This refer­

ence state includes the combined effects of void ratio,

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
129

stress level, mineral composition, size distribution and

shape of the particles on the drained strength charac­

teristics of sands.

3.3.2 Undrained Behavior under Monotonic Loading

In an undrained test, volume change is not permitted.

Consequently, the specimen is forced to remain at a con­

stant (average) void ratio. However, the sand would still

try to reduce its volume, but since this is not possible,

it responds by transferring load from the grain structure

to the pore water. Accordingly, the path in the state

diagram traverses horizontally at a constant void ratio

until the change in effective confining stress and the

process of deformation brings the specimen into a steady-

state condition, such that the particular void ratio at

which the specimen is compelled to remain during shearing

becomes a critical void ratio (131). In principle, the

critical void ratio line from drained tests should be

reached at this stage, since no volume change tendency

after large strains would correspond to no changes in pore

water pressure, and hence a process of continuous deforma­

tion under constant shear stress Is expected to develop.

However, experimental evidence suggests that the pore

pressure response of sand specimens in undrained shear

might not be uniquely related to potential volume changes

during drained tests, as measured by the critical void

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
130

ratio line. Figure 3.18 shows the results of undrained

monotonic load— controlled triaxial tests on "Banding sand"

(14), a uniform, clean quartz sand with subrounded to

subangular grains, at three different relative densities,


2
but at the same consolidation confining stress (4 Kg/cm ).

The paths in the state diagram corresponding to these

three tests are shown in Figure 3.19, together with the

critical void ratio line obtained from drained tests on

the same sand. It is observed that virtually none of the

state paths under undrained shear ends at the critical

void ratio line.

Specimen A exhibits a strain softening type of

behavior. After the peak point of the stress-strain

c curve, which occurs at a small strain,

reduction in strength with increasing strain until the


there is a marked

stress stabilizes at a ultimate or residual strength. As

indicated earlier, this type of deformation is termed

"flow deformation." During the softening process the

specimen underwent very large deformations which were

observed to occur in a time period less than a second.

The specimen reached a "steady-state of deformation" at a

very small residual strength and very small effective con­

fining stress, which remained practically constant between

strains of 6 and 25%. At this strain the test was

automatically stopped. After large strains the original

structure of the sand has been drastically altered. The

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
131

o 8

^ c - 4 . 0 kg/cm '

CM
Eo 6

o> H Test; Drc - 47%


4 7 min. to £ - 1 5 %
?
CO 4
CO R Test; D fg* 4 4 %
UJ
cc 0 .4 sec. from
K Peak to £ - 1 8 %
co
cc
o
H 2
<
R Test; D rc * 3 0 %
>
UJ .0.2 sec. from Peak to £ - 2 5 %
o
0 .3 0 k g /c m 2

0
20

c r
AXIAL STRAIN, £ , %

Initial ^ “C“ 4 .0 kg/cm ' - 0.15 kg/cm


UJ
cc 4 — i— i— r ~ --- ~Y7Xl
3
CO
CO
UJ
cc
O.CM
UJ O 2
o ®
Q.
Q
UJ 3
o
3
3 0
0 5 10 15 20

AXIAL STRAIN, £ , %

Figure 3.18 Undrained Load Controlled Tests on Istropi-


cally Consolidated Specimens of Banding
Sand (After Refs. 13,14).

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
Reproduced with permission of the copyright owner. Further reproduction
c

0.8

S Line: Critical void ratio line from drained 20


axial compression trlaxial tests

RELATIVE DENSITY, Dr (%)


Steady-state deformation -25
0.75
Q> ■oA -30
O
I-
<
cc
Consolidation states -35
9 At end of test
O 40
> 0.7 ,S Line
prohibited without perm ission.

- 45

Limited steady-state -50


deformation

0.65 55

E F F E C T I V E M I N O R PRINCIPAL S T R E S S , CT' (kg/cm2 )

Figure 3.19 Undr ai ne d S t a t e P a t h s C o r r e s p o n d i n g t o T e s t s in Figure 3.18 u>


( D a t a from R e f . 1 3 , 1 4 ) . ro
133

steady— state of deformation is reached only after grain


G orientation and grain breakage, if any, are both com­

pleted. The particles in the shear zone are oriented as

much as they can be under the applied stress state, such

that, the specimen exhibits a minimum resistance to con­

tinuous shear deformation (131).

Specimen B, in Figure 3.18, exhibits also a strain

softening type of response but the reduction in strength

is less significant than for specimen A. For a limited

range of strains, the specimen deforms under a steady-

state condition (continuous deformation at constant shear

stress and constant pore water pressure). However,

further straining causes dilation and a corresponding

C regain in strength, i.e. a strain hardening response.

the contrary, specimen C exhibits only a strain hardening


On

type of response as dilation is soon initiated and the

strength increases with further straining. A steady-state

condition is not reached and dilation is still evident

after an axial strain of 14%. It is important to note

that the stress-strain response changes from strain

softening (specimen A) to strain hardening (specimen C) as

the relative density of the specimen increases from 30% to

47%.

In similar tests performed by Castro (14), it is the

author's interpretation that all the specimens with ini­

tial states above the S line in Figure 3.19 exhibited

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
134

strain softening type of response, reaching in all cases a

steady-state of deformation. The state conditions during

steady-state deformation of all of these specimens were

found to define an approximately unique line in the state

diagram, which is shown in Figure 3.20 (line F) irrespec­

tive of widely different initial states above the S line.

For specimens with initial states below the S line, the

behavior was observed to change gradually from limited

strain softening to strongly dilative as the initial state

moved further away from the S line. The strain hardening

behavior was clearly evident for initial states below the

F line, cases in which the undrained state path was

observed to move towards the right crossing the F line.

Apparently, the state paths corresponding to strongly

dilative specimens would have reached the critical void

ratio line eventually at strains larger than could be

applied in the triaxial apparatus.

Thus, the author concludes that the F line in Figure

3.20 represents the relationship between void ratio and

effective confining stress during steady-state of deforma­

tion for specimens with initial states above the S line.

According to Casagrande (13), the letter F stands for the

critical void ratio in which flow deformation following

the collapse of the structure develops. At the same void

ratio, the undrained steady-state strength is about four

times smaller than the corresponding value during drained

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
135

steady— state of deformation (Fig. 3.20). At this stage,


o the shear stress is related to the effective confining

stress by a function of the steady— state angle of fric­

tion, which (as will be shown later) corresponds to the

large strain drained angle of shearing resistance.

Although some experimental errors could have influenced

the large difference observed between the S and F lines,

the writer attributes part of it to the "collapsiveness"

or brittleness of the sand skeleton (2). In a brittle

skeleton, small shear strains can produce a sudden rear­

rangement of grains and loss of contact points between

neighboring grains. Upon collapse of the structure, load

is suddenly transferred from the soil skeleton to the

water and the pore water pressure increases rapidly in


o spite of the reduction in applied shear stress. Conse­

quently, the shear strength is substantially reduced and

the sand specimen undergoes large deformations in a very

short period of time, reaching a steady-state of deforma­

tion .

Thus, in undrained shear, the very high sudden change

in compressibility and the associated inertial effects at

high rates of deformation produce a drastic change in the

mechanism of pore pressure generation leading to a reduced

steady-state strength. Accordingly, in the author's opin­

ion, the pore pressure response of sands in undrained

shear is not uniquely related to the volume decrease

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
Q.
C
o
CD
Q.
with permission of the copyright owner. Further reproduction prohibited without perm ission.

10
0.8
15

S Line: Critical void ratio line from drained -20


axial compression triaxial tests

25
(D 0.75 F Line: Undrained steady-state line from stress Q
o controlled axial compression triaxial tests ..
I-
30 >-'
< S Line
cc CO
z
o +35 J
o i
> Undrained path UJ
Consolidation state 40 >
^ Z L
- —- _ £ Drained path
45 Ul
cc

F Line + 50

0.65 55
1 2 3 4 5 6 7 8 9 10

E F F E C T I V E M I N O R PRINCIPAL S T R E S S , (kg/cm2 )

Figure 3.20 Steady-State Di ag r a m s for Ba n d i n g Sand (Data from Ref. 14). u>
137

potential determined by the critical void ratio line. The


o pore pressure response is determined not only by the mag­

nitude of the potential volume changes but also by the

variation of the contractive tendencies with respect to

the level of shear strain and the consequent changes in

particle arrangement. The author's views are further sup­

ported by the fact that constant volume tests and

undrained tests may provide different results (139).

The collapsiveness of sand may be determined by a

combination of factors such as composition, size distribu­

tion and shape of the particles, relative density, state

of stress, rate of loading, and possible other factors

such as particle arrangement and orientation, stress-

c strain history, cementation,

example,
and geologic aging.

the influence of the initial state of stress on


For

the "collapsiveness" of banding sand can be illustrated by

comparing the behavior of two test specimens initially

consolidated to points A and B, as shown in Fig. 3.20. In

undrained shear, specimen A collapses completely and —

under controlled loading — will reach the F line at point

A . A steady-state strength corresponding to

observed. On the other hand, subject to the same loading

conditions, specimen B will start to collapse and move

towards the F line (B ). Depending upon the initial state

of stress (cr _) specimen B may, or may not, reach the F

line but in any case dilation is soon initiated (new

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
138

contact points develop) and point B moves back towards the

S line (B"). Thus, only a limited collapse accompanied by

a short-lived flow condition takes place. A steady-state

strength condition (in terms of continuing large strains

at constant shear stress) is not observed. Consequently,

if the initial state of the sand specimen (void ratio and

state of stress) is between the F and S lines, its collap­

siveness will be restricted and it may never reach the

steady-state line (line F) determined from initial states

that fall outside the S line.

Nevertheless, the key factor in the collapse mechan­

ism appears to be an abrupt, very rapid change in compres­

sibility with respect to shear strain, at small strain

amplitudes. The smoother, more rounded and finer the par­

ticles, the higher the collapse potential and the further

apart the F and S lines tend to be. Accordingly, sands

with these characteristics exhibit the highest potential

for "flow deformation" or strain softening behavior, all

other factors being equal. On the other hand, the collapse

of a sand with angular and compressible grains is likely

to be very small. In this case, continuous particle break­

age and hence gradual deformation may develop during

shearing without exhibiting a sudden rearrangement of

grains. Consequently, the drop in strength after the peak

strength, if any, and the corresponding rate of deforma­

tion following the "collapse" are considerably reduced.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
139

T h u s , If a sand is not inherently brittle the S and F

lines would tend to merge. It is clear that the shear

strength at the steady-state of deformation is affected by

the deformation characteristics of the initial structure

of the sand since pore pressure buildup and the resulting

strength drop are dependent upon it.

There is evidence that the pore water pressure gen­

erated, during undrained shear, and hence the undrained

strength, depends on the particular strain rate that is

applied (13,63,150). In load controlled tests such as

those performed by Castro (14), the maximum possible

strain rate is that corresponding to free fall velocity.

However, the test rates of deformation, although high,

were smaller than that corresponding to free fall and

therefore were controlled by the sand properties. On the

other hand, in strain controlled tests it is postulated

that the faster the strain rate the greater the generated

pore pressure until a limiting strain rate is reached

irrespective of how fast we try to increase it. This lim­

iting strain rate is solely a function of the soil charac­

teristics and is observed in load controlled tests subject

to free fall conditions. Consequently, for collapsing

soils the F and S lines will move farther and farther

apart as the applied strain rate is increased until the

limiting F line is reached as controlled by the soil pro­

perties. Thus, a unique steady-state line can be reached

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
140

only if the nature of the applied loads is such that the

initial buildup in pore water pressure is completely over-


l
ridden by the loading conditions. Otherwise, the unique

relationship between void ratio and effective confining

stress at the steady-state of deformation is a function of

the particular velocity of deformation. In an embankment

or slope, if large masses of soil collapse simultaneously,

the soil mass overlying the collapsed zone can move under

free fall. Under these conditions, the characteristics of

the collapsed soil will control the limiting F line as

determined by Castro's tests (14). For non-collapsing

soils the S and F lines tend to merge irrespective of the

strain rate.

Most experimental investigations on the steady-state

strength of sands have been performed by means of axial

compression triaxial tests (14,165,186), which implies

that the behavior has been investigated only along partic­

ular stress paths. However, there is ample experimental

evidence to show that the fabric developed under condi­

tions prevalent during deposition of most natural sand

deposits is highly anisotropic, which causes changes in

the undrained deformation-strength characteristics with

the direction of loading (55,118,119,169). Tests by

Reades (cited in Ref. 9) showed that the undrained

steady-state strength in axial extension for a Kq consoli­

dated specimen was less than half that for an isotropically

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
141

consolidated specimen, at the same void ratio after

consolidation. This was attributed to the 90° rotation of

the planes of principal stress. Accordingly, it is felt

that the steady-state strength of sands may not be

independent of inherent anisotropy and the nature of the

stress path with respect to the orientation of the bedding

planes. The effects of stress path and fabric anisotropy

on the "steady-state" strength have thus far been observed

only in strain controlled tests: it is not yet know

whether they have an influence on the limiting F line.

It Is important to observe that the steady-state line

is a locus of points at which the sand specimen deforms

continuously after reaching a steady-state condition

(131). If the initial state of a specimen is at any point

along the F line, the corresponding state path during

undrained shear will move towards the right in Figure

3.20, as the change in structure occurs, probably trying

to approach the S line. Thus, the initial structure and

state of stress are such that the specimen exhibits no

contractiveness but a strongly dilative behavior. Thus,

if the initial state is on the F line, the F line does not

represent a steady-state condition.

From the foregoing discussion, it can be concluded

that the significant engineering properties, which charac­

terize the behavior of "collapsing" saturated sands under

undrained monotonic loading are:

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
142

1. The peak shear strength, and the corresponding strain

level at which it develops. In this connection, it

is of fundamental interest to define the physical

factors and state conditions controlling the trigger­

ing of the collapse of the structure, that is, the

initiation of the strain softening or flow deforma­

tion response. As discussed later, this is espe­

cially relevant in the evaluation of the response

under cyclic loading.

2. The magnitude of the strength drop from the peak

value to the steady-state strength and the associated

deformations. Once the structure of the sand has

collapsed, the extent of the failure under the acting

shear stresses will depend only on the shear strength

of the sand at the steady-state condition.

These properties are discussed in detail below.

Figure 3.21 presents the peak deviator stress,

(a -a,), corresponding to the tests on Banding sand in


X
which complete or limited strain softening behavior was

observed. As expected, the peak strength increases with

increase in either relative density or consolidation pres-


2
sure. However, at low confining pressures (<1.0 Kg/cm )

the effect of Dr on peak strength is small; at high con-


2
fining pressures (>5 Kg/cm ), Dr has an important effect

on peak strength.

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
Reproduced with permission of the copyright owner. Further reproduction
c

V O I D RATIO, 6
0.8 0.76 0.72 0,70 0.68
%
o
D) Banding sand
*
^ Load controlled triaxial test
b
I Strain softening response: , CT c“ 10 Kg/cm 2

Limited strain
softening response: • a mr
co 5 '
CO
UJ
a:
H 4-.
CO 4
DC
O
£ 3 '
prohibited without perm ission.

fij
O 4.0

UJ
CL

20 40 50
R E L A T I V E DENS I T Y , Dr (%)

143
Figure 3.2 1 Peak Undr ai ne d Shear S t r e n g t h as a F u n c t i o n o f R e l a t i v e
Density ( D a t a from R e f . 1 4 ) .
144

The state parameter approach has been also used for


G evaluating undrained behavior of sands (8). When the peak

deviator stress is normalized with respect to the consoli­

dation confining pressure and plotted versus the state

parameter, i|>, the values fall within a relatively narrow

band as shown in Figure 3.22. For these specimens the

state parameter is defined as the difference between the

consolidation void ratio and the void ratio corresponding

to the undrained steady-state line (line F) at the same

consolidation pressure. The trend is clear with the nor­

malized peak strength decreasing as the state parameter

increases, irrespective of the consolidation pressure.

The relatively well-defined relationship in Figure 3.22

suggests that the normalized peak strength depends on the


c proximity of the initial state with respect to the F line,

irrespective of the absolute values of void ratio and con­

fining pressure at the consolidation state. The scatter

in-Figure 3.22 may reflect unavoidable differences in

specimen preparation procedures and experimental errors.

Moreover, it is not yet know whether the vertical distance

between the F and S lines is constant, irrespective of the

confining stress. If this is not the case, correlations

with a state parameter defined with respect to the F line

may show some scatter because at the same values, some

points may lie initially between the S and F lines whereas

others may be located above the S line, resulting in dif-

f ferent stress-strain behavior. However, results of


I

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
Reproduced with permission of the copyright owner. Further reproduction C

s
0.7.
Bandlng sand

NBA
w N
w N V
UJ N
cc
F a N n °
w 0.6 -
n n *
cc n
o QD
<c
a
a
*
< 0.5 -
LU
CL ° c (Kg/cm2 ) Symbol
O o •
UJ 0.3 N
N N»
prohibited without perm ission.

□ 1.0 A A A 'N
<
5 4.0 □ ■
CC 0.4--
o 10.0 v ▼
2

-I- + + t- -I-
0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1 0.11

STATE PARAMETER, et-Of at <%'

Figure 3.22 No r ma l i z e d Peak Undrai nev’ Shear S t r e n g t h as a F u n c t i o n of

145
S t a t e Pa r a me t e r (Data frim Ref . 1 4 ) .
146

triaxial tests on two different sands reported by Sladen

et a l . (165), suggest that the state line separating the

region of limited strain softening from the region of

strain hardening behavior is essentially parallel to the

steady-state line (line F ) , in a e versus log plot.

The axial strains at peak deviator stress ep , are

plotted in Figure 3.23 versus the state parameter. It is

noted that, in general, the values of ep are relatively

small. Although the normalized peak strength correlated

well with the state parameter, it is observed that ep is

not uniquely correlatable to state parameter. On the con­

trary, for a given state parameter value, the strain at

peak stress increases with increasing confining pressure.

A similar trend can be inferred regarding the undrained

Young's modulus, which is a measurement of the stiffness

of the sand structure, whereas the strain at peak reflects

the ductility of the soil skeleton. Accordingly, both

stiffness and ductility increase with increasing confining

pressure at the same value of state parameter, \|>. Conse­

quently, these two properties, which define the stress-

strain characteristics of the specimen prior to the col­

lapse of the structure, are not uniquely correlatal«le to

state parameter, i|>.

The fact that for a given value the normalized peak

strength reaches a "unique" value irrespective of , sug­

gests that the triggering of specimen collapse might occur

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
Reproduced with permission of the copyright owner. Further reproduction
C

3.0

5 Banding sand
vu
W 2zs5 -
V)
111
DC
I—
CO
g 2.0- 0 -»c = 10 Kg/cm'
K“
<
>
UJ
D 1.5-
*
<
UJ
Q.

1 .0 -
z
prohibited without perm ission.

< 1.0
cc
h-
w 0.5 -
_l
<
X 0.3
<

0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.10 0.11
STATE PARAMETER, = e c - 0 F at crfc
Figure 3.23 A x i a l S t r a i n a t Peak D e v i a t o r S t r e s s as a F u n c t i o n of S t a t e

14 7
P a r a me t e r ( D a t a from R e f . 1 4 ) .
148

o at a particular value of the principal effective stress

ratio, which in turn relates to the mobilized angle of


*
shear resistance. Accordingly, the mobilized <|> at peak

deviator stress were calculated and plotted versus rela­

tive density and state parameter as shown in Figures 3.24

and 3.25. The mobilized angle of shear resistance is

related to the peak deviator stress, ^ ai"°3 ^p the

expression

---------------- (3.3)
c°i— °3 > p « c » c—V

*
where a is the consolidation effective confining pressure
c
and Up is the pore pressure at the peak point of the

c stress-strain curve.

It is seen that the mobilized angle of shear resis­

tance at peak strength for Banding sand increases with

increasing relative density but decreases as the confining

pressure increases in such a way that for initial states

at equal vertical distance from the F line, the mobilized


*
<j> becomes approximately a constant as indicated by the

relationship in Figure 3.25, especially if the confining


2
pressure is lower than 4.0 Kg/cm . The scatter of the

points corresponding to a consolidation confining pressure


2
of 10 Kg/cm is attributed in part to the fact that the F

line in Figure 3.20 is defined only up to a of 3.0


2
Kg/cm and therefore, some extrapolation was required.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
Reproduced with permission of the copyright owner. Further reproduction r

V O I D RATIO, e
0.8 0.78 0.76 0.74 0.72 0.70 0.68 0.66
CO
a>
a>
>_
O) 26- Banding sand
Q)
■a

8 24
LU
CC
f—
CO
22 -

a;
O 4.0
t- 10.0
<
> 20 -

at
Q
*
< 18 -
prohibited without perm ission.

LU
Q_

16 - Strain softening response: o e> □ o


Q
UJ Limited strain softening: • * ■ ▼
N
5 14
O
5 --i------------ +
10 20 30 40 50
R E L A T I V E DENS I T Y , Dr (%)
Figure 3.24 Angl e o f S h e a r i n g R e s i s t a n c e M o b i l i z e d a t Peak D e v i a t o r

149
S t r e s s as a F u n c t i o n o f R e l a t i v e D e n s i t y ( D a t a from R e f . 14)
Reproduced with permission of the copyright owner. Further reproduction

28
w
0)
0)
& 26
2
w
(/> 9
UJ A
*4
DC
F-
(O
DC 22
O
<
>
UJ 20

*
< Banding sand
81 18

Q
UJ ic
16
CTc '(Kg/cm 2) Symbol
prohibited without perm ission.

N 0.3 o •
to 1.0 A A
2 14 4.0 □ ■

10.0 V V
(
...
-I I- H -- -1- -t~- - I—
0.01 0.02 0 .0 3 0.04 0.05
0 .05 oto6
0 .06 0.07 0.08 0.09 0.1 0.11

STATE PARAMETER, v|> ” e c - e F at CTc

Figure 3.25 Angl e o f S h e a r i n g R e s i s t a n c e M o b i l i z e d a t Peak D e v i a t o r


S t r e s s as a F u n c t i o n of S t a t e Par ame t e r ( D a t a from R e f . 14).

150
151

Consequently, for a given state parameter value the ini­

o tiation of strain softening behavior occurs at a well

defined value of the principal stress ratio or mobilized

angle of shearing resistance. As pointed out by Leonards

(108), it is to be noted that the values of mobilized *

are considerably smaller than the maximum angle of shear

resistance, which from drained tests was found to be of

the oreder of 30° to 32°. Although mobilized $ at peak

stress correlates relatively well with *, as the confining

pressure increases, the level of strain required to mobil-


>•
ize * at peak strength also increases. The level of 0£

may have similar implications on the effects of cumulative

residual strain due to cyclic loading.

c It is also of interest to determine the magnitude of

the pore water pressure prior to the collapse of the

structure. From E q . 3.3 it can be readily obtained that

(sin* -1)
"p ' „ < °1 -V p * °c (3-4>
2 s l o *»ob.

By normalizing with respect to the peak deviator stress it

follows that

u_ sin* . -1 a
* mob* + ---- £ _ (3.5)
( al - ff3 )p 2sin *
mob.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
152

The left hand term in E q . 3.5 corresponds to


o Skempton's A^ , the pore pressure parameter at failure. It

can be noted that for a given state parameter, the term

a is constant independently of the level of o .


c 1 3 P c
Consequently, there is a unique relationship between the
*
pore pressure parameter and mobilized $ at peak. This

relationship had already been identified by Leonards (108)

based on the same data and it is reproduced in Figure

3.26. The correlation is remarkably close reflecting the

direct relation between compressibility and . The


*
direct interrelationship between mobilized <p , the pore

pressure parameter A^ and state parameter is clearly evi­

dent. However, it is more convenient to relate mobilized

$ directly to state parameter since A^ is a property


c which has to be measured. Unlike the state parameter, it

is not a measure of initial state conditions of the speci­

men.

After the peak point, the strength decreases drasti­

cally until the specimen reaches a steady-state of defor­

mation. At this stage, the strength is related to the

steady— state conditions defined by the F line and the

angle of shear resistance, <f>_, mobilized after large


r
strains, by the expression

2sin<|>
( °l~°3)F ' . a3F < 3 ’6)
1— sinq>p

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
< 9-
H0- Initial porosity
MOBILIZED

8 -
n« 0 .4 1 5 to 0 .4 5 5

7"
In drained shear
6 -

<jb'-30*
15

0.7 0.8 0.9 1.0 1.1 1.2

P O R E P R E S S U R E P A R A M E T E R A f A T P E A K (O ^ -O s )

Figure 3.26 Relationship between Mobilized <(>' and Pore


Pressure A,— Parameter at Peak Deviator
Stress (Alter Ref. 108).

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
154

where c._ is the effective confining stress at the


3r
o steady— state of deformation. Figure 3.27 shows the mobil-
*
ized <|>_ at the steady-state condition for the tests on
F

Banding sand. It is evident that at the steady-state con­

dition the effective angle of shear resistance is fully

mobilized. For Banding sand, <J> at large strain is found

to be rather independent of the confining pressure for

relative densities below about 50%. However, data

reported by Baldi et al. (7), show that the dependency of


^ *
i|i on o increases with increasing relative density,
c

To quantify the magnitude of the strength drop after

the peak in saturated loose specimens, Bishop (9) defined

a parameter that he called "brittleness index," which is

c the difference between the peak strength and the steady-

state strength normalized with respect to the peak

strength. This parameter is homologous to "sensitivity"

of quick clays. The brittleness index was determined for

the tests on Banding sand in which complete or limited

strain softening response was observed, and plotted versus

relative density and state parameter, as shown in Figures

3.28 and 3.29. It is seen that the degree of softening,

as measured by Iw , increases with decreasing relative den-

sity and increasing confining pressure. For very low

relative densities, the drop in strength may be close to

100% irrespective of the confining pressure (Fig. 3.28).

The combined influence of relative density and confining

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
Reproduced with permission of the copyright owner. Further reproduction
r'.
C

VOID RATIO, e
0,8 0^78 0,76 0^74 0.72 0^74 0.68 0.66
40 — J - - -t
W Undrained Tests
Q)
<1)
i.
Banding sand
D) 38 O-c (Kg/cm 2 ) Symbol
©
TJ 0.3 o •
Z 1.0 A A
o
t- 4.0 a m
Q
Z 10.0 V V
o
0 34
Ul
t-
2
w 32
1 (0.5 ) (0.3) *
V
o * *
(1) *
i * (1)
< (0.3) 0 • * *
UJ 30
F- (0.3) * 14)
CO *
prohibited without perm ission.

(4) (1 0 )
28 -
O
UJ ^ at peak devlator stress
N
□ in drained tests
5
m 26
o () Consolidation confining
2 o? pressure, K g / c m 2
■9- 24 — /V — i— -I
v 10 20 30 40 50
RELATIVE DENSITY, Dr (%)
Figure 3.27 Angl e of S h e a r i n g R e s i s t a n c e M o b i l i z e d a t S t e a d y - S t a t e

155
C o n d i t i o n as a F u n c t i o n o f R e l a t i v e D e n s i t y ( D a t a from R e f . 1 •)
156

m
0)
CM >
XJ
as
a>
0)
ttS
©
«-i
O
e
o
•H
XJ
O
c
5
U-t
CO
CO
CO
CO
CO •
0)
co c *3*
0)
n

c CO
z
UJ
XJ •
XJ U-i
-H 0)
kt p£
CQ
E
Q T5 O
CJ Ul
UJ C <XJ
>
a. I- CO CO
U XJ
cr -- o
CM
< -o CO
cr c Q
UJ
oc
O >>
XJ
X •H
o CO
O' i-i c
CO e 0)
O) l-l O

■o oo
CM
CD'

0)
3
o co CO CM 00
o

aI ‘X 3 Q N 1 SS3N311J.jaa

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
157

pressure on the brittleness or collapsiveness of banding


o sand is such that Ig also correlates well with the state

parameter, as illustrated in Figure 3.29. A plot similar

to that in Figure 3.29 was recently presented by Sladen et

al. (165) based on triaxial tests on two different sands.

Either plot suggests that there exists a state parameter

value below which the brittleness index is zero. This is

a further indication of the fact that specimens with ini­

tial states in the region just above the F line will show

a strain hardening type of behavior.

The state diagram as used by Casagrande and others

(13,14,15) only accounts for the relationship between void

ratio and confining pressure (or mean confining stress)

c during either drained or undrained shearing.

does not permit interpretation of the mobilization of the


However, it

shear resistance of the sand before and after the point of

collapse, if any, until reaching the steady-state condi­

tion. Consequently, either a 3-D diagram, in terms of

mean confining stress, shear stress and void ratio (186)

or a combination of state diagram and stress-path plot

(144,169) is required to fully represent the most impor­

tant stages during undrained monotonic loading of sands;

namely, the initiation of strain softening behavior, the

change for contractive to dilative behavior, when this is

the case and finally the undrained strength at the

steady-state condition.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
Reproduced with permission of the copyright owner. Further reproduction

i.a

Banding Sand

0.8..

CD
M
X
UJ
Q 0.6..
z
CO
CO
UJ
z
UJ
-I
0.4-
H
h Symbol
s
m
prohibited without perm ission.

0. 2..
■ 4.0

10.0

+
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1 0.11
STATE PARAMETER, t|/ = ec- eFat <r£

F ig u re 3.29 Index of U n d r a i n e d B r i t t l e n e s s as a F u n c t i o n of State


Parameter (Data fr om Ref. 14).
159

Based on the behavior of Banding sand, which may be

c / typical of other types of sand with similar characteris­

tics, that is, uniform, clean quartz sands with subrounded

to subangular grains, the relationship between the state

diagram and the effective stress plot will be illustrated

schematically for three typical cases:

1. Constant relative density,

2. Constant consolidation pressure, and

3. Constant state parameter.

1.Behavior at Constant Relative Density. It is of

interest to analyze the undrained behavior of specimens

C with a constant relative density at different levels of

confining pressure. In this case, the state parameter

changes from negative on the dilative side to increasingly

positive as the consolidation pressure increases. The

resulting behavior is illustrated in Figure 3.30. On the

dilative side, as the effective stress path approaches the

failure line dilation prevails and as a result of the

decreasing pore water pressure the stress path exhibits a

turn around point or change in direction and climbs

approximately along the failure line. A peak point is not

observed and the maximum angle of shear resistance is

mobilized at the peak deviator stress, which occurs at

large strains. However, it is to be noted that at the

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
160

o S tate Diagram

S Line

Strain softening region


<D
O

cc o— — o--- o
9
o
> Transition region
Strain hardening region

F Line

c co
Stress Path Plot
S tead y-state
envelope
b
I CM

Locus of
II triggering
cr
states
(CSR line)

Figure 3.30 Undrained Behavior at Constant Relative


Density (Schematic).

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
161

point the stress path changes in direction the mobilized


o angle of shear resistance is only slightly smaller than

the corresponding to the failure line. As the confining

pressure increases the behavior changes from dilative to

limited strain softening behavior. The stress path shows

a peak point which gradually moves away from the failure

line as the confining pressure increases. F in al ly , for

initial states beyond the S line, the behavior becomes

complete strain softening and the mobilized $ at peak

approaches a constant value at high confining pressures in

accord with the trend observed in Figure 3.24. Conse­

quently, the locus of points at which strain softening is

initiated defines a curve which at some point deviates

from the failure line and gradually becomes a straight

c *
line as the mobilized $ decreases to a constant value.

Hence, the locus of peak points defines a critical stress

ratio, CSR, line (186), that is, the line connecting the

points on the stress space surface marking the initiation

of strain softening behavior at different confining pres­

sures. In the tests performed by Sladen et a l . (165) on

Nerlerk and Leighton Buzzard sands, the CSR line was found

to be a straight line. As the relative density increases,

the CSR line branches away from the failure line at a much

higher shear stress level since the behavior becomes con­

tractive at much higher confining pressures.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
162

2. Behavior aC Constant Consolidation Pressure. At


(
low relative densities the state parameter is positive and

becomes negative as the relative density is such*that the

initial state is below the F line. Consequently, as shown

in Figure 3.31, the stress path corresponding to low rela­

tive densities reflects a strain softening behavior with a

large drop in shear stress accompanied by a high pore

water pressure increase. As the relative density

increases the mobilized <J> at peak increases as well as

the steady-state strength. For initial states between the

S and F lines the behavior may correspond to limited

strain softening. Accordingly, the stress path after the

peak moves downwards but at some point moves upwards

approaching the steady-state envelope. As the relative


C density continues to increase, the behavior becomes com­

plete strain hardening (no peak in the stress-strain

curve) and the stress path climbs along the failure line.

An elbow or change in curvature in'the stress path is

observed at a stress ratio corresponding to a mobilized <j>

smaller than the angle of shear resistance, which

increases with increasing relative density but decreases

with increasing confining stress at high relative densi­

ties.

3. Behavior at Constant State Parameter. Since the

normalized peak strength and the pore pressure parameter

at failure are both "unique" functions of state parameter,

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
163

State Diagram

S Line
Strain softening region
-e A

-o B

F Line

Transition region

Strain hardening region

a y + cri
P =
c
Stress Path Plot S tead y-state envelope

co
CM

Locus of triggering states


(CSR line) t
c7
• S teady-state condition

Figure 3.31 Undrained Behavior at Constant Consolidation


Confining Stress (Schematic).

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
the corresponding effective stress paths in the q-p'

space become geometrically similar, as illustrated in Fig­

ure 3*32 for a given state parameter value. Under these

conditions, the locus of peaks point defines a straight

CSR line, irrespective of the level of confining stress.

Although the A^ parameter at peak deviator stress is the

same for all the stress paths shown in Figure 3.32, it is

not to be concluded that the rate of pore pressure buildup

per a given increment of deviator stress is the same. As

illustrated in Figure 3.32, the higher the confining pres­

sure the stiffer the specimen and consequently the lower

the pore pressure generated at the same shear stress

level. When the rates of pressure buildup are compared at

the critical stress ratio they happen to be equal since

all specimens have the same "contraction" potential as

measured by the state parameter. Thus, the higher the

confining pressure, the higher the deviator stress and the

consequent pore pressure at peak strength, such that A ^ ,

which is a function of ip (Figs. 3.25 and 3.26), remains

con st an t.

3.3.3 Effect of Sand Type on Undrained Behavior

As mentioned earlier, the brittleness or "collapsive-

ness" of sand skeletons, that is the susceptibility to an

abrupt, very rapid change in compressibility at small

strain amplitudes, has a crucial influence on the

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
165

o
S Line State Diagram

o A Consolidation states

© • S tead y-state condition


------oB
O
h-
<
cc
o
o F Line
>
^ (constant)

c Stress Path Plot

S teady-state
envelope

e\j
Locus of triggering

S tea d y -s ta te strength - states (CSR line)


cr

Figure 3.32 Ondrained Behavior at Constant State


Parameter (Schematic).

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
undrained behavior of saturated sands. It is clear that,

in addition to relative density and level of stress, the

type of sand, that is the mineral composition, size dis­

tribution, shape and roughness of the particles, play an

important role in the collapse mechanism. Figures 3.33


*
and 3.34 present the mobilized $ at peak deviator stress

versus relative density and the steady-state lines,

respectively, for a uniform, fine to medium, clean sand

with angular, bulky grains composed of basalt, pla-

gioclase, magnetite and olivine (14). The corresponding

values for banding sand are also shown for comparison. It

is seen that at the same relative density the angular sand

mobilizes a higher $ at the peak than the banding sand


*
(Fig 3.33). It is also to be noted that <t> mobilized at

peak is closer to the maximum angle of shear resistance in

the case of the more angular sand. However, the effect of

confining pressure on the mobilized angle of shear resis­

tance at peak deviator stress is significantly larger for

the angular sand, which was observed to be more compressi­

ble. This is likely to result partly from breakage of

sharp edges during both consolidation and shearing stages.

On the other hand, at the same relative density the

steady-state strength of the more angular sand is larger

than for banding sand (Fig 3.34), which exhibited a more

pronounced strain softening behavior (14). It is also

noted that the difference between the F and S lines is

smaller for the more angular sand, which is also

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
Reproduced with permission of the copyright owner. Further reproduction

<n 40
a>
2 38
O) /
a) 1.0 Symbol ( j j drained
~ 36 Angular sand a ■ 3 4 .5 °
0)
a 34 / Banding sand & □ 30.6"
a:
i- U
co 32
DC
o
h
CFc=4.0 K g /cnr
30
< ■ II ■
>
UJ
0 28
X.
<
Hi 26
01
I- 24 -
<
prohibited without perm ission.

Q
UJ 22 -
N

§ 20 4.0

18
-e-
16 Ay-
10 20 30 40 SO
R E L A T I V E DENS I T Y , Dr (%)
Figure 3.33 Angl e o f S h e a r i n g R e s i s t a n c e M o b i l i z e d at Peak D e v i a t o r
S t r e s s f o r Two D i f f e r e n t Sands ( D a t a from R e f . 1 4 ) .
168

i- Triaxial Tests 0 i
1

p 10-!
1 i
Line

--------------------------- 2 C ~
i
N
0.9 ^ 3p at Dr= 2 0 % S Line

Angular sand

DENSITY, Dr (%)
<£'*=35"

©
O

C l-
<
cc

o 0.8

RELATIVE
F Line S Line

0.7 - Banding sand


(Subrounded to subangular)

< £ '- 3 0 .6 °
i i i i i ii11
____ i i i i i iii
0.01 0.1 1 _ 10
E F F E C T IV E M INO R PR IN C IP A L S T R E S S , ^ 3 , k g / s q cm

Figure 3.34 Steady-State Diagrams for Two Different


Sands (Adapted from Ref. 14).

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
169

Indicative of a reduced degree of softening and smaller

collapsiveness of the structure.

Figure 3.35 compares the steady-state diagrams (F

line) for two other types of sand with angular and rounded

grains, respectively, possessing identical gradations

(186). The angular sand is identified as tailings sand,

which is composed of quartz (35%) and feldspar (65%)

grains. The rounded sand corresponds to Ottawa sand.

Both sands have a mean grain size of 0.4 mm and coeffi­

cient of uniformity of about 1.5. For a wide range of

initial states the specimens of both sands only exhibited

limited strain softening response. Therefore, the effec­

tive confining pressure values plotted in Figure 3.35

correspond to the values observed during the short-lived

steady-state condition. For relative densities below

about 38% the steady-state strength of Ottawa sand is con­

siderably smaller than for tailings sand. The reduction

in strength associated with strain softening of Ottawa

sand was found to be much larger than for the tailings

sand under similar initial stress conditions. On the

other hand, at relative densities above about 38% the

steady-state strength of Ottawa sand becomes larger than

that corresponding to tailings sand, at comparable rela­

tive densities. This is attributed to the fact that the

angular sand is still very compressible at relative densi­

ties above 38%, due in part to crushing of grains but also

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
170

O 1.00
Triaxial Tests <f> -38.5 20
0.95 30

DENSITY, Drc (%)


40
0.90
50
O 0.85
60
0.80- ®d 70
80

RELATIVE
> 0.75
S te a d y -s ta te strength
0.70 -

0 .65

E F F E C T I V E C O N F I N I N G S T R E S S , O '3' (kgf/cm2 )
a) A n g u l a r Tailings S a n d
c Triaxial Tests <P =31.5

D ENSITY, Dr (%)
0.751-

- 0.70

5 0.6 5r

O 0.601-
RELATIVE

7 ' v-a
S te a d y -s ta te strengtn

0.1 0 .2 0.5 1.0 5.0 10.0 20.0 50.0 100.0


E F F E C T I V E C O N F I N I N G S T R E S S , 0~3 (kgf/cm^)
b) R o u n d e d O t t a w a S a n d
Figure 3.35 Undrained Steady-State Diagrams for Two
Different Sands (After Ref. 186).

R eproduced w ith perm ission o f the copyright owner. F urth er reproduction prohibited w itho ut perm ission.
171

to a larger volume decrease potential than that of Ottawa

sand, at the same relative density. As discussed earlier,

the volume decrease potential, VD , is the difference

between the initial void ratio and the minimum void ratio,

emin* Additionally, it is clearly evident that a small

change in relative density for Ottawa sand corresponds to

a large change in the steady-state strength, whereas a

similar change in Dr of tailings sand produces a less sig­

nificant change in the undrained strength at the steady-

state condition. It is clear that the differences in

undrained steady-state strength, which reflect large

differences in compressibility, and collapsiveness of the

two sands, must be related to the differences in mineral

composition, size distribution, shape and roughness of the

particles of the the two sands. For Ottawa sand, it is

the relative density that essentially determines the

change from dilative to contractive behavior, since the

steady-state line is rather flat for confining pressures


2
below about 10 Kg/cm . It is apparent that at high rela­

tive densities extremely high effective confining stresses

would be required for the behavior to become contractive

or strain softening. This is attributed to the low

compressibility of Ottawa sand. On the other hand, at any

relative density the behavior of tailings sand may change

from dilative to contractive as the confining stress

increases and the initial state points move to the right

side of the F line, due to the increased compressibility

at higher stress levels.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
172

Based on the test results previously discussed, and

on the results of other Investigations regarding the

steady-state strength of saturated sands (8,14,18,34,186)

the author draws the following conclusions:

a. The slope of the steady-state line reflects the

compressibility of the sand skeleton, which in turn

is related to the mineral composition, size distribu­

tion, shape and surface roughness of the sand grains.

Steady-state lines are steeper for sands with more

angular, rough particles such as mine tailings pro­

duced by rock crushing. This is partly a consequence

of breakage of sharp edges of particles both during

consolidation and undrained shear deformation, and

also of the higher values of e that can be


max
achieved. The flatter lines correspond to sands with

more rounded bulky grains. The slope of the steady-

state lines may also increase with increasing fines

content (< #200 m e s h ) , which is consistent with a

trend towards greater compressibility with increasing

fine content, as compared to cleaner coarser sands.

On the other hand, a rounded sand with friable parti­

cles is likely to behave like an angular sand, hence

exhibiting steeper steady-state lines.

b. The position of the steady-state line at low confin­

ing stresses is related to the maximum void ratio of

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
173

the material, which Is also dependent on the size


O distribution and shape of the particles. e is
r r max
likely to be higher for more angular and well-graded

sands as compared to more uniform coarser sands.

Accordingly, the more angular and well-graded sands

may plot at higher void ratios for the same effective

stress at the steady-state conditions.

c. Sands with smoother, more rounded and finer particles

are more susceptible to sudden "collapse" than sands

with angular or compressible (friable) grains. The

higher the collapse potential the larger the magni­

tude of strain softening after the peak strength, all

other factors being equal. Consequently, sands with

C smoother, more rounded and finer grains may plot in

the state diagram at a lower a, for the same void


vr
ratio or Dr over the range corresponding to loose

states. Sands with these characteristics are also

likely to exhibit flatter steady-state lines.

These conclusions clearly indicate that the effects

of the type of sand on the susceptibility to strain

softening behavior (or collapse) under monotonic loading

and on the liquefaction potential due to sudden reversals

in the direction of shearing, are opposite. As discussed

earlier, liquefaction potential increases with , the

volume decrease potential. On the contrary, as

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
174

Increases i.e. for angular sands, the steady-state lines


G become steeper. Moreover, they will be located at higher
*
a values over the range of Dr corresponding to loose
3F
states, reflecting a reduced degree of softening. As VD

decreases, i.e for rounded sands, the opposite is true.

In summary, sands with flatter steady-state lines are

highly susceptible to flow deformation but they have

higher resistance to liquefaction, at the same Dr. As the

slope of the steady-state line increases, the collapsive-

ness of the sand on loading decreases but its liquefaction

potential increases as a result of the higher volume

decrease potential.

3.3.4 The State Diagram and Cyclic Dndrained Behavior

C During cyclic loading, a sand specimen may be sub­

jected to shear stress amplitudes below or above the

undrained steady-state strength under monotonic loading.

The cyclic undrained behavior for these two conditions is

discussed below.

1. t < t . Consider first a sand specimen with


cy ss
initial state at point A (Fig.3.36), which is subjected to

a series of shear stress applications of magnitude, Tc y *

As cyclic loading proceeds pore pressures are generated

and the effective stress path traverses towards the left,

eventually moving from the contractive to the dilative

region in the state diagram. It Is noted that as the

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
175

o State diagram

S Line

Strain softening region


a>
O
i-
<
0c
9
o
>
F Line Transition region

Strain hardening region

p' _ i-P3

| Stress path plot


i
j
Clo

c Steady-state envelope

Monotomc loading

CSR line

Phase transformation line

Figure 3.36 Undrained Behavior under Cyclic Loading of


Magnitude Smaller than Steady-State Strength
( Schematic).
r

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
176

o effective stress path moves towards the left the principal

stress ratio, that is the mobilized angle of shear resis­

tance gradually increases. However, as the shear stress

amplitude is smaller than the steady-state strength for

that particular void ratio, the stress path cannot reach

the critical stress ratio (CSR) line. Therefore, the

effective stress path traverses below the point

corresponding to the steady-state strength (point S in

Fig. 3.36), and approaches the failure line. Once the

effective principal stress ratio becomes slightly smaller

than the value corresponding to the failure line dilation

prevails and the stress path, corresponding to the cycle

in which this occurs, shows a turn around point or elbow

c and moves back along the failure line as observed during

monotonic loading (stress path starting at point B). Upon

reversal of the shear stress (point R in Fig. 3.36) the

pore water pressure increases significantly (downward

motion of grains) and the effective stress path approaches

the origin of the q-p plot determining the occurrence of

liquefaction.

It is clear that excessive pore water pressure upon

reversal of the direction of shearing only occurs after

the stress ratio reaches a threshold value which marks the

point at which dilation prevails completely on the dila­

tive region of the state diagram. Consequently, the

stress ratio corresponding to the point at which the

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
177

stress path In monotonic loading changes in direction may


o be taken as the threshold separating two different

behaviors (82) . The change in behavior will be manifested

in the subsequent unloading and loading in the other

direction. Unless the stress ratio goes beyond the thres­

hold, irrespective of the cyclic shear stress amplitude,

the specimen preserves the characteristics of a "solid"

body. Conversely, once the stress ratio is greater than

the threshold, excessive pore water pressure develops upon

unloading bringing the specimen to a liquefied state.

Accordingly, the angle defined by the threshold value of

the stress ratio has been termed (82) "angle of phase

transformation" since upon unloading from a stress ratio

c above the threshold value the behavior of the sand speci­

men as a solid is lost and transformed into that of a

liquefied state. Thus, the value of the angle of phase

transformation may be read off from the stress path lines

and for loose specimens it would be slightly smaller in

magnitude than the stress ratio corresponding to the

failure line. In this case, liquefaction is likely to

develop during the first unloading state after the stress

path crosses the line of phase transformation. On the

other hand, as the relative density of the sand increases

the line of phase transformation moves away from the

failure line, since dilation would prevail at much lower

stress ratios. Accordingly, in dense sands many cycles of

loading may be required to reach the zero effective stress

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
178

condition after crossing the line of phase transformation

for the first time. H o w e v e r , as mentioned earlier, the

development of liquefaction in very dense sands does not

really imply a "transformation of phase."

It is interesting to note that once the stress path

and the state path reach the point p =0, the behavior of

the specimen is likely to be strongly dilative, according

to the state diagram. Consequently, after the deforma­

tions which accompany the development of liquefaction

occur, further straining in the other direction permits

dilation to develop, the pore water pressure decreases and

the state point moves to the right towards the steady-

state line, reflecting an increase in the resistance of

the specimen (Point P in Fig. 3.36). If cyclic loading

continues the state path moves back and forth reaching the
*
point p =0 each time liquefaction develops. The

corresponding stress path moves upwards and downwards

along the failure line during the loading stages and

approaches the origin during the unloading phases. If the

steady-state strength of the sand is very low, it is

likely that initial liquefaction is accompanied by the

development of very large strains well beyond the capabil­

ities of any testing equipment. Accordingly, these cases

constitute actual liquefaction. As the relative density

increases, the shear strains required to mobilize the

undrained steady-state strength after liquefaction

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
179

gradually decrease and become Insignificant at very high

relative densities. Dilation develops almost instantane­

ously and the specimen withstands the cyclic shear stresses

without appreciable deformation.

In any case, strains mainly accumulate during the

periods of time the stress path point is within the two

lines of phase transformation, irrespective of the magni­

tude of the applied shear stress, as long as x remains

smaller than the undrained steady— state strength, x .


ss
Consequently, it is expected that as the shear stress

amplitude increases the resistance to liquefaction becomes

less and less dependent on the cyclic stress ratio itself.

This is clearly evident from the cyclic undrained strength

c curves for Toyoura sand shown in Figure 3.37.

that virtually all strength curves for different relative


It is seen

densities have a similar sha^e showing a maximum curvature

at a single value of the cyclic stress ratio, which has

been termed the critical cyclic stress ratio (175). When

the cyclic stress ratio is larger than the critical cyclic

stress ratio, whether cyclic undrained failure is induced

or not in a given sand having a given Dr depends mainly on

the number of cycles and it is less sensitive to the

cyclic stress ratio value. On the other hand, below the

critical value a small change in the cyclic stress ratio

causes a significant change in the number of cycles

required to cause cyclic undrained failure (175). It is

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
Reproduced with permission of the copyright owner. Further reproduction
c

CYCLIC TORSIONAL
SHEAR TEST
TOYOURA SAND
AIR-PLUVIATED
'0 % = 9 8 kPa

15% DA SHEAR STRAIN

Critical 7cy/<7c
prohibited without perm ission.

N U M B E R O F LOADING CYCLES, No

Figure 3.37 Conc ept o f C r i t i c a l C y c l i c S t r e s s R a t i o ( Da t a from R e f . 173).

180
181

also noted that although for each curve the slopes below
o and above the critical stress ratio are different, they

are essentially the same for all the curves, irrespective

of the relative density. Consequently, what seems to

really differentiate each curve is the number of cycles

required to cause undrained failure at the critical cyclic

stress ratio. The author associates the critical cyclic

stress ratio with the cyclic shear stress level which just

induces some dilation in each cycle of loading, that is,

the stress ratio which causes the stress path to cross the

phase transformation line. However, as the threshold

stress ratio at the phase transformation line increases

with decreasing relative density, it may well happen that

c the "elbow" of all the stress paths occurs at approxi­

mately the same cyclic stress ratio, as indicated by Fig­

ure 3.37.

This finding has excellent practical implications.

For cases in which strain softening is not a possibility

(icy smaller than steady-state strength) it would be

enough to determine the undrained strength curve for a

single relative density value. This single curve would

permit determination of the critical cyclic stress ratio.

Once the critical cyclic stress ratio is known, specimens

with different relative densities can be subjected to

stress applications at the critical amplitude in order to

determine the corresponding critical number of cycles. As

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
182

indicated in Figure 3.37 the critical number of cycles may

be of the same order of magnitude as the significant

number of cycles in an earthquake. A c c or di ng ly , the

author strongly supports the idea of Tatsuoka et a l .

(175) of using these two parameters, critical cyclic

stress ratio and critical number of cycles to represent

the resistance to liquefaction of saturated sands. This

would simplify testing work allowing at the same time a

more direct comparison of the cyclic strength characteris­

tics of different sand types at different relative densi­

ties, stress-strain histories, etc. It Is not yet known

whether the slopes of the cyclic strength curves are also

function of factors such as type of sand, etc.

2. TC v ^ Tss* w ^ en a specimen is subjected to cyclic

shear stress amplitudes larger than the undrained steady-

state strength corresponding to the consolidation void

ratio, in the course of cyclic loading the effective

stress path may reach the critical stress ratio line as

indicated in Figure 3.38. Accordingly, at this stage the

specimen will exhibit a strain softening response reaching

finally a steady-state of deformation. In such a case, it

will be of concern to define the following features of the

behavior:

a. the number of cycles of a given stress amplitude

required for triggering the Initiation of the strain

softening response,

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
183

S Line
Strain softening region
®
o
t-
<
oc
o
5
>

F Line Transition region

Strain hardening region

Steady - State envelope


C

Monotonic loading

CSR line
Steady-state condition

b”
CM

Figure 3.38 Undrained Behavior under Cyclic Loading of


Magnitude Larger than Steady— State Strength,

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
184

s b. the position of the critical stress ratio line under

cyclic loading conditions, and

c. the steady-state strength after the collapse of the

structure is triggered by the cyclic loading.

It is important to note that if the specimen reaches

a steady-state of deformation the condition of initial

liquefaction will not develop since the specimen cannot

buildup a strength equal to the stress amplitude at which

the reversal in loading direction is compelled to occur,

and hence it remains at the steady-state condition.

Figure 3.39 presents the values of the angle of shear

resistance (<j>") mobilized at the initiation of strain

softening response during cyclic tests on Banding sand, in

which the specimens were subjected to axial compression

cycles at constant amplitude of deviator stress. The

numbers within parentheses correspond to the effective

confining pressure at the end of the cycle immediately

before the specimen collapsed. The values of mobilized <t>"

at similar initial stress conditions during monotonic

tests are also shown by the solid lines for comparison.

It is seen that the initiation of strain softening

response under cyclic loading occurs at stress ratios

larger than under monotonic loading. For example, under a


2 ~ Q
confining stress of 4.0 Kg/cm , $m o |j increases from 21

to nearly 26°, at a relative density of 32%. In

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
ission of the copyright owner. Further reproduction

(1.4)
26 -
Banding sand ■(1.4) (5.0)
■(1.5)

(1.7)
24 -

Monotonic tests
22 -

CTq= 10 Kg /cnf 10.0


■(1.8 )

20 -
4.0

18 ■-
Cyclic tests
prohibited without perm ission.

°"c (Kg/cm2) Symbol

4.0 ■

10.0 ▼

— +
10 20 30 40 50
R E L A T I V E D E N S I T Y , Dr (%)
Figure 3.39 Angle of Shearing Resistance Mobilized at Initiation of
Strain Softening Response during Cyclic Loading (Data
f rom Ref. 14)•
186

principle, this is in agreement with a trend towards a

more stable structure due to the cumulative strain history

effects during cyclic loading. Cyclic torsional shear

tests by Dobry et a l . (34) also showed that for clean

banding sand, triggering of the flow failure during cyclic

loading occurs when the effective stress path reaches a

line closer to the steady-state envelope ( aT = 26.5 )

while for monotonic tests on the same sand triggering of

strain softening response occurred along a line

corresponding to a lower critical stress ratio

(aT = 21.8 ) On the other hand, in the cyclic triaxial

tests performed by Vaid and Chern (186) on specimens of

angular tailings sand, triggering of strain softening

^ response was observed to occur at the same critical stress

ratio or mobilized <t>^, in both monotonic and cyclic tests,

irrespective of the initial state of the specimens. The

mobilized friction angle at the critical stress ratio line

was very close to <}>u , the interparticle friction angle for

a quartz-feldspar mixture of tailing sand. It is likely

that for angular sands the mobilized friction angle at the

initiation of flow deformation is mainly determined by

grain angularity and hence, it is relatively independent

of the initial state of the specimen.

Consequently, based on the limited number of studies

available, the author considers that cyclic loading may

influence the location of the critical stress ratio line,

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
187

especially in the case of sands with more rounded grains,


c /
where mobilized is more sensitive to changes in the

initial state of the specimen (see Fig. 3.24). However,

the exact magnitude of these effects is not yet known. In

this connection, it would be of particular interest to

establish the level of shear strain required to mobilize

the critical stress ratio (CSR) during cyclic loading.

Under monotonic loading the strain at peak deviator stress

is observed to increase with confining pressure (Fig.


*
3.23). It is yet to be established whether or not a has a

similar effect on the strain accumulated prior to the

specimen collapse during cyclic loading.

Another important engineering property is the magni­

c tude of the steady-state strength following the collapse

of the structure during cyclic loading. Based on his

tests on banding sand, Castro (14) concluded that the

st ea dy — state line (line F) was unique and independent of

flow deformation having been triggered by monotonic or

cyclic loading. Dobry et a l . (34) arrived at the same

conclusion based on their cyclic torsional shear tests on

banding sand and on angular sands with different silt con­

tent. Similarly, Vaid and Chern (186) found that for

tailings sand and Ottawa sand, which exhibited flow defor­

mation over a limited range of strains, the steady-state

strength corresponding to this stage of flow was similar

at the same void ratio, under both monotonic and cyclic

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
188

loading. It is important to note that only the tests by

C }
Castro were load-controlled tests whereas the other tests

were strain controlled tests. In spite of the conclusions

drawn by these investigators, in the author's opinion, it

is too early to assert that the steady— state strength is

independent of the loading conditions. Differences are

expected to arise because of strain rate effects, as pre­

viously discussed, and different changes in fabric induced

by each type of loading, among other factors. Additional

investigations on the steady-state strength of saturated

sands are urgently needed.

For some initial states (e.g. between the F and S

lines in the case of Banding sand), a specimen may exhibit

^ only a limited strain softening response. Consequently,

as shown in Figure 3.40, following the collapse of the

structure triggered by cyclic loading, (point T) the

specimen may reach a steady-state condition over a limited

strain range (point S) . However, further straining stops

the continuous movement as dilation develops and the

specimen is able to regain strength enough for the rever­

sal of direction of loading to occur (point R ) . At this

stage the stress path has already crossed the line of

phase transformation. Consequently, upon unloading the

stress path moves towards the zero effective stress state

and initial liquefaction occurs. From this point on, the

behavior continues as explained earlier with alternating


i
\

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
189

O
S Line

© Strain softening region


o"

QC
O
5
> O A

Transition region
F Line

Strain hardening region

P' =

c b”
Monotonic loading

i eg
b~ Temporary Locus of
ii Steady-state Triggering
cr condition — . states (CSR line)

Phase transformation line

Figure 3.40 Development of Limited Flow Deformation and


Liquefaction during Cyclic Loading (Schematic)

R eproduced w ith perm ission o f the copyright owner. F urth er reproduction prohibited w itho ut perm ission.
190

cycles of liquefaction and large deformations followed by

stiffening upon loading in the opposite direction.

It is important to note that in these cas es , a speci­

fied level of shear deformation during cyclic loading can

be on account of flow deformation or liquefaction, or a

combination of the two. Nevertheless, strain softening

behavior, if any, will always be observed before the ini­

tial liquefaction because the locus of triggering line

corresponds to a critical stress ratio smaller than the

threshold stress ratio at the line of phase transforma­

tion. It is common practice to normalize the cyclic

undrained strength (cyclic stress ratio versus N) with


2
respect to a effective confining pressure of 1.0 Kg/cm

(147). It is likely that at this stress level a sand like


c
Banding sand will undergo unlimited flow deformation only

for relative densities below 30% (see Fig. 3.20). In the

range of Dr between 30% and 40% limited strain softening

may occur in the course of cyclic loading, if the shear

stress amplitude is larger than the steady-state strength.

On the other hand, above Dr of about 45%, strain softening

is not possible and the accumulation of large strains in

the course of cyclic loading will be only on the account

of liquefaction each time the specimen is unloaded to an

isotropic state of stress. Depending on the confining

pressure, it is likely that for most sands the upper limit

for which unlimited flow deformation is a possibility lies

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
191

in the range between about 40% and 50% (13). It can be


o seen that in Figure 3.20 the steady-state strength of

banding sand increases at an increasing rate with increas­

ing relative density, which resembles the cyclic undrained

strength versus relative developed on the density curves

presented in Figure 3.11.

In summary, irrespective of the terminology used by

investigators, it is evident that there is a clear rela­

tionship between the undrained behavior under monotonic

loading and the phenomenon of liquefaction of sands due to

cyclic loading. The author could develop this relation­

ship by understanding that the structure of a "loose" sand

may collapse either because the applied shear stresses are

c too high or as a result of continuous reversals in the

direction of shearing. Under undrained conditions, both

types of "collapse" result in a significant increase in

pore pressure and a consequent reduction in strength.

Hence, "flow deformation" and "liquefaction" are in nature

similar phenomena in the sense that both involve a drastic

reduction in the number of contact points between neigh­

boring grains. Nevertheless, it should be remembered that

liquefaction only develops if in the course of cyclic

loading the sand is unloaded to an isotropic state of

stress. The effects of initial state of stress and in situ

boundary conditions on the cyclic undrained behavior are

discussed in the next section.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
192

3.4 Effect of Initial State of Stress and


In Situ Boundary Conditions
O
The behavior of saturated sands under both nonotonic

and cyclic loading has thus far been discussed based only

on results from tests on isotropically consolidated speci­

mens. However, it is clear that stress isotropy is not

typical of many in situ conditions. On the contrary, soil

elements near structures, in earth structures such as

slopes, embankments, dams, etc., and even under level

ground are generally consolidated under an anisotropic

state stress. Under level ground no static shear stresses

act on horizontal and vertical planes, which then

correspond to the principal planes of stress. On the

other hand, in soil elements under sloping ground condi­


C tions or beneath a structure founded on a level ground

deposit, initial shear stresses will act on horizontal and

vertical planes. The presence of these initial static

shear stresses have a major effect on the response of soil

elements to the superimposed cyclic shear stress applica­

tions during an earthquake, as discussed below.

In the past, it was conventionally assumed that the

test conditions in isotropically consolidated specimens

simulate approximately the stress conditions on an element

of soil at some depth below a horizontal ground surface

because for this condition there are no initial shear

stresses on a horizontal plane, the plane on which the


r

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
193

c cyclic shear stresses were considered to be applied by an

earthquake. Accordingly, the same condition is developed

on the specimen by applying an initial ambient pressure,

that is, by conducting a consolidated-isotropically-

undrained (CIU) test. The author strongly disagrees with

this reasoning because the behavior of a soil element can­

not be expected to be dependent only on the stresses act­

ing on a single plane. Nevertheless, because of fundamen­

tally different reasons, it will be shown later that CID

tests indeed may simulate soil conditions in deposits

under level ground closer than other types of tests, if

these latter tests do not simulate the boundary loading

conditions that exist during an earthquake.

c The problem of undrained stability during cyclic

loading can be studied by considering the way in which the

in situ effective stress path approaches the failure line.

It is this path that determines the failure modes and the

resulting deformations. Starting from an initial aniso­

tropic state of stress, the failure envelope can be

approached following different stress paths, which depend

on the nature and magnitude of the applied loads, and on

the boundary conditions. These effects are discussed

below.

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
194

3.4.1 Level Ground Conditions


o
As mentioned earlier, in this case the initial state

of stress is anisotropic to a degree that depends on the

value of the coefficient of earth pressure at rest, K q ,

which for NC sands is smaller than 1.0 (115). The hor­

izontal and vertical planes are initially the planes of

principal stress. During an earthquake, alternating shear

stresses act on the horizontal and vertical planes causing

a cyclic rotation of the directions of principal stress

about a vertical axis (Fig. 2.6) and corresponding shear

deformations in the soil elements. However, it is clear

that under level ground conditions, which can be approxi­

mately modeled by a half-space, lateral expansion or con­

traction of the soil elements in the horizontal direction

cannot occur. Consequently, the soil elements undergo

shear deformations but without change in the cross section

perpendicular to the direction of shear wave propagation.

The deformations corresponding to the boundary conditions

just described have been termed "simple shear deforma­

tions" (140). Under these conditions, the no lateral

strain boundary plays a major role in the response of soil

elements to cyclic shear stress applications as discussed

below.

The behavior of sand specimens under simple shear

conditions has been studied in the laboratory by means of

cyclic simple shear test or cyclic torsional shear tests

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
195

o In which lateral strain of the specimen is not permitted

(75,81,173,175,197). Typical results from torsional shear

tests in which the specimen cross section remains constant

during cyclic loading are presented in Figure 3.41. The

specimens were subjected to a number of cycles at a con­

stant shear strain amplitude. The following events are

observed: The shear stress amplitude decays progressively

whereas the cell pressure increases gradually as cyclic

loading progresses, and eventually becomes equal to the

initial vertical stress (Fig. 3.41c) at the time the pore

water pressure builds up to the level where liquefaction,

that is, a zero effective stress condition, occurs. At

this stage, the strength of the specimens approaches zero.

c Accordingly, during cyclic loading there is a gradual

change in the ratio of total principal stresses from the

value corresponding to the initial K q value to 1.0,

irrespective of the initial K q value. However, it should

be noted that the rate of pore pressure buildup depends on

the initial K value; the smaller the K value (closer to


o ’ o
failure), the faster the increase in pore water pressure,

and accordingly, the smaller the number of cycles required

for he initiation of liquefaction. Consequently, under an

identical vertical effective stress, a sand deposit with

lower in situ earth pressure coefficient, Kq , will be more

susceptible to liquefaction. Nevertheless, over the range

of Kq in Figure 3.41 the differences are not significant.

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
Reproduced with permission of the copyright owner. Further reproduction
C

1.00 -

S a g a m i River S a n d 0.75
K0= 0.50
Relative density: 4 2 % 0.50
r/'c = 2 0 0 kPa <r3'- K 0 aiJ
1c 0.25 (b) Pore Pressure
Shear Strain ±0.42%
1 3 5 7 10 15 20 30

(a) Variation in Shear Stress Kp= 0.75


prohibited without perm ission.

0.3 Kq= 0.50


K0= 0.33

- (c) Variation in Principal Stress Ratio .

i_ -i-.J— i. i__ j_
20 30 1 3 5 7 10 15 20 30
N U MB E R OF CYCLES N U M B E R OF CYCLES
Figure 3.41 Typical Cyclic Torsional Shear Tests with no Lateral Strain

196
on ^-Consolidated Specimens (After Ref, 75).
197

o Thus, during cyclic undrained tests on anisotropi-

cally consolidated specimens in which lateral displacement

is not permitted to occur, the lateral stress gradually

increases until it becomes equal to the vertical stress.

At this stage, the specimen is subjected to an isotropic

state of stress, a prerequisite for initial liquefaction

to occur. Ishihara and Li (75) stated the problem in

approximately the following terms: 'when a sand element

at a depth z in the deposit is brought to the state of

complete liquefaction, each grain of the sand exists in

the water in the form of suspension, and therefore, the

element apparently behaves as a "heavy" fluid exerting an

ambient pressure, y z both vertically and horizontally.

c Consequently, during the process leading to liquefaction


*
(in which there is a decrease in o by y z and an increase
v
in pore water pressure by the same amount) there needs to

be an increase in total stress by (l-KQ )y'z in the hor­

izontal direction, since the effective horizontal stress

is reduced by KQ y'z but the pore pressure builds up to the

amount y'z. This change in total horizontal stress must

occur for the sand element to reach a state of uniform

compression at liquefaction.' It is clear that this

increase in lateral stress is likely to occur in deposits

under level ground because the horizontal displacement is

prohibited to take place, as observed in laboratory speci­

mens under equivalent boundary and stress conditions.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
198

o It is of interest to compare the behavior of sand

specimens under no lateral strain conditions with the

behavior of sand specimens subjected to cyclic loading

under other boundary and stress conditions. In this con­

nection, Figure 3.42 compares the results of a torsional

shear test with no lateral strain permitted on a sand

specimen anisotropically consolidated, with the results

from two tests, in which the lateral strain of the speci­

mens was not restricted, on specimens consolidated aniso­

tropically and isotropically respectively, at the same

vertical effective stress. It is observed that the gen­

eral behavior of the specimen isotropically consolidated

and the specimen anisotropically consolidated with no

lateral strain during cyclic loading is quite similar in


c terms of shear stress decay, pore pressure buildup and

axial strains. In both tests the pore pressure ratio

approaches 100% while the cyclic strength drops to zero.

On the other hand, when the specimen is anisotropically

consolidated but the lateral strain is permitted during

cyclic shearing, the behavior is completely different; the

measureed shear stress does not change appreciably with

time, and the pore pressure increases much more slowly and

reaches a limiting value smaller than , irrespective of

the number of strain applications. Therefore, in this

case liquefaction does not develop. It is also of

interest to note that the vertical strain it, considerable

when the specimen is permitted to deform laterally.

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
Reproduced with permission of the copyright owner. Further reproduction r

1.00 K0= 0.50

0.75 K= 1.00
S a g a m i River S a n d
Relative density: 4 2 % 0.50
&i'c = 200 kPa " Ko % 0.25 K= 0.50

Shear strain y - + 0 . 4 2 % (b) Pore Pressure


($)No lateral strain 30
© & © Lateral strain permitted
15.0 (c) Vertical Strain

# 12.5
0.5
(a) Variation in Shear Stress

0.4 K= 0.50
prohibited without perm ission.

% r“ <
O 5.0
0.2
I-
oc 2.5
= 1.00 111
>
■ Kn= 0.50 ,

30 1 3 5 7 10 15 20 30
N U MB E R OF CYCLES
Figure 3.42 Cyclic Torsional Shear Tests under Different Boundary and

199
Stress Conditions (After Ref. 75).
200

Because of the constant volume conditions, the lateral

G expansion must he compensated by a contraction in the

vertical direction. The axial strains of the specimen

isotropically consolidated are small since it is not sub­

jected to sustained shear stresses which could induce sig­

nificant lateral expansion.

The effective stress paths for tests with a constant

cyclic shear stress amplitude and the same initial stress

and boundary conditions as in Figures 3.41 and 3.42, are

compared schematically in Figure 3.43. The effective

stress path A represents the process leading to liquefac­

tion of sand deposits under level ground. As the pore

pressure builds up and the lateral stress increases in the

C course of cyclic loading,

down and towards


the effective stress path moves

the left, reflecting continuous reduc­

tions in the static shear stress, until it finally reaches

the zero effective stress state. It is noted that while

the principal stress ratio with respect to total stresses

increases to a value of one during cyclic loading from its

initial value of Kq , the effective principal stress ratio

may remain approximately constant. It is also important

to note that the position of the locus of triggering

states or the reversal of shear stresses (hydrostatic

state of stress) are not relevant to the cyclic undrained

behavior under no lateral strain conditions. Prior to

liquefaction, the specimen remains under an anisotropic

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
Reproduced with permission of the copyright owner. Further reproduction
O

<S> No lateral strain

Lateral strain permitted:


S'
sin <f> -tan Q '
© Isotropic consolidation / a *

© Anisotropic consolidation

Initial state of stress


• for ® & ©
prohibited without perm ission.

<r'
o-(,« k0cr'
'Liquefaction Stress reversal

Figure 3.43 P o t e n t i a l E f f e c t i v e S t r e s s Pa t h s d u r i n g C y c l i c Loadi ng

201
u n d e r D i f f e r e n t Boundary and S t r e s s C o n d i t i o n s (Schematic).
202

stace of effective stress with a stress ratio smaller than

the value corresponding to the critical stress ratio line.

From the foregoing discussion, it is evident that

tests on anisotropically consolidated specimens in which

lateral strain is permitted (stress path C in Fig. 3.43)

cannot be used to determine the cyclic undrained strength

of sand deposits under level ground conditions. As dis­

cussed later, they may simulate the behavior of soil ele­

ments for example close to a slope, a case in which

lateral displacements are not constrained. On the other

hand, tests on isotropically consolidated specimens at the

same vertical effective stress (stress path B in Fig.

3.43) appear to provide a closer but not exact simulation

of the behavior under no lateral strain conditions. The

main differences illustrated in Figures 3.42 and 3.43

between these two types of test may be associated with the

fact that the rate of pore pressure buildup during the

first cycles of shear stress applications may be higher

for the anisotropically consolidated specimen. Moreover,

at the same a , the distance that the stress path has to


v
traverse to reach the origin of the q— p' plot is larger

for the isotropically consolidated specimen. On the other

hand, a complete shear stress reversal occurs each time

the stress cycle through the hydrostatic state in the iso­

tropically consolidated specimen, which favors an increase

in the rate of pore pressure generation. Therefore, due

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
203

o to these counter balancing effects, it is felt that tests

on isotropically consolidated specimens may slightly

overestimate the cyclic undrained strength of sand depo­

sits under level ground at the same effective vertical

stress. This fact is clearly illustrated by the data

shown in Figure 3.44 from stress-controlled torsional

shear tests with no lateral strain permitted on specimens

of Fuji river sand. On the other hand, if the comparison

of strength is made at the same mean confining stress, o ,


in
the cyclic plane strain strength of KQ -consolidated speci­

mens may be larger than the corresponding strength of iso­

tropically consolidated specimens, as indicated by the

data shown in Figure 3.45 for Toyoura sand, which were

c also obtained from stress-controlled cyclic torsional

shear tests. The difference is significant for relative

densities above 70%. This may be related to the effects

of dilation as a result of a higher mobilized friction

angle in the anisotropically consolidated specimens. Con­

sequently, it is very likely that to simulate the behavior

of dense sand elements in deposits under level ground,

simple shear tests or plane strain torsional shear tests

on K -consolidated soecimens cannot be avoided,


o

As shown in Figure 3.41, the rate of pore pressure

buildup and hence the resistance to liquefaction depends

on the initial value of K with increasing susceptibility


o
to liquefaction with decreasing Kq . It is clear that the

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
Reproduced with permission of the copyright owner. Further reproduction P

0.6
Fuji river sand
Dr= 55%
Torsional test
0.5
<r' = 100 kPa

0.4
i

h 0.3
<
cc
CO
CO 0.2
id
prohibited without perm ission.

DC
t—
CO
o 0.1
_J
o
>-
o
1 10 100
N U M B E R O F C Y C L E S T O INITIAL L I Q U E F A C T I O N

Figure 3.44 C y c l i c Undr ai ne d S t r e n g t h from P l a n e S t r a i n T o r s i o n a l Shear

204
Tests ( Ad a p t e d from R e f . 8 1 ) .
Reproduced with permission P

CYCLIC TORSIONAL SHEAR TEST


TOYOURA SAND
of the copyright owner. Further reproduction

a ir - p l u v ia t e d

o Nc= 1 0
1 5 % DA SHEAR STRAIN
bE ^ m c = 9 8 kPa

o
NO LATERAL STRAIN
< Kq-C O N S O LID A TIO N
cc
CO
CO
UJ
DC
h-
CO
O ISOTROPICALLY
_J CONSOLIDATED
o
prohibited without perm ission.

>-
o
40 50 60 70 80
RELATIVE DENSITY, Dr (%)

Figure 3.45 Cyclic Undrained Strength from Torsional Shear Tests at


Constant Mean Confining Stress (Adapted from Ref. 173).

205
206

Kq value in situ is dependent on the stress history of the

deposit. When overconsolidation is induced in a horizon­

tal deposit of sand by the removal of overburden pressure

the lateral deformation is prevented and, therefore, the

coefficient of earth pressure at rest, Kq , increases.

Consequently, the cyclic undrained strength of a sand

deposit which has experienced an overconsolidation process

is likely to be higher than that corresponding to a NC

deposit, not only because of the favorable changes in

fabric and density but also because the Increased K value


J o
(81). It should be noted that a rise in the level of the

ground water table, which also causes overconsolidation of

a sand deposit, does not produce an increase in Kq , as has

, been suggested elsewhere (81).

VJ
Ishihara and Takatsu (81) performed plane strain

cyclic torsional shear tests on overconsolidated specimens

of Fuji river sand with a relative density of 55%. Over­

consolidation was induced by increasing the axial and

radial stresses along stress paths with K0 values of 0.5,

1.0, and 1.5 respectively. The results of these tests are

summarized in Figure 3.46, in terms of the cyclic stress

ratio required to cause initial liquefaction in 20 cycles.

The cyclic undrained strength increase with both increas­

ing Kq and OCR values. When the strength of OC specimens

is normalized with respect to the NC strength for a given

K value, the curves approximately collapse into a single


o

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
207

G >
Initial Liquefaction
Nc- 2 0 Fuji river sand
0.5 Dr - 5 5 %
Torsional tests
6 ^ - 1 0 0 kPa

co
CO
LU
cc
I- 0.2
co
o
0.1 Kg- 0.5
o
>-
o

Overconsolidation ratio, O C R

C 3
Fuji river sand
Initial Liquefaction
NU-20
o D r - 55%

0 V - 100 kPa

Ik 2

S i OCR

0
Overconsoiidation ratio, OCR

Figure 3.46 Effect of Overconsoiidation on Plane Strain


Cyclic Undrained Strength (After Ref. 81).

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
208

one (Fig. 3.46b), which seems to suggest that cyclic


o undrained strength of 0C specimens increase with the

square root of the overconsoiidation ratio, OCR, irrespec­

tive of the K value. It is not yet known whether this


o
relationship is valid at other relative densities or for

other types of sand.

3.4.2 Sloping Ground Conditions

Under sloping ground, sand elements also exist under

an anisotropic state of stress. However, in this case

shear stresses act on horizontal and vertical planes and

hence these planes are not planes of principal stress.

Moreover, close to sloping ground soil elements must

c remain under an anisotropic state of stress in the course

of cyclic loading, due to the need of maintaining equili­

brium of shear stresses imposed by geometry. Lateral

deformations are not constrained and therefore, under the

superimposed cyclic shear stresses the soil elements may

undergo shear deformations and lateral displacement as

well. Consequently, the boundary conditions close to

sloping ground are such that the initial shear stresses

are likely to remain unchanged during the application of

cyclic loading. The presence of these initial shear

stresses have a major effect on the response of soil ele­

ments to the superimposed cyclic shear stresses as dis­

cussed below.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
209

Sloping ground conditions and those where the pres-


C f
ence of a structure produces Initial shear stresses on

horizontal planes in a deposit under level ground have

been simulated in the laboratory by means of cyclic triax-

ial or torsional shear tests on anisotropically consoli­

dated specimens in whi ch lateral deformations are permit­

ted. However, it is important to note that neither one of

these two tests provides a complete simulation of what

might be considered a simplification of the in situ condi­

tions. This is illustrated in Figure 3.47. The idealized

in situ state of stress during cyclic loading is charac­

terized by the following features:

a. The vertical and horizontal planes are not planes of


C principal stress under either cyclic or static load­

ing,

b. The maximum applied cyclic shear stress and the max­

imum static shear stress act on different planes, and

c. When the cyclic shear stress acts in a direction

opposite to the direction in which the static shear

stresses on the horizontal and vertical planes act,

the maximum shear stress will be reduced. Accord­

ingly, the corresponding Mohr's circle becomes

smaller. The reverse is true when the cyclic and

static stresses act in the same direction.

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
210

(
T

Static state of stress

a) Idealized Field Conditions

+■ T cy
T - T cy

Pole//45'

C
Consolidation state

b) Triaxial Test S p e c i m e n s

+T

Consolidation state

c) Torsional Test S p e c i m e n s

Figure 3.47 States of Stress under Cyclic L o a d i n g .

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
211

d. If Che cyclic shear stress amplitude is larger than


o the static shear stress on the horizontal and verti­

cal planes, a shear stress reversal will occur on

those planes. However, at the stress reversal on the

vertical and horizontal planes, the soil element

still remains under an anisotropic state of stress


* *
with a minimum shear stress equal to ( a - o^)/2.

Therefore, the maximum shear stress never reverses in

spite of the rotation of the plane on which it acts.

The corresponding rotation of the planes of principal

stress is limited such that the angle between the

direction of the major principal stress and the vert­

ical axis is always smaller than 45° (Fig. 2.6).

Accordingly, the total rotation of the planes of

c principal stress is much less than 90°, which pre­

cludes a complete shear strain reversal. Note that


^ *
if a = o, stress reversal on the vertical and hor-
v h
izontal planes bring the soil element to an isotropic

state of stress.

In cyclic triaxial tests on anisotropically consoli­

dated specimens, an alternating deviator stress of equal

positive and negative value is added to the initial or

static deviator stress, producing a cyclic shear stress

which is symmetrical about the initial value. It is to be

noted that the horizontal and vertical planes are the

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
212

planes of principal stress throughout the loading stages.


O Moreover, the cyclic and the maximum consolidation shear

stress act on the same planes, which are 145° planes,

irrespective of their magnitudes. Additionally, when the

cyclic deviator stress is equal or greater than the static

deviator stress, the specimen is cyclicly brought to an

isotropic state of stress, which is followed by a 90°

rotation of the planes of principal stress. In situ this


* *
would occur only in the case of a = c^. Consequently, it

is evident that' a great deal of judgment must be exercised

if cyclic triaxial tests are used for evaluating the

cyclic undrained strength of sands close to sloping

ground. The significance of stress reversal or non­

c reversal in cyclic triaxial specimens, which has received

a great deal of attention by many investigators

(17,121,155,184,186,193) is thus irrelevant if the in situ


* ^
state of stress is such that and are not equal, a

case in which an isotropic state of stress cannot be

reached, irrespective of the magnitude of the applied

cyclic shear stresses. Stress reversal in cyclic triaxial

tests leads to an isotropic state of stress. As mentioned

earlier, this is a prerequisite for "initial liquefaction"

and very large deformations to develop. Accordingly,

results from cyclic triaxial tests in which stress rever­

sal occurs are meaningless for evaluating the undrained

stability of a cut slope or the outer part of an embank­

c ment in cases where the in situ state of stress remains

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
213

anisotropic, and therefore initial liquefaction is not a


0
possibility.

On the other hand, stress conditions in torsional

shear tests specimens become closer to the idealized field

loading conditions. The maximum consolidation shear

stress and the applied cyclic shear stress act on dif­

ferent planes. A shear stress reversal on the horizontal

and vertical radial planes does not produce an isotropic

state of stress. The specimen remains subjected to the


^ *
maximum consolidation shear stress given by ( a - a )/2.
z r
Even initial or static shear stresses can be applied on

the horizontal section of the specimen (43,70), which per­

mit a complete simulation of the assumed in situ condi­

c tions. However, in most cyclic torsional shear tests the

torsional loading is superimposed to the anisotropic state

of stress created by the axial and radial stresses, which

results in the stress conditions indicated in Figure

3.47c.

Based on the previous considerations, the cyclic

undrained behavior for sloping ground conditions will be

illustrated for two extreme cases:

1. Case A: The initial state of stresses is anisotropic


*
but a * a,
v h

c
v

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
214

2. Case B: The initial state of stress is anisotropic


*
O with respect to principal stresses but ®v “ •

Nevertheless, it should be kept in mind that changes

in the initial state of stress, including rotation of the

planes of principal stress may occur as a result of shak­

ing during an earthquake. This is attributed to the ten­

dency towards densification and slips at the grain con­

tacts and the corresponding stress changes depend on the

magnitude of the applied loads, the initial state of

stress, and the boundary conditions. For example, the

difference between a and may change even if the static

shear stress remains unchanged. Determination of the way

in which the state of stress actually changes within an


r
earth structure in the course of earthquake loading is a

goal that still remains to be achieved.

Case A: Anisotropic state of stress with o * a, .


* v h
It is clear that the response of anisotropically consoli­

dated specimens to shear stress applications is determined

by the combination of at least the following factors:

a. Initial principal stress ratio or magnitude of static

shear stress

b. Amplitude of the cyclic shear stress applied

c. Location of the locus of triggering states (CSR line)

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
215

d. Magnitude of the steady-state strength


G
Three different conditions can be identified.

1. The magnitude of static shear stress Tg , plus the

cyclic shear stress, TC y> is smaller than the undrained

steady-state strength. Under this condition, the behavior

of anisotropically consolidated specimens during cyclic

loading has been conclusively observed to be as follows

(75,155,184): as cyclic loading proceeds pore pressure

and strains accumulate gradually reaching eventually a

terminal or residual value. The pore water pressure cycles

above and below the terminal value but it never approaches

the confining pressure. The effective stress path

C corresponding to this behavior is illustrated in Figure

3.48. As the pore pressure builds up, the effective

stress path moves horizontally towards the left at a con­

stant distance from the p" axis reflecting the presence of

a constant static shear stress. The rate of strain accu­

mulation increases as the effective stress path crosses

the phase transformation line and approaches the failure

line. However, with further cycling the effective stress

path stabilizes along the failure line and strain accumu­

lation slows down by virtue of dilative tendencies (185).

It can be noted that in Figure 3.48 the residual or termi­

nal pore water pressure, u ^ , depends on the location of

the initial state of stress with respect to the failure

('

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
Reproduced with permission C

sin » tan Q *
Locus of triggering states
of the copyright owner. Further reproduction

(CSR lino)

Steady-state strength

'ss
Consolidation
Phase transformation state of stress
line--------------------- v

static \ ” rs
prohibited without perm ission.

er

Figure 3.48 Torsional Shear Test on Anisotropically Consolidated


Specimen with (t +x ) Smaller than Steady-State

216
Strength (Schematic^.
217

line. For small cyclic shear stress amplitudes, is


O given approximately by the expression

c K -1 1-sin
, 4.
_,, > „ ,„
Ur = 3c " “ 2“ "sinjT > ( 3 *7)

where a. and K are the confining pressure and the prin-


jc c
^ *
cipal stress ratio, a / a , at consolidation, respec-
1c 3c
tively. The residual pore water pressure decreases with

increasing Kc at a constant confining pressure.

As the initial static shear stress increases (larger

K c ) the closer the specimen approaches a failure condi­

tion. Accordingly, more strains accumulate in the initial

stages of loading. However, the higher the initial shear

stress, the shorter the distance the stress path need


c
traverse to reach the failure line, and hence less oppor­

tunity for strain accumulation (121). Moreover, the rate

of strain accumulation after the stress path crosses the

phase transformation line may slow down as the level of

static shear increases (185), which may be due to stronger

dilative tendencies. Consequently, the cyclic undrained

strength in terms of the number of cycles at a given

stress amplitude required to induce a fixed strain level

in the specimen may indeed increase slightly as the ini­

tial shear stress increases as long as (t + x ) remains


s cy
smaller than the undrained steady-state strength. Under

this condition there is no stress reversal, hence the rate

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
218

o of pore pressure generation is considerably smaller than

in isotropically consolidated specimens subjected to simi­

lar loading conditions.

2. The magnitude of the static shear stress Tg is

smaller than the undrained steady-state strength but

(Tg + TCy) larger than the undrained steady-state

strength. C on se que nt ly , under this condition, in the

course of cyclic loading the effective stress path reaches

the locus of triggering states as indicated in Figure

3.49. Accordingly, at this stage the specimen will exhi­

bit a strain softening response reaching finally a

steady-state of deformation. The specimen remains at the

steady-state condition because it cannot build up a

c strength up to the cyclic shear stress amplitude at which

the reversal of direction of deformation is predetermined

to occur. In a corresponding in situ condition, it is

likely that the continuous deformations stop once the

shear stress level drops below the undrained steady— state

strength. It is important to note that as the initial

shear stresses increase, the number of cycles of a given

stress amplitude required to bring the stress state of the

specimen to the critical stress ratio line decreases.

Consequently, the cyclic undrained strength decreases sig­

nificantly as the initial shear stress increases.

3. The static shear stress is larger than the

rV undrained steady-state strength. This is a very critical

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
Reproduced with permission of the copyright owner. Further reproduction r \ n

Locus of triggering states


(CSR line)

S teady-state strength

SS

Consolidation-
state ot stress static \“‘s
prohibited without perm ission.

Figure 3.49 Torsional Shear Test on Anisotropically Consolidated

219
Specimen with (t + t ) Larger than Steady-State
Strength (Schematic^.
220

condition. Small cyclic stress applications may suffice


o to bring the effective stress state of the specimen to the

CSR line causing the initiation of a flow failure (Fig.

3.50). After the collapse of the structure, the specimen

reaches a steady-state of deformation under the driving

static shear stresses. It is to be noted that the

phenomenon is not avoided even if cyclic loading is

stopped. In an equivalent in situ condition, after break­

down of the structure is triggered in the course of cyclic

shear stress applications, the mass of sand will flow,

spreading out until the shear stresses acting within the

soil mass become small enough to be compatible with the

reduced steady-state strength (14,132). It is important

to note that a very small cyclic stress ratio may be


c enough to cause a flow failure if the initial static shear

stress is high enough. Accordingly, this phenomenon is

considered a kind of spontaneous flow failure since the

extent of the failure may appear disproportional as com­

pared to the magnitude of the triggering disturbance

(14,110).

From the foregoing discussion, it is evident that in

sloping ground conditions the main design concern is to

establish whether or not a flow failure can be triggered

in the course of cyclic loading. Therefore, the analysis

Involves the evaluation of the following engineering pro­

perties :

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
Q.
C
o
CD
Q.
with permission of the copyright owner. Further reproduction

sin < fj - tan O t

Locus of triggering states


(CSR line)

Steady-state condition
prohibited without perm ission.

Consolidation - static \ -
state of stress

O ’’ cn

Figure 3,50 Torsional Shear Tests on Anisotropically Consolidated

221
Specimen with Static Shear Stress Larger than Steady-State
Strength (Schematic).
222

a* the steady-state strength,

b. the location of the locus of triggering states (CSR

li n e ) ,

c. the existing static shear stresses needed to sustain

equilibrium, and

d. the flow failure potential.

Flow failure potential is the relationship between

the number of cycles and the cyclic stress (or strain)

amplitude required to trigger the flow failure. This

should be determined whenever the total shear stresses

(cyclic + static) are expected to be larger than the

steady-state strength . The only flow failure potential

curve known to the author is presented in Figure 3.51

obtained from strain controlled cyclic torsional shear

tests on anisotropically consolidated specimens of three

silty subangular sands (34).

The exact determination of the locus of triggering is

extremely important for practical applications, as this

determines how much pore pressure buildup can occur in a


• J ^ A a. 1. _ .C _ _ J a. > J «.U ^
g X V CU O V X X CXCIUCUU UCX.WXG XU 9 0 u i c u g u u u x u p o UV uuc

steady-state strength (34). Once the triggering envelope

is known the maximum allowable residual pore pressure

increase which can be caused by the earthquake itself

before triggering, uT , is given by (34),

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
223

Sand A - 1 3 % silt
05 Sand B - 32 % silt
Sand C - 6 3 % silt
05
05

100 -

Silty Sands A.B.C


CyT - CAU Tests

u. Kc - 2

0.01 0.1 1

Cyclic S h e a r Torsional Strain, yQ , %

Figure 3.51 Flow Failure Potential from Strain


Controlled Cyclic Torsional Shear Tests.

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
where $ is Che mobilized friction angle at the triggering
m
envelope or CSR line. From E q . 3.7» it is seen that

increases as K decreases for a given consolidation con-


c
fining pressure.

An important aspect of the behavior is the effect of

the consolidation stress ratio, K on the locus of


c
triggering line and the corresponding steady-state

strength, which previously were evaluated from tests on

isotropically consolidated specimens. Based on his tests

on Banding sand, Castro (14) concluded that at a constant

void ratio the steady-state strength was not affected by

the consolidation stress ratio with Kc in the range from

1.0 to 2.0. However, this conclusion was based on a lim­

ited number of tests. Castro did not draw any conclusion

regarding the triggering envelope. Vaid and Chern (186)

based on their tests on tailings sand concluded that both

the mobilized friction angle at the peak strength and the

steady-state strength were not affected by the consolida­

tion stress ratio, with K£ also in the range between 1.0

and 2.0. The author finds it rather difficult to visual­

ize these results. A greater rearrangement of grains is

likely to occur with increasing mobilized friction angle

(higher Kc ) causing changes in the contractive charac­

teristics of the sand skeleton. For example, it seems

unlikely that under a very high stress ratio (Kc > 2.0) a

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
225

j sand specimen exhibits the same fabric than a correspond-


( )
ing specimen isotropically consolidated at the same void

ratio. It has been shown conclusively that different

fabrics, at the same Dr. affect the undrained and drained

strength characteristics of sands. As the effects of

anisotropic consolidation on the mobilized friction angle

and the steady— state strength are imperfectly known, in

the author's opinion, it is too early to conclude that

they are independent of the consolidation stress ratio.

More effort should be dedicated to this topic in future

investigations on the undrained behavior of sands.

Although the triggering envelope and the steady-state

strength can be approximately determined from undrained

c monotonic tests, cyclic torsional shear tests on aniso-

tropically consolidated specimens are required to define

the flow failure potential in terms of the number of

cycles of a given stress amplitude for triggering the col­

lapse (34). It is clear that for the occurrence of a flow

failure under a given series of stress applications three

conditions should be satisfied:

1. The initial state of stress is such that the specimen

is susceptible to strain softening behavior,

2. the static shear stress is larger than the undrained

steady-state strength corresponding to the initial

c state conditions, and

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
226

3. the number of stress repetitions is high enough that


C the effective stress path reaches the locus of

triggering states or CSR line.

As the relative density of the sand increases, the

level of confining stress at which the specimen exhibits a

strain softening behavior is significantly higher; so is

the corresponding steady-state strength. Accordingly, the

possibility of a flow failure is greatly reduced, provided

t is smaller than t . The behavior may be limited flow


s ss
deformation at moderate confining pressures or completely

dilative of low confining pressures . As mentioned ear­

lier, it is likely that for dilative sands the cyclic

undrained strength increases with the consolidation stress


C ratio. Consequently, at low confining pressures the

cyclic undrained strength increases as the consolidation

stress ratio increases. At moderate confining pressures,

as the behavior may be limited flow deformation, hence,

the cyclic undrained strength specified at small failure

strain level decreases with increasing initial static

shear stress but increases with xg at large specified

failure strain levels for the range in which xg is larger

than the steady-state strength (186). On the other hand,

at high confining pressures the behavior becomes of com­

plete flow failure and therefore, if xg is larger than the

steady-state strength, the cyclic undrained strength

c always decreases with increasing shear stress ratio since

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
227

the initial state of stress is closer to the triggering

O envelope and a smaller number of cycles of a given ampli­

tude, or smaller cyclic stress amplitudes for a given

number of cycles, will be required to cause the initiation

of a flow failure. Not only the cyclic undrained strength

decreases with increasing initial shear stress if limited

or complete flow failure develops, but also the deforma­

tion associated with the strain softening response as well

as the loss in shear strength until the steady— state is

reached increase with increasing shear stress (186).

Case B: Anisotropic state of stress with a = c, .


v h
In this case, the maximum consolidation shear stress and

the cyclic shear stress act on the same plane. Conse­

C quently, if the cyclic shear stress is larger than the

static shear stress a stress reversal occurs bringing the

specimen momentarily to an isotropic state of stress.

Large strains develop easily when stress reversal occurs,

which is a prerequisite for the occurrence of transient

liquefaction states (states of zero effective stress).

Therefore, in addition to the conditions analyzed for the


^ *
case Ov * a^, there exists the possibility of stress

reversal (isotropic state of stress) when the vertical and

horizontal effective stresses prior to cyclic loading are

equal. This is illustrated in Figure 3.52. It is evident

that the possibility of stress reversal is enhanced as the

initial shear stress becomes smaller. Apparently, this

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
Reproduced with permission of the copyright owner. Further reproduction rs c

Locus of triggering states

(CSR line) \
q

Phase trasnsformation line

•Pole

VH - T
prohibited without perm ission.

Stress reversal

In situ state of stress

228
Figure 3.52 In S i t u E f f e c t i v e S t r e s s Pat h w i t h S t r e s s Reversal during
C y c l i c Loa d i n g (Schematic).
229

might be the only case, although rare, for which the


o cyclic triaxial test is likely to provide some useful

information on the effect of the initial static shear

stress on the cyclic undrained strength.

As the initial static shear stress increases, there

is a decrease in the degree of reversal for the same

cyclic shear stress. This should always result in lower

strains, and thus in an increased cyclic strength. A

further increase in may result in the elimination of

the stress reversal and in a greater increase in the

cyclic strength. However, once the total shear stress

(static + cyclic) becomes larger than the steady-state

strength the cyclic undrained strength decreases with

c increasing static shear stress due to the increased poten­

tial of a flow failure. Consequently, for those cases in

which stress reversal is a possibility, the relationship

between the cyclic undrained strength and the initial

shear stress should have a clear peak or turning point

reflecting first an increased strength due to a reduction

in the degree of reversal and later a decreasing cyclic

strength as the flow failure potential increases. The

effects of initial shear stress on the cyclic strength of

tailings sand is illustrated in Figure 3.53 at two levels

of consolidation confining pressure. It should be noted

that if flow deformation develops strains well in excess

of 2.5% will develop.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
230

O
0.6 -
(a) Tailings Sand
^ 3 C- 2.0 k g f/c m *
0.5
ec « 0.801 ± 0 . 0 0 2
Drc - 6 9 .5 ± 0.5%»
0.4

0.3
Liquefaction
0.2
N -1 0

0.1

0.20 .Liquefaction Lim ited strain softening

(b) Tailings Sand


c ^ 3 C- 16.0 k g f/c m

ec- 0.800 ± 0 .0 0 6
Drc - 7 0 .0 ± 1 . 6 %

N -1 0

1.5 2.0
K,

Figure 3.53 Effect of Static Shear Stresses on the


Resistance to Strain Development under
Cyclic Triaxial Loading Conditions
(After Ref. 186).

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
231

CHAPTER 4

LABORATORY EQUIPMENT AND TESTING PROGRAM

4.1 New Resonant Column/Torsional Shear Test Apparatus

The resonant column and torsional shear tests have

been used separately to obtain the variation in shear

modulus and damping ratio over the range of shear strain

amplitude required for dynamic analysis

(36,38,59,68,88,171). The use of two separate pieces of

testing equipment implies that the dynamic properties of

the tested material, over the range of shear strains of

interest, are determined on different specimens, which are

unlikely to have identical characteristics. Moreover,

very often there is no overlap between the ranges of shear

strains for which results are available from each test.

This is a significant limitation when investigating either

the reduction of shear modulus as a function of shear

strain amplitude or the existence of a critical shear

strain at which the fabric of the sand eventually yields

during a given loading sequence. These drawbacks led to

the design of a new torsional shear apparatus which com­

bines the features of both the resonant column and tor­

sional shear devices and permits the determination of soil

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
232

properties over the whole range of strains of engineering


<J interest as well as a better simulation of the stress

paths associated with earthquake loadings.

The main part of the apparatus was built by Soil

Dynamics Instruments (166), based on ideas and specifica­

tions provided by the author's advisors, Professors G. A.

Leonards and J. L. Chameau. The apparatus, which is

shown in Figure 4.1, has the capability of testing solid

or hollow specimens, consolidated either isotropically or

anisotropically. The solid specimens are 71 mm in diame­

ter. The hollow specimens may be constructed with 71 mm

O.D. and 38 mm I.D., or 100 mm O.D. and 71 mm I.D. The

specimen height is approximately 200 mm. These dimensions

c satisfy the criteria recommended by Saada et a l .

for dimensioning hollow cylinders to reduce the effects of


(140)

end restraints.

The torsional cyclic loading may be either dynamic or

quasi-static, combining into the same device the features

of a resonant column test apparatus and a torsional shear

test apparatus. The dynamic and quasi— static loadings are

controlled independently and may be applied simultane­

ously. In either case, the torsional loading is applied

to the top of the soil specimen which is rigidly fixed at

its base.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
233

Air piston

Stepper
motor system

Hardin
oscillator

A ir piston

c Pressure
control panel
control

Vacuum
regulator

a) Gene r a l view

Figure 4.1 New Resonant Column/Torsional Shear Test


App aratus.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
234

c Air Piston
LVDT
Signal
Cond.

LVDT
Stepper motor
system
Rotating
shaft

Hardin
oscillator
Proximity
i transducers
Platform

c ham b e r

c L o ad cell

Torque
Differential
pressure
> transducer m transducer

Analog control

i 1
MACSYM 350
Stepper motor control Signal conditioning
Data acquisition
system

b) S c h e m a t i c

Figure 4.1, continued.

R eproduced w ith perm ission o f the copyright owner. F urth er reproduction prohibited w itho ut perm ission.
235

^ Confining pressures up to 700 kPa can be applied

inside the Lucite chamber by means of air, water or any

other fluid as a confining medium. The effective confining

pressure is measured by means of a 350 kPa differential

pressure transducer located at the base of the apparatus.

Anisotropic stress conditions are applied by adjusting the

pressure in a double acting air cylinder mounted on a

frame on top of the chamber lid, a modification to the

initial design that was added after proof testing. The

cylinder can apply either an upward or downward force to

the loading shaft. The shaft of the actuator follows any

change in length of the specimen and mates with the rotat­

ing shaft for applying torque through a bearing system

which allows free rotation. The axial load on the speci­


C men is measured by means of a 3.3 kN load cell located

within the chamber, underneath the bottom platen. In this

way friction related problems are avoided and the actual

axial load applied to the specimen is effectively meas­

ured. Changes in the length of the specimen during any

stage of the test are measured by a linear variable dif­

ferential transformer (full scale displacement 25 mm),

LVDT, located on a yoke just above the bearing connection

between the piston shaft and the rotating shaft.

The apparatus has special provisions for volume

change measurements and flushing and backpressuring of the

specimen. Inert ceramic porous discs are used at the face


r
v

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
236

of each platen to distribute or collect pore fluids.

o These porous stones have coarse surfaces and are screwed

into the platens to avoid any slippage during torsional

loading. The bottom platen has two drainage lines, one of

which connects to the differential pressure transducer.

The other line connects to a three-way valve allowing CO^

gas purging followed by water saturation. The top platen

has only one drainage line, which also connects to a

three-way valve, which in turn connects to a vacuum line

and a saturation water line. The vacuuming system for

holding the specimen during preparation is set up in such

a way that CC>2 can be percolated into the specimen as it

is prepared. This procedure facilitates the process of

saturation of the specimens under a reduced backpressure.

c During consolidation, drainage can be allowed through

either or both platens, and the volume change measured in

a burette.

Other features of the apparatus associated with

either the resonant column or the torsional shear modes

are described below.

4.1.1 Resonant Column System

A Hardin type electromagnetic oscillator is used to

apply torsional vibrations that vary sinusoidally with

time. The oscillator is essentially a single-degree of

freedom system in which a spool-shaped spring couples a

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
central mass and the specimen cap attachment to a hollow

cylindrical mass with very large rotational inertia. A

sine wave generator capable of producing a sinusoidal vol­

tage with frequency varying from 5 Hz to 600 Hz is used to

power the coils and permanent magnets of the oscillator,

which are wired to produce a torque about the axis of the

specimen. The large inertia mass of the oscillator essen­

tially provides a fixed reaction such that the forcing

torque produces vibration of the central mass and specimen

cap. The output of the sine wave generator is amplified

electronically to provide enough power to induce vibra­

tions of the specimen with the shear strain, y, ranging


-4
between minimum and maximum values of about 10 % and 1.5

c x 10 respectively. The minimum strain limit is

determined by the sensitivity of the displacement measur­

ing system, whereas the maximum strain limit is controlled

by the stiffness of the soil specimen and the torque capa­

city of the oscillator.

The torsional motion of the soil cylinder is moni­

tored by a piezoelectric accelerometer mounted on the top

platen at a known distance from the axis of rotation. The

accelerometer produces an electric charge which is propor­

tional to acceleration. A charge amplifier is used to

convert the electrical charge to voltage and then it is

read with a conventional voltmeter. Since the motion is

sinusoidal, the displacements are deduced from the meas-

c ured acceleration values.

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
238

The frequency of excitation is adjusted to produce


o
resonance of the system composed of the specimen and its

attached platens and the vibration excitation device. The

phase relationship describing the resonance condition is

established by observing the Lissajous figure formed on an

X —Y oscilloscope by the torque and acceleration signals.

For a fixed bottom specimen, first resonance occurs at the

frequency for which the sinusoidal excitation torque and

the acceleration signals are 90° out of phase (40). In

this case, the trace on the screen of the oscilloscope

corresponds to an ellipse with vertical and horizontal

axes (40,57).

The shear modulus and damping ratio are determined

c from the vibration response of the system. The resonant

frequency of the system is used to calculate the shear

modulus, G. Damping is determined from either the magnif­

ication factor at resonance or the decay in amplitude of

the free vibration of the system. In the first case, the

driving voltage and thus the input torque are known by

measuring the voltage drop across a 5 Q precision (tem­

perature and frequency stable) power resistor in series

with the driving coils. Detailed information on the data

reduction procedure and calibration factors of the

resonant column test is presented in Appendix A. A gen­

eral view of all the electronics associated with the

resonant column test is shown in Figure 4.2 and a wiring

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
Reproduced with permission of the copyright owner. Further reproduction
e

Power am plifier

C h arg e
am plifier

F req u en cy
m eter -------

O scilloscop e
prohibited without perm ission.

C o n tro l box Sine w ave g e n e ra to r

239
Figure 4.2 Electronic S y s t e m for R e so na nt Co lum n Tests.
240

o schematic for the resonant column system is given in Fig­

ure 4.3.

4.1.2 Torsional Shear System

In the torsional shear mode, a quasi-static load is

applied to the specimen by a torque motor system which

consists of a stepper motor and a gear reduction system

with reduction ratios of 200 to 1 and 3200 to 1. The

motor system is mounted on the pressure chamber lid.

Gears of the reduction box match gears on the rotating

shaft which passes through the chamber lid and is rigidly

fixed to the inertial mass of the Hardin oscillator. The

latter in turn is fastened to the top platen by means of a

c special coupling. During assembly, the oscillator remains

attached to the shaft, which is locked to the chamber lid.

After the chamber lid is put into place, the lock is

released, and the oscillator head is lowered to meet the

top platen. The shaft is again locked and the coupling is

completed with a tool inserted through the center of the

piston, which is hollow. Hence, rotation of the shaft

causes the entire Hardin oscillator and the upper platen

to rotate with it. Since the torsional oscillator has

some torsional stiffness of its own, a small amount of

torsional energy will be stored in the torsional oscilla­

tor, which is of little practical consequence (166). A

torque transducer, which is mounted within the chamber

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
241

O
Audio Bangs
Oscillator Oscilloscope

Digital
Freq.Meter Vert. Horiz.
In. In.
Power
Amplifier

Switch Box

Digital
Multimeter

To Drive Coils

Charge To Accelerometer
Amplifier

Figure 4.3 Wiring Diagram for Hardin Resonant


Column Apparatus.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
24 2

base, allows measurement of the torque applied to the


O specimen up to a maximum of 56*5 N-m. The bottom platen

rests on top of the load cell which in turn is screwed

into the head of the torque transducer. This sensing head

is not in contact with the chamber base and hence, the

actual torque applied to the specimen is accurately meas­

ured.

The relative rotation between the bottom and the top

of the specimen is measured by two non-contacting dis­

placement transducers located on a cylindrical platform

which is fastened to the lower platen of the specimen.

The transducers focus on non— concentric circular cams

attached to the top platen and measure the distance

between the cams and the transducer face. As the top pla­
c
ten rotates relative to the bottom, the distance changes,

which causes the transducer output to change. Output from

these transducers is linear for small rotations but varies

with the sine of the rotation angle for larger values of

rotation. This system is designed to be unaffected by

lateral translation of the top platen or by length changes

of the specimen up to approximately 20 mm. It allows

measurement of rotation angles up to a maximum of 30° in

either direction from the at rest position of the speci­

men.

Control for the stepper motor is provided by a motor

js- control unit and an analog control. The stepper motor


V.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
24 3

o control unit has provision for controlling the rate and

direction of rotation of the stepper motor either manually

or automatically. A local speed control allows for mono­

tonic loading in either direction, clockwise or counter­

clockwise, with a rate of rotation adjustable over a wide

range. The torsional loading continues until it is manu­

ally shut off. In the automatic mode the rate of rotation

is selected by a rotary switch. The stepper motor moves

in 1.8 degree steps, or 0.9 degree steps, depending on

whether full or half steps are applied. The rotary knob

selects the number of pulses (or steps) per second (PPS)

which are applied to the stepper motor. Since the gear

reduction ratios are known, these PPS translate directly

c to rates of rotation of the top platen as follows,

steps degrees
Rotation rate -------- », (4.1)
sec step gear ratio

Further, since the geometry of the specimen is known

and will not change very much, the values of PPS also are

related to rates of shear strain. The PPS values can be

varied as a power of 2 from 4 to 2048 pulses per second

allowing changes in the shear strain rate between


-3
4 x 10 % to 2% per second.

In the automatic mode, the motor control unit is used

in conjunction with an analog control, which triggers the

reversal of the rotation of the motor at preset levels of

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
244

either rotation (shear strain) or torque (shear stress).


o The rotation continues through the neutral position to the

same value of rotation or torque in the other direction.

Thus, the apparatus is capable of performing both strain

controlled and stress controlled tests. In the strain

controlled mode, reversal may be preset at different rota­

tion angles by adjusting a rotary knob on the analog con­

trol. The maximum amplitude of the twist angle, 6DaX ,

depends on the gear ratio. It is 2.88° for the reduction

ratio of 3200 to 1 and 28.8° for the gear ratio of 200 to

1. The rotary knob on the analog control selects a vol­

tage divider, VD, to reduce the peak rotation as a power

of 2 from 1 to 32 times. The corresponding period of

c. loading will depend on the gear ratio, the number of

pulses per second, and the voltage divider and can be

obtained as

(4)(0 )
_ . , , sec N max Gear Ratio
Period (cy c l e ) - (VD) x steps ..de gr ee ,( 2)
sec step

For example, in order to obtain a period of about 10

to 12 seconds per cycle, the PPS and VD switches should be

set as indicated in Table 4.1. Depending on the specimen

size, the rotation range in Table 4.1 corresponds approxi­

mately to a single amplitude shear strain spanning from


_2
1.8 x 10 % to 10%. Thus, the Hardin oscillator and the

torsional shear system essentially overlap in their strain

R eproduced w ith perm ission o f the copyright owner. F urth er reproduction prohibited w itho ut perm ission.
24 5

Table 4.1 Switch Settings for Peak Rotation

Gear ratio » 3200


Motor deg/step = 0.9 (half-step)
Cycle time = 10 sec.

Peak Rotation PPS Voltage Divider

2.88 _
1
1 .44 2048 2
0.72 1024 4
0.36 512 8
0.18 256 16
0.09 128 32

Gear ratio = 200


Motor deg/step = 0.9 (half step)
Cycle time = 12.5 sec.

Peak Rotation PPS Voltage Divider

28.8 2048 1
14.4 1024 2
7.2 512 4
3.6 256 8
1 .8 128 16
0.9 64 32

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
24 6

, amplitude ranges and provide a range in strain amplitudes

from 10_ 4 % to 10%. However, the frequency of loading is

very different in the two tests. In the torsional shear

mode, the maximum frequency is on the order of 0.2 Hz,

whereas the frequency at resonance in the resonant column

test varies with shear strain amplitude and confining

pressure and in general is greater than 100 Hz.

In the stress controlled mode, the output of the

torque transducer is amplified electronically and input

back into the analog control. A torque limit set dial on

the analog control unit is calibrated to cause torque

motor reversal at any level of voltage between zero and

ten volts. The dial of the ten turn potentiometer is

^ divided into 1000 units. Hence, each division corresponds

to 0.01 volts. By setting the dial equal to the voltage

level associated with any desired stress level within the

limits of the apparatus, motor reversal will occur at

approximately those levels.

Figure 4.4 displays the main components of the tor­

sional shear system. The stress and deformation limits

that can be achieved with the apparatus are summarized in

Table 4.2, according to specimen size. Detailed informa­

tion on the data reduction procedures is presented in

Appendix B.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
R eproduced w ith perm ission o f the copyright owner. F urther reproduction prohibited w itho ut
perm ission.
248

Table 4.2 Stress and Deformation Limits of New Apparatus

O
Specimen Solid Hollow Hollow

Outer diameter (mm) 71.1 71.1 101 .6

Inner diameter (mm) 38.1 71.1

Length (mm) 203.2 203.2 203.2

Maximum cell pressure, or (kPa) 700 700 700

Maximum deviator stress


( °z ” °r^ * (kPa> 840 1178 807

Maximum axial strain (%) 12.5 12.5 12.5

Maximum shear stress,


tz8 (kPa) 533 690 310

Maximum single amplitude


shear strain, y 0 (%) 6.1 7.3 11.2
z0

c Table 4.3 Properties of Ottawa 20-30 Sand*

Composition 100% quartz (sj°2 ^

Specific gravity 2.65

Particle shape Rounded

0.71 mm
D 10
0.81 mm
°9 0
Uniformity coefficient cu = 1.1

Minimum void ratio 0.489

Maximum void ratio 0.732

* After McDonald and Raymond (111)

R eproduced with perm ission o f the copyright owner. Further reproduction prohibited w ith o u t perm ission.
24 9

4.1.3 Data Acquisition System


O
Output signals from all the transducers are

transferred directly to a data acquisition system and

recorded synchronously. The system is a MACSYM 350 and

consists of a set of measurement and control devices, con­

figurated around the MACSYM 150 work station, which is a

16-bit microcomputer with integral disk storage and inter­

nal RAM up to 1 megabytes. Specialized Input/Output cards

(Dual Serial Interface cards) plug directly into a six

slot internal backplane of the MACSYM 150 work station.

The purpose of these cards, which are two channel cards,

is to provide serial communication capabilities from the

MACSYM 150 to peripherals, printers, terminals, plotters,

C etc. They also provide analog and digital input and out­

put, RAM memory expansion and a Winchester disk interface.

The other major elements of the MACSYM 350 system are the

MACSYM 200, and the ADIO (Analog/Digital Input/Output)

cards. The MACSYM 200 is an intelligent, general purpose,

precision measurement and control front-end. When using

the MACSYM 200 with the MACSYM 150 work station, the sys­

tem is known as the MACSYM 350. The major components of

the MACSYM 200 consist of a single control board, a six­

teen slot card cage with backplane and a power supply.

The control board has all the power and features of a sin­

gle board computer, combined with precision analog meas­

urement capabilities. Processing power is derived from a

r
V

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
250

5 MHz, 16-bit 8088 microprocessor with up to 64 Kilobytes


G of on-board memory. The system features analog/digital

conversion, variable sampling rates and a capacity of up

to 512 channels of input/output. Analog/Digital

Input/Output (ADIO) cards plug directly into the system

backplane and provide signal conditioning for a wide range

of sensors and transducers. All the cards can be readily

addressed through software.

The output from the sine wave generator is connected

to a frequency input card (FIN01 card) which allows MACSYM

users to interface directly to variable frequency sources

and measure the corresponding frequency in the range from

1.6 Hz to 500 KHz. The outputs from the rotation trans­

ducers (provided by the analog control), the torque trans­


C
ducer (after being amplified) and the LVDT (after being

demodulated), are read through a multichannel analog/input

card (AIM03 card), which handles up to 32 channels of high

or low level electrical signal from sensors, transducers,

transmitters and various instruments. A four— channel

analog/input card (AIM05 card) is used to interface the

system to the differential pressure transducer, the load

cell and the torque transducer which are all four-arm

bridges. This card allows MACSYM users to interface to a

wide variety of strain-gage-type transducers requiring

constant voltage excitation. The board contains four

independent channels providing bridge completion

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
251

c amplification, filtering and excitation of 5 or 10 volts

DC for various bridge configurations. The card allows

direct Interfacing of one, two or four-arm bridges to

MA CSYM main frame systems. The excitation voltage was set

at 10 volts DC for calibration of all the transducers.

The system has been complemented with a dot matrix

printer and a multi-pen plotter. The data acquisition

system allows for immediate data reduction and plotting of

the results of each test, a feature highly advantageous

for comprehensive testing programs.

4.2 Materials and Testing Procedures

Tests were performed on reconstituted specimens of


c standard Ottawa 20-30 sand, which has the properties sum­

marized in Table 4.3. Specimen preparation procedures for

solid cylindrical specimens were similar to those in con­

ventional triaxial testing by using a membrane covered

split mold which forms the outer diameter of the specimen.

For hollow cylindrical specimens a membrane covered man­

drel was used on the interior of the specimen to form the

inner diameter. This interior membrane is sealed to the

upper and lower platens by means of packers which squeeze

an 0-ring against the membrane and hold it to the interior

wall of the platens. The elements used for construction

of hollow cylinder specimens are shown in Figure 4.5.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
252

Figure 4.5 Elements Used for Construction of Hollow


Cylinder Specimens.

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
253

Solid and hollow specimens were prepared by pluviat-


O ing the sand in air through a funnel attached to a glass

tube which extended to the bottom of the mold. Loose

specimens were obtained by keeping the tube full of sand

and slowly refracting it from the mold. Denser specimens

were prepared by reducing the diameter of the glass tube

ahd/or increasing the height of free fall from the tube.

Before taking off the forms, the specimen was subjected to

a low effective confining stress of about 25 kPa by apply­

ing vacuum to the porous space.

A thin steel circumferential tape called a Pi tape,

which can be wrapped around the specimen and is calibrated

to read the diameter directly with an accuracy of 0.001

c inches (0.0254 mm), was used to measure the outside diame­

ter near the top, near the middle, and near the bottom of

the specimen. This technique is easier and is considered

to be a more accurate measurement technique (79,159).

These measurements were averaged to obtain the gross out­

side diameter. The net outside diameter of the specimen

was obtained by subtracting the average double thickness

of the exterior membrane. The Pi tape was also used to

measure the inside diameter in the case of hollow cylinder

specimens. The specimen length was determined by measur­

ing the distance between the edges of the top and bottom

porous stones with a vernier caliper with an accuracy of

0.0005 inches (0.0127 mm).

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
254

4.2.1 Drained Test


O
Drained tests were performed to determine the varia­

tion of shear modulus with shear strain amplitude given

certain initial conditions regarding density, state of

stress, and stress-strain history. After preparation,

each dry specimen was subjected to the desired initial

static state of stress by adjusting the cell pressure and

the pressure applied to the bellofram piston. For some of

the specimens this step included prestressing along a

predetermined stress path.

The specimen was first subjected to dynamic loading

in the resonant column mode. The resonant frequency of

c the oscillator/soil system was measured for oscillations

as small as possible, i.e. for the power applied to the

oscillator as low as practical. The shear modulus of the

soil specimen at very low strains was determined from this

frequency. Then, the applied power was doubled and the

corresponding modulus measured. This operation was

repeated until the maximum capacity of the oscillator was

reached. These experiments provided shear moduli at sin-

gle amplitude shear strains ranging form 10 % to about


_o
1.4 x 10 %. Following the resonant column tests, the

specimens were subjected to strain controlled torsional

cyclic loading with a loading period of about 10 to 20

seconds per cycle. The tests were performed in stages,

c with the amplitude of rotation being doubled between

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
255

, successive loading stages in accord with the values listed

in Table 4.1. During each stage, several cycles (10 in

average) were applied at the predetermined amplitude of

the twist angle. For the first stage, a minimum single


-2 -2
amplitude shear strain between 1.2 x 10 % to 2.0 x 10 %

could be applied, depending upon the size of the speci­

mens. The maximum single amplitude shear strain applied

during the last stage ranged between 3% and 10%.

In some cases, the variation of shear modulus with

shear strain amplitude was determined after the specimens

had been subjected to previous cyclic strain histories of

varying magnitude. In other cases, "virgin" specimens

were subjected directly to cyclic loading in the torsional

^ shear mode in order to quantify the effects of stage test­

ing associated with performing resonant column tests at


-4 -2
shear strains ranging from 10 % to about 10 %. In these

cases, each specimen was subjected to more than 50 cycles

at the chosen shear strain amplitude. Therefore, the

variation of shear modulus with number of cycles at a con­

stant shear strain amplitude could also be evaluated. A

few tests were performed to study the effects of consoli­

dation stress ratio and stress history on the maximum

shear modulus, G .
max

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
256

o 4.2.2 Undrained Tests

Stress controlled cyclic undrained tests were per­

formed on saturated sand specimens under different Initial

stress conditions to evaluate the deterioration of the

sand stiffness as a result of the pore pressure buildup

and consequent shear strain development. For these tests,

CC>2 was percolated through the specimen by setting up a

small difference in absolute pressure between the vacuum

line connected to the top of the specimen and the CO^

purging line at the bottom. Following CO^ purging,

deaired water was allowed to seep through the specimen

under a very small gradient. Then, with the top drainage

line open to atmospheric pressure, seepage was allowed to

c occur for at least half an hour. Finally, to ensure com­

plete saturation, a back pressure of 300 kPa was applied

for at least two hours before applying the desired conso­

lidation stresses. The saturation procedure resulted in

values of the B parameter always larger than 0.98.

In spite of keeping the effective stress constant at

about 25 kPa, some axial strains occurred during the

saturation stage. These strains were carefully monitored

and the void ratio of the specimen was corrected by assum­

ing a volumetric strain equal to three times the axial

strain. During application of consolidation stresses,

volume changes were monitored with the burette to deter­

c mine the changes in void ratio of the specimen. Volume

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
257

o changes measured in Che burette are generally subject to

some error due to penetration of the membrane into the

voids between soil grains. These errors were estimated

based on previously published results (165,188). However,

the corrections were in general very small since confining

pressures of only 100 kPa or less were used in most of the

tests .

After the consolidation stage, the shear modulus of

the specimen was determined by performing resonant column


-4
tests at very small shear strains on the order of 10 %.

The agreement of Gmax was used as an index for specimen

replication throughout the testing program. Finally, the

drainage lines were closed and specimens were subjected to

c a number of torsional loading cycles, with reversal occur­

ring at predetermined levels of shear stress. The loading

period was of about 10 to 20 seconds per cycle. This low

frequency loading permits detailed examination of the

phenomenon of pore water pressure generation and conse­

quent strain development within each cycle of loading.

In some cases, specimens were reconsolidated after

initial liquefaction in order to evaluate the effect of

prior strain history on the cyclic undrained strength. A

few specimens were sheared monotonically or were subjected

to cyclic loading with reversal occurring at predetermined

levels of shear strain. Shear strains larger than those

corresponding to the rated capacity of the transducers


c

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
258

o used for measuring rotation developed in most undrained

tests. Therefore, beyond the range of the transducers,

shear strains were estimated by using the elapsed time and

the rate of straining which for each test was essentially

co n s t a n t .

4.3 General Purposes and Scope of the Investigations

Preliminary tests were performed under both drained

and undrained conditions to evaluate the whole range of

capabilities of the new apparatus and to compare these

results with those available in the literature, which had

been obtained using other devices and similar (or dif­

ferent) types of sand. After these tests, a comprehensive

testing program was initiated in accord with the main


c
goals.of the investigation, as described earlier. How­

ever, the testing program was from time to time adjusted

as the author developed a better understanding of response

of sands to cyclic loading under both drained and

undrained conditions or as a result of unexpected or espe­

cially interesting results. Nevertheless, the main

aspects of the behavior of sands on which attention was

concentrated the most were as discussed below.

Some researchers have recommended the development of

a relationship between shear wave velocities and liquefac­

tion potential of sands (29,32,33,89,147). It is con­

sidered that shear wave velocities can be readily


c

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
259

determined by means of geophysical tests in situ and that

all the factors known to affect the maximum shear modulus

of dry sand will also affect the stress-strain behavior

and liquefaction potential of the materials when subjected

to undrained loading in a saturated condition. In this

connection, stiffer soil deposits are considered to be

less susceptible to liquefaction (32), whereas lower shear

modulus values will be reflected by higher pore water

pressures during undrained loading (160). On the other

'land, it is claimed that elastic wave velocities respond

to fabric differences (29,36,38,39).

Consequently, the determination of shear moduli by

means of resonant column tests, at strain levels

corresponding to those produced in geophysical tests for


C
measuring in situ shear wave velocities, would be a power­

ful way of monitoring the effects of factors such as

stress-strain history and aging on the strength-

deformation characteristics of sand. As discussed ear­

lier, these effects are reflected in the sand structure,

since they affect the predominant grain orientation or the

conditions at the grain contacts. To evaluate the vali­

dity of these views shear moduli at low strain levels

corresponding to those produced in geophysical tests were

determined on dry specimens which had been prestressed

along different stress paths or subjected to previous

cyclic strain histories of various magnitudes. In most of

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
260

the cases, the complete variation of shear modulus with


o shear strain amplitude was determined over the whole range
^ *
of strains of engineering interest, from 10 % to 10%.

Accordingly, the effects of the factors listed above on

the small strain and large strain behavior could be

evaluated.

As discussed in Chapter 3, there is experimental evi­

dence which indicates the existence of a threshold shear


_2
strain for sands on the order of 10 % (24,32,33,43). If

cyclic strains below this value are induced in the sand

deposit, it is considered that there is neither densifica—

tion of dry sand nor pore pressure buildup in saturated

sands. It is claimed that the threshold shear strain is a

constant characteristic of each type of sand for a wide


c
range of particle arrangements, relative densities and

confining pressures. In the author's view these findings

appear to conflict with the results of a significant

amount of research, which shows that the mechanical

behavior of granular materials is affected significantly

by factors such as fabric and its associated anisotropy,

stress-strain history, stress path, etc. among others, as

discussed throughout this dissertation. It was thought

that under drained conditions the threshold shear strain

could be identified as the shear strain marking the ini­

tiation of degradation in the shear modulus, that is, the

shear strain at which the behavior of sands theoretically

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
261

changes from linear elastic to nonlinear and inelastic.


O Accordingly, all the results from the tests described in

the previous section were carefully analyzed in the light

of the threshold shear strain concept. A few resonant

column tests and torsional shear tests were performed on

saturated specimens in order to validate the existence of

the threshold shear strain under undrained conditions and

establish its relationship with the modulus reduction

curves from the drained tests.

As discussed earlier, there is ample experimental

evidence to show that the fabric developed under condi­

tions prevalent during deposition of most natural deposits

is highly anisotropic, which causes changes in the

C undrained deformation-strength characteristics with the

direction of loading (55,118,119,120,169). However, most

investigations on the undrained steady-state strength of

sands have been based on axial compression triaxial tests,

and hence they have been limited to a particular stress

path. Moreover, only recently some researchers had

attempted to establish a relationship between the mono­

tonic and cyclic undrained behavior of sands

(34,121,165,184,185,186). In fact, most of these research

reports became available to the author after developing

the concepts and the detailed frame of understanding

presented in Chapter 3. Some of these results support the

author's views but there are still conflicting opinions on

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
262

o critical aspects of the undrained behavior of sands such

as:

a. the effect of initial fabric on the large st.rain

deformation characteristics

b. the effect of consolidation shear stresses (anisotro­

pic consolidation) on the subsequent behavior under

undrained shearing, and

c. the state conditions marking the initiation of strain

softening behavior and hence the relationship between

monotonic and cyclic loading of sands.

Consequently, it was of great interest to study some


C of these aspects in the new torsional shear apparatus.

Actual steady-state conditions were difficult to develop

since the strain rate capabilities of the new apparatus

are far from the strain rate levels which have been

observed to develop at the steady-state limiting condi­

tions (2). Therefore, investigation of these effects was

limited to a particular strain rate.

Saturated loose specimens isotropically consolidated

were subjected to monotonic and cyclic loading under dif­

ferent cyclic shear stress amplitudes. In some cases, a

static shear stress was applied during the consolidation

stage. In other cases, specimens were reconsolidated

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
after initial liquefaction in order to evaluate the
o effects of prior strain history, and hence of changes in

fabric, on the monotonic and cyclic undrained strength.

c.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
3 5 on
Me, i . r e

c
CYCLIC STRESS-STRAIN AND LIQUEFACTION

CHARACTERISTICS OF SANDS

VOLUME II

A Thesis

Submitted to the Faculty

of

Purdue University

by
(
Adolfo Alarcon— Guzman

In Partial Fulfillment of the

Requirements for the Degree

of

Doctor of Philosophy

August 1986

<r

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
264

CHAPTER 5

TESTS RESULTS AND ANALYSES

The types of tests performed are listed in Table 5.1.

The results from these tests are presented in the follow­

ing sections. In evaluating the tests results, considera­

tion is given to the effects of shear strain level, ini­

tial state of stress, prestressing, and prior stress his­

tories along different stress paths on the cyclic drained

stress-strain behavior of sands; the effects of prior

strain history on the undrained behavior are also exam­


c ined. Emphasis is placed on the relationship between

monotonic and cyclic undrained behaviors, as it reflects

on a physical appreciation of the complex response to

loading of granular soils.

5.1 Drained Cyclic Shear Tests

5.1.1 Effects of Specimen Geometry

Figure 5.1 presents a typical relation between shear

modulus and shear strain amplitude, obtained from tests

run on dry, solid and hollow sand specimens with a rela­

tive density of about 55% (e = 0.598). For both cases,

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
Reproduced with permission of the copyright owner. Further reproduction
r

Table 5.1 Summary o f T e s t s P e r f o r me d .

a) Drained Tests

Test Specimen 0 0 0 Results


z r m
No. type1 e kPa kPa kPa OCR In Fig. Descript inn

2 S 0.592 100.0 100.0 100.0 1 .0 5.1 G versus y


3 II 0.549 108.0 108.0 108.0 1 .0 5.2 G versus y
4 II 0.529 108.0 108.0 108.0 1 .0 - G versus y
5 II 0.599 100.0 100.0 100.0 1 .0 5.1/5.2 G versus y
15 II 0.631 106.5 101 .0 102.8 1 .0 5.2/5.4 G versus y
16 H 0.626 107.0 101 .6 103.4 2.0 5.13 G versus y
17 H 0.593 211.1 205.7 207.5 1 .0 5.8 G versus y
18 H 0.585 104.0 50.1 68.0 1 .0 5.8 G versus y
IP H 0.619 103.5 50.1 67.9 4.0 - G versus y
20 II 0.624 101.1 50.3 67.2 1 .0 5.11 G versus y
21 H 0.620 100.0 50.0 66.7 4.0 5.14 G versus y
22 II 0.641 71.5 66.7 68.3 1 .0 5.4 G versus y
23 H 0.623 101.0 50.0 67.0 4.0 5.14 G versus y
24 H 0.643 205.4 200.0 201.8 1.0 5.4 G versus y
25 II 0.622 100.6 50.1 66.9 1.0 5.11 G versus y
prohibited without perm ission.

26 H 0.629 100.4 50.2 66.9 ** 5.18 G versus y


after prestresslng
27 11 0.620 100.4 50.1 66.9 — 5.18 G versus y
after prestressing
28 II 0.612 100.5 50.2 67.0 *" 5.17 G versus y
after prestraining
29 II 0.632 100.2 50.0 66.7 5.17 G versus y
after prestralnlng
30 II 0.534 109.0 100.6 103.4 1 .0 - G versus y
31 S 0.591 103.9 50.0 68.0 1-10 5.15 G versus OCR
5.16 Gm a* versus o
32 H 0.622 32.4 27.3 29.0 1 .0 - Grav§rsus y m
33 H 0.596 103.9 50.0 68.0 1.0 - G versus N
34 II 0.632 32.0 26.9 28.6 1 .0 5.4 G versus y
35 II 0.614 105.0 50.0 68.3 - 5.19 G versus y
36 II 0.619 105.6 50.9 69.1 I .0 5.20 G versus N
37 II 0.624 104.9 50.3 68.5 1.0 5.20 G versus N
38 H 0.633 104.3 50.0 68.1 1 .0 5.20 G versus N
II 0.518 106.7 100.0 102.2. 1.0 5.2 G versus y

265
39
40 II 0.509 106.9 100.2 102.4 1 .0 - G versus y
60 S 0.626 - - 100.6 1 .0 5.12 Gmax versus K
266

e 0 0 0 0 0
a. a* 0 a*
0 a* a. 0
M 0 .* a. X
0 X X <0
Cm X 0 0
cn j£ © 0 in O' © 0 3
• • CM •
0 • • a*
•* 0O’
«0 m eo •0 *6*m
CM 44 © ** •* 44
in 0 m0
■■ M 44
0 M 44 — a. -4
0 X 0
1 1 « aX a 1 a • * *4
0 03 0
>» X >s X X X X X* *4
0 0
CJ 0V 0 cr> e
H 0 44 00 0 0 0000 -4 44
a « a 0 0 44 0 0 0 0
a* • ex. a* * a* Om • 0 O* • • a* • a* a. • «i i 0
-3 J* •3 K K K •o *3 0 X *o X X *3 0 •3 0 0
0 <C « O 0 0 0 0 0 a. 0 0 a* 0 (D0 a* X X X
ec u c ?> 0 CM^ 03 X X 41 © 0 X m 0 m © 0 X 0 0 0 10 X
e • 0 • • 0 • •m© © a 0 0 6 • 0 • 0 • • 0 a. 0 0 0 0 o
o « < O N x •3 eo © • • • 3 •3 3 u *6* *3 CMX 73 © m *o © x -o • CM 3 3 3
X 0 X 44 — © © © 0 0 O *»4 0 0 • 44 *p4 44<0 • •0 0 0 • 0 0 0
0 •M 44 b» 0 w M 44 •m 44 m
0 cn •4 *H m 0 0 0
A •H O *H -H 4t 4t +t 0 o 0 «4 <44 e e c rnm+1 O 44 44 o — a* o c e mm 0 0 0
a 0 > 0 > e 0 o e 1 0 0 0 o o > > >
0 i e i ■ e i 1 1 1 1 c 3 a c 00 a e a a c a 1«e 0 0 a
XJ e o 1 ua 01 o o e a o e o o o «« E
a x j :n x j :►»X X >» >o 0 O e X J e c XJ K X J X XJ c e M3 O O
0 00 (J CJ0 J O XJ XJ0^(N O 0 o 00 o o 0 0 0 J 0 0 (J 00 O o ex. X.
a *- OS 0 0 0C H 0 ^ 3 S 3 z 0 0£ z z KZ z 0 3 3 3

H 0 bC

3 •*• CM cn © tt © © 44 44 © cn < m 44 X CO X © X O' c-> X X X
0 3 to* x

t CM

X CMCM
• •
CMm
• •
<n cn cn -0 < X ■0 «0 cn cn *0 cn cn X) X X X
• • • • • • • • • • • • • • • • •
•O m
5
5
m Wi m m in m in m m in m m m m in m m in m m m m
0) Q0£ 4e4
9
C C

c
o
© x c*>
• • • •
O G
o © © o ©

c

O«o mm •4
© © © C

©

o ©
M o cn © © CMm m © cn

o © O o o o © © © ©
X C

Mcn *0 X _ ©

© © © o
•m
© ©
• •

© ©
X X

©
**
©
_


*
o> m O

o m mm -4 44 © © o o o O © o o © o © © © © © © © © © •4* C"» o
X © x in ©

o —
E
E
o © ci o © © o CM © in m C MCMm in © Cl CMX cn © X © X X o ©
• • • 4 • • • • • • •
© © 40 44 mm © r i mm 44 CMCM CM© © © © O o © © © © o © © © o © © X) -0 44 •
44
m © © C O O © m o
mm mm
o o o o o © o

C
*•
o c ©
44
o O ©
44
© © © ©
**
© © © © X m o
* X

0)
••H O' — O' X sO *0 CM «0 X © m © © n © X X © c o m X *0 © cn m O' m X cn O' ©
"
X) X © o 44 © 44 © 44 o ON CM •M CMeM CM o CMX CMCM3* CM X © X X o X X X en X X
eO m m ©
• • •
© in © © ©

© m © © © © © © © © © © in © © ©

© © ©

© © © © © ©
©

H
0

© o © © © ©
0

© o C © o © © o © o o O O © © O © o © © © © o o o

Cl
«eH ~ 3
o a C 6 O 3 O 3 O 3 O 3 » 3 O 3 X = Z O :O 3 O 3 O 3 0 3 3 3 O 3 0 3 C /3 O 3 0 3 O 3 0 :O 3 O 3 O 3 O 3 O 3 O 3 O 3 C /3 O 3 O 3
0 >s
c. 44
03 I3
03 S

a
o•
< CO < fl D < 6 3 < 0 3 < 0 3 < 6 3 < CO < B W
V N N ® 0v0vO"<Nn*J-<-^Nrt<<in*eNB«O“ —
• ( ■ - t ^ N < 9 4 « < 9 < 4 4 < 7 < ^ < i n i n i n i n u ,,i i n i n i A i A ( n i n i r )
t- z

R eproduced w ith perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
Reproduced with permission of the copyright owner. Further reproduction
r

Ottawa 20-30 sand


120
e - 0* 5 9 6
Dr - 5 5 %
cr'z = O r - 100 K p a
a. 100

Solid cylinder s p e c im e n
BO
(D=71mm)

Hollow cylinder specimen A ft


60 I O D = 7 1 m m , I D “3 8 m m )

Resonant column tests

40 Torsional shear tests


prohibited without perm ission.

20

10-1 1 10

s h e a r ; STRAIN, y ,%J

F i g ur e 5.1 Ef fect ; o f Speci men Geometry on Shear Modulus - Shear

267
Strain Relationship.
268

Che average shear strain (as obtained by integraCing an

assumed linear variation over the specimen cross-section)

was used in the comparison. The shear modulus values from

torsional shear tests presented in Figure 5.1 and in most

of the succeeding Figures correspond to the average value

in approximately 10 cycles. As discussed later, when the

stage testing procedure described in section 4.2.1 was

used, the shear modulus, G, did not vary appreciably with

number of cycles (see Fig. 5.20). The modulus reduction

curves obtained with the solid and hollow cylinder are

comparable, and the results are in general agreement with

those previously published (22,88). Chung et a l . (22)

also concluded that the type of specimen used (solid or

hollow cylinder) had no significant effect on the shear


c moduli and damping ratios obtained from resonant column
-4
tests for shear strain amplitudes between 2 x 10 % to 4 x
2

10 %. However, differences on the order of 5% to 15%

were found in this investigation when comparing shear

moduli from dry hollow cylinders and saturated solid

cylinders which the author attributes to the inevitable

differences resulting from the preparation procedures of

the specimens. It is noted that a 10% variation in shear

modulus may be reflected in a larger difference in pore

pressure response. The results in Figure 5.1 cover the

full range of shear strains of engineering interest, from


-4
10 % to 10%, reflecting the capabilities of the new

apparatus.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
269

5.1.2 Effects of Relative Density

Figure 5.2 presents the relation between shear

modulus and shear strain amplitude obtained from tests run

on dry specimens at four different relative densities,

isotropically consolidated under the same mean confining

stress, a , of 100 kPa. Although the shear modulus


in
increases as the void ratio decreases, especially in the

low strain range, it is important to note that a signifi­

cant change in relative density (from 40 to 90%) is

reflected by only a modest change in shear modulus. This

result has to be taken into account when comparing the

effects on the shear modulus of other factors such as

stress path or stress history. As the shear strain

increase the modulus reduction curves seem to collapse

into a single curve suggesting that in the process of

cyclic loading at increasing shear strain amplitudes the

density or void ratio of the specimens was approaching a

limiting or residual value, independently of the initial

void ratio.

The maximum shear modulus values corresponding to


—4
shear strains of about 10 % (from Fig. 5.2) are plotted

versus relative density in Figure 5.3. The modulus values

determined using two previously published empirical equa­

tions obtained from test results on different sands

(58,88) are also shown for comparison. The measured

values closely follow the trend predicted by the empirical

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
Reproduced with permission of the copyright owner. Further reproduction

140

Ottaw a 2 0 - 3 0 sand
120 ^
H ollow cylinder specimens

cr'z =»<r’r s ioo kPa


100

Dr,%

0.513 90

0.549
CD
0.598 55
w 60-
_i
D Resonant column tests • * 0.635 40
Q
0
prohibited without perm ission.

2 Torsional shear tests


a: 40-
<
ui
1
c/>

2 10 1 10

SHEAR STRAIN, y i% )

270
Figure 5.2 E f f e c t o f R e l a t i v e D e n s i t y on Shear Modulus - Shear S t r a i n
Relationship.
271

Empirical Equations:
c
(A) Hardin (1978) G - 700 [F(e)] cr *m round grains
y <io-4
e < 0.8
Iwasaki et al (1978) G - 9 0 0 [F(e>] cr’m 1'2 clean sands
X * 10 ~6
0.61 < e < 0,86
F(e) * (2.17 - e)2 / (1+e)
G, in K g / c m 2
s
Q.
1 180'
x
a
E
® 160-
co*
3
=5
a 140-
o
cr
< 120
in -

x
co
Eq.®
=> 100 -
s
x
< o Data from Fig. 5.1
80- Ottawa 2 0 - 3 0 sand
e m a x = 0-732
e min 1 = 0,4891
u 1 ------------------------------- — -

40 50 60 70 80 90 100

RELATIVE DENSITY, Dr (%)

Figure 5.3 Effect of Relative Density on Maximum


Shear Modulus.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
272

equations although the absolute values differ by about


c 30%. In some cases, this difference can be even higher

(see Table 1 in Ref. 22). Therefore, empirical equations

should be calibrated for the particular type of sand being

studied. It is also noted that, for sands with differ­

ences between e and e . larger than that corresponding


max min °
to Ottawa 20-30 sand, the variation of shear modulus with

relative density will be more significant than the varia­

tion observed in Figures 5.2 and 5.3. Accordingly, a

unique correlation between shear modulus and relative den­

sity for all sands is not possible. As discussed in

Chapter 3, at the same relative density, different sands

will not have the same drained or undrained stress-

deformation characteristics.
(
5.1.3 Effects of Mean Confining Stress

*
The influence of mean confining stress, a , on the

modulus reduction curve is illustrated in Figure 5.4, for

specimens with a relative density of about 40%. Unlike

the limited influence of void ratio, the considerable


*
effect of c on the shear modulus is clearly evident. The
m
shear modulus values corresponding to different shear

strain levels (from Fig. 5.4) are plotted versus the mean

confining stress in Figure 5.5. At small shear strains

the shear modulus values are proportional to the square

root of the mean confining stress whereas at large shear

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
Reproduced with permission of the copyright owner. Further reproduction r ' n

160

140
Ottawa 20-30 sand
Hollow cylinder specimens
120 e ° 0.635
Dr - 40%

Test
100
202
103
80
68.3

28.6
60
prohibited without perm ission.

40
Resonant column tests ■ a • *
Torsional shear tests o a on

r4 10-1 1
SHEAR STRAIN, /(%)

273
Figure 5.4 E f f e c t o f Co n f i n e me n t on Shear Modulus - Shear S t r a i n
Relationship.
274

L
-3-
o
CM m

Fig
o
CM

from
©
CO

(Data
©
CO

Stress
co
a.
JC

E
b

Confining
co
o CO
UJ
QC
h-
CO

C C5

Mean
o

Versus
co O
<
UJ

Modulus
<0
^
CM © Shear

o
CM
5.5

N.
Figure

o O O o O o
CD CM O ® co rr CM

( V d i w d ‘s m n a o w u v s h s

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
275

strain amplitudes the shear moduli become a linear func-

c tion of o .
m
These results confirm previous test results

on various sands (22,58,59,88,89). Therefore, the


*
exponent of with which the modulus varies is a function

of the shear strain amplitude. In Figure 5.5 this


-4
exponent changes from about 0.5, at shear strains of 10 %

or less to 1.0 at shear strains of about 0.65%. It is

considered that at small shear strains the modulus depends

not only on frictional forces but also on particle inter­

locking, whereas at large strain amplitudes the modulus

depends mainly on the frictional component of the shear

strength of the sand, which in turn is proportional to the

confining stress level.

Figure 5.6 presents the modulus reduction curves from


c Figure 5.4 normalized with respect to the shear modulus

values determined at a shear strain amplitude of about


-4 "
10 %. It is noted that the mean confining stress, a
in
affects not only the values of G , but also the shape
J max* r
and location of the modulus degradation curves with

respect to the shear strain axis. The reduction in

modulus, expressed as a percentage of Gm a x , is larger for

lower o at the same shear strain level. Although the


m
curves look close because of the log scale, the largest

difference between the upper and lower curves is on the

order of 60%. Moreover, it has been shown that the shape

and location of the modulus degradation curves

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
276

CM co CD

Reduction
CO so CO
CM CD CM

CM

Modulus
the
of
Shape
a?

(Data from Fig. 5 .A).


on the
CM <
cr
c o
co
cr

of Confinement
<
UJ
I
CO

co
o Effect
Curve
5.6
Figure

o CO CO ■M- CM
o d d

XBUJ0/9 ‘o u va sm naow h v h h s

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
277

(G/G versus y ) depend also on the void ratio and on the


max
c soil type (59). It will be shown later that stress his­

tory and the number of prior stress cycles also affect the

modulus degradation curves. Consequently, the simplified

procedure, which in common practice utilizes a single

average modulus ratio, G/G , versus shear strain curve


6 * max*
for cohesionless soils (146,154), can result in signifi­

cant errors in the interpreted values of shear modulus.

Hardin and Drnevich (59,60) suggested that the varia­

tion of shear modulus with shear strain amplitude could be

closely represented by a hyberbolic relation of the form

(5.1)
G 1+yJyr
max
(
where yr was defined as the reference shear strain and

equals

Yr = (5 .2 )
max

where t corresponds to the maximum shear strength of


max
the soil. In other words, the reference strain, Yr,

represents the level of shear strain that the soil would

experience under an applied shear stress equal to t , if


’ max
the behavior of soil were "elastic" with a shear modulus

equal to G . Hardin and Drnevich (59,60) found that, by


d dX
representing test data in the normalized form of

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
278

G / Gjjax v e r s u s Y/Yr, the data fell along a single curve,


c Irrespective of a
*
m or Dr. In order to determine the

reference strain, both t and G should be known.


* max max
However, according to E q . 5.1, at a shear strain equal to

the reference strain the shear modulus ratio becomes equal

to 0.5. Therefore, the reference shear strain, Yr, can be

estimated directly from the modulus degradation curves as

the shear strain level at which the shear modulus becomes

a half of G , even without knowing t . Using this


max * & max
approximation, if the the data in Figure 5.6 are plotted

with respect to the normalized shear strain, y / y r , they

do, indeed, tend to collapse into a single curve, as

illustrated in Figure 5.7. It is noted that the data

points essentially follow the hyperbolic relationship pro­


c posed by Hardin and Drnevich (59,60), who adjusted

slightly the definition of the reference strain in order

to get a perfect match between the theoretical and experi­

mental values. However, for cases in which only approxi­

mate shear modulus values are required, the values given

by E q . 5.1 seem to be close enough. The values of the

reference strain as a function of the mean confining

stress are also presented in Figure 5.7. It is interest­

ing to note that yi was found to increase with the square

root of o . This is not completely surprising since in


m
*
Eq. 5.2, t increases proportionally to a whereas G
* max m max
increases with the square root of o (see Fig. 5.5).
m

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
Reproduced with permission

Ottawa 20-30 sand


of the copyright owner. Further reproduction

x
CO Hollow cylinder specimens
E 1.0
0 0*R. Dr = 40%
0 I'A.

0.8
< \Q G 1
cc
to Gmax 1+ X/Xr
D V4
_J 7
D 0.6 \
Q
0
2
CC
<
cr’m xr
Ui 0.4 Symbol A \
1 (kPa) (%) \
to □ V
prohibited without perm ission.

ON
□ 202 0.0470
A 103 0.0335 \
0.2 o 68.3 0.0273
V 28.6 0.0173
Remark: Xr = 0.00331 m)1/2

10 -3 10’2 10'1 1 10 100


NORMALIZED SHEAR STRAIN, //X r

Figure 5.7 Dimensionless She ar Modulus - Shear S t r a i n R e l a t i o n s h i p .

279
280

From the foregoing discussion, it follows that the

c modulus reduction curve for a given o can be readily


Cl
obtained simply by knowing G and yr. Moreover, if G
r J J max * max
and yr both increase with the square root of a a single
Cl
shear modulus versus shear strain curve is all that would

be required to estimate the modulus degradation curves at


*
different levels of Om . These findings support the sim­

plified procedure recommended by Drnevich (39), which con­

sists in determining the maximum shear modulus G , at


& max *
*
increasing levels of c and then determining the complete

variation of shear modulus with shear strain amplitude on

the same specimen but only at the highest value of o^

required for the case under consideration. Since the

variation of G with o is known, the values of yr can


max m
( be scaled in accord with E q . 5.2. This procedure was

checked by using the data obtained from tests run on dry

specimens with a relative density of about 60%, consoli­

dated under mean confining stresses of 68.0 and 207.5 kPa,

respectively. G and yr for a of 68 kPa were estimated


max m
*
from the values obtained from the test with a equal to
m
207.5 kPa, by assuming a scaling factor proportional to
1/2
a . The comparison between the predicted and measured
m
values is presented in Figure 5.8. The remarkable agree­

ment needs no additional comments. Nevertheless, it

should be noted that stress/strain histories also affect

the modulus degradation curves, and hence, the reference

strain, y ^ , as discussed in a later section.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
Reproduced with permission of the copyright owner. Further reproduction r

180

160
Ottawa 2 0 -3 0 sand
« Hollow cylinder specimens
o.
5 140
Dr ■= 60%
O
co Test fr'm, |<pa
O 120
Predicted curve \
=> 207.5
□ (Assumed Y t - 0.03%)
0 100 if
68.0
5
cc
<
UJ 'max
1
CO
\A
prohibited without perm ission.

Resonant column Tests *


Torsional Shear tests o &

SHEAR STRAIN. (%)

281
Figure 5.8 P r e d i c t i o n o f Modulus R e d u c t i o n Cur ve.
282

The hysteretic damping ratio values corresponding to


c the tests reported in Figures 5.4 to 5.7 are presented in

Figure 5.9. Damping values from resonant column tests

were determined by using the magnification factor at reso­

nance (see Appendix A), whereas damping values from tor­

sional shear tests were determined by measuring the area

of the hysteretic stress-strain loops (see Appendix B).

The damping ratio, D, increases as the mean confining

stress decreases, especially for shear strain amplitudes


_2
above 10 %. However, all the curves approach a limiting

value as the shear strain amplitude increases. The sharp

increase in damping values above a shear strain of about


-2
10 % seems to correspond to the degradation in shear

modulus in Figure 5.7. In fact, in the author's opinion,

the equivalent shear modulus and damping values are not

necessarily two independent variables. As loading takes

place under increasing mobilized friction (larger shear

s t r a i n s ) , the plastic nature of the behavior gradually

becomes dominant. This is primarily due to irrecoverable

displacements and permanent changes in the fabric or grain

arrangements, which cause a reduction in the shear modulus

and a consequent increase in the size of the loop result­

ing from the unloading and reloading stages. In this con­

nection, Hardin and Drnevich (59,60) attempted a direct

relationship between the modular ratio, G/G , and damp-


r max
ing values. By assuming that the initial slope of the

stress-strain curve upon reversal is equal to G ,


e max

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
Reproduced with permission of the copyright owner. Further reproduction

Ottawa 2 0 -3 0 sand
50 • Hollow cylinder specimens

Dr - 4 0 %
40 -
Symbol
(kPa)
30- 202 Range of results
O
F*
<
103
o: 68.3
| 20-
28.6
prohibited without perm ission.

CL
5
<
O
10 - Resonant column tests Torsional shear tests

10-' 1

S H E A R STRAIN, / <%)
Figure 5.9 E f f e c t of Conf i n e me n t on Damping R a t i o - Shear S t r a i n hO
00
Relationship. lo
284

Irrespective of the shear strain amplitude, Hardin and

Drnevich concluded that D is approximately related to

6/6 by the expression


max 3

D = D (1 - T-2-) (5.3)
max G
max

where D is the limiting damping value. At large shear


HIclX

strain amplitudes, 6 becomes almost zero and hence, D

equals D . O n the other hand, at very small shear


n max * J
strains, 6 is close to Gmax> resulting in a damping ratio

of approximately zero. If the data from Figure 5.9 are

plotted with respect to the shear modulus ratio, they col­

lapse approximately into a simple curve, as illustrated in

Figure 5.10. However, the relationship is not linear as


( implied by E q . 5.3. This is attributed to the fact that

the initial slope of the stress-strain curves after rever­

sal becomes smaller than 6 as the shear strain (or


max
stress ratio) increases, as will be discussed later on.

An evaluation of previous investigations on the

nature of damping in sands (54,56,59,89,160,171,172) per­

mits drawing the following general conclusions:

a. Although the energy losses within cyclicly loaded

sand masses are caused by interparticle slip and

friction at particle contacts, which develop hys­

teretic damping, the decay of free vibrations of a

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
Reproduced with permission of the copyright owner. Further reproduction

60

50
Ottawa 20-30 sand
Hollow cylinder specimens
Dr = 4 0 %
Q
O Symbol c r’m.kPa
E 202
£ 30
0 103
z Range of results
a 68.3
< 20 28.6
a
prohibited without perm ission.

D = D m a x (1- G m a x

0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0

S H E A R M O D U L U S RATIO, G / G m a x

285
Figure 5.10 Damping Ratio Versus Shear Modulus Ratio.
286

specimen of granular materials resembles the vibra­


c tions of a viscoelastic system;

b. Damping ratios increase significantly with increasing


*
shear strain amplitude but decrease as cr^ or the

number of cycles increases;

c. The effects of void ratio and methods of specimen

preparation seem to be small;

d. The maximum damping ratio. D , seems to be a


* max’
characteristic of the type of sand. Damping ratios

of clean sands seem to be slightly larger than those

of other natural sands, especially if they include

particles finer than 0.074 mm. Damping ratios appear

to decrease as the average grain size decreases.

Accordingly, the damping ratio values for standard

Ottawa 20-30 sand are observed to be higher than

those of other typical sands used for dynamic studies

such as Toyoura or Monterrey No. 0 sands, which are

finer. Therefore, the use of a single damping ratio

versus shear strain relationship for all sands

(146,154) is likely to introduce significant errors;

e. Damping ratio values are likely to be affected by

stress history as well as by consolidation stress

ratios. However, these effects are much less impor­

tant than the effects of shear strain amplitude and

mean confining stress;

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
287

f. It is difficult to obtain accurate damping ratio

c values. Accordingly, damping values measured by dif­

ferent investigators using different testing devices

may show a considerable scatter.

5.1.4 Effects of Frequency

The results presented in Figures 5.1 to 5.10 cover

the full range of shear strains of engineering interest,


—4
from 10 % to 10%, and the shear moduli at a shear strain
—2
amplitude of about 10 % obtained from the resonant column

(high frequency tests, f > 100 Hz) and torsional shear

tests (low frequency tests, f Z 0.1 to 0.2 Hz) agree

approximately, in spite of the great disparity in fre—

^ quency. This agreement has also been reported by previous

investigators (42, 88,136), leading to the conclusion that

the frequency of loading does not affect the response of

sands to cyclic loading. However, in the author's opin­

ion, it is too early to assert that frequency has no

effects on the stress-strain response of sands. Besides

frequency, the number of loading cycles is also very dif­

ferent between the two tests. The relative contribution

of each one of these factors to the response of sands in

the resonant column tests is not yet fully established,

although the effects of changes in frequency say be impor­

tant in clayey soils (31). Accordingly, the torsional

shear test apparatus is being modified to expand its

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
288

frequency range, and to reduce the lower limit of shear


c strain amplitude that can be measured accurately. More­

over, it will be shown later that when stage testing pro­

cedures are not used, the shear modulus values from the

two tests may differ significantly, depending on the

number of cycles in the torsional shear test at which the

results are compared.

5.1.5 Effects of Shear Strain Level

As discussed previously, shear modulus values

decrease as the shear strain amplitude increases and the

corresponding modulus reduction curve is not unique with

respect to shear strain. Differences of up to 60% in the

G/Gmax values are observed in Figure 5.6 depending on the

shear strain amplitude and on the level of mean confining


*
stress, o . If the values of G/G in Figure 5.6 were
m max °
plotted against shear stress instead of shear strain, the

horizontal distance between two adjacent curves would

increase considerably, resulting in differences of up to

250%. Therefore, at a given G/G value, the range in


m oX
shear strains is smaller than the corresponding range in

shear stresses. However, it is not to be concluded that

the applied shear strain is the only factor controlling

the response of sands under different loading conditions,

as claimed by some researchers (24,34,70). For example,


*
the effect of a on the degradation of shear modulus at
m
c

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
289

the same strain amplitude is evident in Figure 5.6. More­


c
over, it is well known that different types of sand or

sands with different fabrics may mobilize different

strength values under both drained and undralned condi­

tions, even if they are compared at the same shear strain

level. Although, shear stresses and strains are depen­

dent, their relationship is not unique. It is affected by

factors such as mean stress, fabric, aging, etc.

It should be noted that all the modulus reduction

curves in Figures 5.2 and 5.4 show some degradation in the

shear modulus even at small shear strain amplitudes in the


-4 -3
range from 10 % to 10 %. This was also observed by Har­

din et a l . (59). From this point of view, if a threshold


-4
( shear strain exists, its value is less than 10 %. How­

ever, the results in Figure 5.9 suggest that a sharp

increase in damping, which reflects the non-linear and

inelastic properties of the sand, only occurs for shear


-2
strain amplitudes slightly above 10 %. This strain level

has been suggested as the threshold shear strain, if , i.e.


— 2 -2
the strain level (10 % to 2 x 10 %) below which the pore

pressure buildup during cyclic loading of saturated sands

is negligible (24,32,33,43). This seems to be the case

also for Ottawa 20-30 sand, as will be illustrated later

with results from undrained tests. Nevertheless, it is

interesting to observe that a considerable reduction in

shear modulus has already occurred at a strain level of

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
290

_2
10 %. At this strain level, the corresponding shear
c modulus ratio, G/G , in Figure 5.6 is observed to range
max °
between 0.70 to 0.85. The reasons for this behavior are

not clear to the author. At a first glance, one tends to

think that there is a significant increase in the rate of

degradation in shear modulus for shear strains slightly


_2
above 10 % (see Figs. 5.4 and 5.6), which could reflect

in higher pore pressures under undrained shear. However,

when the shear modulus values are plotted versus shear

strain in an arithmetic scale, it is noted that the rate

of degradation of shear modulus essentially decreases with

increasing shear strain amplitude. Accordingly, it should

be kept in mind that the semilog scale, although con­

venient, may obscure the relative effects of shear strain


(
v amplitude on the shear modulus and damping values of

sands.

5.1.6 Effects of Stress Ratio

Two replicate hollow cylinder specimens with the same

relative density of about 45% were tested in the resonant

column/torsional shear apparatus. One specimen was iso-


*
tropically consolidated with a mean confining stress,

of 68.3 kPa. The o.her specimen was anisotropically con-

solidated along a stress path with a stress ratio, a2 / ar *

of 2 .0 , but with approximately the same mean confining

stress as the first specimen. The shear modulus values

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
291

obtained for these two specimens were plotted as a func­


c tion of shear strain and resulted in essentially the same

modulus degradation curve, as illustrated in Figure 5.11.

Hence, at the same mean normal stress an initial static

stress ratio of 2.0 had little effect on the shear

modulus-shear strain relation in drained shear, for shear


—4
strain amplitudes ranging from 10 % to 1%.

The effects of stress ratio on the shear modulus are

further illustrated by the results presented in Figure

5.12. A solid cylinder specimen with a relative density

of about 44% was isotropically consolidated to a mean con­

fining stress of 100.6 kPa. After determining G by


max
means of resonant column tests, the axial and radial

( stresses were gradually changed keeping constant the mean

confining stress, until a stress ratio of 1.5 was

obtained. At this stage, the maximum shear modulus was

determined again. The procedure was repeated at higher

stress ratios until failure developed, providing che vari­

ation of G with stress ratio. The maximum shear


max
modulus decreased as the stress ratio increased above 1.5.

The apparent increase in G at the failure condition can be

due to an increase in the circumferential stresses associ­

ated with end restraint effects. At this stage, bulging

of the specimen was significant. Nevertheless, it is

interesting to observe that the reduction in G from an


6 max
isotropic state of stress to the failure condition was

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
Reproduced with permission of the copyright owner. Further reproduction C'

Ottawa 2 0 -3 0 sand
120 Hollow cylinder specimens

e = 0.623
'max’ 95,3 Dr = 45%
| 100
O Test

^ 'z.k P a 100.2 71.5


80
D ° ” r,kPa 50.2 66.7
O
O
5
<J"m,kPa 67.1 68.3
< 60
LU Kc = o " 2 /cr*r 2.0
1 100 1.07
<n

ry
prohibited without perm ission.

Resonant Column
o A Torsional Shear
20
20

20 40 60 80 100
c'r.kPa
r‘' 1(T3
SHEAR STRAIN %

Figure 5.11 E f f e c t of C o n s o l i d a t i o n S t r e s s R a t i o on Modulus R e d u c t i o n

292
Curve.
Reproduced with permission of the copyright owner. Further reproduction /■“"N

Ottawa 2 0 -3 0 sand
Solid cylinder specimen
Dr = 4 3 .6 %
= 100.6 kPa
120 -

a f
Q. \
1 100
o
CO
D Axial compression
^ 80 max
o <T'z /°"r
o
S
E 60 -
<
UJ
X
CO Secant / Torsional loading
40 '
prohibited without perm ission.

> o-'z = cr'r


Tangent J T z0 increasing
20

1.0 1.5 2.0 2.5 3.0


S T R E S S RATIO, <r'i/0"3

Figure 5.12 E f f e c t of S t r e s s R a t i o on Maximum Shear Modul us .

293
294

only of 15%. In the author's opinion, this behavior

implies that Gm a x is not significantly affected by changes

in fabric since the particle arrangement certainly changes

in a drastic manner as the stress ratio increases. It Is

considered that under anisotropic states of stress, fabric

anisotropy increases resulting in a larger number of nor­

mals at the contact points in the direction of the major

principal stress (5,126,127,178). This increased aniso­

tropy is reflected by only a modest reduction in stiffness

under torsional loading.

Yu and Richart (198,199) reviewed previous investiga­

tions on the effects of stress ratio, and then presented

the results of a comprehensive series of resonant column

tests in which the effects of static stress ratio, o /a ,


\ * z r’
and stress paths on the shear modulus at a shear strain
-4
amplitude of about 6 x 10 % were evaluated. Their

results and those of others (172) also showed that

increasing the stress ratio decreases the maximum shear

modulus, G . However, it is interesting to note that


m ax
stress ratios as high as those corresponding to failure

only decrease G up to a maximum of 30%. The effects of


incix
stress ratio on G seem to be relatively unimportant for
max
stress ratio below the range 1.5 to 2.5. The reduction in

modulus is also dependent on the stress path as a result

of the relation between the direction of the major princi­

pal stress with respect to the orientation of the principal

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
295

axes of anisotropy. Thus the fact that G is not


e max

affected significantly by stress ratio reflects the step­

wise nature of the stress-strain characteristics of sands,

which in turn is associated with their frictional nature.

In other words, the stress-strain curve of a granular

material is really composed of small "steps," which essen­

tially seem to have the same initial slope, irrespective

of the shear strain amplitude. These initial slopes are

only up to 30Z smaller at strains approaching shear

failure.

The effects of stress ratio and hence of initial

shear stress, in principle, should be more evident as the

level of shear strain increases. In this connection, also


/
( shown in Figure 5.12 are the values of the secant and

tangent shear moduli for the specimen as a function of

stress ratio under torsional loading. The effects of

stre%: ratio on the small strain parameter and on the

large strain behavior (tangent modulus) are significantly

different. Accordingly, as the initial shear stress

increases, smaller cyclic shear strains should be required

to produce a given reduction in the secant shear modulus

with respect to G (59). The effects of stress ratio


r m ax
cannot be independent of the effects of shear strain

level. For low stress ratios (below about 2.0), induced

shear strains are small and hence, neither the maximum

shear modulus nor the shape of the modulus curve is

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
296

affected significantly by the initial anisotropic static


L stress conditions, as observed in Figure 5.11. However,

at higher stress ratios the modulus reduction curve may

not be independent of the initial static shear stress.

The actual relationship is still imperfectly known.

On the other hand, it is felt that resonant column

tests cannot reflect stress ratio effects even at rela­

tively large shear strain amplitudes because of the large

number of cycles applied by the oscillator to the speci­

men. Shortly after the first few cycles, the stress-

strain loops due to the superimposed cyclic shear stresses

correspond to unloading and reloading stages, regardless

of the initial shear stress (or strain) level. Conse­

quently, the resonant column tests measures the

unloading-reloading secant modulus after several hundred

cycles, which is considerably higher than, for example,

the tangent modulus during the stage of virgin loading in

the first cycle. The unloading-reloading modulus at very

small shear strains after unloading from the virgin curve

will vary with stress ratio in a similar fashion as the

G values presented in Figure 5.12.


max

In the torsional shear test, the number of applied

cycles is significantly smaller than in the resonant

column test. However, it is still unlikely that an

equivalent secant shear modulus, which only considers the

changes of stress and strain at the maximum amplitude

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
297

p o i n t s , be completely appropriate to account for stress

c ratio effects. If an initial static shear stress is act­

ing on the sand specimen, the difference in behavior

between the first cycle and the subsequent ones cannot be

measured by a single secant shear modulus. On the con­

trary, the stress-strain model should consider the changes

of stress and strain throughout each loading and unloading

stage, especially if correlations are to be made for exam­

ple with behavior under undrained cyclic loading. Conse­

quently, more attention should be given to the effects of

stress ratio on the first few cycles of loading.

In view of the foregoing discussion, the author con­

cludes that G values which inherently cannot account


max

( for stress ratio effects will not reflect the Initial

buildup in pore water pressure in an undrained cyclic test

under anisotropic stress conditions, which are charac­

teristic of most in situ conditions. Accordingly, the

attempted correlation between shear wave velocities meas­

ured at very small shear strain amplitudes and liquefac­

tion potential (29,32,33,89,147) becomes also question­

able. Furthermore, It is felt that not even drained

behavior can be correlated to G . There is experimental


max
evidence which suggests that the compressibility of sands

is controlled by the magnitude of the major principal

stress for a wide combination of o and stress ratio


m
values (102,105,186). In this connection, two specimens

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
298

under different o values will exhibit different G


/ m max
values but their compressibility may still be similar

depending on the relative magnitude of o^ and stress

ratio.

5.1.7 Effects of Stress History

Figure 5.13 presents the variation of shear modulus

with shear strain amplitude obtained from a hollow

cylinder specimen with a relative density of about 43.5%,

which was prestressed along the stress path ABA shown in

the Figure (overconsolidation ratio, OCR, of 2.0). The

results from a normally consolidated specimen at the same

mean confining stress are also shown for comparison. The

two modulus reduction curves are essentially the same.

Therefore, at the same mean confining stress, a , isotro-


m
pic prestressing with an OCR of 2.0 had a negligible

effect on the shear modulus - shear strain relation in

drained shear, for shear strain amplitudes ranging from


—4
10 % to 2%. Tatsuoka et a l . (172) also found that the

effects of isotropic prestressing on shear modulus values

were negligible.

Figure 5.14 shows the shear modulus values obtained

from two replicate specimens with the same void ratio and

initial state of stress, which were prestressed along the

stress path CBC shown in the Figure (OCR of 4.0). This

stress path corresponds approximately to the K q line

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
Reproduced with permission of the copyright owner. Further reproduction r

OT = 101.6 kPa
:> 80
_i
r>
Q
O
2
oc 60
< Test results after prestressing
UJ
X along ABA (OCR = 2.0)
to B /--J
200 • Resonant column
prohibited without perm ission.

40f «
£ 150 o Torsional Shear

50 100150 200
^ 'r.kP a
10 -« 101 1 10
SHEAR STRAIN. (%)

Figure 5.13 E f f e c t o f ' I s o t r o p i c P r e s t r e s s i n g on Shear Modulus - Shear

299
Strain Relationship.
Reproduced with permission of the copyright owner. Further reproduction

O ttaw a 2 0 -3 0 sand
Hollow cylinder specimens
NC Specimens
e = 0.623
G m ax=95.3
100 (awcTr =1&2 Dr = 45%
at same cr')
/ 101 kPa
50 kPa
= 67 kPa

n
0.
Gmax = 84
5
</> OC Specimen
r>
-i Test results a fte r prestressing
o
o along path CBC (OCR*4):
5
• Resonant column
oc
< ° Torsional Shear
UJ
prohibited without perm ission.

X
V)

100 200 300

SHEAR STRAIN, / ( % )

300
F ig u re 5.14 E f f e c t of A k i a l Co mp r e s s i o n P r e s t r e s s i n g on Shear Modulus -
Shear S t r a i n R e l a t i o n s h i p .
301

(axial compression) for Chls specimen. The results from a


G
normally consolidated specimen with corresponding void

ratio and state of stress (curve 1 from rig. 5.1i) are

also shown for comparison. Frestresslng along the K=2

line causes a degradation of the shear modulus of the

specimen at low shear strains, but the degradation is not


_2
significant for shear strain amplitudes above 2 x 10 %.

The unfavorable effect of the applied prestressing can be

attributed, in part, to the fact that the stress path dur­

ing prestressing differs from the one followed during

cyclic loading. As mentioned earlier, the prestressing

effect is considered to be more significant when the

reloading stress path is the same as the one followed dur-

^ ing the prior stress history, as a result of the stiffen­

ing effect associated with the orientation of normals at

the contacts in the direction of the major principal

stress. For other stress paths the effect may be more or

less favorable than the case illustrated in Figure 5.14.

It is important to observe that although both the NC

specimen and the prestressed specimen have been subjected

to the same maximum stress ratio of about 2 .0 , the maximum

shear modulus values are different (95.3 and 84 MPa,

respectively) due to the stress history effect. In addi­

tion to the stress path effect, it is felt that the reduc­

tion in modulus could be related to the level of shear

strain and consequent disruption of contacts induced

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
302

during the stress history. This level of shear strain

depends not only on the stress ratio but also on the mag­

nitude of the applied stresses.

Yu and Richart (198) also investigated the effects of

stress history on the maximum shear modulus. In their

study, prestressing consisted of loading and unloading by

increasing the stress ratio. K. Degradation of the shear

modulus, G , due to the applied stress histories was


max
observed for both compression and extension stress paths.

Yu and Richart (198) concluded that "it is the maximum

value of stress ratio that governs the modulus reduction,

and the effect of stress histories, involving lower stress

ratios is negligible." The results shown in Figure 5.14

show that prestressing along a stress path with constant K

also causes a degradation of the shear modulus hence

stress ratio alone cannon account for stress history

effects, as proposed. Consequently, the effects of over­

consolidation along a stress path with constant K must

also be considered. This fact is clearly illustrated by

the data shown in Figure 5.15, which show the variation of

G with overconsolidation ratio, OCR, obtained from a


max
solid cylinder specimen with a relative density of about

58%. The reduction in G in Figure 5.15 is attributed


max
only to the prestressing effect. The modulus values were

corrected to account for void ratio variations during

prestressing by using a void ratio function (58,88,198).

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
Reproduced with permission of the copyright owner. Further reproduction

No lateral strain
unloading (estimated)

(0
« 110"
a.
x
at Ottawa 2 0 -3 0 sand
E
0 1 0 0 -,at A' Solid cylinder specimen
CO e = 0.591
ID
_l
D Dr = 5 8 %
Q 90--
Path ABA© °"z = 103.9 kPa
O
2 Path ACA <7"r = 5 0 kPa
a: 1000 •

<
UJ 80 S. 800+
x
CO
prohibited without perm ission.

600- 2 /c
Path ABA
ID
2 70"
X
<
5
<>
200 400 6008001000
cr'r, kPa

6 8 9 10 11

OVERCONSOLIDATION RATIO

303
F igu re 5.15 Effect of Axial Compression Prestressing on Maximum Shear
Modulus of Medium Dense Sand Specimen.
304

Although prestressing decreased the void ratio, there was

also a reduction of up to 20% in the uncorrected shear

modulus. Hence, the effects of stress history on the

shear modulus are of the same order of magnitude than

stress ratio effects discussed in the previous section.

The shear modulus values measured on the same speci­

m e n at normally consolidated stages (points A, B, C, and D

in Fig. 5.15) are shown in Figure 5.16. These results

confirm that the relationship between shear modulus at

small strains and mean confining stress presented in Fig-


*
ure 5.5 is also valid for a as high as 600 kPa. However,
m
a slight decrease in the exponent of <?m is observed for a

change in relative density from 40 to 58%. This is attri­

( buted to the increased effect of particle interlocking as

the relative density increases, resulting in a relative


*
reduction in the contribution of a on the shear modulus.
I&
This is consistent with prior test results, as discussed

elsewhere (109).

The difference between prestressing and cyclic load­

ing stress paths must be considered when comparing the

cyclic stress-strain behavior and liquefaction potential

of normally consolidated and prestressed sands. Results

in Figures 5.14 and 5.15 reflect the effects of prestress­

ing along a stress path with constant stress ratio, which

result in G decreasing with increasing OCR. It is not


max
yet known whether this reduction in G will be reflected
J max

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
305

Speci men
(A
to

Dense
Medium
o
to
^3*

of
o

Modulus
to
CO CO
CL

Shear
E
CO
b

( CO
co

on
UJ
ac
t-

Confinement
CO
O
CNJ

O
O

CO
UJ
of
Ef f ect
CNJ

CD cc
5.16

to
Figure

o o
to to to
CM
1
.
smnaow u v 3h s

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
306

c in higher pore water pressures in undrained shearing.

This is considered to be unlikely. In fact, prestressed

sands have been found to exhibit a lesser tendency to pore

pressure buildup than NC sands (6,145). On the other

hand, a stress path with a constant stress ratio during

unloading may not be representative of actual in situ con­

ditions. For example, it is felt that the effect of OCR

on G may be more pronounced for no lateral strain tests.

It is well known that prestressing without permitting

lateral strain during the unloading stage produces an

increase in the lateral stress, which might become even

larger than the vertical stress for OCR values higher than

about 5 (67,115). Under these conditions, the fabric of

the sand deposit is likely to get stiffer in the horizon­

tal direction, which is the direction of particle motion

under the earthquake induced shear stresses. Moreover,

the increase in mean confining stress, itself causes an

increase in the stiffness o f 'the sand. Based on the

correlations developed by Mayne et a l . (115), the coeffi­

cient of earth pressure at rest, Kq , for Ottawa 20-30 sand

was estimated as a function of OCR, by using the expres­

sion

(Kq )oc = (1 " sin*') 0CRS±n<r (5.4)

At a void ratio of 0.591, the angle of shear resistance

was estimated to be on the order of 32.2° (115). The

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
307

estimated (K . values were 0.92, 1.26, and 1.62 which


° OC
correspond to OCR values of about 3.4, 6.1 and 9.8,

respectively. These values were used to estimate the

plane strain Gffiax values by correcting the values in Fig-


*

ure 5.15 according to the increase in o^ corresponding to

plane strain conditions under the same vertical stress, a


r * 2
of 103.9 k P a . The estimated values are also shown in Fig­

ure 5.15, resulting in an increase in G with OCR of up


° max
to about 16%. This trend does agree with the tendency to

higher liquefaction resistance with increasing OCR. These

values could be considered a lower bound of the effects of

plane strain prestressing on the maximum shear modulus,

G
max

Nevertheless, it should be noted that in either case,

stress history effects only change Gmax on the order of

15% to 20%. The changes in compressibility in drained

shear or in pore pressure response in undrained tests are

known to vary within a much wider range (23,82,202,209).

Hence, G , which reflects the behavior of sands at very


* max
small shear strains does not seem to be sensitive enough

to account for stress history effects on pore pressure

generation in undrained shear. The results of this inves­

tigation and those from previous studies on the effects of

prestressing in sands (7,23,101) permit to draw the fol­

lowing general conclusions:

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
308

c a. Prestressing affects significantly the stress-strain

characteristics of sands. These effects are related

to changes in void ratio, and in the conditions at

the contact points between neighboring grains.

Accordingly, the actual effects depend on the stress

and boundary conditions during prestressing, the

level of mean stress and relative density.

b. Prestressing along a particular stress path does not

necessarily result in stiffening of the sand specimen

when reloaded along other stress paths. When loading

and reloading occurs along a particular stress path,

the increase in stiffness is more significant for OCR

values in the range from 1 to 2 (see data in Refs.

7,23,109). For OCR values higher than 2 the addi­

tional increase in stiffness is moderate.

c. It is considered that a modular ratio, MR (ratio

between the overconsolidated and the normally conso­

lidated moduli) is strongly dependent on the type of

modulus and the strain conditions during unloading,

and that MR increases as the mean confining stress

decreases or as the relative density decreases (109).

In terms of Young's moduli for axial compression, the

modular ratio may be much larger when the stress

ratio is held constant during unloading, compared to

the condition of no lateral strain (23,109). This is

attributed to the effects of rotation of the planes

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
of principal stress, and hence in Che direction of

shearing, under plane strain conditions at high OCR

values C(Ko )0C > 4.0).

5.1.8 Effects of Cyclic Prestraining

The effects of cyclic prestraining on the shear

modulus are illustrated by the results presented in Figure

5.17. These results were obtained from specimens with the

same initial conditions as in some of the previous tests


^ »»
(D Z 45% and stress ratio, o /a » 2.0), but which had
r * z r *
been cyclicly prestrained in the resonant column mode.

The modulus degradation curve for a specimen without a

prestraining history (solid curve) is also shown for com­

parison. Before prestraining the specimen, a maximum

shear modulus, G , of about 95.3 MPa was determined in


* max
stage A. This value compares very well with that for the

normally consolidated specimen. In stage B, the specimen

was prestrained cyclicly using the resonant column mode


-2
with a shear strain amplitude of about 1.3 x 10 %. At

this strain level, the measured shear modulus (about 71

MPa) also agrees with the value for the NC sand. Follow­

ing stage B, the complete modulus degradation curve was

determined by means of resonant column tests (stage C) and

torsional shear tests (stage D ) .

The torsional prestraining (stage B) caused the shear

modulus to Increase above the value corresponding to the

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
Reproduced with permission of the copyright owner. Further reproduction

Test stages:
120
Low strain RC *

Large strain RC ■
NC specimen (Prestraining)
100
RC after B •

TS after C

a 80
2
(9
W RC3 Resonant column
D TS= Torsional shear
60
D
O
O
2
OC
40
prohibited without perm ission.

iu
X Ottawa 20-30 sand
(0 Hollow cylinder specimens
e=0.623
Dr=45%
20 °"z =100.2 kPa
cr'r =50 kPa

-4 -3 -2

SHEAR STRAIN, /( % )

Figure 5.17 E f f e c t of T o r s i o n a l P r e s t r a i n i n g on Shear Modulus - Shear


Strain Relationship.
311

NC specimen without any prestraining, for shear strains

below the shear strain amplitude used during cyclic pres­

training. The maximum shear modulus, G , increased from


max
95 to 100 MPa. For shear strain amplitudes larger than
-2
the previous maximum of 1.3 x 10 %, the shear modulus was

insensitive to the previous strain history. A similar

behavior has been observed in previous investigations

(42,160,172). The increase in shear modulus due to cyclic

prestraining has been considered to result from the elimi­

nation of local instabilities at the contact points, and

consequently better interlocking of the particles, even

without any general rearrangement or densification taking

place (36,42,47,196). The stiffening effect is greatest

as the number of cycles increases (42,160). For the


( present tests, the increase in shear modulus can also be

attributed, in part, to the fact that the loading paths

during prestraining (stage B) and reloading (stage C) are

the same, which favors an increase in shear modulus. This

was not the case when the two loading paths differed (Fig.

5.14). Nevertheless, it should be noted that the differ­

ence between the G values at stages A and C is rela-


max
tively small.

The effects of torsional prestressing and prestrain­

ing and of the level of shear strain associated with them

are further illustrated by the results obtained from tests

on two hollow cylinder specimens presented in Figure 5.18.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
:eproduced with permission of the copyright owner. Further reproduction

120

Test stages:
Gmax = 102
NC Specimen (A) Low strain RC A

(5) TS: 2 cycles D


* CD)
(Prestessing

Gmax = 91.8 (§ ) RC a fte r B

(D) RC a fte r c
(E) TS a fte r D

«
Q.
2 60
O
Increase in shear
<0
modulus due to
_J
3 prior strain history
D 40 -
prohibited without perm ission.

O O ttaw a 2 0 -3 0 sand
5 Hollow cylinder specimens
DC
< e = 0.623
UJ
Dr = 45 %
w 20
CT’z = 100.4 kPa
<7"r = 50.1 kPa

-4 -1
10 10 10 " 2 10

SHEAR STRAIN. X(%)

Figure 5.18 Effect of Torsional Prestressing on Shear Modulus - Shear


Strain Relationship.
313

For reference the modulus reduction curve for a normally

consolidated specimen without prior stress or strain his­

tory, but under the same initial conditions, is also given

in Figure 5.18 (continuous line). In stage A, the initial

maximum shear modulus, G , was measured to be about 94.7


max
MPa. In stage B, each specimen was subjected to two
-2
cycles of shear strain with amplitude of about 2.2 x 10 %

using the torsional shear mode. The secant shear modulus

values measured in the second cycle for the two specimens

were 32 and 36 MPa, respectively. These values correspond

to shear modulus ratios, G/G , of 0.34 and 0.38, respec-


nioX
tively. The influence of these two strain cycles, stage

B, on the stress-strain response at smaller shear strains

was evaluated during stage C using resonant column tests.


(
The two cycles of shear strain with an amplitude of
_2
2.2 x 10 % caused the shear modulus at low shear strains

(stage C) to decrease slightly below the initial Gmax of-

the specimen (G reduced from 94.7 to 91.8 MPa). This


r max
observation can be interpreted as the result of two coun­

terbalancing effects: an increase in G value due to pres­

tressing effects and a degradation of G associated with

the larger levels of prior shear strain. For the sequence

of stages A, B, and C, the latter effect is slightly dom- -

inant. After stage C, the specimen was subjected to a

second series of resonant column tests starting again at

very low strains (stage D ) . The prior strain history

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
314

ing stage C produced a regeneration of the shear modulus


L.
during stage D, giving modulus values higher than the

shear modulus of the specimen prior to any loading history

(G increased from 91.8 to 102 MPa). Therefore, the


max
prestraining effect due to the cyclic loading with

increasing shear strain amplitude during stage C overrode

the degradating effects of prestressing involving rela­

tively large shear strains (stage B ) . However, it should

be noted that stage C (resonant column tests) included a

significant number of cycles at small strains with

increasing amplitude, whereas in stage B (torsional shear

tests) only two cycles of strain were applied. Accord­

ingly, stage B has been described as a "prestressing"

stage.

Finally, the variation of shear modulus in the large

shear strain range was determined using torsional shear

tests (stage E ) . The effect that the prior strain history

(stages C and D) had on the shear modulus values at large

strain is considerable: the shear modulus increased by

about 65% between stages B and E at a shear strain ampli-


_ 2
tude of 2.2 x 10 %. At this strain level, the shear

modulus of the specimens prior to any cyclic strain his­

tory (results obtained during stage B) is much smaller

(32-36 MPa) than the values (55-60 MPa) obtained after the

specimens have been subjected to the strain histories

corresponding to stages C and D. Although the shear

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
315

_2
moduli at Y * 2.2 x 10 % are very different, it is impor-

tant to note that G of the prestrained specimens (102


m sx
MPa) and Gmax of the specimens prior to any strain history

(94.7 MPa) only differ by about 7%. Therefore, specimens

with essentially the same Gmax may exhibit a drastically

different "large strain" behavior. In the light of these

results, Gma x does not seem to be an appropriate parameter

to be correlated with the response of sands to cyclic


—2
loading involving shear strains larger than about 10 %.

Tests with various strain levels at stage B (two

cycles at constant shear strain amplitude) gave similar

results to those presented in Figure 5.18. These results

are summarized in Figure 5.19, which fully illustrates the

effects of prior strain history. The magnitude of these

effects seems to conflict with those observed by previous

investigators (160,172). For example, Silver and Park

(160) reported that the shear modulus values for fresh or

"virgin" dry specimens were slightly larger than for stage

tested specimens, in the shear strain range from 0.08 to

0.44%. Since staged testing involves the application of

repeated shear strains at increasing amplitude, the

effects of this technique on shear modulus correspond to

strain history effects. According to Figure 5.19, at y =

0 .1%, or greater, prestraining seems to have little

effect. On the other hand, Silver and Park (160) obtained

their modulus values by means of cyclic triaxial tests,

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
Reproduced with permission of the copyright owner. Further reproduction

120
Resonant column tests Torsional shear tests

100 Stage tested


specimen

80

«
I 60

Increase in shear
modulus due to
_i prior strain history
D
prohibited without perm ission.

O 40 O ttaw a 2 0 -3 0 sand
o
5 Hollow cylinder specimens oV
oc
< e « 0.623
UJ
m 20 Dr = 45%
c 'z = 100.4 KPa
’Virgin" specimens sO
cr'r = 50.1 KPa
(2nd cycle)

10 " 4 10 _1 1 4

SHEAR STRAIN, / (%)

316
Figure 5.19 E f f e c t o f S t r a i n H i s t o r y on Shear Modulus - Shear S t r a i n
Relationship.
317

which involve a limited number of cycles even if stage

L testing procedures are used. For these reasons, Silver

and Parkis results did not reflect effects of a large

number of prior strain cycles. However, if staged testing

includes the determination of shear moduli by means of

resonant column tests, it can lead to a significant

overestimation of shear modulus for sand with no prior

cyclic strain history even at shear strain amplitudes as

low as 1.1 x 10"2% (Fig. 5.19).

It is interesting to note that in Figure 5.17 a bene­

ficial effect of staged testing is not observed (compare

stages B and C ) . This is due to the fact that stage B

already includes stage testing effects. Since the meas­

urement of G in the resonant column mode implies the


(
application of a considerable number of cycles, which pro­

duces a significant stiffening of the specimen. Accord­

ingly, the measured shear modulus corresponds to the

unloading-reloading secant modulus after several hundred

cycles. It is not yet known at which shear strain level

the stress-strain characteristics of sands as measured by

resonant column tests, change as a function of the number

of cycles. According to the data in Figure 5.18, resonant


-4
column tests at shear strains of about 10 % do not pro­

duce stiffening of the specimen (compare stages A and B ) .

Most of the modulus reduction curves in this investigation

were obtained by using stage testing procedures, and hence

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
318

, the results are affected by strain history effects. This

does not invalidate any of the conclusions regarding the

maximum shear modulus. G . However, it should be


max
pointed out that the stiffening effects associated with

performing resonant column tests in the strain range from


_A —2
10 % to 10 % can override the degradating effects of

stress ratio (Fig. 5.11), prestressing (Fig. 5.14), and of

large strain prestraining (Fig. 5.18), in the range of


_3
medium to large shear strains (10 % to 0 .1%).

From the foregoing discussion, it is evident that the

number of stress (or strain) repetitions has an important

effect on stress-strain behavior even for "virgin" speci­

mens, which results in an increase in G with increasing

( number of cycles. This is attributed to the elimination

of local instabilities at the contact points and to the

changes in density. To quantify these effects, specimens

with no prior strain history were subjected to cyclic

loading in the torsional shear mode. Each specimen was

subjected to more than 50 cycles of constant shear strain

amplitude. Figure 5.20 shows the variation of shear

modulus with number of cycles obtained from these tests.

The values for prestrained specimens (at the noted shear

strain levels) with similar density and stress conditions

are also shown for comparison. Although G increases with

the number of cycles, it is noted that the stiffening

effect is more significant during the first 10 to 20

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
Reproduced with permission of the copyright owner. Further reproduction

60 Prestrained spec i m e n
r
000
5 0 --0 0 '
0 0©
O <,0 o o O o ° ® ° 0 ° ° 0 o G G
cd
Q.
2 40 00 00 O
O .0 ®
•»
CO S p e c i m e n with n o
ZD
_J 30 -- o prior strain history
D
Q
O Ottawa 2 0 - 3 0 sand
5
Hollow cylinder spec i m e n
cc 20
prohibited without perm ission.

<
UJ e = 0.633
X Dr = 4 0 . 7 %
CO
10 a 'z = 104.3kPa
-
a) Sh ea r strain, y = ±2.21 x 10 “2
°"'r = 50.0kPa
G = 94.1MPa
max
4- + + + + + +
5 10 15 20 25 30 35 40 45 50 55 60
N U M B E R OF CYCLES

319
Figu re 5.20 Effect of N um b e r of Cycles at C on sta nt Shear Strain Amplitude
on Shear M od ulu s,
Reproduced with permission of the copyright owner. Further reproduction

60

50 -f
Cyclic prestraining
73
CL Prestrained s p e c i m e n (Resonant c o l u m n m o d e )
§
0 40 - / /.
q Q O q Q 0 0 O0
co" 3O00000000
D O O 0 0 0 0 00 0® ® 0000O°

30 - 0 0 °°
S p e c i m e n with n o Ottawa 2 0 - 3 0 sand
prior strain history Hollow cylinder s p e c i m e n s
DC
<
HI 20 •*

X e = 0.62
prohibited without perm ission.

CO Dr = 4 6 %
°"z = 105.6kPa
10 -
°"r = 50.9kPa
b) S h ea r strain, Y = ±4.4 x 10 "2 %
G m a x = 9 5 -8 M P a

10 15 20 25 30 35 40 45 50 55
N U M B E R OF CYCLES, N

320
Figure 5 . 2 0 , continued.
321

c in
in

c
in

o
Oi co
CO
HI
in
CM o
>
o
li.
o
cc
111
in CD
eo "o S
Ql 3
Z

continued
CM CM <0 CO CM
CO Tf' 0 ©

in
5.20,
Figure

o o o o
co in CO CM

(Bdv\i)0 ‘s m n a o w dvaHS

R eproduced w ith perm ission o f the copyright owner. F urth er reproduction prohibited w itho ut perm ission.
322

cycles, depending on the shear strain amplitude. This

range corresponds approximately to the significant number

of cycles in an earthquake. Figure 5.20 also permits

visualization of the tremendous effects of prestraining

due to small shear strain cycles of increasing amplitude

(resonant column tests). However, it is clearly noted

that the effects of small strain prestressing decrease as

the shear strain amplitude increases (compare Fig. 2.20a

and 2.20c). Moreover, it is interesting to note that the

variation of G with number of cycles is considerably

smaller for the prestrained specimen than for the "virgin"

specimen.

These findings have an important implication regard­

ing commonly used numerical methods for dynamic analysis.

The simplified procedure which consists of using a single

modulus degradation curve (146,154) or defining an

equivalent shear modulus at the tenth cycle (8 8 ) is not

conceptually satisfactory since the strain history effects

resulting from the previous cycles or from the relative

order in which the stress cycles are applied in a given

loading sequence, are not accounted for.

In summary, it can be concluded that the influence of

prior stress or strain histories is controlled by three

different factors:

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
323

1. the prestress effect itself, which is dependent on

c stress path and strain level,

2. the destructuration that occurs as shear strain

increases, and

3. the number and magnitude of prior cyclic strain or

stress repetitions.

The second factor has an effect opposite to that of

the other two. In this regard, the results in Figs. 5.14

to 5.20 clearly illustrate the important difference

between prestressing effects from those due to cyclic

strain history.

The test results previously discussed show that in

addition to shear strain level, the stress-strain charac­

teristics of sands under drained cyclic loading are

affected by relative density, mean confining stress,

stress ratio, stress history and prior number of strain

applications. However, the behavior at very small shear

strains, as measured by G is mainly affected by the


* J max J
mean confining stress and Dr. The effects of stress

ratio, stress history and cyclic prestraining on Gmax are

not significant. Consequently, a direct correlation

between G and "large" strain behavior such as liquefac-


max 6
tion potential which is affected by those factors, is not

appropriate. For the same reasons, the attempted

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
3 24

correlation between shear wave velocities measured at very

small shear strains and liquefaction potential of natural

sand deposits becomes also questionable.

5.2. Dndrained Cyclic Shear Tests

5.2.1 Results of Typical Stress and Strain Controlled Tests

Figure 5.21 presents the results of a test on a solid

specimen with Dr = 63% (e = 0.578) consolidated isotropi-


*
cally to a mean confining stress, a , of 100 k P a . A
m
cyclic shear stress with a maximum amplitude of 26.8 kPa

was applied at a constant strain rate. The residual shear

strain that developed at the end of each of the first two

cycles is very small (Fig. 5.21a) while the pore pressure

Increased progressively at a relatively rapid rate (Fig.

5.21b). Then, the residual shear strains become larger

with the largest net increment occurring during the first

half of the 4th cycle. It is important to note that after

the 3rd cycle, the mechanism of pore pressure generation

changes drastically. Peak pore pressures develop whenever

the applied shear stress approaches zero, while a decrease

in pore pressure is observed after large cyclic strains.

At the end of the 4th cycle the pore pressure becomes

equal to the confining pressure, that is, "initial

liquefaction" develops. The pattern repeats itself

thereafter with continual reduction in the net increment

of the residual strain at the end of each cycle (Fig.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
325

c
D r- 63%
Crz -CTr -1 0 0 kPa O tta w a 2 0 - 3 0 sand
T z e - 2 6.8 kPa Solid cylinder specimen
STRAIN, /< % >

4—

2 .. Load
cycle

\2Q 40 <60 80 100


SHEAR

2 --

4 --

H alf
cycle

TIM E. SEC .

C onfining p ressu re
PRESSURE, kPa
PORE

50-L
H alf
cycle
Load
.. cycle
• Points o f z e ro shear s tre s s

20 40 60 80 100 120

TIM E, SEC.

Figure 5.21 Undrained Torsional Shear Test on Isotropi-


cally Consolidated Specimen (Dr=63%).

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
326

20

SHEAR STRAIN.X(%)

40 ,
P h a s e tra n s fo rm a tio n lin e
(

o 10*0
1 3 kPa

J
>. (2 -3 7 ° (^>-490)
-4 0

EFFECTIVE CONFINING STRESS. P' <*Pa)

Figure 5.21, continued.

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
327

5.21a). As discussed In Chapter 3, this behavior Is

characteristic of dense sands.

Figure 5.21 shows that the stress-strain curves dur­

ing the first two cycles of loading resemble a simple

closed loop. However, as loading progresses the stress-

strain relationship becomes complex and simplification of

the behavior in terms of an equivalent secant shear

modulus is no longer appropriate. It is noted that the

residual strains in one direction of rotation are larger

than those in the opposite direction, although the

corresponding rotation of principal stresses is symmetri­

cal with respect to fabric anisotropy. It is thought that

this may be due to the fact that a significant increment

c in residual strain first occurs in that direction, which

corresponds to the direction in which loading was ini­

tiated. It is also important to note that during the last

loading cycles large cyclic shear strains occur as the

shear stress reverses sign (e.g. goes from positive to

negative.). At these stages of the test the specimen

exhibits a very low strength since the pore pressure

reaches a value close to the confining pressure.

The effective stress path plot (Fig. 5.21d) permits

visualization of the relationship between pore pressure

buildup and strain development during cyclic loading. As

the number of cycles increases the effective stress path

gradually moves towards the left. However, a drastic

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
328

change in curvature, or elbow, in the shape of the stress


c path is observed in the second half of the third cycle

(point A in Fig. 5.21d). Dilation develops and the stress

path moves back towards the right, reflecting a reduction

in pore pressure. Hence, as the effective confining pres­

sure decreases and the stress path approaches the failure

line, the behavior changes from contractive to dilative on

loading. Larger shear strains are also observed at this

stage (Figs. 521a and 5.21c). Upon reversal in the direc­

tion of shearing (point B in Fig. 5.21d), a sharp increase

in the pore water pressure is observed and the stress path

moves again towards the left. The author associates this

behavior to the fact that the sudden reversal in the

c direction of shearing causes a downward motion of the

grains in the "dilated” structure of the sand specimen and

probably a reduction in the number of contact points

between neighboring grains. Consequently, a large strain

increment occurs after the shear stress reverses sign

(point C in Figs. 5.21c), reflecting a reduction in the

strength of the specimen. However, further straining

causes dilation to develop (new contact points develop)

and the stress path moves back towards the right (point D

in Fig. 5.21d) as the pore pressure decreases. The

effects of dilation are clearly reflected on the steepen­

ing of the stress-strain curve between points D and E in

Figure 5.21c. The contractiveness increases in the next

unloading stage and the stress path moves finally towards

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
329

a state of zero effective st re ss , corresponding to the

occurrence of Initial liquefaction. As the shear stress

reverses (point F in Fig. 5.21c), the stress-strain curve

becomes very flat as a result of the loss of strength.

However, once again, after further straining dilation

develops, the pore pressure decreases and the stress path

moves back along the failure line. The development of new

contact points after large shear strains allows dilation

of the sand skeleton since the effective confining pres­

sure is extremely low. Thereafter, the specimen experi­

ences cycles of limited liquefaction on unloading and

large deformations with gradual stiffening on loading in

either direction. However, it is noted that the increment

of residual strain at the end of each cycle apparently


( d e c r e as es .

The states of stress at which the behavior changes

from contractive to dilative on loading defines a locus .of

points that have been called lines of phase transformation

(82), which mark a drastic change in the mechanism of pore

pressure generation and in the consequent stress-strain

characteristics of sands. As the loading stress paths

cross the phase transformation line, dilation develops

causing a reduction in pore pressure and in the rate of

shear strain accumulation. On the other hand, as the

unloading stress paths move into the area limited by the

two phase transformation lines contractiveness increases,

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
330

which is reflected in high pore water pressures and in the

development of larger shear strains. Accordingly, the

corresponding stress-strain curves become flatter. It

will be shown later that once the stress path is at a

state of zero effective stress (liquefaction), the amount

of shear strain required to mobilize a given shear resis­

tance, in the opposite direction, increases significantly

as the relative density of the sand decreases. After the

collapse of the structure due to the reversal in the

direction of shearing, the looser the specimen the larger

the shear strains required for new contact points to

develop.

Figure 5.22 presents the results of another tests on

a looser solid specimen with Dr = 50% (e = 0.61) consoli-

dated isotropically to a mean confining stress, a of


m
104.3 kPa. A cyclic shear stress with a maximum amplitude

of about 23.0 kPa was applied at a constant strain rate..

As in the previous test (Fig. 5.21), although the pore

pressure increased progressively, the shear strains that

developed during the first two cycles of loading are rela­

tively small (Fig. 5.22a) and the corresponding stress-

strain curves resemble a simple close loop (Fig. 5.22c).

However, a large shear strain increment accompanied by an

additional large increase in pore pressure (Fig. 5.22b)

developed in the first loading stage of the 3rd cycle.

This stage of deformation and pore pressure buildup

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
331

Ottawa 2 0 -3 0 sand
Solid cylinder specimen o"z - o"r - 104.3 kPa
e - 0.61 Dr - 50 .% r cy - 22.6 kPa

Load cycle
z
<
cc
I-
co
cc 20 30 40 ,50 60 70 80
<
til
I -2
CO

-4 Half cycle

-6 ■

TIME, sec.
-8 ■

Confining pressure

100 ■

( 90

80 ■ Half
CO
CL cycle
. 70 •
w
cc
3 60 ■
CO
CO
£ 50 ■

B 4.:
o
a
30 ■

20
Load cycle

20 30 40 50 60 70 80

TIME, sec.

Figure 5.22 Undrained Torsional Shear Test on Isotropi-


cally Consolidated Specimen (Dr=50%).

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
332

30

25

20
STRESS, kPa

-10 -9
SHEAR

4th cycle -5

-10

3rd cycle

-20

-2 5

SHEAR -3 0 STRAIN, %

40 ■

30 ■

20 ■
SHEAR STRESS, kPa

IU ‘

■JO 20 40 50 80 90 100 T t = 8.7

-1 0 •
t kPa

-20 ■

Phase transform ation


^-30 • line a -3 2 .5 (0 1 = 39.6)

EFFECTIVE CONFINING STRESS.kPa


-4 0 •

Figure 5.22, continued.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
333

^ occurred under an approximately constant shear stress

amplitude, which is reflected by the fact that the stress

path moves almost horizontally towards the failure line

(between points A and B in Fig. 5.22d). It should be

noted that this type of behavior was not observed in the

specimen with Dr of 63% (Fig. 5.21).

After a shear strain of about 3% a slight dilation is

apparent, which slows down the rate of deformation and

allows the specimen to buildup strength up to the stress

amplitude at which reversal of rotation occurs (point B in

Fig. 5.22d). At this point the specimen collapses almost

completely causing a sharp increase in pore pressure,

which becomes close to the confining pressure. As the

( shear stress reverses sign the stress-strain curve becomes

flat (point C in Fig. 5.22c) reflecting the loss of

strength. This state practically corresponds to initial

liquefaction. As a result of dilation with further

straining the pore pressure decreases, the specimen builds

up strength at an increasing rate while the stress path

climbs along the failure line from point C to point D

(Fig. 5.25d). Reversal of the direction of shearing at

point D causes a second collapse of the structure bringing

again the specimen to a liquified state-. It is interest­

ing to note that after "liquefaction" at point C a shear

strain of about 8% is required for the specimen to build

up a strength of 23 k P a . However, after the second state

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
334

of liquefaction the strength mobilized after a shear

strain of 12% is only 10 kPa. Therefore, in this case,

the shear strains which accompany the development of tran­

sient states of liquefaction increase as cyclic loading

progresses. This is the behavior characteristic of medium

loose s a n e s • -

The behavior of medium loose specimens is further

illustrated by the results presented in Figure 5.23, which

were obtained from a test on a solid specimen with a Dr of

48%. The specimen was subjected to cyclic loading with a

smaller shear stress amplitude (18.2 kPa) than in the pre­

vious tests in order to obtain more detailed information

on the mechanisms of pore pressure and strain development

with increasing number of cycles. The residual shear

strain that developed at the end of each of the first six

cycles is very small (Fig. 5.23a) while the pore pressure

increased progressively with each loading cycle (Fig.

5.23b). During the loading stage of the second half of

the 7th load cycle, large shear strains suddenly develop

together with a sharp increase in pore pressure (point A

in Figs. 5.23a and 5.23b) and a small drop in the strength

of the specimen (Fig. 5.23c). It is evident that at this

stage the structure of the sand has partially collapsed,

thereby imparting the brittle (strain softening) behavior

observed in Figure 5.23c. Once the sand structure starts

to collapse (at - 19.6°), the specimen becomes

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
335

5t
L 4
O tta w a 2 0 - 3 0 san d
S olid c y lin d e r specim en
=<r-r » 100.8 kPa
= 18.2 kPa
3 e - 0.614 G m ax ■ 116.6 MPa
Dr - 4 8 %
2

* 1 L o a d c y c le
X 1/ 2 3

< 80 100 180 200 220 240


cc 1
H
CO H a lf c y c le
cc 2
<
UJ
X 3
CO

5 TIM E SEC.

110 T
C o n fin in g p re s s u re
100
( 90 -

80 -
€ axial
I«L 7 0 ..

6 0 -■
UJ
cc
^ 5 0 ■■
CO
CO
g 40
Q. H a lf c y c le
UJ 30 ■ L o a d cycle
CC
o
0. 20-- I 2
1
10 -

20 40 60
TIM E, SEC.

Figure 5.23 Undrained Torsional Shear Test on Isotropi-


cally Consolidated Medium Loose Sand
Specimen (Dr=48%).

R eproduced w ith perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
336

25 t
7th c y c le
20 .
T (kPa)
STRESS,

-5
SHEAR

8th c yc le
-1 0

S H EA R STRA IN , X %
40
STRESS, q(KPa

30-
Phase transform ation
line „
20 --

10 - L o ad cycle
SHEAR

..C 20 30 40 50 BD 9 0,

-1 0 •
M A X IM U M

-20

E F F E C T IV E C O N FINING STR E S S , p ' (KPa)

Figure 5.23, continued.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
337

unstable, and the stress path moves quickly towards the

failure line. It is noted that the almost horizontal

movement of the stress path towards the failure line

(between points A and B in Figure 5.23d) is larger than

that observed during the 3rd cycle In the previous test

(Fig. 5.22d), reflecting a larger increase in pore pres­

sure even though the shear stress amplitude is smaller.

The effective stress path in Figure 5.23d also shows that

the specimen becomes unstable when the mobilized friction

angle is only a fraction of the maximum angle of shear

resistance. The mobilized friction angle at peak strength

corresponds to a critical principal stress ratio of 2.0

which is reached at a shear strain of about 0.3% (Fig.

5.23c), whereas the pore pressure at the same strain

corresponds to 45% of the confining pressure. As dis­

cussed in Chapter 3, this behavior is characteristic of

medium loose to loose sands. The conditions determining

the collapse of the structure under cyclic torsional load­

ing will be discussed in more detail in a later section of

this Chapter.

After a shear strain of about 1.3%, dilation becomes

apparent resulting in a slight gain in strength with

Increasing strain as reflected by some steepening of the

stress-strain curve. However, further straining is

abruptly interrupted at a shear strain of 2 .2% when the

reversal of rotation occurs. At this point (B in Fig.

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
338

5.23) the pore water pressure has increased to about 60%

of the confining stress. Reversal of the direction of

rotation and the sudden rearrangement of grains associated

with it (collapse) cause once more a drastic change in the

mechanism of pore pressure generation, which is clearly

evident by the change in curvature (elbow) in the shape of

the stress path. The pore pressure continues to increase

as the shear stress is reduced to zero. At an early stage

of reloading the pore pressure reaches a value close to

the confining pressure (point C in Figs. 5.23b and 5.23d).

The specimen then exhibits a very low strength, and very

large shear strains develop (8 th cycle in Fig. 5.23c).

After a shear strain of about 8% the specimen has only

mobilized a strength of 7 kPa. At this strain level,

dilation is again apparent (a second "elbow" is observed

at point C in Fig. 5.23d) and the stress path is moving

back along the failure line. The corresponding stress-

strain curve is becoming slightly steeper (Fig. 5.23c).

Figure 5.23c shows that the stress-strain behavior

during the first six cycles of undrained loading resemble

a simple closed loop. The equivalent secant shear modulus

values are plotted on Figure 5.24 as a function of the

average mean confining stress during each load cycle. The

shear modulus values calculated from results of stage-

tested dry specimens at the same mean confining stress and

shear strain level are also shown for comparison. The

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
Reproduced with permission of the copyright owner. Further reproduction

e = 0.614
Dr = 4 8 %
G m a x = 116.6MPa
t eye = 18.2kPa
40-

o Measured values
• Estimated from results of
30- staged tests on dry specimen
<
Q.
5
Sw*

(D
20-
CO
D
_l
D
Q
prohibited without perm ission.

O
5 10-- Sample collapse
CC
(7=0.176%)
<
UJ
L o ad cycle
X
CO
6 5 4 3 2 1 0

20 30 40 50 60 80 90 100 110

E F FE C T I V E C O N F I N I N G S T R E S S , p' (kPa)

'Igor. 5.24 Shear M o d u l u s - E f f e c t i v e C o n f i n i n g St ress R e l a t i o n s h i p d uri ng


u>
Undrained Cyclic Loading (Data fron Fig. 5.23c and 5.23d). U)
VO
340

measured and estimated shear modulus values resulted in

essentially the same curve, except at small shear strains.

This difference is attributed to staged testing effects

which are likely to be more pronounced in the dry speci­

men. However, as discussed earlier, prestraining has lit­

tle effect for shear strain amplitude larger than about

0.1%. Consequently, over this range the modulus values

agree very well. It is interesting to observe that the

equivalent shear modulus during undrained loading

increases slightly during the first few cycles reflecting

also strain history effects. These effects also cause a

reduction in the rate of pore pressure buildup with

increasing number of cycles during the first six cycles

(Fig. 5.23b), which will be discussed further in a later


c section. In the second half of the 7th cycle the specimen

collapses and the stress-strain relationship becomes com­

plex. As pointed out earlier, simplification of the

behavior in terms of an equivalent secant shear modulus is

no longer appropriate.

Figure 5.25 illustrates the behavior of a loose

specimen with Dr = 44%, which was subjected to cyclic

loading with a maximum shear stress amplitude of 16.5 kPa.

The residual shear strain that developed at the end of

each of the first nine cycles is very small (Fig. 5.25a)

whereas the pore pressure increased progressively,

although at a decreasing rate. In the first loading stage

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
341

10
V -
Ottawa 20-30 sand o-'z ,o " r - 100.3 kPa
Said cyfnder specimen Tcy ■ 16.5 kPa
e - 0.625
Dr - 44 %
60 120 180 240 300 3 60
ye 0
S -2
<
« —4
CO
oc -6 10 th cycle
<
£ -8
co
-10

-1 2

-1 4
-1 6
TIME, sec.
-18
-20
Confining pressure
100
'S te a d y - state'
deformation
( 90 ■

80
m
CL
JC 70
a.
3 60
CO
CD
UJ
tr 50
CL
UJ 40 Load cycle
CC
o
CL
30
Collapse
20

60 120 180 240 300 360

TIME, sec.

Figure 5.25 Undrained Torsional Shear Tests on Isotropi-


cally Consolidated Loose Sand Specimen
(Dr=44%).

R eproduced w ith perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
342

20
STRESS, kPa

-6 -4 -2

S tead y-state
SHEAR

deformation -4

-8

-16
SHEAR STRAIN, %
■ -20

40

30 ■

20 ■
STRESS, kPa

10 •

20 30 40 50 60 83 90
SHEAR

(fr'mob = 15.5"

.* S tea d y -s ta te "
-20
deformation
vff'= 30.7"(<£’ =36.41
-30

EFFECTIVE CONFINING STRESS,kPa

Figure 5.25, continued

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
343

of the tenth cycle, the specimen becomes unstable exhibit­

ing a marked strain softening behavior after a cumulative

residual shear strain on the order of 0.3%. The collapse

of the structure is manifested by a significant increase

in pore pressure and a consequent reduction in strength.

The stress path moves quickly towards the failure line

reflecting the increase in pore water pressure and the

reduction in strength (between points A and B in Fig.

5.25d). The mobilized friction angle at the point of col­

lapse (A in Fig. 5.25) is only 15.5°, whereas the cumula­

tive pore pressure corresponds to only 39% of the confin­

ing pressure. However, at point B the pore water pressure

has increased up to 88% of the confining pressure, whereas

the mobilized friction angle corresponds to 36°, the large

c strain angle of shear resistance. These results clearly

support the author's view that loose sands show a strain

softening behavior because at some stage in the loading

process there is a sudden rearrangement of grains or loss

of contact points between neighboring grains, thereby

transferring load from the soil skeleton to the water.

This is the only plausible explanation for the drastic

change in the rate of pore pressure buildup that is

observed at point A in Figure 5.25b. It should be noted

that the results presented in Figure 5.25 were obtained

from a strain controlled test. Under load-controlled con­

ditions, the "collapse" and consequent strain softening

behavior are likely to develop in a considerably shorter

period of time (14).

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
344

The process of continuous deformation after the "col-


<*■ *
L lapse" at point A ($ , Z 15.5 ) in Figure 5.25d causes a
mo d
reduction in the rate of pore pressure buildup (Fig.

5.25b) and consequently a reduction in the rate of loss of

strength (Fig. 5.25c). It is thought that this is due to

the fact that the tendencies to contract on shearing gra­

dually decrease as the arrangement of particles approaches

the "critical state," as discussed in Chapter 3. After a

shear strain of about 7% the pore pressure becomes essen­

tially constant and the specimen reaches a steady-state of

deformation (131) at a residual strength of about 7 kPa

(point B in Fig. 5.25d). Further straining beyond a shear

strain of about 12% causes dilation of the sand skeleton,

the pore pressure decreases and the stress path moves back

( along the failure line. Steepening of the stress-strain

curve is also evident (Fig. 5.25c). It should be noted

that in the test presented in Figure 5.25 a condition of

initial liquefaction did not develop because after the

collapse the specimen deformed undirectionally without

mobilizing a strength equal to the shear stress amplitude

at which the reversal of rotation was programmed to occur

during the test (T 16.5 kPa) . After a shear strain of

about 19% in the clockwise direction the test was manually

stopped. On the other hand, strain softening behavior did

develop because the cyclic shear stress amplitude was

apparently larger than the "steady-state" strength

(T 7 k P a ) , as discussed in Chapter 3. This relationship

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
345

will be discussed further later on.

L
The previous tests results provide a clear picture of

the undrained behavior of Ottawa 20-30 sand under cyclic

loading over the range in relative density corresponding

to initial states from medium dense to loose. The results

confirm the behavior illustrated schematically in Figures

3.36, 3.38 and 3.40 in Chapter 3. From the detailed

description of these tests it is evident that under cyclic

loading loose sands may experience very large deformations

even without the development of a condition of initial

liquefaction (Fig. 5.25). This is attributed to the col­

lapse of the structure oil l o ad in g, which leads to a sharp

increase in pore pressure and a consequent reduction in

^ strength. Since this collapse occurs at a stress ratio

significantly smaller than that corresponding to the phase

transformation line, it will always occur before a condi­

tion of initial liquefaction can develop, provided the

shear stress amplitude is larger than the steady-state

strength. Consequently, a specified level of shear strain

in loose sands under cyclic loading may result from either

the shear strains that develop before the stress path

reaches the phase transformation line (including strain

softening beha vi or ), or the strains accompanying the

development of initial liquefaction, or a combination of

the two mechanisms of strain accumulation, as observed in

Figures 5.22 (3rd cycle) and 5.23 (7th and 8th cycles).

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
346

On the other hand, as the relative density increases, the


c strains that develop before the stress path reaches the

phase transformation line decrease (see Fig. 5.21) and

therefore, a given level of shear strain is mainly on the

account of the strains associated with transient states of

liquefaction after the stress path crosses the phase

transformation line. However, the denser the sand the

lower the rate of strain accumulation after initial

liquefaction.

Strain controlled tests have also been used to evalu­

ate the behavior of saturated sands under cyclic loading

(34,43,75,161). Typical results from this type of tests

are presented in Figure 5.26, which were obtained from a

test on a hollow cylinder specimen with a Dr of 55.6%.

The specimen was consolidated isotropically to a mean con­

fining stress of 101 kPa and subjected to cyclic loading

in the strain controlled mode of the new apparatus, with a

shear strain amplitude of about 0.'67%. It may be seen

that under strain controlled conditions the mobilized

shear stress gradually falls off (Fig. 5.26a) as the gen­

erated pore water pressure increases (Fig. 5.26b). How­

ever, it should be noted that the reduction in the mobil­

ized shear stress produces a reduction in the rate of pore

pressure buildup until a condition of initial liquefaction

develops. This condition is manifested by the fact that

the specimen does not mobilize any strength over the range

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
40
V Ottawa 20-30 sand e - 0.597
Hollow cylinder specimen Dr - 55.6 %
30
o”z =o-’r « 101 kPa
Xcy - = 0.67 %
20
a
a.
JC
co
CO
Ui
cc

cc I240 306 36
< 60 1210
UI
X
» -1 0

-20
Load cycle

-30 TIME, sec.

-40

Confining pressure

c CO
CL
JC
100

uj 80
or
D
CO
CO Load cycle
Ui
8
&F 60
UJ
cc
o
40

o Points of zero shear strain


20

60 120 180 240 300 360 420 48 0

TIME, sec.

Figure 5.26 Typical Results of a Strain Controlled


Torsional Shear Test on Isotropically
Consolidated Specimen.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
348

40

30

20
Q.

= - 1.0 - 0.8 '0.6 0.8 1.0 1-2

-20

-30
SHEAR STRAIN. %
-40
50

c 40

30
o
a.
20

3 0 / 40 \' 50
cc t=11
ui -1 0 kPa

-20

-30

-40
EFFECTIVE CONFINING STRESS.kPa
-50

Figure 5.26, continued.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
349

within which the shear strains are controlled. The hys­


c teresis loops showing the stress-strain relationship for

the saturated sand become progressively flatter with

increasing number of cycles (Fig. 5.26c). At initial

liquefaction, the slope of the stress-strain curve is

practically zero. These results clearly indicate that

once initial liquefaction develops shear strains much

larger than 0.67% are required for Ottawa 20-30 sand at a

Dr of about 56% to be able to mobilize any strength.

As cyclic loading progresses the effective stress

path moves towards the left reflecting not only the

increase in pore pressure but also the reduction in the

mobilized shear stress (Fig. 5.26d). The change from con-

^ tractive to dilative behavior on loading is clearly evi­

dent. However, as the strains are controlled, the

increase in pore pressure upon reversal in the direction

of shearing gradually decrease (Fig. 5.26b). It is

interesting to observe that the decaying peak points of

the effective stress path lie along a slightly curved line

which passes through the origin, reflecting a clear rela­

tionship between the shear strain amplitude and the mobil­

ized angle of shear resistance, <j> , , which decreases


BOD

slightly as the effective confining stress increases. In

other words, strain controlled tests are approximately

equivalent to tests in which the principal stress ratio

(which depends on <j> , ) is maintained constant. A


moo

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
350

c comprehensive discussion on the relation between $

pore pressure development was presented in Chapter


nob
. and

3.

The discussion of the results in Figure 5.26 permits

the author to emphasize once more the significant effects

of the boundary conditions on the response of saturated

sands to cyclic loading. First of all, the drastic

difference between stress controlled (Figs. 5.21 to 5.25)

and strain controlled tests (Fig. 5.26) is evident. In

stress controlled tests the principal stress ratio and the

consequent shear strains increase gradually as a result of

the pore pressure buildup. Moreover, the rate of pore

pressure buildup suddenly increases as the specimen col­

lapses or as it is brought close to a state of liquefac­

tion. On the other hand, in strain controlled tests the

principal stress ratio remains approximately constant for

a given shear strain amplitude, whereas the rate of pore

pressure accumulation decreases with number of cycles. In

the author's opinion, the results of these two types of

tests cannot be correlated directly, as suggested by

Sil\rer and Park (161). They concluded that the number of

cycles and the corresponding mobilized shear stress at

which the pore pressure in a strain controlled test

becomes equal to 50% of the initial confining pressure can

be taken as the equivalent undrained strength (in terms of

cyclic stress ratio and number of cycles) under stress

controlled conditions. The results presented in Figure

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
351

5.26 clearly do not meet this criterion, since after the

first cycle the pore pressure is already close to 50% of

the confining stress. It is evident that more than one

cycle at a constant shear stress amplitude equal to the

shear stress mobilized in the first cycle would be

required to develop a condition of initial liquefaction.

(However, it should be noted that the strain controlled

triaxial tests performed by Silver and Park (161) only

covered the range of axial strains from 0.03 to 0.31%).

In addition, since the stress ratio remains constant

throughout the strain controlled test, it cannot reflect

some features of stress controlled behavior especially in

the case of loose sands. For example, if the stress ratio

associated with the maximum shear strain amplitude is


c smaller than that corresponding to the critical stress

ratio (CSR) line (see Figs. 3.30 to 3.32), strain soften­

ing behavior cannot develop in a strain controlled tests.

Nevertheless, a basic relationship must exist between

stress and strain controlled tests as a result of the

direct interrelationship between shear strain amplitude,

pore pressure buildup and mobilized angle of shear resis­

tance, as discussed in Chapter 3. However, in the

author's view, the issue is not whether results from

stress and strain controlled tests can be correlated but

to delineate the conditions to which the conclusions based

on results from each type of tests are applicable. Each

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
352

in situ situation is characterized by an initial state of

stress and by specific strain boundary conditions, which

affect the response under cyclic loading. For example, as

discussed earlier, in a soil deposit under level ground

lateral strains are constrained. Consequently, the cyclic

loading due to an earthquake is applied under conditions

corresponding to plane strain. It was shown that under

these ci ditions the total lateral stress increases with

number of cycles in such a way that the effective stress

path moves down and towards the left whereas the principal

effective stress ratio at the end of each cycle remains

approximately constant. It is evident that this stress

path resembles the envelope of decaying peak points of the

stress path in Figure 5.26, which also corresponds to an

approximately constant stress ratio. Moreover, it should

be noted that cyclic loading under plane strain conditions

may lead to the development of initial liquefaction,

irrespective of the initial state of stress or strain. In

other words, a wide range of initial conditions may have

the same limiting condition (liquefaction). The initial

state of stress changes during cyclic loading as much as

required by the imposed boundary conditions. It is clear

that under these conditions the effect of the initial

state of stress (or strain) on the behavioral trend

decreases significantly. (Nevertheless, it should be

noted that the further the initial state of stress is from

the origin of the q-p' plot, the larger the number of

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
353

cycles at a given shear stress (or strain) amplitude

required for a sand to reach a condition of initial

liquefaction under undrained conditions.) In fact, this'

conclusion has been drawn based on results from strain

controlled cyclic tests (69,70). However, it is not

applicable to cases in which strain controlled conditions

do not prevail such as in the case of soil elements close

to sloping ground. Lateral strains are not constrained

and the Initial shear stresses are sustained during the

application of cyclic loading due to the need to maintain

equilibrium imposed by geometry (2). These initial shear

stresses have a significant effect on the cyclic undrained

behavior (see discussion in section 3.4.2).

A remark on the advantages of the torsional shear


c
tests can be drawn from the test results presented in Fig­

ures 5.21 to 5.26. In the first few cycles subsequent to

the first cycle, the stress paths are essentially symmetr­

ical about the normal stress axis since the mean confining

stress and the orientation of the principal stresses is

such that the response of the specimen is expected to be

the same whether it Is rotated right or left. This condi­

tion cannot be obtained in cyclic triaxial tests.

5.2.2 Threshold Shear Strain

First of all, it is noted that in the test results

presented in Figures 5.21 to 5.26 pore pressure buildup

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
354

becomes apparent during the first loading cycle at a shear

o stress level, t , which ranges from about 5.0 to 13 kPa,

depending upon the relative density of the sand. Below

this shear stress level the stress paths are essentially

vertical. The range in t corresponds to a shear strain

of about 0.012% to 0.018%. It may be expected that the

accumulation of pore water pressure under cyclic loading

at a shear stress (or strain) amplitude below the levels

given above is likely to be negligible. Accordingly,

these ranges may constitute practical threshold shear

stresses (or strains) marking the initiation of signifi­

cant pore pressure buildup In specimens of Ottawa 20-30

sand at a relative density ranging from about 45% to 65%.

A threshold shear strain on the order of 0.012% to 0.018%

agrees with the values reported by other investigators

(24,32,33,43). On the other hand, as discussed previ­

ously, a sharp increase in damping ratios, which reflect

the non-linear and inelastic properties of the sand, was

also observed at shear strain amplitudes slightly above

0.01% (see Fig. 5.9), reflecting the direct relationship

between pore pressure buildup and the development of plas­

tic deformations.

A few tests were performed to investigate further the

existence and magnitude of the threshold shear strain.

Figure 5.27 presents the results obtained from a test per­

formed on a solid cylinder specimen with an initial relative

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
355

V...
120 -

O ttaw a 2 0 - 3 0 sand
S o lid c y lin d e r sp ecim en ,
as
X
2 100 -

5
CO
X
_J
X
a
o
2
EC
<
LLJ
X
03

40-

Test D r.% ^ ’m, kPa


20 ”

E © 4 0 .8 22.C
( b 0.15-
3 © 4 2 .3 5 3 .4
©
O
K © 4 3 .6 9 9 .8
<
X 0 .10 --
LU
X
X
03
03
UJ
X 0 .0 5 -
X
UJ
X
X
o
10 10-3 10-2 10 _1

SH EA R S T R A IN , (% )

Figure 5.27 Results of Undrained Resonant Column Tests


on Isotropically Consolidated Specimen.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
356

density of about 41%. The specimen was first conso-

lidated isotropically to a mean confining stress, o , of


m
22 kPa. After consolidation, the specimen was subjected

to dynamic loading in the resonant column mode under

undrained conditions over the range of shear strains from


—4 -2
10 % to about 1.6 x 10 %. The procedure was repeated

after reconsolidating the specimen to mean confining

stress of 53.4 and 99.8 kPa, respectively. The results in

Figure 5.27 show the variation of shear modulus and pore


*

pressure ratio, u/a , as a function of shear strain ampli-


m
tude for the three stages of the test. It may be seen
-3
that for shear strain amplitudes below 10 % the pore

pressure buildup was essentially zero. Above this strain

level, the relative increase in pore pressure as measured


c by the ratio u/a
Cl
increases with increasing shear strain

amplitude. However, it is noted that a shear strain


—2
amplitude of 10 % the maximum pore pressure ratio is only
^ —

5% (fo-r a = 2 2 kPa), irrespective of the several hundred


m
cycles applied by the Hardin oscillator at each strain

level.

Although the tests results in Figure 5.27 are influ­

enced by the changes in density and the effects of staged

testing (strain history), it is apparent that the lower

the mean confining stress the higher the pore pressure

ratio, u/o , which is consistent with a trend towards


m
greater reduction in stiffness (Fig. 5.6) and higher

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
357

damping (Fig. 5.9) with decreasing confining stress at a

given shear strain amplitude. However, it should be noted

that absolute value of the pore water pressure actually

increases with increasing confining stress at the same

shear strain amplitude. This is attributed to the greater

tendency to contract under higher confining stresses.

Hence, it is clear that the generated pore water pressure

depends not only on the magnitude of the plastic deforma­

tions (which reflects in reduction of stiffness) but also

on the tendency towards densification (which increases

with a ). Although these counterbalancing effects may


m
reduce the effects of the mean confining stress, it cannot

be concluded that the relationship between pore pressure

ratio and shear strain is independent of stress level, as


c proposed elsewhere (24). The compressibility of sands

increases with a at a rate that is less than linear


m
(102,105). A similar trend can be inferred regarding the

magnitude of the pore pressure generated under undrained

conditions at the same shear strain (or stress ratio)


*
level. Consequently, the ratio u / v e r s u s shear strain

relationship cannot be independent of a , as illustrated

by the results in Figure 5.27.

Two other solid cylinder specimens with a relative

density of about 44% were tested in the resonant

column/torsional shear apparatus under undrained condi­

tions. Both specimens were consolidated isotropically to

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
a mean confining scress of about 101 kPa. One specimen (A

in Fig. 5.28) was first subjected to cyclic loading in the

resonant column mode over the range of shear strains from


-4 -2
10 % to 1.4 x 10 %. Thereafter, the specimen was sub­

jected to a series of cyclic loading stages at constant

shear stress amplitude in the torsional shear mode. The

shear stress amplitude was increased after each sequence

of five cycles. The other specimen (B in Fig. 5.28) was

directly subjected to a series of five cycle loading

stages in the torsional shear mode in order to evaluate

the effects on the pore pressure response of the pres­

training associated with performing resonant column tests


-4 -2
in the range of shear strains for 10 % to 10 %. The

average shear moduli and the pore pressure ratio values at

the end of each loading stage obtained from these two

specimens are plotted as a function of shear strain in

Figure 5.28. It is seen again that below a shear strain


_3
of about 10 % the pore pressure buildup is essentially

zero. Thereafter, the generated pore pressure increases

with increasing shear strain amplitude. However, it is


-2
noted that at a shear strain of 10 %, the pore pressure

ratio is only 3% for specimen A and 2% for specimen B.

The pore pressure ratio becomes higher at shear strains

above 10 ^%.

It is surprising to observe that the pore pressure

buildup curves are essentially the same for the two

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
359

140'
O tta w a 2 0 - 3 0 san d
S o lid c y lin d e r sp ecim en s 0 .6 2 6 0 .6 2 4
SHEAR MODULUS, G (MPa)

120 " Or (% ) 4 3 .6 4 4 .6

°" z (k P a ) 1 0 2 .5 102.2
o ^ r(k P a ) 1 0 0 .3 1 0 0 .4
100 "

80"

6 0 --

40"
PORE PRESSURE RATIO, “ /^r

R e s o n a n t co lu m n te s ts
Stress controlled torsional
S h e a r te s ts (5 c y c le s )
0.3- 19.3
T c y = 17.8kP a

0.2 - 16.2
14.6 17.4

1 1 .7 \
10l\
14.3

10.2
5.7

S H E A R STRAIN, (%)

Figure 5.28 Pore Pressure Ratio Versus Shear Strain


Relationship.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
specimens even Chough specimen A has been subjected to a

number of cycles several hundred times large than specimen

B. Hence, the rate of pore pressure buildup per cycle is

larger in the torsional shear mode than in the resonant

column test at the same strain amplitude. In undrained

cyclic tests performed by Drnevich (36) pore pressures

were also observed to be higher in torsional shear tests

than in resonant column tests when compared at the same

shear strain amplitude. Although the effects of frequency

are still unknown the author attributes this behavior par­

tially to the prestraining of the specimen due to resonant


-2
column tests at shear strains below 10 %. As discussed

earlier, it causes stiffening of the specimen, which

apparently reflects in an increased resistance to pore

pressure accumulation. Accordingly, it is thought that

undrained cyclic loading which involves small shear strain

amplitude also causes strain history effects even though

the pore pressure may increase continuously with number of

cycles. Strain history effects are clearly manifested by

an increase in shear modulus and a reduction in the rate

of pore pressure buildup, as discussed in more detail in a

later section. In this connection, it is interesting to

note that average shear modulus values for specimen A are

slightly larger than for specimen B at a shear strain


-2
amplitude of about 10 %, although the pore pressure

ratios are slightly larger for specimen A (Fig. 5.28).

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
361

Figure 5.29 shows the actual variation with number of

cycles in the shear modulus values for the two specimens

at different loading stages with approximately the same

shear stress amplitude in the torsional shear mode. It is

noted that for specimen A (prestrained) the shear modulus

values remain relatively constant whereas for specimen B

(no prior strain history) the shear moduli increase with

number of cycles. This behavior was also observed in

strain controlled drained tests (see Fig. 5.20).

Another solid cylinder was subjected to a large

number of cycles (N = 520) in the torsional shear mode

with a shear stress amplitude of about b.2 kPa. This

stress level is similar to the shear stress amplitude

applied to specimen B (Fig. 5.28) during the first stage

of cyclic loading. However, in this case, before applying

the cyclic loading the new specimen was consolidated under

a mean confining stress of 101 kPa and a static shear

stress of 11.2 kPa with rotation occurring in the clock­

wise direction. Hence, the initial state of stress in the

specimen was anisotropic. Results from this test are sum­

marized in Figure 5.30. First of all it is noted that

larger shear strains (Z 0.08%) and pore pressures (; 5%)

accumulated during the first five cycles of loading than

in the case of specimen B in Figure 5.28, reflecting the

effect of the initial static shear stress. Secondly, it

is evident that as cyclic loading progressed the residual

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
362

Specimen® ^ © & a (prestrained)


S p e c i m e n (B) • a ■ (no prior strain history)
70

65 - TCy = ± 10.2 kPa

60
10.6
M O D U L U S , G (MPa)

14.6
5 5 --

5 0 --
14.3

45 " 17.4
SHEAR

•?
4 0 -- —a -Q-

17.8

30
1 2 3 4 5

N U MB E R OF CYCLES

Figure 5.29 Effect of Number of Cycles at Constant


Shear Stress Amplitude on Shear Modulus
(Data from tests reported in Fig. 5.28).

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
363

Ottawa 20 - 30 sand ° " z - o " r - 100.1 kPa


0.08
Solid cylinder specimen Ts - - 11.2 kPa
e » 0.627 Tcy - = 6.2 kPa
0.04
Dr - 4 3 .2 %

60 120 180 240 300 360 420 480

i -0 .0 4

fc -0 .0 8 5th cycle

50th cycle 120th cycle


-0.16

- 0.20
iimii
-0.24

TIME. sec.
-0 .2 8 ’

100

c a
90 •

80
Confining pressure

Q.
70 •
UJ
cc
3 60
CO
CO
UJ
cc 50 ■
Q.
UJ 40 •
CL
30 ■
120th cycle
50th cycle
20 ■
5th cycle

60 120 180 240 300 360 420 480

TIME, sec.

Figure 5.30 Torsional Shear Tests with Small Cyclic Shear


Stress Amplitude ( t =11.2 k P a ) .

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
e.& <
CO
.30»
tig

o\\oe proP''
(fodo
rep’
f u ^ et
COPV^
0U^ e
;\ss\o°
e rff"!
dM M^ '
duce
ptepf0 '
365

shear strain in the clockwise direction increased,

o although at a decreasing rate. However, it is noted that

after the first few cycles, the peak to peak cyclic shear
_o
strain remained on the order of 0 .02% (10 % single ampli­

tude). Strain accumulation and pore pressure buildup vir­

tually stopped after 120 cycles (Figs. 5.30a and 5.30b).

Figure 5.30c displays the stress-strain curves correspond­

ing to approximately the first 40 cycles of loading.

These results confirm that if an initial static shear

stress is acting on the sand specimen the stress-strain

behavior in the first cycle and in the subsequent ones

cannot be represented by a single secant shear modulus, as

discussed in section 5.1.6. On the other hand, it is evi-

dent that after the first few cycles, the stress-strain

"loops" due to the superimposed cyclic shear stresses

correspond to unloading and reloading stages, regardless

of the initial shear stress. As the number of cycles

increases, the accumulation of residual strains gradually

decrease until the behavior becomes essentially elastic

with no additional increase in pore water pressure (Fig.

5.30b) .

Although based on a limited number of tests the

author has to recognize that the accumulation of signifi­

cant pore pressures seems to be unlikely at shear strain


_o
amplitudes smaller than 10 %. However, whether signifi­

cant pore pressures accumulate or not in the course of

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
366

cyclic loading depends not only on the level of applied

stress (or strain) but also on the initial state of stress

(see Fig. 5.30). Moreover, it is dependent on the strain

boundary conditions, as discussed throughout this disser­

tation. As the initial static shear stress increases

(anisotropic consolidation), it is very likely that cyclic

loading produces greater shear strains and pore pressures

during the initial stages of loading as compared with iso-

tropically specimens. However, the larger the initial

shear stress, the smaller the degree of stress reversal,

if any, in such a way that after a number of cycles at a

given stress amplitude the rates of pore pressure buildup

and of the consequent strain development become negligi­

c ble, as observed in Figure 5.29 provided that CTs+ T Cy)

remains small enough such that a strain softening behavior

cannot be initiated. On the other hand, the rate of pore

pressure buildup in isotropically consolidated specimens

may be smaller in the first cycles of loading as compared

to anisotropically consolidated specimens but as cyclic

loading proceeds the rate may remain steady or even

increase due to the continuous stress reversals, which

enhance the tendency to contract until liquefaction

develops. The relationship between initial state of

stress and pore pressure buildup is still imperfectly

known.

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
367

In the author^s point of view, the concept of a

threshold shear stress (or strain) should be defined not

as the level of shear stress (or strain) below which there

is no accumulation of pore pressure but as the level below

which the rate of pore pressure buildup becomes zero after

a number of stress (or strain) repetitions under certain

stress and strain conditions. This condition may be


_2
reached even if shear strains larger than 10 % develop

(Fig. 5.29). Some accumulation of pore pressure occurs at

virtually any shear strain level (see Figs. 5.27 and

5.28). It is claimed that the shear strain level at which

pore pressure buildup is initiated, that is, the so called

"threshold shear strain" is independent of fabric, stress

level and relative density (24). It is felt that the

threshold shear strain as defined may be another "small

strain" parameter of characteristics similar to G and


max
hence, relatively independent of some of these factors.

This is again attributed to the stepwise nature of the

stress-strain characteristics of granular soils (see dis­

cussion in section 5.1.6). However, this conclusion can­

not yet be extended to the behavior at larger shear

strains. Additional research in this area is needed in

order to clarify these conflicting views. In this connec­

tion, it becomes Important to be able to operate the tor­

sional shear test apparatus at low frequencies in the


—2
range of shear strains below 10 % in order to avoid the

undesirable prestraining effects produced by any type of

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
368

oscillator of high frequency. Only by means of these

tests, will it be possible to determine at which strain

level the stress-strain characteristics of sand and the

consequent pore pressure buildup start changing as a func­

tion of the number of cycles. This strain level may

correspond to a "true" threshold shear strain, which seems

to be smaller than 10” ^% (see Figs. 5.27 and 5*28).

On the other hand, in addition to the number of

cycles, the pore pressure response may be affected by the

frequency of loading itself. In this regard, the conclu­

sions drawn by previous investigators are conflicting

since experimental evidence has been presented showing

either that frequency does not affect the undrained

behavior of sands at least in the frequency range from

0.05 to 5 Hz (175,194) or that the dilatancy tendencies of

sands and hence the consequent pore pressures increase

significantly at higher rates of deformation (63,150).

The only information the author has on the effects of fre­

quency is based on a test performed on a replicate solid

cylinder spec men which was subjected to about the same

loading sequence as in the test performed on specimen B

reported in Figure 5.28. However, in this case, the rate

of straining was increased four times which resulted in

higher pore water pressures by about 25% at the end of the

first two stages of loading, and 70% at the end of the

third stage of five cycles. The specimen collapsed

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
369

o completely in the fourth stage. The results of this test

are presented in a later section (see Fig. 5.36). More

tests are needed before we can draw definitive conclusions

on the effects of frequency. In this connection, as men­

tioned earlier, the torsional shear apparatus is being

modified to expand its capabilities to the frequency range

more characteristic of earthquake loadings (0.25 to 0.50

H z ) . and to reduce the lower limit of shear strain ampli­

tude at which reversal of the direction of rotation can be

controlled automatically.

5.2.3 Monotonic and Cyclic Strength Relationship


under Torsional Loading

In order to evaluate the relationship between the

monotonic and cyclic undrained behaviors of Ottawa 20-30

sand under the stress and strain loading conditions of the

new torsional shear apparatus, a series of tests were per­

formed on replicate solid cylinder specimens with a rela­

tive density of about 43 to 46%. Most of the specimens

were first consolidated isotropically to a mean confining

stress of about 100 kPa and then they were subjected to

either monotonic or cyclic loading under undrained condi­

tions. In a few cases, a static shear stress (torsional

loading) was applied under drained conditions before the

stage of undrained shearing.

Figure 5.31 presents the results of a monotonic load-

.
— ing test with rotation occurring in the clockwise

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
370

Ottawa 2 0 - 3 0 sand
20 e - 0.627 a"z « o*'r«ioo.3 kPa
Dr - 4 3 .3 % Clockwise rotation

'S teady state'


deformation

30 60 90 120 150 180 210 240 270 300


TIME, sec.

Confining pressure
100

90 ■

80
as
a. 'S teady state'
70 ■
UJ deformation
cc 60 ■
=5
CO
CO
UJ 50 ■
cc
a.
UJ 40 •
cc
O
a. 30

20
Collapse ( / = 0.42 %)

30 60 90 120 150 180 210 240 270 300

TIME, sec.

Figure 5.31 Monotonic Torsional Shear Test on Isotropi-


cally Consolidated Specimen.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
371

STRESS, kPa
SHEAR 20

’Steady state'
deformation

y - 0.44 %

20 24
SHEAR STRAIN, %
50

40 ■

30 a ‘= 30 « £ ’ = 35*)
'Steady state'
STRESS, kPa

deformation
20 •

<£' mob = 13*


SHEAR

20 30 40 50 60 70 80 90 100

-1 0

-20

EFFECTIVE CONFINING STRESS.kPa


-3 0

Figure 5.31, continued

R eproduced w ith perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
372

direction. After a shear strain of about 0.44% (Fig

5.31c) the specimen exhibits a marked strain softening

behavior with the strength decreasing from 18 kPa to about

8 kPa whereas the pore pressure increases up to about 83%

of the confining stress (Fig. 5.31b). After the peak

shear stress (point A in Fig. 5.31d) is reached, the

stress path moves quickly toward the failure line reflect­

ing both the reduction in strength and the increase in

pore water pressure. Thereafter, the specimen exhibits a

short lived steady-state condition over the range of shear

strains from about 7.5% to 12.5% during which the shear

stress (Fig. 5.31a) and the pore water pressure (Fig.

5.31b) remain essentially constant. Further straining

c causes dilation, the pore pressure decreases and the

specimen regains strength (Fig. 5.31c). At this stage,

the stress path climbs back along the failure line (Fig.

5. 3Id) .

The author wants to call attention to the fact that

shortly before the specimen mobilizes the peak strength

there is an increase in the rate of pore pressure buildup

as reflected by the change in curvature of the pore pres­

sure versus time plot (Fig 5.31b), in spite of the small

scale used. As in the case of strain softening behavior

induced by cyclic loading (Fig. 5.25), the author postu­

lates that small shear strains may be enough for the meta­

stable structure of a loose sand to collapse, that is, to

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
373

experience a "sudden" tendency to rearrangement of grains

L or a reduction in the number of contact points between

neighboring grains. This structural instability* which

occurs close to the mobilization of the peak strength,

results in a significant increase in pore water pressure

and a consequent reduction in strength. To the author's

knowledge, this feature of the behavior of loose sands has

not been reported by any of the previous studies on the

undrained behavior of sands. All the test results avail­

able to the author, which basically were obtained from

aXxax uvuipicooxvu wt xaAxax lco uo i*rj } ouuw a nj

bolic" type of relationship between pore pressure and

axial strain, which implies that the rate of pore pressure

buildup is reported to be higher in the initial stage of

C loading and smaller close to the peak point of the

stress-strain curve. In the author's view, this pore

pressure response was not consistent with the initiation

of a softening process after the peak strength.

After the author improved his understanding of

apparatus and techniques for shear testing of soils he

could realize that the "hyperbolic" type of pore pressure

response was a consequence of the limitations of the

triaxial test. As discussed in Chapter 2, in the triaxial

test, pore pressure buildup is a consequence of both the

applied shear stresses and changes in the mean normal

stress. The pore pressure increase due to a change in

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
374

c mean normal stress Is equal to 1/3 of the applied devlator

stress, Ao, and hence becomes higher as the peak point Is

approached, overriding the changes in pore pressure due to

the dilative tendencies of soil. On the other hand, after

the peak Ao/3 decreases and the tendency to contract due

to shear stresses takes over the pore pressure response.

When the two components are combined the pore pressure

versus axial strain curve resembles a "hyperbolic" rela­

tionship, which is not representative of the actual rela­

tionship between pore pressure buildup and mobilized shear

strength, unless the loading conditions involve changes in

the mean confining stress. Consequently, in the author's

opinion, pore pressure curves from triaxial tests should

c: be corrected to account for changes in mean confining

stress. When this is done, it becomes also evident that

the rate of pore pressure buildup due to the applied shear

stresses on loose sand specimens is higher close to the

peak strength and not at the initial stage of loading. The

sharp change in the rate of pore pressure buildup result­

ing from the "collapse" of the structure of loose sands

becomes more evident under cyclic loading at shear stress

amplitudes smaller than the peak strength, as illustrated

conclusively in the succeeding sections.

Figure 5.32 presents the results of another monotonic

loading test performed on a solid cylinder specimen, which

after consolidation was subjected to a static shear stress

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
375

c Ottawa 2 0 -3 0 sand
20 ■ e - 0.629 c ’z =o"'r * 100.2 kPa
Dr - 42.5 % Clockwise rotation

ffl
a
.sc Starting point of
CO undrained shear
CO
Ui
CE
K
CO
CE
<
ui
X
CO

' Steady state


deformation

30 60 90 120 150 180 210 240 270 300

TIME, sec.

Confining pressure
100
r
80
'Steady state*
tu deformation
cc
=3 60
CO
CO
UJ
cc
CL
£ 40
O
a.

30 60 90 120 150 180 210 240 270 300

TIME, sec.

Figure 5.32 Monotonic Torsional Shear Test on Specimen


with an Initial Static Shear Stress.

R eproduced w ith perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
376

20

Starting point of
undrained shear
STRESS, kPa
SHEAR

Steady state*
deformation

20 24
SHEAR STRAIN, %

50

40 •

30 •
* S te ad y-S tate *
STRESS, kPa

deformation
20 ■
SHEAR

20 30 40 50 60 70 80 90 100

-10

-20

EFFECTIVE CONFINING STRESS,kPa


-30

Figure 5.32, continued

R eproduced w ith perm ission o f the copyright owner. F urth er reproduction prohibited w itho ut perm ission.
377

of about 13.8 kPa under drained conditions, with rotation

occurring again in the clockwise direction. This shear

stress level was chosen to be larger than the strength of

the previous specimen at the short lived steady-state con­

dition. Undrained shearing was continued in the clockwise

direction. It is noted that strain softening behavior was

initiated after applying a shear stress increment of just

4 kPa. The pore pressure buildup during the first second

of loading (At Z 3 kPa) was negligible. Thereafter, the

specimen collapsed and pore pressure increased up to about

8 8 % of the confining pressure. At this stage the specimen

exhibited also a short lived steady— state condition fol­

lowed by stiffening due to dilation, in a similar fashion

c as the previous specimen.

It is interesting to note that the initial shear

stress essentially had no effect on the peak stress, the

pore pressure and the "steady-state" strength as compared

to the isotropically consolidated specimen. It is also

Interesting to note that in either case, the trace of the

stress— strain curve after the peak point shows minute

fluctuations reflecting the stepwise (frictional) nature

of the stress-strain characteristic of sands. The same

behavior was also observed whenever strain softening was

caused by cyclic loading. The results in Figure 5.32 show

that if the initial static shear stress is larger than the

steady-state strength small stress increments may be

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
378

enough to cause the initiation of a flow failure, which


L under stress controlled conditions such as close to slop­

ing ground is of catastrophic consequences.

Once it was established that specimens of Ottawa 20-

30 sand exhibited strain softening behavior at a relative


*
density of about 44% (o *= 100 kPa), replicate specimens
m
were subjected to cyclic loading at different shear stress

amplitudes larger than the steady-state strength of about

8 kPa. The main issue was to determine the conditions at

which strain softening behavior, if any, developed under

cyclic loading and their relationship with strain soften­

ing under monotonic loading. As discussed in Chapter 3,

there are conflicting opinions on this aspect of the

c undrained behavior of loose sands. Accordingly, it was

thought that the collapse could occur:

a. at the same stress ratio ( <j> , = 13 ) as in the moiuo-


mob
tonic test as concluded by Vaid et a l . (186) based on

triaxial tests on Ottawa and tailings sands; or

b. at a higher unknown value of $ as suggested by the


mob
results for banding sand presented in Figure 3.39

(data from Ref. 14), which the author attributed to

strain history effects; or

c. when the stress path reaches a line close to the

steady-state envelope as reported by Dobry et a l .

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
379

(34) based on strain controlled torsional shear tests


a on anisotropically consolidated specimens.

Typical results from one of the present cyclic tests

at a constant shear stress amplitude of about 15.2 kPa are

presented in Figure 5.33. It became apparent pretty soon

that none of the predictions was right. As illustrated in

Figure 5.33c , strain softening behavior was initiated at a

stress ratio ($ . ) corresponding to a point on the mono­


mob
tonic stress path (from Fig. 5.31) at the same shear

stress amplitude. In other words, collapse occurred at

the instant the stress state during cyclic loading reached

the effective stress path from the monotonic test at the

same Dr. After the collapse the two stress paths are

essentially the same. Up to the point of collapse the

cumulative shear strain was on the order of 0.38% (Fig.

5.33a) whereas the pore water pressure increased progres­

sively (although at a decreasing rate) up to about 50% of

the confining pressure (Fig. 5.33b). The initiation of

strain softening behavior is clearly characterized by a

sharp increase in pore pressure (Fig. 5.33b) and the

development of large shear strains (Fig. 5.33a). The

stress path moves quickly towards the failure line (Fig.

5.33c), reflecting the reduction in strength due to the

increase in pore water pressure and the reduction in the

number of contact points between adjacent grains (col­

lapse) .

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
Reproduced with permission of the copyright owner. Further reproduction

Ottawa 2 0 -3 0 sand
°"z - <rr = 100.I kPa

z
<
cc
I—
w 100 200 300 400 500 600 700 800 900 1000
oc
<
Ui Collapse
I
CO -2
prohibited without perm ission.

85th cycle
-4

-6
TIME, sec.

Figure 5.33 Cyclic Torsional Shear Test with Shear Stress Amplitude

380
Larger than Steady-State Strength.
381

continued
5.33,
Figure

Bd>l ‘aunssaad 3 dOd

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
382

continued
5. 3 3 ,
Figure

R eproduced w ith perm ission o f the copyright owner. F urth er reproduction prohibited w itho ut perm ission.
383

c Tests performed at different shear stress amplitude

yielded similar results, which are summarized in Figures

5.34 and 5.35. The numbers within parentheses in Figure

5.34 correspond to the cumulative shear strain up to the

point of collapse. It is noted that the locus of stress

states triggering the collapse during cyclic loading

defines an envelope which essentially corresponds to the

stress path under monotonic loading (Fig. 5.34). Accord­

ingly, it is concluded that a given relative density the

monotonic stress path constitutes a "collapse" surface

which determines the initiation of strain softening

behavior under cyclic loading, provided the cyclic shear

stress amplitude is larger than the steady-state strength.

Consequently, the angle of shear resistance mobilized at


(
the triggering state is not a constant value but a vari­

able one depending upon the cyclic shear stress amplitude.

It is also interesting to note that the shear strain accu­

mulated up to the collapse is essentially the same (I

0.3%), irrespective of the amplitude of the cyclic shear

stress.

It is surprising to the author that cyclic loading

seems to affect neither the stress conditions nor the

cumulative shear strain at the initiation of strain

softening behavior. It was thought that a specimen with a

more "stable" structure resulting from the prior strain

cycles could mobilize a larger angle of shear resistance

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
Reproduced with permission

30
of the copyright owner. Further reproduction

Ottawa 20 - 30 sand ( ) Cumulative shear strain


Dr - 43 - 45 %
a ' z - o - r - ioo kPa
a.

20 30 40 50 60 70 80 90 too

<f)’mob - 18.4'
Collapse surface
-10
prohibited without perm ission.

,? /(0 .2 9 )
(0.29) (0.29)
~ — ft-.
-20

<f>'mob - 13
-3 0

EFFECTIVE CONFINING STRESS,kPa

Figure 5.34 S t r e s s C o n d i t i o n s fit I n i t i a t i o n o f S t r a i n S o f t e n i n g B e h a v i o r

384
under C y c l i c Loadi ng.
(.0
"O
3
Q.
C
o
CD
Q.
with permission of the copyright owner. Further reproduction

(0
CL
J*
18-
Ui
Q
D
H
_i
Q. 16 -
5
<
<0
Ui
Ui
a:
t~
Ui 14 -
tr
<
Ui ?
I
Ui Ottawa 2 0 -3 0 sand
prohibited without perm ission.

□ 12 + Dr 4 3 -4 6 %
O 0"z =cr'r = lOOkPa o Collapse
>-
o
End of test

10 20 50 100 200 500

NUMBER OF CYCLES
Figure 5.35 Number o f C y c l e s t o I n i t i a t i o n o f S t r a i n S o f t e n i n g B e h a v i o r .
386

before collapsing. In other words, it was thought that


L cyclic loading could push outwards the corresponding col­

lapse surface. This seems not to be the case. Neverthe­

less, it should be noted that cyclic loading does affect

the rate of pore pressure buildup in the subsequent load­

ing stages. This is clearly observed in Figures 5.33b and

5.33c. The rate of pore pressure buildup is higher in the

first few cycles of loading. As cyclic loading progresses

the rate of pore pressure buildup decreases significantly

as reflected conclusively by the gradual reduction of the

spacing between subsequent cycles in the stress path plot.

This is attributed to a stiffening effect associated with

the elimination of local instabilities at the contact

points and the collapse of looser particle groups. How­


C ever, as the stress path approaches the yield surface, the

rate of pore pressure buildup increases again up to the

point where the collapse and consequent sharp increase in

pore pressure occur. This pore pressure response confirms

the behavior illustrated schematically in Figure 3.6.

Since the collapse surface is relatively flat after the

peak point (A in Figure 5.34) the number of cycles to the

initiation of the strain softening behavior is very sensi­

tive to small variations in the cyclic shear stress ampli­

tude, as illustrated in Figure 5.35. As discussed ear­

lier, this type of plot represents the resistance of the

specimen to develop a "flow deformation" type of failure.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
387

It is considered that the behavioral trends observed

in Figures 5.31 and 5.35 should be valid for other stress

and density conditions under which Ottawa 20-30 exhibits

strain softening behavior. The relationship between the

collapse surfaces defined by means of torsional shear

tests (Fig. 5.34) and cyclic triaxial tests is yet to be

established. The effects of initial static shear stresses

on the position of the collapse surface are also imper­

fectly known at this point. Only one specimen was

cyclicly loaded after being subjected to an initial static

shear stress of about 11.2 kPa. The cyclic shear stress

amplitude was 6.2 k P a . However, at this stress amplitude

pore pressure buildup completely stopped after 120 cycles

and the stress path never approached the collapse surface,

irrespective of the facts that more than 500 cycles were

applied to the specimen (see data in Fig. 5.30) and the

total shear stress (xg + TCy) was periodically larger than

the steady-state strength.

The collapse surface presented in Figures 5.33c and

5.34 has characteristics very similar to those of the

state boundary surface used by Symes et a l . (169) to

account for the effects of principal stress rotation on

the undrained behavior of sands. Symes et a l . (169) per­

formed a test on a hollow cylinder specimen in which the

orientations of the principal stresses, and , were

cyclicly rotated between ± 22.5° from their initial vertical

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
338

and horizontal directions, respectively, while keeping

c the deviator stress, ( - <*3 )* constant. Cyclic rotation

of the principal stresses caused continuous accumulation

of pore pressure even though the applied shear stresses

remained constant during the test. It was observed that

the specimen became "unstable", that is, it developed

larger pore pressures and consequently larger shear

strains, once the pore pressure accumulation caused the

stress state to reach the post peak part of the effective

stress path corresponding to a monotonic test in which the

orientation of the major principal stress was 22.5° from

the vertical.

Although Symes et al. (169) reported a single test,

the results in Figure 5.34 are full evidence of the rela­


c
tionship between the stress path from monotonic tests and

the behavior under cyclic loading. It is interesting to

note that Symes et al. (169) concluded that the shear

strains appear to more directly related to the position of

the effective stress state with respect to the failure

envelope rather than with respect to the state boundary

surface. However, the data in Figure 5.34 suggest a clear

relationship between the effective stress state and shear

strains at the collapse surface, which mark a rapid change

in the stress-strains relation. Of course, the absolute

magnitude of shear strain increases as the failure surface

is approached. Figure 5.33c also illustrates the

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
389

convenience of using stress path plots in order to define

behavioral trends. Unfortunately, effective stress paths

from cyclic triaxial tests are not available for com­

parison.

It should be noted that in torsional shear tests on

isotropically consolidated specimens the shear stress

increases gradually and the axis of the major principal

stress changes its orientation from -45° to +45° from the

vertical (Fig. 2.5) in the course of reversing the shear

stress from clockwise to counterclockwise. Hence, the

collapse surface in Figure 5.34 corresponds to a specific

orientation of the principal stresses (±45° from the vert­

ical). As the orientation of the major principal stress

changes it is likely that the collapse surface or state

boundary surface also changes as discussed by Symes et a l .

(169). From this point of view, triaxial tests and tor­

sional shear tests will yield different collapse surfaces.

For the same reasons, the cyclic undrained strength cannot

be independent of the initial state of stress since it

influences the magnitude of the rotation of the planes of

principal stress (Fig. 2.6) and the orientation of the

principal axes of anisotropy. The pore pressure buildup

depends not only on the magnitude of the applied shear

stresses but also on the degree of rotation of the planes

of principal stress.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
A few specimens were subjected to cyclic loading in

the torsional shear apparatus with non— uniform shear

stress amplitudes. Results from one of these tests are

presented in Figure 5.36. The specimen was subjected to a

series of loading stages at constant shear stress ampli­

tude, which was increased after each sequence of five

cycles, as shown in Figure 5.36a. Strain softening

behavior was triggered also in accord with the collapse

surface presented in Figure 5.34. It is interesting to

note that in each one of the loading stages prior to the

collapse, the rate of pore pressure buildup increases with

shear stress amplitude but decreases with number of cycles

(Fig. 5.36c and 5.36d), reflecting once more strain his­

tory effects. After a shear strain of about 4% following

the collapse at point A (Fig. 5.36c) the direction of

shearing was suddenly reversed at point B (Fig. 5.36b),

stopping the unidirectional deformations of the specimen.

Upon reversal of the direction of shearing the specimen

experience a second collapse which reflects in a further

increase in the rate of pore pressure buildup (Fig.

3.36c). It Is noted that at point B the stress path was

already close to the failure line and hence, near the

phase transformation line. After point B, the stress path

moves towards the failure line and as the shear stress

reverses sign (Fig. 5.36a) the specimen exhibits a reduced

strength. However, further straining causes dilation, the

pore pressure decreases and the stress path moves back

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
391

20 Ottawa 2 0 - 3 0 sand
e - 0.62 Dr - 46 % o"z « o " r 100.2 kPa

12 •

eo
O
J.£
CO
CO
LU
cc

c 120 240 300 360


<
HI -4
X
CO

-8

-1 2

-1 6

-20
5r

4 •

2
<
CC
i- 60 120 180 240 360
CO
cc
<
HI
X
CO

-2 ■

-3

-4

-5

Figure 5.36 Cyclic Torsional Shear Test with Non-Uniform


Shear Stress Amplitude.
f

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
392

C onfining pressure
C 100

Collapse on
reversal —
80
«
Q.
UJ
CC
3 60
CO
CO
UJ
cc
Q.
UJ Collapse on
cc 40
O loading —
a.

120 180 240 360


TIME, sec.
40

c 30
d)

a 20
CL
JC
CO
CO 10
UJ
cc
V-
co
cc 10 2 40 50 60 70 80 100
<
UJ
X
CO
-1 0 •

Forced
reversal
-2 0 -

-30

-40L EFFECTIVE CONFINING STRESS.kPa

Figure 5.36, conCinued.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
393

along Che failure line, reflecting a gradual regain In


o strength. Figure 5.36c clearly illustrates the main

stages in the process of pore water pressure generation

under cyclic loading; namely

a. the pore pressure buildup during the first cycles of

loading due to the initial tendency to contract,

which proceeds at a decreasing rate as the fabric

becomes more stable due to the prior cycles,

b. a sharp increase in pore pressure upon the collapse

of the structure on loading, which also proceeds at a

decreasing rate as the specimen approaches the criti­

cal state,

r
c. an additional increase in pore pressure as the struc­

ture collapses again upon reversal in the direction

of loading, and

d. reduction in pore pressure due to dilation after

large strains under very small confining stresses.

These results further confirm the behavior illus­

trated schematically in Figure 3.6.

In the previous tests, reversal of the direction of

shearing was forced after a shear strain of 4% following

the initiation of strain softening behavior (Fig. 5.36).

It was of great Interest to study the effect of reversal

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
394

in the direction of shearing at smaller strains after the


C — ^)' collapse. These effects are illustrated by the results

presented in Figure 5.37. A solid cylinder specimen was

cyclicly loaded until collapse (point A in Fig. 5.37).

After a shear strain of about 1% a reversal in the direc­

tion of shearing was forced at point B and the specimen

was subjected to half a cycle in the opposite direction.

It is seen that as the stress path moves away from the

state boundary surface the pore pressure buildup slows

down (between points B and C in Fig. 5.37b). As the

stress path approaches again the collapse surface, the

specimen experiences a second collapse on loading (point C

in Fig. 5.37d), and the rate of pore pressure buildup

c increases again (Fig. 5.37b).

experiences a second stage of strain softening as


Thereafter the specimen

reflected by the stress-strain curve in Figure 5.37c.

After a unidirectional deformation of about 10.5% a second

reversal of the direction of rotation was compelled to

occur and the specimen was brought to a state of liquefac­

tion (Fig. 5.37b). The specimen was rotated in the

counter-clockwise direction until reducing the residual

strain to about 3.8%. At this stage the specimen was

reconsolidated under the initial state of stress and sub­

jected to a second stage of undrained cyclic loading. The

results of this second loading are presented in the next

se ct io n.

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
395

o Ottawa 2 0 - 3 0 sand
e - 0 .62 3 Dr - 4 5 % o -’z - ° " r - 100 kPa

60 120 180 240 300 360 420 480

z
< -2
cc
t-
CO

cc -4
<
UI
X
CO
-6

-8

-10 ■
TIME. sec.

-12

Confining pressure
100

c 80
Collapse on
reversal —
Liquefaction

Second collapse
on loading
cc
Z) 60
CO
CO
UJ
CE
First collapse
0. on loading —
UJ
CE 40
o
CL

20

60 120 180 240 300 360 420 480


TIME, sec.

Figure 5.37 Cyclic Torsional Shear Test with Reversal


Shortly after Initiation of Strain
Softening Behavior.
Is

R eproduced w ith perm ission o f the copyright owner. F urth er reproduction prohibited w itho ut perm ission.
396

20

STRESS, kPa

-1 2 -11
SHEAR

-4

-8

-12

-16
SHEAR STRAIN, %
-20
40

30 ■

20
STRESS, kPa

20 30 40 [50 60 311 90 100


SHEAR

-1 0

Forced
-20 reversal Forced
reversal
-30

EFFECTIVE CONFINING STRESS,kPa


-40

Figure 5.37, continued.

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission
397

The results in Figure 5.37 and from other replicate


o tests show that after the triggering of strain softening

behavior high pore pressures continuously develop if the

stress path remains on the collapse surface. However, if

the loading conditions cause the stress path to move away

from the collapse surface the process of significant

increase in the pore pressure slows down, provided the

shear strains after the collapse are not too large and the

stress path is far away from the phase transformation

line. A sudden interruption of the softening process is

clearly evident in Figure 5.37c (point B ) . However, it

should be noted that Figure 5.37 reflects only the effects

of the cyclic shear stresses. It static shear stresses

/ exist such as close to a slope, the softening process will

be interrupted only when the applied shear stresses become

compatible with the reduced strength, as discussed in

Chapter 3. This may imply very large deformations.

In summary, based on the results previously discussed

it can be concluded that the undrained behavior of sands

susceptible to strain softening behavior under both mono­

tonic and cyclic loading is characterized by the following

f ea t u r e s.

a. A "collapse" surface or "state boundary" surface,

which depends on the orientation of the principal

stresses, defines a region outside which no stress

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
398

state Is possible, although it is not a failure cri­

terion. The effective stress path approaches the

failure line along the state boundary surface. At a

constant void ratio and for a given orientation of

the major principal stress, the state boundary sur­

face seems to be defined by the effective stress path

from a monotonic test. In addition to the contrac­

tive region, the stress path may show a dilative

region.

b. The state boundary surface also provides the boundary

between "stable" (small strains) and "unstable"

(strain softening) states. If the stress state of a

specimen is on the pre-peak region of the state boun­

C dary surface it will be in a "stable" state.

ever, If the stress state of a specimen is on the


How­

post-peak region of the state boundary surface it

will show an "unstable" behavior. Accordingly, when

a specimen is subjected to cyclic loading at a shear

stress amplitude smaller than that corresponding to

the peak point of the collapse surface, strain

softening behavior will be initiated when the effec­

tive stress state reaches the post-peak region of the

collapse or state boundary surface.

c. The collapse surface seems to constitute a kind of

strain contour corresponding to the cumulative shear

strain causing the initiation of strain softening

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
399

behavior, irrespective of the cyclic shear stress

amplitude. After the collapse, the magnitude of

shear strain increases as the effective stress path

approaches the failure surface along the state boun­

dary surface.

d. Cyclic loading conditions may be such that the stress

state of the specimen is compelled to remain on the

state boundary surface or to move away from it. The

pore pressures that develop when the stress state of

the specimen remains on the state boundary surface

are substantially larger than when the stress path

moves away from such a collapse surface provided the

stress state is far from the line of phase transfor­

mation. Unloading in the region of the state boun­

dary surface beyond the phase transformation line

(dilative region) also produces significant pore

pressure buildup which may lead to a state of

l iq ue fac ti on.

e. The effects of initial shear stresses on the collapse

surface are still imperfectly known. Also, the col­

lapse surface in stress space is probably strain rate

d e p e nd en t.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
400

r- 5.2.4 Effects of Prior Cyclic Stress/Strain History

O
In the author's view, cyclic stress/strain history

effects are of concern regarding the following aspects of

the undrained behavior of sands:

1. The effects of prior stress (or strain) cycles on the

pore pressure and strain development during the sub­

sequent cycles within the same loading sequence,

2. The undrained strength-deformation characteristics of

the sand under monotonic loading either immediately

after the end of the stage of cyclic loading or after

reconsolidation, and

^ 3. The effects of prior strain history on the resistance

to liquefaction under subsequent cyclic loading after

dissipation of the induced pore pressure has taken

place.

1. Strain History Effects within a Loading Sequence.

As mentioned earlier, cyclic strain history effects are

reflected by a reduction in the rate of pore pressure

buildup with increasing number of cycles, provided that

neither the state boundary surface nor the phase transfor­

mation line are reached in the course of cyclic loading.

At either of the latter state conditions, the rate of pore

pressure buildup increases again leading to either strain

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
A 01

softening behavior or liquefaction. The reduction in the

rate of pore pressure buildup before collapse in the

course of cyclic loading was conclusively illustrated in

Figures 5.23,5.25,5.30.5.33,5.36 and 5.37. This was attri­

buted to a kind of "stiffening" of the sand structure,

which is also reflected in a reduction in the magnitude of

shear strain increments that develop during the initial

stages of subsequent loading cycles under stress con­

trolled conditions, irrespective of the fact that the pore

pressure continuously increases. This behavior is illus­

trated in Figure 5.38, which shows the variation in secant

shear modulus with number of cycles from some of the tests

performed at different shear stress amplitudes on solid

cylinder specimens with a Dr in the range from 44 to 46%.


c As pointed out earlier, the stress-strain curves during

the first cycles of undrained loading resemble simple

closed loops. Accordingly, simplification of the behavior

in terms of an equivalent secant shear modulus is con­

sidered appropriate. The results in Figure 5.38 clearly

indicate that the variation in shear modulus with number

of cycles under undrained cyclic loading is the result of

two counterbalancing effects: an increase in G due to the

"prestraining" caused by the prior cycles and a degrada­

tion in G associated with the progressive increase in pore

water pressure. Accordingly, it is seen that the shear

modulus first increases with number of cycles but after a

number of cycles it starts decreasing again. Hence,

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
Reproduced with permission of the copyright owner. Further reproduction

Ottawa 2 0 - 3 0 sand
45 Solid cylinder specimens
Dr « 4 4 - 4 6 %
i~cy = 13.0 kPa
40 - Peak

(0 1 o
OL 0 ,i>-o-0 -o-o^-ooO"Oo-o-o-o-o-A oo-0-- 0 0 o “ °‘£>"0'0'00-Cl-0--0 0-(j0-0-00.0-0-<i N = 165
5 Q-O'o'o
35 15.2
0 I.

CO
/ » s 4<T \
D 0 a'
_l / A'
ZD 30 1 'L
Q I * * 16.5
0
5 0/n/Q d'od /
DC
25
/ n/
< " \ N = 80
prohibited without perm ission.

111 K o
1 □ \ A ►
CO \
□ (Imminent collapse)
Collapse
20

t
10 15 20 25 30 35 40 45 50 55
NUMBER OF CYCLES, N

Figure 5.38 E f f e c t of Number o f C y c l e s on S e c a n t Shear Modulus under

402
Und r a l n e d C y c l i c L o a d i n g .
403

during the first few cycles the prestraining effects com­

pletely override the degradating effects of the pore pres­

sure buildup. It is noted that the number of cycles

corresponding to the peak value of shear modulus decreases

significantly as the shear stress amplitude increases.

Nevertheless, it should be noted that the shear modulus

becomes smaller than the value corresponding to the first

cycle only when the specimen is about to collapse. There­

fore, the cumulative effects of prior strain history on

the shear modulus are essentially dominant over most part

of the loading sequence prior to the initiation of strain

softening behavior.

2. Monotonic Undrained Strength after Cyclic Load­

c ing. At the state boundary surface there is a rapid

change in the shear stress-shear strain relation, which

reflects a drastic change in the particle arrangement or

structure of the sand. The magnitude of this change

depends on the level of shear strain accumulated after the

collapse. As the shear strain increases the arrangement

of particles approaches the "critical” state. As dis­

cussed in Chapter 3, at the critical state, a soil may

exhibit "continuous" deformations at a constant shear

stress, depending on the initial density and stress condi­

tions. If, after the collapse, the sand is subjected only

to monotonic loading (see Fig. 5.25), the subsequent

undrained behavior is controlled by the steady-state

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
404

strength, which apparently is not affected by the prior

strain history. This is based on the fact that the stress

paths from cyclic tests and mcnotonic tests defined the

same collapse surface (see Fig. 5.34). The same conclu­

sion has been drawn by other investigators (34,132,186).

On the other hand, if after the initiation of strain

softening behavior, a sudden reversal in the direction of

shearing occurs, it may lead to a condition of initial

liquefaction, provided the effective stress path has

already approached the phase transformation line (see

Figs. 3.36 and 3.37). Once this condition develops, it is

of interest to establish the strength-deformation charac­

teristics of the sand under subsequent monotonic loading

in either direction. This aspect of the behavior is

illustrated in Figure 5.39.

A solid cylinder specimen was cyclicly loaded until

collapse was initiated (point A in Fig. 5.39). A uni­

directional shear strain in the clockwise direction of

about 10.5% was applied until the pore pressure stabilized

(point B ) . A reversal in the direction of shearing was

then forced at point B, bringing the specimen to a state

of initial liquefaction. Thereafter the specimen was

rotated monotonically in the counter-clockwise direction.

It is noted that after initial liquefaction the specimen

undergoes large deformations without mobilizing appreci­

able strength (Fig. 5.39c). However, further straining

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
405

28
e 24 •

20 ■
Ottawa 2 0 - 3 0 sand
e - 0.625 Dr - 4 4 %
o-'z - o - r -1 0 0 .2 kPa
£
z
<
cc

oc
<
HI 60 4 20 480 540 600 660 720
z
CO

-4

-8
TIME, sec.
-12

Confining pressure
100

c 80 Liquefaction

60 •
Collapse on reversal
o- 40
2C .

Ill
Sj 20 Collapse on
C/3 loading
C/3 -
Ui 0
GC
0.
ui -2 0
cr
o
“■ - 4 0

-6 0

-8 0
TIME, sec.
-100

Figure 5.39 Monotonic Torsional Shear Test after Develop­


ment of Initial Liquefaction Due to Cyclic
L o ad in g.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
406

•120

o
100

80

60

40

20

-12 -8 -4
20 24 28
-7.2
"20 S H E A R STRAIN, %

120r

c 100 •

80 ■
cs
CL.

CO 60
CO
LL1
CE ' = 30.5 (<£' = 36*)

a. 40
<
m
x
CO
20

20 40 100 120 140 160 180

-20 Forced
reversal
EFFECTIVE CONFINING STRESS.kPa

Figure 5.39, continued.

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
407

allows dilation to d e v e l o p , the pore water pressure


o decreases (Fig. 5.39b) and the specimen starts mobilizing

strength at an increasing rate with the stress path climb-*

ing along the failure line (point C ) . It is interesting

to note that the mobilized strength increases continuously

up to 100 kPa and dilation is still evident after a resi­

dual shear strain in the counter-clockwise direction of

about 21%. At this strain level the test was terminated.

The increase in strength due to dilation will continue

until cavitation develops. In this connection it should

be noted that if the magnitude of the back pressure

applied to a laboratory specimen is significantly higher

than the anticipated initial pore pressure in the field,

c then the magnitude of the induced negative pore pressure

before cavitation occurs will cause excessively high meas­

ured strengths which most likely will not occur in the

field (140).

Some researchers (18,132) claim that after 100% pore

pressure buildup (initial liquefaction) under cyclic load­

ing the undrained steady-state strength from monotonic

tests still remains. The results in Fig. 5.39 show that

this is not the case. Upon reversal of the shear stress a

shear strain of about 15% developed before the specimen

again mobilized a shear stress equal to the "steady-state"

strength in the opposite direction (point B'). In gen­

eral, a certain amount of shear strain under essentially

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
408

zero shear strength will be required to mobilize any


L. appreciable shear resistance after initial liquefaction.

This shear strain is clearly a function of the relative

density of the sand as discussed in Chapter 3. However,

it should be noted that after large strains the specimen

in Figure 5.39 mobilizes a strength considerably higher

than the "steady-state” strength, without showing strain

softening behavior. The author attributes this behavior

in the first place to the fact that after the collapse the

"brittleness” of the sand skeleton on loading disappears.

In other words, the residual particle arrangement is less

metastable than the original one. On the other hand, at

initial liquefaction the state point is located in the

c hardening or dilative region of the state diagram (see

Figs. 3.36 and 3.40). Consequently, upon monotonic load­

ing the state path will move towards the right probably

approaching the S line, as discussed in Chapter 3. More­

over, as indicated earlier, at a relative density of 44%

and under a of 100 kPa, Ottawa 20-30 sand only exhibits


m
limited strain softening behavior (see Fig. 5.31), that

is, only a limited collapse accompanied by a short-lived

flow condition takes place under monotonic loading.

Therefore, the state of stress moves first towards the

left in the state diagram but once dilation is initiated

it moves back also towards the S line, in a similar

fashion as shown schematically in Figure 3.20. Unfor­

tunately, the S and F lines for Ottawa 20-30 sand are not

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
409

s yet available. The effects of prior strain history on the

cyclic undrained strength after reconsolidation are dis­

cussed next.

3. Cyclic Undrained Strength after Reconsolidation.

Based on the results discussed in the previous sections,

the author considers that if cyclic loading stops before

the effective stress path reaches the state boundary sur­

face, the prestraining effect will reflect in a higher

resistance to pore pressure buildup under similar

undrained loading conditions after reconsolidation. Of

course, for a given shear stress amplitude, the prestrain­

ing effect increases as the number of cycles (or residual

shear strain) increases. The increase in cyclic undrained

c strength will be due not only to the prestraining effect

but also to the increase in density that occurs during the

process of dissipation of pore water pressure. However,

it should be noted that prior strain history apparently .

does not influence the location of the state boundary sur­

face at a given Dr (see Fig. 5.34). More tests are needed

to confirm whether this conclusion is valid or not even

after reconsolidation.

On the other hand, if cyclic loading stops after the

effective stress path has reached the state boundary sur­

face, the fabric (and hence the undrained strength)

remaining after reconsolidation depends on the magnitude

of shear strains that accumulate between the point of

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
410

collapse and the end of the cyclic loading. The larger

the shear strain, the closer the fabric after reconsolida­

tion is expected to be to the fabric corresponding to the

critical state and consequently the higher the "deforna-

bility" or compressibility of the sand skeleton at the

same level of applied shear stress. In other words,

"yielding" of the sand structure, that is, the development

of large pore pressures and consequently large shear

strains, is triggered by smaller shear stress amplitudes.

This is attributed to a lesser degree of "interlocking"

between neighboring grains within a particle arrangement

which is close to the "critical" fabric. This behavior is

illustrated by the results presented in Figure 5.40, which

c were obtained from a second stage of undrained cyclic

loading on a solid cylinder specimen that experienced

large shear strains (2 10.5%) during the first loading.

The results of the first undrained cyclic loading were

presented in Figure 5.37.

During reconsolidation after initial liquefaction the

relative density of the specimen increased from 45% to

about 52.2%. It is noted that the residual shear strain

after reconsolidation was of about 3.8% in the clockwise

direction. The cyclic shear stress amplitude during the

first and second loadings was essentially the same (Z 15

kPa). It is interesting to note that a shear strain

increment of about 5.5% (also in the clockwise direction)

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
411

Ottawa 2 0 -3 0 sand
e - 0.605 a-z «o"r«ioo.1 kPa
Dr - 52.2 % First loading in Fig. 5.37

60 120 180 240 300 360

After reconsolidation

Z -4
<
c
Hc
-
CO
oe -8 •
<
LU
X
CO
-1 2 •

-1 6 -
TIME, sec.

-20
Confining pressure
100

Initial liquefaction
80
CB
a
til
•3 60
CO
CO
UJ
cc
a.
LU 40
a:
o
a.

20

60 120 180 240 300 360


TIME. sec.

Figure 5.40 Cyclic Torsional Shear Tests on Reconsolidated


Specimen with Residual Shear Strain (— 3.8%).

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
412

20

STRESS, kPa

-18
SHEAR

-4

-8

-1 2

-16

S H E A R STRAIN, %
-20

30-

20 ■
STRESS, kPa

30 40 50 60 70 80 90
SHEAR

Phase transformation line


-20-
v£Z'= 32.2 «#= 39*)

EFFECTIVE CONFINING STRESS.kPa

Figure 5.40, continued.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
413

developed in the first loading stage of the first cycle


c (Figs. 5.40a and 5.40c ), whereas the pore pressure

increased to about 70% of the confining pressure (Figs.

5.40b), which is reflected by the fact that the stress

path moves quickly towards the failure line in the first

stage of loading (Fig. 5.40d). These values are signifi­

cantly higher than the pore pressure and shear strain that

develop at the same stage during the first loading (see

Fig. 5.37), at the same shear stress amplitude, even

though the relative density of the reconsolidated specimen

was higher. It is also noted that the "threshold" shear

stress is smaller for the reconsolidated specimen than for

the specimen with no prior strain history. On the other

c hand, the maximum shear modulus, G


max
reconsolidated specimens was essentially the same (102 and
of the virgin and
°

98 MPa, respectively). These results indicate that the

degradating effects of large strains during the first

loading on the deformability of the fabric existing after

reconsolidation completely overrode the increase in

"stiffness" associated with the increase in density of the

specimen. However, the reconsolidated specimen only exhi­

bits a slight strain softening behavior (after point A in

Fig. 5.40c). This is attributed to the increase in the

residual strength of the specimen. Hence, after reconso­

lidation the structure of the sand is more deformable but

less brittle.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
414

Towards, the end of the first loading stage in the

first cycle, a slight dilation is apparent, which slows

the rate of deformation (Fig. 5.40c) and allows the speci­

m e n to build up strength up to the shear stress amplitude

(2 15 kPa) at which the reversal of rotation occurs (point

B in Fig. 5.40d). It is evident that at this point the

effective stress path has already crossed the phase

transformation line and therefore the specimen collapses

as the direction of shearing reverses. The pore pressure

increases again and the stress path moves towards the ori­

gin of the q-p" plot (Fig. 5.40d). A drastic reduction in

strength is observed as the shear stress reverses sign

(point C in Fig. 5.40c). Thereafter, the behavior is as

observed in other previous tests. As a result of dilation

with further straining the pore pressure decreases, the

specimen builds up strength at an increasing rate while

the stress path moves backwards close to the failure line.

Reversal of the direction of shearing at point D causes a

complete collapse of the specimen, which reaches a state

of initial liquefaction (Fig. 5.40b). As the shear stress

reverses again the specimen undergoes very large deforma­

tions (; 18%) before mobilizing a strength equal to the

shear stress amplitude of about 15 k P a .

The most important feature of the behavior shown in

Figure 5.40 is that the reconsolidated specimen, which is

denser, develops a condition of initial liquefaction at

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
415

y the end of the first cycle, as compared to more than 10

^ cycles that were required to bring the specimen into a

"liquefied" state during the first loading (Fig. 5.37).

Then, it is clear that the pore pressure buildup and con­

sequent strain development under cyclic loading depend on

the deformation characteristics of the initial structure

which in turn depends not only on the relative density but

also on the initial particle arrangement. The latter is

drastically altered by prior strain histories which

involve large shear strains. It is considered that the

drastic change in behavior due to different initial parti­

cle arrangements cannot be accounted for by the small

difference in the G values of the two specimens.


max

^ In the previous test, after a unidirectional deforma­

tion of about 10.5% in the clockwise direction, the speci­

men was rotated in the counter-clockwise direction until

the residual shear strain was reduced to about 3.8% (Fig.

5.37c) before reconsolidating it under the initial confin­

ing stress of about 100.1 kPa. However, the reduction in

resistance to liquefaction due to prior strain histories

with large shear strain amplitudes is observed even if the

residual shear strain at the end of the first loading is

smaller. This is illustrated by the results presented in

Figures 5.41 and 5.42. Another solid cylinder specimen

was first cyclicly loaded until collapse was initiated

(point A in Fig. 5.41). After a unidirectional shear

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
416

o Ottawa 2 0 - 3 0 sand
e - 0^622 Dr - 45.2 % ° " z - o " r - 100.4 kPa

60 120 180 240 300 360 420 480

z
<
DC -2
H
CO
DC
<
Ui -4
z
CO

-6

-8
TIME, sec.

Confining pressure —^
100
Collapse on
c reversal Liquefaction
80 ■
a
a.
UJ
tr
=> 60
CO Collapse on
CO
UJ loading —
a.
a.
Ui 40
cc
o
CL

20

60 120 180 240 300 360 420 480


TIME, sec.

Figure 5.41 Cyclic Torsional Shear Test on Virgin


Specimen (t =15.6 kPa).
cy

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
417

20

STRESS, kPa

-9 -8 -6 -5 -4 -3 -2
SHEAR

-4

-8

-12

-16
SHEAR STRAIN, %
-20
40 r

30 ■

20 ■
STRESS, kPa

10 •

20 30 40 50 60 90 100
SHEAR

Forced
-20 reversal 11th cycle

-30

EFFECTIVE CONFINING STRESS.kPa


-40

Figure 5.41, continued

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
418

c 8

6
e - 0.604
Dr - 52.7 %
o~z - 0 " r - 100.2 kPa
First loading in Fig. 5.41

4
After reconsolidation
2

* 0
60 120 80 240 300 360 420 480
-2
<
GC
h -4
CO

GC
< -6
U]
X -8
CO

-10

-1 2

-1 4

-1 6 TIME, sec.

-1 8

Confining pressure
100

Liquefaction
80 •

UJ
GC
X 60
CO
CO
UJ
GC
CL

UJ
ac 40
o
CL

20

60 120 180 240 300 360 420 480


TIME, sec.

Figure 5.42 Cyclic Torsional Shear Test on Reconsolidated


Specimen with Residual Shear Strain (-1.0%).

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
419

20

After reconsolidation
STRESS, kPa

-18 -16 -14 -1 2 -10 -8 -6

—4
SHEAR

-8

-12

-16
SHEAR STRAIN, %
-20
40

30

20
STRESS, kPa
SHEAR

30 40 50 60 70 80 90

-1 0

-20 Phase transformation line

-30

EFFECTIVE CONFINING STRESS,kPa


-40*-

Figure 5.42, continued

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
420

strain of about 7% in the clockwise direction, a reversal


o in the direction of shearing was forced at point B and the

specimen was rotated back in the counter-clockwise direc­

tion until the residual strain was reduced to less than

1%. It in noted that after reversal at point B initial

liquefaction develops (Figs. 5.41b and 5.41d) and the

specimen exhibits essentially zero strength over a wide

strain range (Fig. 5.41c). After this stage, the specimen

was reconsolidated under the initial confining stress and

subjected again to cyclic loading under undrained condi­

tions at the same shear stress amplitude (I 15.6 k P a ) .

The results of the second loading are presented in Figure

5 .42.

r
During reconsolidation the relative density of the

specimen increased from 45.2% to about 52.7%. It is seen

again that the reconsolidated specimen (Fig. 5.42)

develops larger pore water pressures and larger shear

strains in the first loading stage of the first cycle than

the "virgin" specimen (Fig. 5.41) and that initial

liquefaction also develops at the end of the first cycle

(Fig. 5.42) compared to 11 cycles before the development

of initial liquefaction during the first ioading (Fig.

5.41). Therefore, the behavior of the reconsolidated

specimens in Figures 5.40 and 5.42 is essentially the

same, although the pore pressure and shear strain during

the first cycle and the shear strain required to mobilize

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
421

a strength equal to the shear stress amplitude after ini­


o tial liquefaction are slightly larger in Figure 5.42.

This attributed in part to the fact that the shear stress

amplitude in Figure 5.42 (2 15.6 kPa) is slightly larger

than in Figure 5.40 (T 15 kPa). However, it should be

noted that both the maximum shear strain during first

loading and the residual shear strain after reconsolida­

tion were larger in the test presented in Figure 5.40. In

any case, the main issue is that the resistance to

liquefaction of the reconsolidated specimens with a rela­

tive density of about 52 to 53% is significantly smaller

than that of the specimens with no prior strain history,

which have a lower Dr (2 45%). This is attributed to

c degradating changes in fabric resulting from the develop­

ment of large strains during the first loading stage.

In the previous two tests, the specimens were tested

after reconsolidation with a residual shear strain towards

the clockwise direction. The first loading stage in the

first cycle was also in the clockwise direction. There­

fore, it was of great interest to study also the effects

of a residual shear strain in the direction opposite to

the direction of first loading. These effects are illus­

trated by the results presented in Figures 5.43 and 5.44

which were obtained from a cyclic test on another solid

cylinder specimen with a shear stress amplitude of 17.5

kPa. The specimen collapsed in the second cycle of the

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
422

10 r
Ottawa 2 0 - 3 0 sand
e - 0.62 Dr - 45.9 %
o - z - c r r - 100.2 kPa
SHEAR STRAIN,

30 90 120 150 180 210

-2

-4

-6

-8 TIME, sec.

-10

Confining pressure
100

First Second
80 ■ liquefaction iquefaction
PORE PRESSURE, kPa

60

Collapse on
40
loading

20

30 60 90 120 150 180 210 240 270 300 330

TIME, sec.

Figure 5.43 Cyclic Torsional Shear Tests on Virgin


Specimen ( t =17.5 kPa).
r cy

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
423

/ 20

—4

-8

-1 2

SHEAR STRAIN, %
-20
40

c 30

20 •

as
Q.
■X
CO
CO
Ui
IT 40 100
I- 50 60 70 80 90
co 0
cc
<
UJ
^ -10
Forced
-20 reversal

-30

EFFECTIVE CONFINING STRESS,kPa


-40

Figure 5.43, continued

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
4 24

u Ottawa 2 0 - 3 0 sand
e - 0.598 <r-z - c r ’r = io 0.2 kPa
Dr - 55.2 % First loading in Fig. 5.43

After reconsolidation
5 2
<
cc ■i
6 0‘ .60 120 240 300 420 480
cc
ui -2 •
x
CO
-4 •

-6

-8 •
TIME. sec.

-10

Confining pressure
100

c to
80 •
Liquefaction

a£.
J
UJ
CC
3 60 •
CO
CO
UJ
cc
a.
UJ
cc 40 •
o
a.

20 •

60 120 180 240 300 360 420 480


TIME, sec.

F ig u r e 5 .4 4 C y c lic T o r s io n a l S h e a r T e s ts on R e c o n s o li­
d a t e d S p e c im e n w i t h R e s i d u a l S h ear S tr a in
(+0.5%).

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
425

O 20

-4

-8

-12

-16
SHEAR STRAIN, %
-20
40
c
30 •

20 ■

co

30 40 50 60 70 80 90 1D0

til
-1 0
co

-20 Phase transform ation line

-30

EFFECTIVE CONFINING STRESS,kPa


-40
Figure 5.44, continued.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
426

first loading (point A in Fig. 5.43d). After a unidirec­

tional shear strain of about 8 .2% in the clockwise direc­

tion, a reversal in the direction of loading was forced at

point B, bringing the specimen to a state of initial

liquefaction. Thereafter, the specimen was rotated until

a residual shear strain of about 4.7 developed in the

counter-clockwise direction (Fig. 5.43c). It is noted

that after a peak-to-peak shear strain of about 13%, the

specimen mobilized a shear stress of only 4 kPa (point C

in Fig. 5.43c). A second reversal in the direction of

shearing was forced at point C, bringing the specimen

again to a state of liquefaction. Rotation in the clock­

wise direction continued until the residual shear strain

was reduced to about 0.5%. It should be noted that the

residual shear strain remained towards the counter­

clockwise direction. At this stage, the specimen was

reconsolidated under the initial confining stress of about

100.2 kPa, and subjected to cyclic loading at the same

shear stress amplitude (” 17.5 k P a ) . The results of the

second loading are presented in Figure 5.44.

During reconsolidation the relative density of the

specimen increased from 45.9 to about 55.2%. Contrary to

what was observed in the previous two tests, it is noted

that the reconsolidated specimen (Fig. 5.44) developed

smaller pore pressures and smaller shear strains in the

first half of the first cycle than the virgin specimen

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
427

(Fig. 5.43). However, during the second half of the first

G cycle the opposite was the case. Hence, the prior strain

history reverted the stress-straln characteristics of the

specimen, which became stiffer on loading in the clockwise

direction and softer on loading in the counter-clockwise

direction. This behavior is clearly reflected by the

stress— strain curves (compare Figs. 5.43c and 5.44c) and

the stress path plots (compare Figs. 5.43d and Fig.

5.44d). During the first loading of the second cycle the

reconsolidated specimen exhibits a slight strain softening

behavior, (after point A in Figs. 5.44c and 5.44d) which

is arrested by a reduction in the pore pressure due to

dilation after a unidirectional shear strain of about 5%.

Thereafter, the specimen rebuilds strength up to the level

of the applied shear stress amplitude at which the rever­

sal of the direction of rotation occurs (point B in Figs.

5.44c and 5.44d). Upon reversal the specimen collapses

and a state close to initial liquefaction develops. As

the shear stress reverses sign (point C in Fig. 5.44c) the

stress-strain curve becomes flat reflecting the loss of

strength. However, as in previous cases, the specimen

mobilizes strength with further straining due to a reduc­

tion in the pore pressure caused by the dilative tenden­

cies. After a peak-to— peak shear strain of about 11% the

reconsolidated specimen mobilizes a strength of 17.5 kPa

(Fig. 5.44c). It should be noted that after a unidirec­

tional shear strain of about 13% following initial

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
428

liquefaction the virgin specimen only mobilized a strength


o of 4 kPa (Fig. 5.43c). In relative terms, it can be con­

cluded that the reconsolidated specimen exhibited a larger

resistance to pore pressure and strain development than

that of the virgin specimen. This is based on the fact

that the reconsolidated specimen is less susceptible to

strain softening behavior and that small shear strains

develop upon the occurrence of initial liquefaction. The

author attributes in part the more favorable behavior of

the reconsolidated specimen to the fact that the first

loading at a shear stress amplitude of 17.5 kPa caused a

higher increase in density than in previous tests at

smaller shear stress amplitudes. Hence, it is considered

that the hardening effect due to densification overrode

the weakening effect due to changes in fabric caused by

large strains during the previous cyclic loading.

The tests results presented in Figures 5.40 to 5.44.

clearly show that prior strain histories which involve

large shear strain amplitudes create a particle packing

which, although denser, may be more susceptible (Figs.

5.40 and 5.42) or less susceptible (Fig. 5.44) to pore

pressure buildup and strain development during subsequent

cyclic loading at the same shear stress amplitude. In

addition to the overall changes in Dr and fabric, another

factor controlling the behavior seems to be the relative

direction and orientation of previous cyclic loading,

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
especially of the last loading stage, as compared to the

orientation and direction of the subsequent cyclic loading

after reconsolidation (77, 78, 178» 182‘, 196) . For example,

it is noted that the first loading in Figures 5.37 and

5.41 was terminated with rotation of the specimen in the

counter-clockwise direction. As illustrated in Figures

5.40 and 5.42, the second loading of these two specimens

was initiated with rotation in the clockwise direction.

Therefore, the initial direction of rotation in the second

loading was opposite to the direction of rotation in the

final stage of the first loading. It was pointed out that

regardless of the residual shear strain towards the clock­

wise direction, the reconsolidated specimen was softer and

more compressible than the virgin specimen and that the

stress path approached the failure line in the first stage

of loading in the clockwise direction. On the other hand,

it is noted that the first loading in Figure 5.43 was ter­

minated with rotation of the specimen in the clockwise

direction while the second loading in Figure 5.44 was ini­

tiated also with rotation in the clockwise direction. It

was pointed out that in this case the specimen became

stiffer in the clockwise direction but softer in the

counter-clockwise direction.

Consequently, it is considered that prior strain his­

tories involving large shear strains create a fabric or

particle arrangement which exhibit a highly anisotropic

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
430

stress-strain behavior, that is, a behavior very sensitive

to changes in the direction of loading (77,78,178,182).

Apparently, if the second loading is initiated with rota­

tion in the direction opposite to that during the terminal

stage of the previous cyclic loading, the resistance to

pore pressure buildup and strain development in the first

loading cycle decreases significantly. This is attributed

to the fact that under very small confining pressures such

as after initial liquefaction the particles become

strongly oriented in the direction of rotation with most

of the normals at the contact points, if any, oriented in

the direction of the major principal stress. Moreover,

after large unidirectional shear strains dilation develops

since the effective confining stresses are very low.

Accordingly, at the end of the first cyclic loading the

group of particles could be left in a "dilated" condition

(198). It is clear that under these conditions reversal

in the direction of shearing in the second loading causes

an immediate collapse of the structure and the consequent

development of large pore pressures and shear deformations

in the first loading cycle, as observed in Figures 5.40

and 5.42. In other words, the residual fabric after

reconsolidation requires a significant rearrangement of

grains in order to withstand the load in the opposite

direction, which result in a greater tendency to contract,

larger pore pressures and lower stiffness (178).

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
431

On the contrary, it follows that if the second load­


o ing is initiated with rotation in the same direction that

during the last portion of the first cyclic loading the

resistance to pore pressure buildup and strain development

in the first loading cycle increases as observed in Figure

5.44. It should be noted that it is the direction of

shearing during termination of the first loading and not

the direction of the maximum shear strains that seems to

control the behavior in the first loading cycle under sub­

sequent cyclic loading. In this connection, it is noted

that in Figure 5.43 the specimen experienced larger defor­

mations in the clockwise direction than in the counter­

clockwise direction. Nevertheless, the specimen became

f stiffer when reloaded in the clockwise direction (Fig.

5.44), because the previous cyclic loading was also ter­

minated with rotation in the clockwise direction under a

very small effective confining stress.

In summary, it can be concluded that the effects of

prestraining involving large strains on the undrained

strength of sands after reconsolidation are controlled by

two counterbalancing factors: an increase in density and a

drastic change in fabric. In the test results presented

in Figures 5.40 and 5.42 the latter effect was dominant.

On the other hand, in the case of Figure 5.44, the

increase in density controlled the subsequent undrained

behavior under cyclic loading. It must be pointed out

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
that the highly anisotropic stress-strain characteristics

of the fabric created by prior stress histories involving

large strains have practical implications only in the case

of monotonic loading or of a single application of cyclic

shear stress after reconsolidation. Under continuous

cyclic loading the direction of shearing continuously rev­

erses and , thus, the relative importance of stiffening in

a particular direction disappears. The main issue in this

case is that prestraining at large strains may create a

fabric which is very susceptible to collapse on either the

first loading or the first reversal, depending on the

relative orientation and direction of the first cycle of

the subsequent cyclic loading with respect to the final

stage of the prior loading sequence.

The discussion of the results presented in Figures

5.38 to 5.44 permits the author to emphasize once more the

fact that the undrained stress-strain characteristics of

sands at a given Dr are not independent of the initial

fabric or particle arrangement, as suggested elsewhere

(18,132). The initial fabric may be significantly altered

by prior cyclic strain histories, which in turn is

reflected by drastic changes in the tendency towards den-

sification, and the resulting pore pressures under

undrained loading.

In summary, the results of this investigation and

those from previous studies on the effects of prior strain

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
433

history on the cyclic undrained behavior of sands

(47,77,78,151, 167,178,182,196) permit to draw the follow­

ing conclusions:

a. Undrained cyclic prestraining involving small shear

strain amplitudes, with subsequent drainage of excess

pore pressures, causes the resistance to liquefaction

to be higher than that of a virgin specimen. Cyclic

prestraining with very small shear strain amplitudes

tends to eliminate local instabilities at the grain

contacts and thus, to produce a fabric with more

resistance to densification and pore pressure

buildup. These effects are independent of the direc­

tion of loading.

b. On the other hand, if a sand specimen has experienced

large deformations in the course of prior cyclic

undrained loading, its resistance to liquefaction,

after reconsolidation under the initial state of

stress, may be considerably diminished although the

density of the reconsolidated specimen may have

increased. Undrained prestraining that generates a

condition of liquefaction or flow failure changes the

initial fabric to one more closely corresponding to

the critical or steady-state of deformation which,

although it may be denser, offers the minimum resis­

tance to shearing. However, these effects depend not

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
434

only on the magnitude of the prior shear strains but

largely on direction of loading during the terminal

stage of the previous cyclic loading especially if

very low effective confining stresses develop at this

stage. Consequently, sands which have experienced

large deformations during prior undrained cyclic

loading exhibit a highly anisotropic stress-strain

behavior under subsequent undrained loading after

reconsolidation. The resistance to pore pressure

buildup is especially diminished if the direction of

shearing reverses between the terminal stage of the

first loading and the the initiation of the subse­

quent undrained loading.

From the foregoing discussion, it is evident that the

relationship between the resistance to reliquefaction and

the level of shear strain during prior prestraining must

have a peak or turning point (167). Previous investiga-'

tions (47,167) suggest that when the shear strain during

prestraining exceeds the range from 0.5 to 3%, the

liquefaction resistance after reconsolidation becomes

lower than that of virgin specimens, although it depends

also on the relative orientation and direction of shearing

in the previous and subsequent loadings. As discussed

earlier, under certain conditions, large shear strains

only develop when the effective stress path crosses the

phase transformation lines. Accordingly, a stress/strain

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
435

o history for which the range of variation in stress states

is restricted to the domain bounded by the two lines of

phase transformation has been termed "small preshearing"


"7 *7\ _ _ j i n ^ _ — .r * pm . ; C;
' —i i a « V « ••
( / / /£ 63 X XX U3wi 6 u Xu £Xgui ^ J • 4^ • £1O UIOJ.1 pi COU C 61*

i n g ” stress history causes an increase in the resistance

to liquefaction after reconsolidation. On the other hand,

a stress history where the change in stress crosses the

lines of phase transformation has been termed "large

preshearing" (77), which is very likely to cause a reduc­

tion in the liquefaction resistance. In the author's

view, the criterion shown in Figure 5.45 may be valid if

the cyclic shear stress amplitude is smaller than the

steady-state strength or when the relative density is

c higher than the critical (Fig.

sands shear strains on the order of 0.5 to


3.11). For loose to medium

3Z may be suf­

ficient to trigger the collapse of the sand structure and

the initiation of a flow failure, provided the cyclic

shear stress is larger than the steady-state strength

(Fig. 5.34). As discussed earlier, the initiation of

strain softening behavior occurs before the stress path

reaches the phase transformation. Consequently, it is

considered that if the sand is susceptible to strain

softening behavior it is the state boundary surface and

not the phase transformation lines that separates the

"small preshearing" and the "large preshearing" regions in

the stress space.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
Reproduced with permission of the copyright owner. Further reproduction n
v,...

q
Failure line

Line of p h a s e
transformation q

(a) Small preshearing


p' = \ (aa+20Y)
R a n g e of small
\ preshearin g q
prohibited without perm ission.

Range of Large
preshearing

(b) Large preshearing

Figure 5.45 "Small Preshearing" and "Large Preshearing" Stress Paths


under Cyclic Loading of! Magnitude Smaller than Steady-State

436
Strength (After. Ref. 24).
437

CHAPTER 6

SUMMARY AND CONCLUSIONS

6.1 General Considerations

An integrated framework for understanding the

stress-strain behavior of sands under both monotonic and

cyclic loading was developed and is presented in this

dissertation. Emphasis was placed on the collapsiveness

of the sand structure, which is associated with sudden

rearrangement of grains and loss of contact points between

^ neighboring grains, either because the applied shear

stress approaches a threshold value on loading or as a

result of reversals in the direction of shearing under low

levels of confining pressure. Under undrained conditions,

both types of "collapse" result in a significant increase

in pore water pressure and a consequent reduction in

strength. The first collapse mechanism leads to "flow

deformation" whereas the second one results in "liquefac­

tion," as defined in this dissertation. Hence, in the

author's view, "flow deformation" and "liquefaction" are

in nature similar phenomena in the sense that both involve

a drastic reduction in the number of contact points

between neighboring grains and, consequently, a drastic

reduction in strength.

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
438

Flow deformation is a consequence of strain softening

behavior, which is Initiated when the state of effective

stress (stress path) reaches a boundary surface, which is

referred to as the "critical stress ratio" or "state boun­

dary" surface. On the other hand, liquefaction develops

during an unloading stage (reversal in the direction of

shearing), from an initial stress state beyond what has

been called "phase transformation" lines, at the instant

the state of stress becomes isotropic. If the loading

conditions do not bring the sand specimen to a hydrostatic

state of stress, liquefaction, that is, a condition of

zero effective stress cannot develop. The factors con­

trolling the location in stress space of the critical

stress ratio and phase transformation lines (see Fig.

c 6.1), were discussed in detail in Chapters 3 and 5.

Under cyclic loading, a sand specimen is subjected to

continuous reversals in the direction of shearing, that

is, it experiences stages of loading, unloading, and

reloading. Under these conditions, it was shown that

depending on the initial state of stress and the cyclic

shear stress amplitude, the corresponding effective stress

path may reach the critical stress ratio (CSR) surface

during any loading stage causing large shear strains to

develop. If the stress path moves beyond the phase

transformation line dilatancy will develop and further

pore pressure development will be attenuated. However, in

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
439

this case, large pore pressure and consequent strains will

develop in the subsequent unloading stages. Consequently,

during cyclic loading of saturated loose sands, the level

of deformation after a sufficient number of cyclic stress

applications could be due to strain softening behavior (or

flow deformation) or to a combination of flow deformation

followed by liquefaction. Nevertheless, strain softening

behavior, if any, will always be observed before the

development of a condition of initial liquefaction because

the "collapse" surface corresponds to a critical stress

ratio smaller than the threshold stress ratio marking the

change in the behavior of sands from contractive to dila­

tive on l o a d i n g , at the phase transformation line, as

illustrated in Figure 6.1

As discussed in Chapter 3, the problem of undrained

stability of a contractive saturated sand under earthquake

loadings can be studied by considering the way in which

the in situ effective stress path approaches the failure

envelope. It is this path that determines the failure

modes and the resulting deformations. Starting from an

initial anisotropic state of stress, which is typical of

most in situ conditions, the failure line can be

approached following different stress paths which depend

not only on the initial state of stress, and the nature

and magnitude of the applied loads, but also on the boun­

dary strain conditions. For example, under plane strain

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
Reproduced with permission of the copyright owner. Further reproduction

State b o u n d a r y surface
P h a s e transformation line ( C S R line)

Failure line

Liquefaction Collapse

X
prohibited without perm ission.

Figure 6.1 Development of Limited Flow Deformation and Liquefaction under


Cyclic Loading.

440
441

conditions such as in deposits under level ground, initial

liquefaction may develop after a sufficient number of

cycles, irrespective of the initial state of stress, but

strain softening behavior is unlikely. On the other hand,

close to the surface of sloping ground, the lateral

strains are not constrained, and the susceptibility to

strain softening increases with the inclination of the

slope. In this case, liquefaction potential is very low

because the state of stress must remain anisotropic due to

the shear stresses needed to maintain equilibrium imposed

by geometry.

In summary, there are two extreme unfavorable

behaviors regarding undrained strength of saturated sands:

flow deformation and liquefaction. It is considered that

a combination of these "failure" mechanisms may contribute

to the failure of a complex structure, such as an earth

dam. It is noted, however, that determination of the way

in which the state of stress actually changes within an

earth structure during an earthquake and hence of the way

in which the in situ effective stress path approaches the

failure line in the course of cyclic loading, is a goal

that still remains to be achieved.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
442

6.2 New Apparatus for Cyclic Shear Testing

o
As described In Chapter 4, a new apparatus was

designed and developed* which combines the features of

resonant column and torsional shear devices into a single

apparatus. In the past, one of the main advantages of the

resonant column test was considered to be the low shear

strain levels that can be applied to the specimen

corresponding to those produced in geophysical tests in

situ* thereby allowing one to judge if laboratory speci­

mens have been replicated, or to examine to what extent

they have been reconstituted to an "equivalent fabric"

which exhibits a shear modulus in agreement with the

values determined from in situ seismic techniques. (In

this connection* see section 6.3.1, item 11.) On the other


c
hand, the torsional shear test is now recognized as having

superior capabilities for determining soil behavior under

earthquake loading, including liquefaction phenomena.

Particularly, it is considered that:

a. the test allows a better simulation of the cyclic

shear stresses and the resulting rotation of princi­

pal stresses associated with earthquake loadings*

b. the orientation of the major principal stress with

respect to the potential planes of anisotropy is the

same in both halves of the loading cycles, and

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
443

c. Che pore pressure buildup in undrained tests is a

consequence only of Che applied shear stresses since

Che mean normal stress remains constant throughout

each cycle.

Some problems still remain regarding stress variation

across the section of a soil cylinder subject to torsional

loading as a result of curvature of the specimen and end

restraint at the specimen caps but they can be reduced to

acceptable limits with a properly selected specimen

geometry. The new apparatus can be used to determine the

variation of shear modulus, G, with shear strain ampli­

tude, Y , on a single solid or hollow cylinder specimen

over the entire range of shear strain amplitudes of


-4
engineering interest, i.e. from 10 % to 10% under drained

conditions, and to monitor pore pressures, shear strains,

and effective confining pressures during cyclic shear

tests under undrained conditions. These capabilities are

not duplicated by any other apparatus known to the author.

6.3 Test Results

Torsional shear tests were performed with the new

hybrid resonant column/torsional shear apparatus on recon­

stituted specimens of Ottawa 20-30 sand. Within the

design objectives, the apparatus performed in a superior

manner and provided test results which illustrate the

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
444

, stress-strain relationships under both drained and


V
^
J undrained conditions for a wide range of shear strain

amplitudes. The results of the tests performed were

presented and discussed in Chapter 5. The main conclusions

and findings of this investigation can be summarized as

follows:

6.3.1 Drained Conditions

1. For uniform Ottawa 20— 30 sand with bulky grains,


*
at the same o , a significant change in relative density
m
(from 40 to 90%) had only a relatively modest effect on

the value of the initial tangent shear modulus, G , and


6 * max’
on the pattern of its subsequent degradation with increas­

c ing strain amplitudes.

modulus
As the shear strain increases,

reduction curves at different relative densities,


the

Dr, seem to converge into a single curve suggesting that

in the process of cyclic loading under drained conditions

at increasing shear strain amplitudes, the void ratio of

the sand approaches a limiting or residual value, indepen­

dent of the initial void ratio.

2. The initial tangent shear modulus, G , of nor-


° max
mally consolidated dry specimens was found to be essen­

tially proportional to the square root of the mean confin-


*
ing stress, o for relative densities in the range of 40
m
to 60%. However, at large shear strain amplitudes the

shear moduli were found to be a linear function of o .


m

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
445

Therefore, the exponent of a with which the modulus


in
o varies is a function of the shear strain amplitude.

3. At the same Dr and shear strain level, Y» the

reduction in modulus, expressed as a percentage of 6 ,


Q a X
*
is larger for lower values of mean stress, a . The shape
m
and location of the modulus degradation curves (G/G
luoX

versus y ) depend also on the void ratio and on the soil

type. Consequently, the simplified procedure, which in

common practice utilizes a single average curve of G/G


max
versus y , can result in significant errors in the inter­

preted values of shear modulus. Nevertheless, if the

G/G values are plotted with respect to the normalized


max
shear strain, Y/Y » where Yp is a reference strain, the

c curves tend to collapse into a single curve, which can be

closely represented by a hyperbolic relation of the form

(59,60)

G = 1 + Y/Y (6.1)
max r

as y was found also to increase with the square root of


r
a . For a given soil, this relationship appears to be
m
independent of void ratio but the actual shape can vary

with different soils types (59).

4. At the same Dr, the damping ratio, D, of dry

specimens increases as the mean confining stress

decreases, especially for shear strain amplitudes above

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
446

—2 "
10 %. However, all the curves at different a approach a
m
limiting value at large strains. The increase in damping

values with increasing shear strain seems to correspond to

the degradation in shear modulus. In fact, if the D

values are plotted with respect to G/G , the curves at


max
*
different a collapse approximately into a single curve,

suggesting that the equivalent shear modulus and damping

ratio are not independent variables. However, the rela­

tionship between D and G/Gmax is not linear, as suggested

elsewhere (59,60).

5. Degradation in the shear modulus of dry specimens

was observed even at small shear strain amplitudes in the


-4 -3
range from 10 % to 10 %. From this point of view, if a

threshold shear strain exists, its value is less than


-4
10 %. However, a sharp increase in damping, which

reflects the non-linear and inelastic properties of the

sand only occurs for shear strain amplitudes above .10 %.

On the other hand, the pore pressure buildup in undrained


_2
tests at shear strain amplitudes below 10 % was observed

to be small, in spite of a considerable reduction in

m o d u l u s.

^ *
6. As the stress ratio during consolidation, <r2 /°r >

increases, the initial shear modulus, G , decreases.


max
However, stress ratios as high as those corresponding to

failure only decrease G up to a maximum of 20 to 30%.


013 X

This behavior implies that G is not sensitive to


max

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
447

, changes in fabric since the particle arrangement certainly

changes in a drastic manner as the stress ratio increases

to failure. The fact that 6 is not significantly


max
affected by stress ratio also reflects the slip/stick

frictional nature of the stress-strain relation for granu­

lar materials, which is actually composed of small

"steps," which seem to have essentially similar initial

slopes (G ), irrespective of tv e stress ratio (or shear


max
strain amplitude). Moreover, it was concluded that

resonant column tests cannot reflect stress ratio effects

even at shear strains larger than that at which G is


° max
measured. This is because resonant column tests measure

the unloading-reloading secant shear moduli after several

c hundred cycles, which is considerably higher than, for

example, the tangent modulus during the stage of virgin

loading in the first cycle. Although in torsional shear

tests the number of applied cycles is significantly

smaller than in resonant column tests, it is still true

that under anisotropic stress conditions the difference in

behavior between the first cycle and subsequent ones can

be very significant. Therefore, stress-strain models

should consider the changes of stress and strain

throughout each loading and unloading stage, especially if

correlations are to be made, for example, with behavior

under undrained cyclic loading conditions.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
7. At the same a , isotropic prestressing with an
m
OCR of 2.0 had a negligible effect on the shear modulus-

shear strain relation in drained shear, for shear strain


—4
amplitudes ranging from 10 % to 2%. Prestressing at

increasing stress levels (or increasing OCR) but along a

stress path with constant stress ratio (other than one)

was found to cause a degradation in the shear modulus sub­

sequently measured at low shear strains. Consequently,

prestressing along a particular stress path to higher

values of OCR does not necessarily result in stiffening of

the sand specimen when reloaded along other stress paths.

In this connection, it was also estimated that G could


max
increase with OCR for tests with no lateral strain during

unloading. In any case, it is noted that stress history

effects only change G on the order of 15 to 20%.


J ° max
Accordingly, it was concluded that G is relatively
° max
insensitive to stress history effects.

8.
Cyclic prestraining at low shear strain amplitude
_o
(1.3 x 10 %) in the resonant column mode (several hundred

cycles) caused the shear modulus at lower strain ampli­

tudes to increase slightly above the value corresponding

to that of a specimen without any prestraining. On the

other hand, two cycles, which can be considered as pres­

tressing, at a constant shear strain amplitude (2.2 x


_2
10 %) in the torsional shear mode caused the shear
-4 —2
modulus at strains ranging from 10 % to 1.3 x 10 % to

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
449

decrease slightly below the values corresponding to the


o specimen with no previous stress history. It is evident

that the prestraining effects depend on the number and

magnitude of prior strain or stress repetitions.

Nevertheless, it is noted that in either case the change

in G was relatively small (I 5%), in spite of the fact


max J
that one case included a significant number of cycles dur­

ing the prestraining stage. On the other hand, cyclic

prestraining in the resonant column mode at increasing


-4 -2
shear strain amplitudes ranging from 10 % to 10 %,

caused a significant increase in the shear moduli at shear


_2
strains larger than 10 %. Consequently, it was observed

that specimens with essentially the same G may exhibit


inaX
very different "large strain" moduli depending on the
c cyclic strain history. In the light of these results,

G^ax does not seem to be an appropriate parameter to be

correlated with the response of sands to cyclic loading


-2
involving shear strains larger than 10 %.

9. The determination of G in the resonant column

mode implies the application of a considerable number of

cycles, which produce a significant stiffening of the

specimen. It is not yet known at which strain level the

stress-strain characteristics of sands as measured by

resonant column tests, change as a function of the number

of cycles. Consequently, if staged testing includes the

determination of shear moduli by means of resonant column

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
4 50

tests at increasing shear strain amplitudes, it can lead

to a significant overestimation of shear modulus for sand

with no prior cyclic strain history even at shear strain


-2
amplitudes as low as 1.1 x 10 %.

10. In general, it can be concluded that the influ­

ence of prior stress or strain histories is controlled by

three different factors:

a. the prestress effect itself, which is dependent on

stress path and strain level (OCR),

b. the destructuration that occurs as the shear strain

increases, and

c. the number and magnitude of prior cyclic strain or

stress repetitions.

The second factor has an effect opposite to that of

the other two. On the other hand, under earthquake load­

ings, the planes of principal stress continuously rotate,

and, thus, the relative importance of prestressing along a

particular stress path may also change.

11. The stress-strain characteristics of sands under

drained cyclic loading are affected by relative density,

mean confining stress, stress ratio, stress history, and

prior number of strain applications. However, the

behavior at very small strains, as measured by G , is


1Q3X

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
451

mainly affected by o and Dr. The effects of stress


TO
o ratio, stress history and cyclic prestraining on Gmax are

modest. Consequently, a direct correlation between Gmax

and "large strain" behavior such as liquefaction potential

which is affected by the latter factors, is not possible.

For the same reasons, correlations between shear wave

velocities measured at very small shear strains and

liquefaction potential of natural sand deposits is ques­

tionable.

6.3.2 Undrained Conditions

12. Once the effective stress path corresponding to

an undrained test is at a state of zero effective stress

(liquefaction), the amount of shear strain required to


c
mobilize a given shear resistance, in the opposite direc­

tion, increases significantly as the relative density of

the sand decreases. After the collapse of the structure

due to the reversal in the direction of shearing, the

looser the specimen the larger the shear strains required

for new contact points to develop.

13. Under cyclic loading loose sands may experience

very large deformations even without the development of a

condition of initial liquefaction. This is attributed to

the collapse of the structure on loading, which leads to a

sharp increase in pore pressure and a consequent reduction

in strength. Since this collapse occurs at a stress ratio

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
452

significantly smaller than that corresponding to the phase

transformation line, it will always occur before a condi­

tion of initial liquefaction can develop, provided the

cyclic shear stress amplitude is larger than the steady-

state strength. Consequently, a specified level of shear

strain in loose sands under cyclic loading may result from

either the shear strains that develop before the stress

path reaches the phase transformation line (including

strain softening behavior), or the strains accompanying

the development of initial liquefaction, or a combination

of the two mechanisms of strain accumulation.

14. In strain controlled undrained tests the princi­

pal stress ratio remains approximately constant for a

given shear strain amplitude, whereas the rate of pore

pressure accumulation decreases with number of cycles.

Such tests cannot reflect some features of stress con­

trolled behavior, in which case the principal stress ratio

increases as a result of the pore pressure buildup.

Therefore, the results of these two types of tests cannot

be correlated directly, as suggested elsewhere (161). For

example, if the stress ratio associated with the maximum

shear stress amplitude is smaller than that corresponding

to the critical stress ratio (CSR) line, strain softening

behavior cannot develop in a strain controlled tests

whereas it may develop in a stress controlled test.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
453

15. In Che first few cycles subsequent to the first

cycle, the effective stress paths corresponding to tor­

sional shear tests are essentially symmetrical about the

normal stress axis since the mean confining stress and the

orientation of the principal stresses is such that the

response of the specimen is expected to be the same

whether it is rotated right or left. This condition can­

not be obtained in cyclic triaxial tests.

16. The accumulation of significant pore pressures

seems to be unlikely at shear strain amplitudes smaller


—2 "
than 10 %, irrespective of Dr and o . However, at larger
m
shear strains the pore pressure ratio u /a cannot be
m
independent of these factors, as suggested elsewhere (24).
-
( It was found that the lower a the higher the pore pres-
V ®
sure ratio, u/o , which is consistent with a trend towards
m
greater reduction in stiffness and higher damping with

decreasing confining stress at a given shear strain ampli­

tude. However, the absolute value of pore pressure actu­

ally increases with confining stress at the same shear

strain amplitude, due to the greater tendency to contract

under higher confining stresses. Hence, the generated

pore pressure depends not only on the magnitude of the

plastic deformations (which is reflected in a reduction in

stiffness) but also on the tendency towards densification


*
(which increases with a ). Whether significant pore pres-
m
sures accumulate or not in the course of cyclic loading

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
454

depends not only on the level of applied stress (or


o strain) but also on the initial state of stress and on the

boundary strain conditions.

17. Shortly before a sand specimen mobilizes the

monotonic peak strength there is an increase in the rate

of pore pressure buildup as reflected by a change in cur­

vature of the pore pressure versus time (or shear strain)

plot. The author postulated that small shear strains may

be enough for the metastable structure of a loose sand to

collapse, that is, to experience a "sudden" tendency to

rearrangement of grains or a reduction in the number of

contact points between neighboring grains. This struc­

tural instability, which occurs close to the mobilization

c of the peak shear strength, results in a significant

increase in pore water pressure and a consequent reduction

in strength. To the author's knowledge, this feature of

the behavior of loose sands has not been reported by any

of the previous studies on the undrained behavior of

sands. All the test results available to the author,

which basically were obtained from axial compression

triaxial tests, suggest a "hyperbolic" type of relation­

ship between pore pressure and axial strains, which

implies that the rate of pore pressure buildup is highest

in the initial stage of loading, which is not consistent

with the initiation of a strain softening process at the

peak point. The sharp increase in the rate of pore

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
455

pressure buildup resulting from the "collapse" of the


o structure of loose sands becomes more evident under cyclic

loading at shear strain amplitudes smaller than the peak

monotonic strength.

18. The trace of the stress-strain curve after the

peak point under monotonic loading shows minute fluctua­

tions reflecting again the stepwise (frictional) nature of

the stress-strain characteristics of sands. The same

behavior was also observed whenever strain softening was

initiated by cyclic loading.

19. The locus of stress states triggering the col­

lapse during cyclic loading defines an envelope which

c essentially corresponds to the stress path under monotonic

loading at the same relative density. Accordingly, it is

concluded that the monotonic effective stress path consti­

tutes a "collapse" or state boundary surface, which deter­

mines the initiation of strain softening behavior under

cyclic loading, provided the cyclic shear stress amplitude

is larger than the steady-state strength. Consequently,

the angle of shear resistance mobilized at the triggering

state, <f> , is not a constant value but a variable one


* mob
depending upon the cyclic shear stress amplitude.

The state boundary or collapse surface seems to con­

stitute a kind of strain contour corresponding to the

cumulative shear strain causing the initiation of strain

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
456

softening behavior, irrespective of the cyclic shear .

G stress amplitude and/or number of cycles. After the col­

lapse, the magnitude of shear strain increases as the

effective stress path approaches the failure envelope

along the state boundary surface, which is postulated to

depend on the orientation of the principal stresses

(stress path). From this point of view, triaxial tests

and torsional shear tests are expected to yield different

collapse surfaces. Cyclic loading conditions may be such

that the stress state of the specimen is compelled to

remain on the state boundary surface or to move away from

it. The pore pressures (and shear strains) that develop

when the stress state of the specimen remains on the state

boundary surface are substantially larger than when the

C stress path moves away from such a collapse surface, pro­

vided the stress state is far from the line of phase

transformation. Unloading in the region of the state

boundary beyond the phase transformation line (dilative

region) also produces significant pore pressure buildup

which may lead to a state of liquefaction.

20. Four main stages in the process of pore water

pressure generation in "loose" sands under cyclic loading

were clearly identified, namely:

a. the pore pressure buildup during the first cycles of

loading due to the initial tendency to contract,

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
457

which proceeds at a decreasing rate as the fabric


C I
becomes more stable due to the prior stress cycles,

b. a sharp increase in pore pressure upon the collapse

of the structure on loading, which then proceeds at a

decreasing rate as the specimen approaches the

"steady-state" condition,

c. an additional increase in pore water pressure as the

structure collapses again upon reversal in the direc­

tion of shearing, and

d. reduction in pore pressure due to dilation after

large strains under small effective confining

stresses.

c 21. The prior stress (or strain) cycles during

undrained cyclic loading cause strain history effects on

the subsequent cycles within the same loading sequence.

These effects are reflected by a reduction in the rate of

pore pressure buildup with increasing number of cycles,

provided that neither the state boundary surface nor the

phase transformation line are reached in the course of

cyclic loading. This is attributed to a kind of "stiffen­

ing" of the sand structure which is also reflected in a

reduction in the magnitude of shear strain increments

(increase in G) that develop during the initial stages of

subsequent loading cycles under stress controlled

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
458

conditions, irrespective of the fact that the pore pressure

continuously increases. Accordingly, the variation in

shear modulus with number of cycles under undrained cyclic

loading is the result of two counterbalancing effects: an

increase in 6 due to "prestraining" caused by the prior

cycles and a degradation in G associated with the progres­

sive increase in pore pressure. However, the state boun­

dary surface apparently is not affected by this prestrain­

ing.

22. Dndrained cyclic prestraining involving small

shear strain amplitudes, with subsequent drainage of

excess pore pressures, causes the resistance to liquefac­

tion to be higher than that of a virgin specimen. These

effects are independent of the direction of loading. On

the other hand, if a sand has experienced large deforma­

tions in the course of prior cyclic undrained loading, its

resistance to liquefaction, after reconsolidation under

the initial state of stress, may be considerably dimin­

ished although the density of the reconsolidated specimen

may have increased. Moreover, sands which have experi­

enced large deformations during prior undrained cyclic

loading exhibit a highly anisotropic stress-strain

behavior under subsequent undrained loading after reconso­

lidation. It follows that the resistance to liquefaction

after drainage of excess pore pressures depends on the

prior level of shear strain. If the sand is susceptible

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
459

/- to strain softening behavior, it is the state boundary

surface that separates the "small preshearing” and the

"large preshearing" regions in the stress space, and not

the phase transformation lines, as has been suggested

elsewhere (77).

6.4 Commentary

The most important practical implications of this

study can be summarized as follows:

1. It Is considered that "flow deformation," or strain

softening behavior, and "liquefaction," i.e. a condi­

tion of zero effective stress, are in nature similar

phenomena in the sense that both involve a drastic

reduction in the number of contact points between

neighboring grains and, consequently, a drastic

increase in pore water pressure and corresponding

reduction in shear strength.

2. The maximum shear modulus, G , of sands is mainly


max
affected by the mean confining stress and to a lesser

extent by the relative density. The effects of stress

ratio, stress history and cyclic prestraining on G


SlSX

are very modest. Consequently, a close correlation

between G and "large strain" behavior such as


max
liquefaction potential, which is strongly affected by

the latter factors, is not possible. The same can be

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
460

said about correlations with shear wave velocities

measured at very small shear strains.

3. The determination of shear moduli by means of

resonant column tests involves the application of a

considerable number of stress cycles. This means

that the measured shear modulus corresponds to the

unloading— reloading secant modulus after several hun­

dred cycles. This shear modulus is considerably

higher than, for example, the tangent modulus during

the stage of virgin loading in the first cycle. It

is not yet known at which strain level the stress-

strain characteristics of sands as measured by

resonant column tests, change as a function of the

c number of cycles. Consequently,

tests at shear strains larger than 10


resonant column
—4
% may lead to

significant overestimations of shear modulus for

sands with no prior cyclic strain history.

4. The conditions leading to the different modes of

undrained failure under both monotonic and cyclic

loading have been clearly delineated throughout this

dissertation. It is considered that the work will

serve as a useful point of departure for further

research on sand liquefaction.

The author hopes that in the future geotechnical

engineers will systematically study the problem of

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
461

undrained stability under cyclic loading by considering

the potential ways in which the in situ effective stress

path approaches the failure envelope. It is this path

that determines the failure modes and the resulting defor­

mations. Starting from an initial anisotropic state of

stress, which is typical of most in situ conditions, the

failure line can be approached following different stress

paths which depend not only on the initial state of

stress, and the nature and magnitude of the applied loads,

but also on boundary restraint conditions. In this con­

nection, we are still far from being able to predict accu­

rately the way in which the state of stress changes within

earth structures during an earthquake.

The number of cycles of torsional shear stress

required to reach a stress ratio corresponding to strain

softening behavior appears to be very sensitive to the

applied shear stress ratio. As the number of cycles of

large shear stress amplitude during an earthquake is

decidely limited - generally ranging between 5 and 20

cycles - this sensitivity is a significant impediment to

the development of practical tools to predict flow defor­

mation and/or liquefaction potential due to earthquake

loadings.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
462

o
CHAPTER 7

RECOMMENDATIONS FOR FURTHER RESEARCH

7.1 Improvement of Laboratory Equipment

The new hybrid resonant column/torsional shear

apparatus is considered to have enhanced available tech­

niques for measuring the response of granular soils to a

wide range of cyclic loading. However, its capabilities

can be further improved as discussed below.

One of the main advantages of the torsional shear

C test is that the cell pressure, axial stress and torsional

shear stress can be changed as desired to obtain any

orientation of the planes of principal stress. Conse­

quently, the torsional shear test apparatus can be used in

fundamental investigations of anisotropy and principal

stress rotations in soils. To fully accomplish this,

automatic controls are needed to control simultaneously

the axial and torsional stresses and the cell pressure, a

capability that is not yet available in the new apparatus.

At present, only the torsional loading can be automati­

cally controlled; however, computer controlled solenoid

valves can readily be added to control fully the entire

stress path.

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
463

Plane strain cyclic tests are required to properly


G simulate the conditions to which an element of soil within

deposits under level ground is subjected during an earth­

quake. The new apparatus does not have this capability.

In principle, it would be possible to impose no lateral

strain by preventing any change in the volume of cell

water during torsional loading (75,81). This is not pos­

sible in the new apparatus because the Hardin oscillator

cannot be submerged. Besides, the cross sectional areas

of the rotating shaft and of the specimens are not the

same. Under undrained shearing, a condition of plane

strain can be achieved by fixing the length of the speci­

men (173). Alternatively, proximity transducers could be

used to measure radial displacements of the cylinder


C walls* However, this alternative is possible only for the

solid specimen. Proximity transducers sense the gap

between the transducer face and a target such as aluminum

foil attached to the membrane. Computer controlled

solenoid valves would then be used to change the cell

pressure automatically to prevent any radial displacement.

Shear strains larger than those corresponding to the

rated capacity of the transducers used for measuring rota­

tion developed in most of the undrained tests performed by

the author. Therefore, beyond the range of the transduc­

ers, shear strains were estimated by using the elapsed

time and the rate of straining, which for each test was

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
464

constant. This is a significant limitation, especially

under cyclic loading. An alternative system for measuring

rotations as the top platen cams rotate out of the range

of the proximity transducers is desirable. A rotary vari­

able differential transducer, RVDT, could be a possibil­

ity.

As described in Chapter 4, the new apparatus is

equipped with a differential pressure transducer, and,

therefore, independent measurements of the cell and pore

water pressures cannot be made. This is a limitation,

especially for accurate measurements of the pore pressure

B-parameters.

The volume change measurement system of the apparatus

is relatively crude. Accurate measurements of volume

changes can be made by monitoring the changes in the

height of water in a burette with an extremely sensitive

differential pressure transducer under close temperature

control conditions. This could allow continuous measure­

ments of the volume change during cyclic loading under

drained conditions. In this way, the effects of changes

in void ratio of the specimen on the drained behavior can

be accounted for. On the other hand, if the cell pressure

is automatically controlled, a constant volume test can be

performed. Constant volume drained tests versus undrained

tests at the same Dr may be a way to check the differences

between the S and F lines. It should be noted that with

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
465

o automatic measurement of volume changes, K q conditions

during consolidation can be also imposed by automatically

adjusting the cell pressure so that the volumetric strain

remains equal to the axial strain.

As discussed in Chapter 6 , determination of G in the

resonant column tests implies the application of a consid­

erable number of cycles, which produce a significant

increase in stiffness of the specimen. It is not yet

known at which strain level the stress-strain characteris­

tics of sands as measured by resonant column tests, change

as a function of the number of cycles. Consequently, it

becomes important to be able to operate the new apparatus


_2
in the torsional shear mode at strains below 10 %, in

order to avoid the prestraining effects produced by any

type of oscillator of high frequency. Currently, the

lower limit of shear strain amplitude at which reversal of

the direction of rotation can be controlled automatically

is about 2 x 10

Besides the number of cycles, the frequency is also

very different between the resonant column and the tor­

sional shear tests. The relative contribution of each one

of these factors to the response of sands in the resonant

column test is not yet fully established, although the

effects of changes in frequency are known to be important

in clayey soils (31). On the other hand, the maximum fre­

quency in the torsional shear mode is currently limited to

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
466

about 0.1 to 0.2 Hertz. Consequently, the torsional shear


c apparatus must be modified to expand Its capabilities to

the frequency range more characteristic of earthquake

loadings, i.e. at least to 0.5 Hz.

7.2 Topics for Further Research

Throughout this dissertation, the author identified

topics which require further investigation. They can be

summarized as follows:

1. The effects of frequency of loading on both

drained and undrained behavior of sands are still imper­

fectly known. The agreement in shear moduli at a shear


_2

c strain amplitude of about

column (high frequency tests)


10 % obtained from the resonant

and torsional shear tests

(low' frequency tests) has led some researchers to the con­

clusion that the frequency of loading does not affect the

response of sands to cyclic loading under drained condi­

tions (42,88,136). However, as discussed previously,

besides frequency, the number of cycles is also very dif­

ferent between the two tests. The relative contribution

of each one of these factors is not yet known. In regard

to undrained behavior, the conclusions drawn by previous

investigators are conflicting since experimental evidence

has been presented showing either that frequency does not

affect the undrained behavior of sands, at least in the

frequency range from 0.05 to 5 Hz (175,194), or that the

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
467

dilatancy tendencies of sands and hence the consequent


o pore pressures increase significantly at higher rates of

deformation (63,150).

2. The effects of initial shear stresses, that is of

stress ratios during consolidation other than one, are

also subject to conflicting opinions. Although 6 may


IDa X

not be significantly affected by stress ratio, the modulus

degradation curve must be dependent on the initial shear

stresses. This relationship is still to be established.

Particular attention should be paid to the changes in

stress and strain during the first few cycles of loading,

especially if correlations are to be made with behavior

under undrained conditions. Simplification of the

behavior in terms of a single secant shear modulus is not


c
desirable. Of course, resonant column tests are ruled out

for this purpose. On the other hand, some researchers

have concluded that the compressibility of sands under

axial compression triaxial tests is controlled by the

major principal stress, for a wide combination of confin­

ing pressure and stress ratio values (102,105,186). These

results conflict with the author's understanding of the

factors controlling the shear resistance of granular

soils. Accordingly, additional research is recommended in

this area. The effects of initial shear stresses on sub­

sequent undrained behavior are not clear either at this

point. It is felt that stress ratio values other than one

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
468

could affect the location and shape of the state boundary

o surface, and, consequently, the values of the peak and

steady-state strengths* It is evident that stress ratio

effects cannot be independent of inherent anisotropy and

of the nature of the stress path with respect to the prin­

cipal axes of anisotropy. Most investigations on the

steady— state strength of sands have been based mainly on

triaxial tests which correspond to a particular stress

path. Investigations to include other stress paths are

recommended.

3. It is claimed that the shear strain level at

which pore pressure buildup is initiated, that is the so

called "threshold shear strain" is independent of fabric,

stress level, stress path, and relative density


c
(24,34,70). These findings appear to conflict with a sig­

nificant amount of research which shows that the mechani­

cal behavior of granular materials is affected signifi­

cantly by factors such as fabric and it associated aniso­

tropy, stress-strain history, stress path, etc., among

others. The author concluded that the threshold shear

strain as defined may be another "small strain" parameter

with characteristics similar to G , and hence, rela-


max
tively independent of some of these factors. However, it

is felt that as the shear strain increases, the undrained

behavior is not independent of fabric, stress level and

relative density. Accordingly, a research program can be

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
469

o initiated to study the effects of these factors on the

relationship between the pore pressure ratio, u/a


*
and the

level of shear strain. The author recognizes that it is

the level of shear strain more than the level of shear

stress that seems to determine the mechanical behavior of

soils. However, it is not to be concluded that the

applied shear strain is the only factor controlling the

pore pressure response of saturated sands, as claimed by

some researchers. At the same shear strain level, volume

change tendencies, which determine the pore water pressure

buildup, are dependent on fabric, stress level, and the

like.

4. Other recommendations for further study relate

( to:

a. Additional tests, in many areas, at very low relative

densities, i.e. in the range from 20 to 40%. To

fully accomplish this, the specimen preparation pro­

cedures need to be improved.

b. Additional studies of the effects of prestressing on


—2
shear moduli at shear strain larger than 10 %.

c. Investigations of the usefulness and limitations of

correlations between state parameter (8 ) and drained

and undrained behavior of sands.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
470

o d. Effects of strain rate on the state boundary surface,

and hence, on the steady-state strength (F line).

(
v

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
c BIBLIOGRAPHY

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
471

BIBLIOGRAPHY

1. Alarcon, A., Chameau, J. L., and Leonards, G. A., "A


New Apparatus for Investigating the Stress-Strain
Characteristics and Liquefaction Potential of
Saturated Sands," submitted for publication to the
Geotechnical Testing Journal, November 1985.

2. Alarcon, A. and Leonards, G. A., "Liquefaction


Evaluation Procedure," Discussion, submitted for pub­
lication to Journal of Geotechnical Engineering,
ASCE, January 1986.

3. American Society for Testing and Materials, "Standard


Test Methods for Modulus and Damping of Soils by the
Resonant— Column Method," Annual Book of ASTM Stan­
dards, Vol. 4.08, D4015-81, Philadelphia, Pennsyl-
f . vania, pp. 677-701.
V 4. Arthur, J.R.F., Chua, K. S., and Dunstan, T.,
"Induced Anisotropy in a Sand," Geotechnique 27, No.
1, March, 1977, pp. 13-30.

5. Arthur, J.R.F., Chua, K. S., Dunstan, T., and Rodri­


guez del C, J. I., "Principal Stress Rotation: A -
Missing Parameter," Journal of the Geotechnical
Engineering Division, ASCE, Vol. 106, No. GT4, April
1980, pp. 419-433.

6. Baldi, G. , "Static and Cyclic Test on Sand,"


Geotechnical Engineering in Italy - An Ov erv iew ,
1985, A.G.I., Published on the occasion of the ISSMFE
Golden Jubilee, XI ICSMFE, San Francisco, 1985, pp.
233-238.

7. Baldi, G . , Bellotti, R . , Crippa, V., Fretti, C.,


Ghionna, V., Jamiolkowski, M . , Ostricati, D., Pascu-
alini, E., and Pedroni, S., "Laboratory Validation of
In Situ Tests," Geotechnical Engineering in Italy -
An O ve rv ie w, 19 85 , A.G.I., Published on the occasion
of the ISSMFE Golden Jubilee, 11th International
Conference on Soil Mechanics and Foundations
Engineering, San Francisco, 1985, pp. 251-270.
c

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
472

8. Been, K. and Jefferies, M. G . , "A State Parameter for


o Sands," Geotechnique, Vol. 35, No. 2, June 1985, pp.
99-112.

9. Bishop, A. W. , "Shear Strength Parameters for Undis­


turbed and Remoulded Soil Specimens," Stress-Strain
Behavior of Soils, Edited by R.H.G. Parry, 1972.
Proceedings of the Roscoe Memorial Symposium, Cam­
bridge University, March 29-31, 1971, pp. 3-58.

10. Bjerrum, L. Kringstad, S. and Kummeneje, 0., "The


Shear Strength of a Fine Sand," Proceedings 5th
International Conference on Soil Mechanics and Foun­
dation Engineering, Vol. 1, 1961, pp. 29-37.

11. Bouckovalas, G . , Whitman, R. V., and Marr, W. A.,


"Permanent Displacements of Sand with Cyclic Load­
ing," Journal of Geotechnical Engineering, ASCE, Vol.
110, No. 11, November 1984, pp. 1606-1623.

12. B ro ms , B. B, "Soil Sampling in Europe: State-of—


the-Art," Journal of the Geotechnical Engineering
Division, ASCE, Vol. 106, No. GT1, January 1980, pp.
65-98.

13. Casagrande, A., "Liquefaction and Cyclic Deformation


of Sands- A Critical Review," Harvard Soil Mechanics
Series No. 8 8 , January 1976, Presented at Fifth
Panamerican Conference on Soil Mechanics and Founda­
tion Engineering, Buenos Aires, Argentina, November
1975.

14. Castro, G., "Liquefaction of Sands," Harvard Soil


Mechanics Series No. 81, Ph.D. Dissertation, Harvard
University, Cambridge, Mass., 1969.

15. Castro, G., "Liquefaction and Cyclic Mobility of


Saturated Sands," Journal of the Geotechnical
Engineering Division, ASCE, Vol. 101, No. GT6 , June
1975, pp. 551-569.

16. Castro, G. and Christian, J. T., "Shear Strength of


Soils and Cyclic Loading," Journal of the Geotechni­
cal Engineering Division, ASCE, Vol. 102, No. G T 9 ,
September 1976, pp. 887-894.

17. Castro, G. and Poulos, S. J., "Factors Affecting


Liquefaction and Cyclic Mobility," Journal of the
Geotechnical Engineering Division, ASCE, Vol. 103,
No. GT6 , June 1977, pp. 501-516.

(
\

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
473

18. Castro, 6 ., Poulos, S. J., and Leathers, F. D., "Re-


Examination of Slide of Lover San Fernando Dam,"
Journal of Geotechnical Engineering, ASCE, Vol. Ill,
No. 9,. September 1985, pp. 1093-1107.

19. Chang,C. S., "Residual Ondrained Deformation from


Cyclic Loading," Journal of the Geotechnical
Engineering Division, ASCE, Vol. 108, No. G T 4 , April
1982, pp. 637-646.

20. Chang, C, S., Kuo, C. L . , and Selig, E. T., "Pore


Pressure Development During Cyclic Loading," Techni­
cal Note, Journal of Geotechnical Engineering, ASCE,
Vol 109, No. 1, January 1983, pp. 103~107.

21. Charlie, W. A., Muzzy, M. W . , Tiedemann, D. A., and


Doehring, D. 0., "Cyclic Triaxial Behavior of Mon­
terey Number 0 and Number 0/30 Sands," Geotechnical
Testing Journal, GTJODJ, Vol. 7, No. 4, December
1984, pp. 211-215.

22. Chung, R. M . , Yokel, F. Y. and Drnevich, V. P.,


"Evaluation of Dynamic Properties of Sands by
Resonant Column Testing," Geotechnical Testing Jour­
nal, GTJODJ, Vol. 7, No. 2, June 1984, pp. 60-69.

23. Clayton, C.R.I., Habasa, M. B., and Simons, E.,


"Dynamic Penetration Resistance and the Compressibil­
ity of a Fine Grained Sand — A Laboratory Study,"
Geotechnique, Vol. 35, No. 1, March 1985, pp. 19-31.

24. Committee on Earthquake Engineering, Commission on


Engineering and Technical Systems, National Research
Council, Liquefaction of Soils During Eart hqu ake s,
National Academy Press, Washington, D.C., 1985, 240
PP •

25. Committee on Soil Dynamics of the Geotechnical


Engineering Division, "Definition of Terms Related to
Liquefaction," Journal of the Geotechnical Engineer­
ing Division, ASCE, Vol. 104, No. G T 9 , September
1978, pp. 1197-1100.

26. Cuellar, V., Bazant, Z. P., Krizek, R. J., and


Silver, M. L., "Densification and Hystereris of Sand
under Cyclic Shear," Journal of the Geotechnical
Engineering Division, ASCE, Vol. 103, No. G T 5 , May
1977, pp. 399-416.

27. Daramola, 0., "Effect of Consolidation Age on Stiff­


ness of Sand," Technical Note, Geotechnique, Vol.
30, No. 2, June 1980, pp. 213-216.

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
474

28. Das, B. M . , Fundamentals of Soli D y n a m i c s , Elsevier


Science Publishing Company, New York, 1983, 399 pp.

29. De Alba, P., Baldwin.;, K. , Janoo, V.., Roe, G. and


Celikkol, B., "Elastic-Wave Velocities and Liquefac­
tion Potential," Geotechnical Testing Journal,
GTJODJ, Vol. 7, No. 2, June 1984, pp. 77-87.

30. De Alba, P., Seed, H. B. and Chan, C. K . , "Sand


Liquefaction in Large-Scale Simple Shear Tests,"
Journal of the Geotechnical Engineering Division,
ASCE, Vol. 102, No. G T 9 , September 1976, pp. 909-927.

31. Devenny, D. W . , "Strength Mechanisms and Response of


Highly Sensitive Soils to Simulated Earthquake Load­
ing," Thesis, presented to Purdue University in par­
tial fulfillment of the requirements for the degree
of Doctor of Philosophy, West Lafayette, Indiana, May
1975.

32. Dobry, R . , Powell, D. J . , Yokel, F. Y., and Ladd,


R.S., "Liquefaction Potential of Saturated Sand-The
Stiffness Method," Proceedings of the 7th World
Conference on Earthquake Engineering, Istanbul, Sep­
tember, 1980, Vol. 3, pp. 25-32.

33. Dobry, F*. , Stokoe II, K. H. , Ladd, R. S., and Youd,


T. L., "Liquefaction Susceptibility from S-Wave Velo­
city," Preprint 81-544, In Situ Testing to Evaluate
Liquefaction Susceptibility, ASCE, National Conven­
tion, St. Louis, Missouri, October 26-31, 1981.

34. Dobry, R . , Vasques-Herrera, A., Mohamad, R . , and


Vucetic, M. , "Liquefaction Flow Failure of Silty Sand
by Torsional Cyclic Tests," Advances in the Art of
Testing Soil under Cyclic Conditions, Edited by Vijay
Khosla, Proceedings of a Session at the ASCE Conven­
tion, Detroit, Michigan, October 1985, pp. 29-50.

35. Dobry, R . , Yokel,F. Y., and Ladd, R. S., "Liquefac­


tion Potential of Overconsolidated Sands in Areas
with Moderate Seismicity," Earthquakes and Earthquake
Engineering: The Eastern United States," J. E.
Beavers, editor, Ann Arbor Science Publishers, Ann
Arbor, Michigan, 1981, pp. 643-664.

36. Drnevich, V.P., "Undrained Cyclic Shear of Saturated


Sand," Journal of the Soil Mechanics and Foundations
Division, ASCE, Vol. 98, No. SM8 , August 1972, pp.
807-825.

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
475

37. Drnevich, V. P., "Resonant Column - Problems and


Solutions," Dynamic Geotechnical Testing, ASTM Spe­
O cial Technical Publication, STP 654, October 1978,
pp. 384-398.

38. Drnevich, V. P., "Evaluation of Sample Disturbance on


Soils Using the Concept of 'Reference Strain'1," Soil
Mechanics Series No. 26, College of Engineering,
University of Kentucky, May 1979.

39. Drnevich, V. P., "Recent Developments in Resonant


Column Testing," Richart Commemorative Lectures,
Edited by R. D. Woods, Proceedings of a Session spon­
sored by the Geotechnical Engineering Division in
conjunction with the ASCE Convention, Detroit, Michi­
gan, October 23, 1985, pp. 79-107.

40. Drnevich, V. P., Hardin, B. 0., and Shippy, D. J.,


"Modulus and Damping of Soils by the Resonant Column
Method," Dynamic Geotechnical Testing, ASTM Special
Technical Publication, STP 654, October 1978, pp.
91-125.

41. Drnevich, V. P. and Jent, J. P., "Response of


Saturated Sands to Cyclic Shear at Earthquake Ampli­
tudes," Research Report No. 87, Water Resources
Research Institute, University of Kentucky, Lexing­
C 42.
ton, Kentucky, October 1975.

Drnevich, V.P., and Richart, Jr., F.E., "Dynamic


Prestraining of Dry Sand," Journal of the Soil
Mechanics and Foundations Division, ASCE, Vol. 96,
No. SM2, March, 1970, pp. 453-469.

43. Dyvik, R. , Dobry, R . , Thomas, G. E . , and Pierce, W.


G., "Influence of Consolidation Shear Stresses and
Relative Density on Threshold Strain and Pore Pres­
sure During Cyclic Straining of Saturated Sand," U.S.
Army Corps of Enginers, Miscellaneous Paper GL-84-15,
July 1984.

44. Erguvanli, A., "Effect of Anisotropic Consolidation


on Liquefaction," Proceedings of the Seventh World
Conference on Earthquake Engineering, Istanbul, Tur­
key, September 8-13, 1980, Vol. 3, pp. 163-170.

45. Ferrito, J. M . , Forrest, J. B., and W u , G . , "A Compi­


lation of Cyclic Triaxial Liquefaction Test Data,"
Geotechnical Testing Journal, GTJODJ, Vol. 2, No. 2,
June 1979, pp. 106-113.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
476

46. Finn. W.D.L., "Liquefaction Potential: Developments


r \ Since 1976," Proceedings, International Conference on
Recent Advances in Soil Dynamics and Earthquake
Engineering, St. Louis, Missouri, May 1981, Vol. II,
pp. 655-681.

47. Finn, W.D.L., Bransby, P.L., and Pickering, D.J.,


"Effect of Strain History on Liquefaction of Sand,"
Journal of the Soil Mechanics and Foundations Divi­
sion, ASCE, Vol. 96, No. SM6 , November, 1970, pp.
1917-1934.

48. Finn, W.D.L., Byrne, P. M . , and Martin, G. R . ,


"Seismic Response and Liquefaction of Sands," Journal
of the Geotechnical Engineering Division, ASCE, Vol.
102, No. GT8 , August 1976, pp. 841-856.

49. Finn, W.D.L., Lee, K. W . , and Martin, G. R . , "An


Effective Stress Model for Liquefaction," Journal of
the Geotechnical Engineering Division, ASCE, Vol.
103, No. GT6 , June 1977, pp. 517-533.

50. Finn, W.D.L., Pickering, D. J., and Bransby, P. L.,


"Sand Liquefaction in Triaxial and Simple Shear
Tests," Journal of the Soil Mechanics and Foundations
Division, ASCE, Vol. 97, No. SM4, April 1971, pp.

c 51.
639-659.

Frydman, S., Melnik, J. and Baker, R . , "The Effect of


Prefreezing on the Strength and Deformation Proper­
ties of Granular Soils," Soils and Foundations,
JSSMFE, Vol. 19, No. 4, December 1979. pp. 31-42.

52. Fukushima, S. and Tatsuoka, F., "Deformation and


Strength of Sand in Torsional Simple Shear," Proceed­
ings of the IUTAM Conference on Deformation and
Failure of Granular Materials, Delft, August-
September 1982, pp. 371-379.

53. Gilbert, P. A., "Investigation of Density Variation


in Triaxial Test Specimens of Cohesionless Soil Sub­
jected to Cyclic and Monotonic Loading," Technical
Report GL-84-10, U.S. Army Engineer Waterways Experi­
ment Station, September 1984, 100 pp.

54. Hall, J. R . , Jr. and Richart, F. E., Jr., "Dissipa­


tion of Elastic Wave Energy in Granular Soils," Jour­
nal of the Soil Mechanics and Foundations Division,
ASCE, Vol. 89, No. SM6 , November 1963, pp. 27-56.

55. Hanzawa, H . , "Undrained Strength and Stability of a


Quick Sand," Soils and Foundations, JSSMFE, Vol. 20,
No. 2, June 1980, pp. 17-29.
r

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
477

56. Hardin, B. 0., "The Nature of Damping in Sands,"


Journal of the Soil Mechanics and Foundations Divi­
O sion, ASCE, Vol. 91, No. SMI, January 1965, pp. 63-
97.

57. Hardin, B. 0., "Suggested Methods of Test for Shear


Modulus and Damping of Soils by the Resonant Column,"
ASTM Special Technical Publication, STP 479, 1970,
pp. 516-529.

58. Hardin, B. 0., "The Nature of Stress— Strain Behavior


of Soil," Proceedings of the ASCE Geotechnical
Engineering Division Specialty Conference on Earth­
quake Engineering and Soil Dynamics, Pasadena, 1978,
Vol. 1, pp. 3-90.

59. Hardin, B. 0., and Drnevich, V.P., "Shear Modulus and


Damping in Soils: Measurement and Parameter
Effects," Journal of the Soil Mechanics and Founda­
tion Division, ASCE, Vol. 98, No. SM6 , June 1972, pp.
603-624.

60. Hardin, B. 0., and Drnevich, V. P., "Shear Modulus


and Damping in Soils: Design Equations and Curves,"
Journal of the Soil Mechanics and Foundations Divi­
sion, ASCE, Vol. 98, No. SM7, July, 1972, pp. 667-
692.

C' 61. Hardin, B. 0. and Richart, F. E., Jr., "Elastic Wave


Velocities in Granular Soils," Journal of the Soil
Mechanics and Foundations Division, ASCE, Vol. 89,
No. SMI, February 1963, pp. 33-65.

62. Hardin, B. 0. and Scott, G. D., "Generalized Kelvin-


Voigt Used in Soil Dynamics Study," Journal of the
Engineering Mechanics Division, ASCE, Vol. 92, No.
EMI, February 1966, pp. 143-156.

63. Healy, K. A., "The Dependence of Dilation in Sand on


Rate of Shear Strain," Thesis, presented to the Mas­
sachusetts Institute of Technology in partial ful­
fillment of the requirements for the degree of D.Sc.,
1963.

64. Heiniger, C. and Studer, J. A., "Resonant-Column


Apparatus for Coarse Grained Materials," Technical
Note, Geotechnical Testing Journal, GTJODJ, Vol. 8 ,
No. 3, September 1985, pp. 132-136.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
478

65. Hight, D. W . , Gens, A. and Symes, M. J . , "The


Development of a New Hollow Cylinder Apparatus for
Investigating the Effects of Principal Stress Rota­
tion in Soils," Geotechnique 33, No. 4, December
1983, pp. 355-383.

66. Hight, D. W . , Gens A., and Symes, M. J . , Correspon­


dence on "The Development of a New Hollow Cylinder
Apparatus for Investigating the Effects of Principle
Stress Rotation in Soils," and "Undrained Anisotropy
and Principle Stress Rotation in Saturated Sand,"
Authors' Reply, Geotechnique 35, No. 1, March 1985,
pp. 82-85.

67. Holtz, R. D. and Ko v a c s , W. D., An Introduction to


Geotechnical Engineering, Prentice-Hall, Inc., Eagle-
wood Cliffs, NJ, 1981.

68. Isenhower, W. M., "Torsional Simple Shear/Resonant


Column Properties of San Francisco Bay Mud," Thesis
presented to the University of Texas at Austin,
Texas, in partial fulfillment of the requirements for
the degree of Master of Science in Engineering, 1979.

69. Ishibashi, I., Kawamura, M., and Bhatia, S. K . , "Tor­


sional Simple Shear Apparatus for Drained and
Undrained Cyclic Testing," Advances in the Art of
Testing Soils under Cyclic Conditions, ASCE Fall Con­
vention, Detroit, Michigan, October 1985, pp. 51— 73.

70. Ishibashi, I., Kawamura, M., and Bhatia, S., "Effect


of Initial Shear on Cyclic Behavior of Sand," Journal
of Geotechnical Engineering, ASCE, Vol. Ill, No. 12,
December 1985, pp. 1395-1410.

71. Ishibashi, I. and Sherif, M. A., "Soil Liquefaction


by Torsional Simple Shear Device," Journal of the
Geotechnical Engineering Division, ASCE, Vol. 100,
No. GT8 , August 1974, pp. 871-887.

72. Ishibashi, I. Sherif, M. A., and Tsuchiya, C.,


"Pore— Pressure Rise Mechanism and Soil Liquefaction,"
Soils and Foundations, JSSMFE, Vol. 17, No. 2, June
1977, pp. 17-27.

73. Ishihara, K . , "Evaluation of Soil Properties for Use


in Earthquake Response Analysis," Proceedings of the
International Symposium on Numerical Models in
Geomechanics, Zurich, 13-17 September, 1982, pp.
237-259.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
479

74. Ishihara, K. , "Stability of Natural Deposits during


Earthquakes," Theme Lecture No. 7, Proceedings of the
o 11th International Conference on Soil Mechanics and
Foundation Engineering, San Francisco, California,
1985, Vol. 1, pp. 321-376.

75. Ishihara, K. and Li, S.-I., "Liquefaction of


Saturated Sand in Triaxial Torsion Shear Test," Soils
and Foundations, JSSMFE, Vol. 12, No. 2, June 1972,
pp. 19-39.

76. Ishihara, K. and Okada, S., "Yielding of Overconsoli­


dated Sand and Liquefaction Model under Cyclic
Stresses," Soils and Foundations, JSSMFE, Vol. 18,
No. 1, March 1978, pp. 57-72.

77. Ishihara, K. and Okada, S., "Effects of Stress His­


tory on Cyclic Behavior of Sand," Soils and Founda­
tions, JSSMFE, Vol. 18, No. 4, December 1978, pp.
31-45.

78. Ishihara, K. and Okada, S., "Effects of Large


Preshearing on Cyclic Behavior of Sand," Soils and
Foundations, JSSMFE, Vol. 22, No. 3, September 1982,
pp. 109-125.

79. Ishihara, K . , Silver, M. L., and Kitagawa, H.,

c "Cyclic Strengths of Undisturbed Sands Obtained by


Large Diameter Sampling," Soils and Foundations,
JSSMFE, Vol. 18, No. 4, December 1978, pp. 61-76.

80. Ishihara, K . , Silver, M. L., and Kitagawa, H.,


"Cyclic Strength of Sands Obtained by a Piston
Sampler," Soils and Foundations, JSSMFE, Vol. 19, No.
3, September 1979, pp. 61-76.

81. Ishihara, K. and Takatsu, H . , "Effects of Overconso­


lidation and K Conditions on the Liquefaction
Characteristics of Sands," Soils and Foundations,
JSSMFE, Vol. 19, No. 4, December 1979, pp. 60-68.

82. Ishihara, K . , Tatsuoka, F., and Yasuda, S.,


"Undrained Defo TmatIon and Liquefaction of Sand under
Cyclic Stresses," Soils and Foundations, JSSMFE, Vol.
15, No. 1, March 1975, pp. 29-44.

83. Ishihara, K. and Towhata, I., "Sand Response to


Cyclic Rotation of Principle Stress Directions as
Induced by Wave Loads," Soils and Foundations,
JSSMFE, Vol. 23, No. 4, December 1983, pp. 11-26.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
480

f 84. Ishihara, K. , Troncoso, J . , Kawase, Y., and


I Takanashi, Y., "Cyclic Strength Characteristics of
Tailing Materials," Soils and Foundations, JSSMFE,
Vol. 20, No. 4, December 1980, pp.. 127-142.

85. Ishihara, K. and Yasuda, S., "Sand Liquefaction Due


to Irregular Excitation," Soils and Foundations,
JSSMFE, Vol. 12, No. 4, December 1972, pp. 65-77.

86. Ishihara, K. and Yasuda, S., "Sand Liquefaction in


Hollow Cylinder Torsion under Irregular Excitation,"
Soils and Foundations, JSSMFE, Vol. 15, No. 1, March
1975, pp. 45-59.

87. Ishihara, K. and Watanabe, T., "Sand Liquefaction


through Volume Decrease Potential," Soils ad Founda­
tions, JSSMFE, Vol. 16, No. 4, December 1976, pp.
61-70.

88. Iwasaki, T., Tatsuoka, F. and Takagi, Y., "Shear


Moduli of Sands under Cyclic Torsional Shear Load­
ing," Soils and Foundations, JSSMFE, Vol. 18, No. 1,
March 1978, pp. 40-56.

89. Kokusho, T., "Cyclic Triaxial Test of Dynamic Soil


Properties for Wide Strain Range," Soils and Founda-
^ tions, JSSMFE, Vol. 20, No. 3, June 1980, pp. 46-60.

90. Kolbuszewski, J. J . , "An Experimental Study of the


Maximum and Minimum Porosities of Sands," Proceed­
ings, Second International Conference on Soil Mechan­
ics and Foundation Engineering, 1948, Vol. 1, pp.
158-165.

91. Kolbuszewski, J. J., "General Investigation of the


Fundamental Factors Controlling Loose Packing of
Sands," Proceedings, Second International Conference
on Soil Mechanics and Foundation Engineering, 1948,
Vol. 7, pp. 47-49.

92. Kolbuszewski, J. and Frederick, M. R . , "A Contribu­


tion Towards a Universal Specification of the Limit­
ing Porosities of a Granular Mass," Third European
Conference on Soil Mechanics and Foundation Engineer­
ing, Wiesbaden, 1963.

93. Ladd, C. C . , Foote, R., Ishihara, K . , Schlosser, F.


and Poulos, H. G . , "Stress-Deformaticn and Strength
Characteristics," State-of-the-Art Report, Proceed­
ings of the Ninth International Conference on Soil
Mechanics and Foundation Engineering, Tokyo, 1977,
Vol. 2, pp. 421-494.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
481

94. Ladd, R. S., "Specimen Preparation and Liquefaction


C } of Sands," Technical Note, Journal of the Geotechni­
cal Engineering Division, ASCE, Vol. 100, No. GT10,
October 1974, pp. 1180-1184.

95. Ladd R. S., "Specimen Preparation and Cyclic Stabil­


ity of Sands," Journal of the Geotechnical Engineer­
ing Division, ASCE, vol. 103, No. GT6 , June 1977, pp.
535-547.

96. Ladd, R. S., "Preparing Test Specimens Using Under-


Compaction," Geotechnical Testing Journal, ASTM, Vol.
1, No. 1, March 1978, pp. 16-23.

97. Lade, P. V., "Interpretation of Torsion Shear Tests


on Sand," Proceedings of the Second International
Conference on Numerical Methods in Geomechanics,
Blacksburg, Virginia, June 1976, Vol. 1, pp. 381-389.

98. Lade, P. V., "Prediction of Undrained Behavior of


Sand," Journal of the Geotechnical Engineering Divi­
sion, ASCE, Vol. 104, GT6 , June 1978, pp. 721-735.

99. Lade, P. V. and Duncan, J. M., "Stress-Path Dependent


Behavior of Cohesionless Soils," Journal of the
Geotechnical Engineering Division, ASCE, Vol. 102,
^ No. GT1, January 1976, pp. 51-68.

100. Lade, P. V. and Hernandez, S. B . , "Membrane Penetra­


tion Effects in Undrained Tests," Journal of the
Geotechnical Engineering Division, ASCE, Vol. 103,
No. G T 2 , February 1977, pp. 109-125.

101. Lambrechts, J. R. and Leonards, G. A., "Effects of


Stress History on Deformation of Sand," Journal of
the Geotechnical Engineering Division, ASCE, Vol.
104, No. G T 11, November 1978, pp. 1371-1387.

102. Lee, K. L. and Farhoomand, I., "Compressibility and


Crushing of Granular Soil in Anisotropic Triaxial
Compression," Canadian Geotechnical Journal, Vol. 4,
No. 1, February 1967, pp. 68-99.

103. Lee, K. L. and Seed, H. B., "Cyclic Stress Conditions


Causing Liquefaction of Sand," Journal of the Soil
Mechanics and Foundations Division, ASCE, Vol. 93,
No. SMI,January, 1967, pp. 47-70.

104. Lee, K. L. and Seed, H. B., "Dynamic Strength of


Anisotropically Consolidated Sand," Journal of the
Soil Mechanics and Foundations Division, ASCE, Vol.
93, No. SM5, September 1967, pp. 169-189.

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
482

o 105. Lee, K. L. and Seed, H. B., "Drained Strength Charac­


teristics of Sands," Journal of the Soil Mechanics
and Foundations Division, ASCE, Vol. 93, No. SM6 ,
November 1967, -pp. 117— 141.

106. Lee, K. L. and Seed, H. B., "Undrained Strength of


Anisotropically Consolidated Sand," Journal of the
Soil Mechanics and Foundations Division, ASCE, Vol.
96, No. SM2, March 1970, pp. 411-428.

107. Lee, K. L. ad Vernese, F. J . , "End Restraint Effects


on Cyclic Triaxial Strength of Sand," Journal of the
Geotechnical Engineering Division, ASCE, Vol. 104,
No. GT6 , June 1978, pp. 705-719.

108. Leonards, G. A., "Stability of Slopes in Soft Clays,"


Conference Special Lecture, Proceedings, of the Sixth
Panamerican Conference on Soil Mechanics and Founda­
tion Engineering, Lima, Peru, 1979, Vol. I, pp. 225-
274.

109. Leonards, G. A., Alarcon, A., Frost, J. D.,


Mohamedzein, Y. E., Santamarina, J. C., Thevanayagam,
S., Thomaz, J. E., and Tyree, J. L., "Dynamic Pene­
tration Resistance and the Prediction of the Compres­
sibility of a Fine-Grained Sand — A Laboratory
Study,” Discussion, submitted for publication to
G eo te chn iq ue , December 1985.

110. Lindenberg, J. and Koning, H. L . , "Critical Density


of Sand," Geotechnique 31, No. 2, June 1981, pp.
231-245.

111. McDonald, L. M. and Raymond, G. P., "Repetitive Load


Testing: Reversal or Rotation," Canadian Geotechnical
Journal, Vol. 21, Number 3, August 1984, pp. 456-
474.

112. Marr, W. A. and Hoeg, K . , "Stress-Strain Behavior


from Stress Path Tests," Norwegian Geotechnical
Institute Publication No. 125, Oslo, 1979.

113. Martin, G. R . , Finn, W.D.L., and Seed, H. B., "Funda­


mentals of Liquefaction under Cyclic Loading," Jour­
nal of the Geotechnical Engineering Division, ASCE,
Vol. 101, No. G T 5 , May 1975, pp. 423-438.

114. Martin, G. R . , Finn, W.D.L., and Seed, H. B.,


"Effects of System Compliance on Liquefaction Tests,"
Journal of the Geotechnical Engineering division,
Vol. 104, No. G T 4 , April 1978, pp. 463-479.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
483

115. Mayne, P. W. and Kulhawy, F. H., "Ko - OCR Relation­


ship in Soi 1, " Journal of the Geotechnical Engineer­
ing Division, ASCE, Vol. 108, No. GT6 , June 1982, pp.
851-872.

116. Mitchell, J. K . , Chatoian, J. M . , and Carpenter, G.


C., "The Influences of Sand Fabric on Liquefaction
Behavior," Contract Report S-76-5, College of
Engineering, University of California, Berkeley, June
1976.

117. Mitchell, J. K. and Solymar, Z. V., "Time-Dependent


Strength Gain in Freshly Deposited or Densified
Sand," Journal of Geotechnical Engineering, Vol. 110,
No. 11, November 1984, ASCE pp. 1559-1576.

118. Miura, S. and Toki, S., "A Sample Preparation Method


and Its Effect on Static and Cyclic Deformation-
Strength Properties of Sand," Soils and Foundations,
JSSMFE, Vol. 22, No. 1, March, 1982, pp. 61-77.

119. Miura, S. and Toki, S., "Anisotropy in Mechanical


Properties and Its Simulation of Sands Sampled from
Natural Deposits," Soils and Foundations, Vol. 24,
No. 3, September 1984, pp. 69-84.

120. Miura, S., Toki, S., and Tanizawa, F., "Cone Penetra­
tion Characteristics and its Correlation to Static
and Cyclic Deformation-Strength Behavior of Anisotro­
pic Sand," Soils and Foundations, JSSMFE, Vol. 24,
No. 2 June, 1984, pp. 58-74.

121. Mohamad, R. and Dobry, R., "Effect of Static Shear on


Resistance to Liquefaction," Discussion, Soils and
Foundations, JSSMFE, Vol. 23, No. 4, December 1983,
pp. 139-143.

122. Mori, K . , Seed, H. B., and Chan, C. K . , "Influence of


Sample Disturbance on Sand Response to Cyclic Load­
ing," Journal of the Geotechnical Engineering Divi­
sion, ASCE, Vol. 104, No. GT3, March 1978, pp. 323-
339.

123. Mulilis, J.P., Seed, H.B., Chan, C.K., Mitchell,


J.K., and Arulanandan, K . , "Effects of Sample
Preparation on Sand Liquefaction," Journal of the
Geotechnical Engineering Division, ASCE, Vol. 103,
No. GT2, February, 1977, pp. 91-107.

124. Nemat-Nasser, S. and Takahashi, K . , "Liquefaction and


Fabric of Sand," Journal of Geotechnical Engineering,
ASCE, Vol. 110, No. 9, September 1984, pp. 1291-1306.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
484

- 125. Oda, M., "Initial Fabrics and Their Relations to '


( Mechanical Properties of Granular Material," Soils
and Foundations, JSSFME, Vol. 12, No. 1, March 1972,
pp. 17-36.

126. Oda, M . , "The Mechanism of Fabric Changes during


Compressional Deformation of Sand," Soils and Founda­
tions, JSSMFE, Vol. 12, No. 2, June 1972, pp. 1-18.

127. Oda, M . , Nemat-Nasser, S., and Konisbi, J., "Stress-


Induced Anisotropy in Granular Masses," Soils and
Foundations, JSSMFE, Vol. 25, No. 3, September 1985,
pp. 85-97.

128. Park, !. K. and Silver, M. L., "Dynamic Triaxial and


Simple Shear Behavior of Sand," Journal of the
Geotechnical Engineering Division, ASCE, Vol. 101,
No. GT6 , June 1975, pp. 513-529.

129. Peacock, W. H. and Seed, H. B., "Sand Liquefaction


under Cyclic Loading Simple Shear Conditions," Jour­
nal of the Soil Mechanics and Foundations Division,
ASCE, Vol. 94, No. SM3, May 1968, pp. 689-708.

130. Peck, R. B., "Liquefaction Potential: Science Versus


Practice," Journal of the Geotechnical Engineering
Division, ASCE, Vol. 105, No. GT3, March 1979, pp.
393-398.
c
131. Poulos, S.J., "The Steady State of Deformation,"
Journal of the Geotechnical Engineering Division,
ASCE, Vol. 107, No. GT5, May, 1981, pp. 553-562.

132. Poulos, S. J . , Castro, G. and France, J. W.,


"Liquefaction Evaluation Procedure," Journal of
Geotechnical Engineering, Vol. Ill, No. 6 , June 1985,
pp. 772-792.

133. Prakash, S., Soil D y n a m i c s , McGraw-Hill Book Company,


New York, 1981, 426 pp.

134. Prange, B., "Parameters Affecting Damping Proper­


ties," Proceedings of DMSR 77, Karlsruhe, Vol. 1, 5-
16 September 1977, pp. 61-77.

135. Rad, N. S. and Clough, G. W . , "New Procedure for


Saturating Sand Specimens," Journal of Geotechnical
Engineering, ASCE, Vol. 110, No. 9, September 1984,
pp. 1205-1218.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
485

136. Richart, Jr., F. E., "Field and Laboratory Measure­


ments of Dynamic Soil Properties," Proceedings of
o DMSR 77, Karlsruhe, Vol. 1, September 1977, pp. 3-35.

137. Roscoe, K. H . , Schofield, A. N., and Wroth, C. P. "On


the Yielding of Soils," Geotechnique, Vol. 8 , No. 1,
March 1958, pp. 22-52.

138. Saada, A. S., Fries, G . , and Ker, C.-C., "Stress


Induced in Short Cylinders Subjected to Axial Defor­
mation and Lateral Pressures," Soils and Foundations,
JSSFME, Vol. 23, No. 1, March 1983, pp. 114-118.

139. Saada, A. S., Fries, G . , and Ker, C.-C., "An Evalua­


tion of Laboratory Testing Techniques in Soil Mechan­
ics," Soils and Foundations, JSSFME, Vol. 23, No. 2,
June 1983, pp. 98-112.

140. Saada, A. S. and Townsend, F. C., "State of the Art


Laboratory Strength Testing of Soils," Laboratory
Shear Strength of Soil, ASTM, STP 740, R. N. Yong and
F. C. Townsend, Editors, American Society for Testing
and Materials, 1981, pp. 7-77.

141. Schimming, B. B., Haas, H. J . , and Saxe, H. C.,


"Study of Dynamic and Static Failure Envelopes,"
Journal of the Soil Mechanics and Foundations Divi­
sion, ASCE, Vol. 92, No. SM2, March 1966, pp. 105-
124.

142. Schmertmann, J. H . , "Stress Ratio Effects on Shear


Modulus of Dry Sands," Discussion, Journal of
Geotechnical Engineering, ASCE, Vol. Ill, No. 9, Sep­
tember 1985, pp. 1153-1154.

143. Schneider, H . , "Undisturbed Sampling of Saturated


Cohesionless Soils by Chemical Impregnation," Ph.D.
Dissertation, Purdue University, West Lafayette, IN,
December 1985.

144. Schofield, A. and Wroth, P., Critical State Soil


M e c h a n i c s, McGraw-HiJi, New York, 310 pp.

145. Seed, H. B., "Soil Liquefaction and Cyclic Mobility


Evaluation for Level Ground during Earthquakes,"
Journal of the Geotechnical Engineering Division,
ASCE, Vol. 105, No. G T 2 , February 1979, pp. 201-255.

146. Seed, H. B. and Idriss, I. M., "Soil Moduli and Damp­


ing Factors for Dynamic Response Analyses," Report
No. EERC 70-10, University of California, Berkeley,
1970.

r
V

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
147. Seed, H. B., Idriss, I. M . , and Arango, I., "Evalua­
tion of Liquefaction Potential Using Field Perfor­
mance Data," Journal of Geotechnical Engineering,
ASCE, Vol. 109, No. 3, March 1983, pp. 458-482.

148. Seed, H. B. and Lee, K. L., "Liquefaction of


Saturated Sands during Cyclic Loading," Journal of
the Soil Mechanics and Foundations Division, ASCE,
Vol. 92, No. SM6 , Proceedings Paper 4972, Nov. 1966,
pp. 105-134.

149. Seed, H. B. and Lee, K. L., "Undrained Strength


Characteristics of Cohesionless Soils," Journal of
the Soil Mechanics and Foundations Division, ASCE,
Vol. 93, No. SM6 , November 1967, pp. 333-360.

150. Seed, H. B. and Lundgren, R . , "Investigation of the


Effect of Transient Loading on the Strength and
Deformation Characteristics of Saturated Sands,"
Proceedings, ASTM, Vol. 54, 1954, pp. 1288-1306.

151. Seed, H. B., Mori, K . , and Chan, C. K . , "Influence of


Seismic History on Liquefaction of Sands," Journal of
the Geotechnical Engineering Division, ASCE, Vol.
103, No. GT 4 , April, 1977, pp. 257-270.

152. Seed, H. B. and Peacock, W. H . , "Test Procedures for


Measuring Soil Liquefaction Characteristics," Journal
of the Soil Mechanics and Foundations Division, ASCE,
Vol. 97, No. SM8 , August 1971, pp. 1099-1119.

153. Seed, H. S., Tokimatsu, K . , Harder, L. F., and Chung,


R. M., "The Influence of SPT Procedures in Soil
Liquefaction Resistance Evaluations," Report No.
UCB/EERC-84/15, Earthquake Engineering Research
Center, University of California, Berkeley, October
1984.

154. Seed, H. B., Wong, R. T., Idriss, I. M . , and Tok­


imatsu, K. , "Moduli and Damping Factors for Dynamic
Analyses of Cohesionless Soils," Report No.
UC B/E ERC-84/14, Earthquake Engineering Research
Center, University of California, Berkeley, September
1984.

155. Selig, E. T. and Chang, C. S., "Soil Failure Modes in


Undrained Cyclic Loading," Journal of the Geotechni­
cal Engineering Division, ASCE, Vol. 107, No. GT5,
May 1981, pp. 539-551.

with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
487

156s Shannon, W. L., Yamane, G., and Dietrich, R. J.,


"Dynamic Triaxial Tests on Sands," Proceedings of the
c First Pan American Conference on Soil Mechanics and
Foundation Engineering, Mexico City, 1959, Vol. 1,
pp. 473-489.

157. Shibata, T., Yukitomo, H., and Miyoshi, M.,


"Liquefaction Process of Sand During Cyclic Loading,"
Soils and Foundations, JSSMFE, Vol. 12, No. 1, March
1972, pp. 2-16.

158. Silver, M. L., "Laboratory Triaxial Testing Pro­


cedures to Determine the Cyclic Strength of Soil,"
Report No. NUREG-31, U.S. Nuclear Regulatory Commis­
sion, Washington, D.C., December 1976, 70 pp.

159. Silver, M. L . , Chan, C. K . , Ladd, R. S., Lee, K. L.,


Tiedemann, D. A., Townsend, F. C., Valera, J. E., and
Wilson, J. H . , "Cyclic Triaxial Strength of Standard
Test Sand," Journal of the Geotechnical Engineering
Division, ASCE, Vol. 102, No. G T 5 , May 1976, pp.
511-523.

160. Silver, M. L. and Park, T. K . , "Testing Procedure


Effects on Dynamic Soil Behavior," Journal of the
Geotechnical Engineering Division, ASCE, Vol. 101,

c 161.
No. GT10, October 1975,pp. 1061-1083.

Silver, M. L. and Park, T. K . ,"Liquefaction Poten­


tial Evaluated from Cyclic Strain-Controlled Proper­
ties Tests on Sands," Soils and Foundations, JSSMFE,
Vol. 16, No. 3, September 1976, pp. 51-65.

162. Silver, M. L. and Seed, H. B., "Deformation Charac­


teristics of Sands under Cyclic Loading," Journal of
the Soil Mechanics and Foundations Division, ASCE,
Vol. 97, No. SM8 ,August 1971, pp. 1081-1098.

163. Silver, M. L. and Seed, H. B., "Volume Changes in


Sands during Cyclic Loading," Journal of the Soil
Mechanics and Foundations Division, ASCE, Vol. 97,
No. SM9, September 1971, pp. 1171-1182.

164. Skoglund, G. R . , Marcusson III, W. F., and Cunny, R.


W., "Evaluation of Resonant Column Test Devices,"
Journal of the Geotechnical Engineering Division,
ASCE, Vol. 102, No. GT11, November 1976, pp. 1147-
1158.

165. Sladen, J. A., D'Hollander, R. D . , and Krahn, J.,


"The Liquefaction of Sands, A Collapse Surface
Approach," Canadian Geotechnical Journal, Vol. 22,
No. 4, November 1985, pp. 564-578.

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
488

166. Soil Dynamics Instruments, Inc., "Quasi-Static Tor­


sional Simple Shear/Resonant Column Apparatus,"
Operation Manual Apparatus Serial No. 252, Built for
the Geotechnical Engineering Laboratory, School of
Civil Engineering, Purdue University, West Lafayette,
Indiana, July 1983.

167. Susuky, T., and Toki, S., "Effects of Preshearing on


Liquefaction characteristics of Saturated Sand Sub­
jected to Cyclic Loading," Soils and Foundations,
JSSMFE, Vol. 24, No. 2, June, 1984, pp. 16-28.

168. Symes, M. J. and Burland, J. B., "Determination of


Local Displacements on Soil Samples," Geotechnical
Testing Journal, GTJODJ, Vol. 7, No. 2, June 1984,
pp. 49-59.

169. Symes, M.P.R., Gens, A. and Hight, D. W., "Undrained


Anisotropy and Principal Stress Rotation in Saturated
Sand," Geotechnique 34, No. 1, March, 1984, pp. 1-27.

170. Tatsuoka, F., "Stress Ratio Effects on Shear Modulus


of Dry Sands," Discussion, Journal of Geotechnical
Engineering, Vol. Ill, No. 9, September 1985, pp.
1155-1157.

171. Tatsuoka, F., Iwasaki, T. and Takagi, Y., "Hysteretic


Damping of sands under Cyclic Loading and its Rela­
tion to Shear Modulus," Soils and Foundations,
JSSMFE, Vol. 18, No. 2, June 1978, pp. 26-40.

172. Tatsuoka, F., Iwasaki, T., Fukushima, S., and Sudo,


H., "Stress Conditions and Stress Histories Affecting
Shear Modulus and Damping of Sand under Cyclic Load­
ing," Soils and Foundations, JSSMFE, Vol. 19, No. 2’,
June, 1979, pp. 29-43.

173. Tatsuoka, F., Muramatsu, M. and Sasaki, T . , "Cyclic


Undrained Stress-Strain Behavior of Dense Sands by
Torsional Simple Shear Test," Soils and Foundation,
JSSMFE, Vol. 22, No. 2, June 1982, pp. 55-70.

174. Tatsuoka, F., Muramatsu, M. and Sasaki, T., "Cyclic


Undrained Stress-Strain Behavior of Dense Sands by
Torsional Simple Shear Test," Closure, Soils and
Foundations, JSSMFE, Vol. 23, No. 3, September 1983,
pp. 142-145.

175. Tatsuoka, F., Ochi, K . , end F u j i i , S., "Effect of


Sample Preparation Method on Cyclic Undrained
Strength of Sand in Triaxial and Torsional Shear
Tests," Bulletin of Earthquake Resistant Structure
Research Center, No. 17, March 1984, pp. 29— 61.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
489

176. Tatsuoka, F. and Silver, M. L., "New Method for the


Calibration of the Inertia of Resonant Column Dev­
C ices," Geotechnical Testing Journal, GTJODJ, Vol. 3,
No. 1, March 1980, pp. 30-34.

177. Tiedemann, D. A., Kaufman, L. P., and Rosenfield, J.,


"Determining Dynamic Properties for Embankment Dams
from Laboratory Testing," U.S. Department of the
Interior, Bureau of Reclamation, Engineering and
Research Center, Report No. REC-ERC-34-17, December
1984, 34 pp.

178. Tobita, Y . , "Effects of Fabric Change on Deformation


Characteristics of Sand in Cyclic Shearing," Advances
in the Mechanics and *the Flew cf Granular Materials,
Mohsen Shahinpoor, Editor, Trans. Tech. Publications,
1983, Vol. II, pp. 885-901.

179. Tokheim, 0., "Deformation Behavior of Soils in Terms


of Soil Moduli," The Third Laurits Bjerrum Lecture
presented 27th February, 1978, Norwegian Geotechnical
Institute, Publication No. 152, pp. 1-19.

180. Tokimatsu, K. and Yoshimi, "Criteria of Soil


Liquefaction with SPT and Fines Content," Proceed­
ings, 8th World Conference on Earthquake Engineering,
San Francisco, July 1984, Vol. 3, pp. 255-262.

c 181. Towhata, I. and Ishihara, K . , "Response of Sand in


Cyclic Torsional Loading Influenced by Rotation of
Principle Stress Directions," Advances in the Mechan­
ics and the Flow of Granular Materials, Mohsen
Shahinpoor, Editor, Trans Tech Publications, 1983,
Vol. II, pp. 903-928.

182. Towhata, I. and Ishihara, K . , "Undrained Strength of


Sand Undergoing Cyclic Rotation of Principle Stress
Axes," Soils and Foundations, JSSMFE, Vol. 25, No. 2,
June 1985, pp. 135-147.

183. Towhata, I, and Ishihara, K . , "Shear Work and Pore


Water Pressure in Undrained Shear," Soils and Founda­
tions, JSSMFE, Vol. 25, No. 3, September 1985, pp.
73-84.

184. Vaid, Y. P. and Chern, J. C., "Effect of Static Shear


on Resistance to Liquefaction," Soils and Founda­
tions, JSSMFE, Vol. 23, No. 1, March, 1983, pp. 47-
60.

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
490

185. Vaid, Y. P. and Chern, J. C., "Effects of Static


Shear on Resistance to Liquefaction," Closure, Soils
and Foundations, JSSMFE, Vol. 25, No. 3, September
1985, pp. 154-156.

186. Vaid, Y. P. and Chern, J. C., "Cyclic and Monotonic


Undrained Response of Saturated Sands," Advances in
the Art of Testing Soils under Cyclic Conditions,
Session 52, ASCE Annual Convention and Exposition,
Detroit, October 21-25, 1985, pp. 120-147.

187. Vaid, Y. P. and Finn, W.D.L., "Static Shear and


Liquefaction Potential," Journal of the Geotechnical
Engineering Division, ASCE, Vol. 105, No. GT10,
October 1979, pp. 1233-1246.

188. Vaid, Y. P. and Negussey, D., "A Critical Assessment


of Membrane Penetration in the Triaxial Test,"
Geotechnical Testing Journal, GTJODJ, Vol. 7, No. 2,
June 1984, pp. 70-76.

189. Vaid, Y. P. and Negussey, D., "Relative Density of


Pluviated Sand Samples," Technical Note, Soils and
Foundations, JSSMFE, Vol. 24, No. 2, June 1984, pp.
101-105.

190. Wang, J.— N. and Kavazanjian Jr., E. "Pore Water Pres­


sure Development in Non-Uniform Cyclic Triaxial
Tests," Report No. 73, The John A. Blume Earthquake
Engineering Center, Department of Civil Engineering,
Stanford University, April 1985, 109 pp.

191. Woods, R. D., "Parameters Affecting Elastic Proper­


ties," Proceedings of DMSR 77, Karlsrune, Vol. 1, 5-
16 September 1977, pp. 37-59.

192. Woods, R. D., "Measurement of Dynamic Soil Proper­


ties," Proceedings, ASCE Specialty Conference on
Earthquake Engineering and Soil Dynamics, Pasadena,
California, June 1978, Vol. 1, pp. 91-178.

193. Yoshimi, Y. and Oh-oka, H . , "Influence of Degree of


Shear Stress Reversal on the Liquefaction Potential
of Saturated Sand," Soils and Foundations, JSSMFE,
Vol. 15, No. 3, September 1975, pp. 27-40.

194. Yoshimi, Y . , Tokimatsu, K . , Kaneko, 0., and Makihara,


Y., "Undrained Cyclic Shear Strength of a Dense Nii­
gata Sand," Soils and Foundations, JSSMFE, Vol. 24,
No. 4, December 1984, pp. 131-145.

r\

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
491

195. Youd, T. L . , "Compaction of Sands by Repeated Shear


Straining, " Journal of the Soil Mechanics and Foun­
o dations Division, ASCE, Vol. 98, No. S M 7 , July 1972,
pp. 709-725.

196. Youd, T. L. , "Packing Changes and Liquefaction Sus­


ceptibility," Technical Note, Journal of the
Geotechnical Engineering Division, ASCE, Vol. 103,
No. GT8 , August 1977, pp. 918-S23.

197. Youd, J.L. and Craven, T. N . , "Lateral Stress in


Sands During Cyclic Loading," Technical Note, Journal
of the Geotechnical Engineering Division, ASCE, Vol.
101, No. GT2, February 1975, pp. 217— 221.

198. Yu, P. and Richart, Jr., F. E . , "Stress Ratio Effects


on Shear Modulus of Dry Sands," Journal of Geotechni­
cal Engineering, Vol. 110, No. 3, March 1984, ASCE,
pp. 331-345.

199. Yu, P. and Richart, Jr., F. E . , "Stress Ratio Effects


on Shear Modulus of Dry Sands," Closure, Journal of
Geotechnical Engineering, Vol. Ill, No. 9, September,
1985, pp. 1157-1159.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
c APPENDICES

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
492

Appendix A

Data Reduction Procedures for Resonant Column Tests

The principle of the resonant column test is to

determine the dynamic properties of sands (and of other

materials) from observation of the steady— state response

of a cylindrical specimen subjected to forced vibrations.

The measurement of properties of a material in the range

of dynamic loading is complicated by the fact that inertia

forces must be considered. It is clear that in order to

excite and measure the vibration of a system, it is

altered by the attachment of the vibration excitation dev­

ice and the required transducers (54,56). The problem is

further complicated for soils since the specimen must usu­

ally be encased in a membrane which needs to be sealed '

against the end platens by using O-rings. Consequently,

the observed vibration is not the response of the soil

specimen alone but of the system composed of the specimen

and its attached apparatus. In order to obtain an analyti­

cal solution for the system, it is necessary to assume

some model for the soil material, so that the response may

be interpreted in terms of material constants (G and D ) ,

which could then be used with boundary conditions other

than those existing in test specimens.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
493

In the new resonant column/torsional shear apparatus,


o the Hardin type electromagnetic oscillator applies tor­

sional vibrations to the top of the specimen. For small

amplitude vibrations of generally high frequency, the

response of sand specimens is essentially elastic. Hence,

the theory of propagation of shear waves in elastic rods

can be used to interpret the resonant column tests

results. However, as the amplitude of vibration

increases, the stress— strain characteristics of sands

become nonlinear and inelastic. Accordingly, the theory

of wave propagation alone is not sufficient to model the

stress— strain response, and the energy dissipating proper­

ties of the soil should be accounted for. Analytical

solutions for both small and large amplitude vibrations


c and the corresponding data reduction procedures are dis­

cussed herein.

Torsional Vibrations in a Rod

Consider a cylindrical rod subjected to torsional

vibrations (Fig. A.la). The shear modulus and unit weight

of the material which constitutes the rod are G and t ,

respectively. At a distance x from the origin of coordi­

nates, the rod is subjected to a torque T, which

corresponds to an angle of rotation 6. The torque at a

section located at a distance x + Ax will be given by T +

(3T/3x)Ax and the corresponding rotation by 0 t (30/3x)Ax.


(

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
X ■H AXH

e>T
T+ --- AX
) £>x

AX

c a)

Twisted zone
T acting for At

Wave front
ax = Vs At

b)

Figure A.l Torsional Vibrations of Elastic Rods.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
495

Applying Newton's second law for torsional motion, the


G equation for the systems becomes,

2
-T + (T +-|^Ax) = pi Ax — | (A.la)
3t2

or

_
ai =
T a2e
pi —
,. .
(A.lb)

where p * Yt /g is the density of the material in the rod

and I is the polar area moment of inertia of the cross

section of the rod. The torque is related to the angle of

rotation, 6 , by the expression,

C T = GI | | (A.2)

Substituting E q . A.2 into E q . A.lb and differentiating

gives

32 6 G a2 6
(A.3a)
at2 p ax2

or

a2 e 2 a2 e . .
5- = V 5- (A.3b)
atz s 3x

where Vg = \|G/p is the velocity of the torsional waves.

Eq. A. 3 constitutes the wave equation for torsional vibra­

f tions in an elastic rod.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
496

The fact that the term \|G/p corresponds to the velo­


c city of propagation of the shear waves, Vg , can be demon­

strated by using the principle of conservation of momen­

tum. Consider that a torque T is applied at the end of

the rod for a small time interval At at time t * 0 (Fig.

A.lb) This torque will twist a small zone of the rod, and

the torsional wave will be transmitted to successive zones

as time increases. If the torsional wave is transmitted

at a velocity V , the length of the twisted zone in the


s
rod, Ax, will be given by Vg At and the corresponding angle

of rotation, A 6, by (T/GI) Ax. Consequently, A 6 at the

end of the time interval At is given by

A 6 - (T/GI) V At (A.4a)
s
c
or

TV
A6 ’s TT /,
it * — * VP ( A -41>>

where V is the rotational velocity (or particle velocity)


P
in radians per unit time. On the other hand, from momen­

tum equilibrium it follows that

TAt = pIAxVp (A.5)

Substituting E q . A.4b into E q . A.5 and replacing Ax by

V g At gives

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
497

o or
P
(A.6a)

vs = \|g77 (A.6b)

Analytical Solution for Small Amplitude Vibrations

The Hardin oscillator applies torsional vibrations

that vary sinusoidally with time. It Is well known

(28>133) that for steady-state forced vibrations with a

sinusoidal forcing function, the solution of the wave

equation (Eqs. A.3) can be assumed of the form

0(x,t) « (Aj sin(wt) + A2cos(wt) ) ( B 1sin— + B^os— ) (A.7)


s s

where w is the circular frequency of the applied vibra-

tions and A, jA - j B, and B, are constants which depend on


1 Z 1>
% to
the boundary conditions of the problem under considera- "

tion.

The theoretical model used for the resonant column

system and the boundary conditions in the new apparatus

are shown schematically in Figure A.2. It is noted that

for small amplitude vibrations, damping in either the soil

specimen (internal damping) or in the oscillator (external

damping) is not considered. The base of the specimen is

attached rigidly to a pedestal with large rotational iner-

tia (bottom platen) so that it is essentially fixed.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
A 98

(Schem atic).
Vi

System
Column
c

Resonant
A .2
Figure

Cl -3 O □ CO

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
499

Consequently, 0(0,t) equals zero at any time t. In order


o to satisfy this boundary condition, the coefficient in

Eq. A. 7 should be also zero. It follows that,

0(x,t) - (AjSinCwt) + A 2cos(wt) ^ ^ s i n (A.8 )


s

As mentioned earlier, the Hardin oscillator applies a

torsional vibration to the top of the specimen. The

oscillator is essentially a single-degree-of-freedom sys­

tem with negligible damping, in which a spool-shaped

"spring" couples a central mass (consisting of four solid

arms) and specimen top platen attachments to a hollow

cylindrical mass with very large rotational inertia. A

c sine wave generator capable of producing a sinusoidal

current is used to power the coils and permanent magnets

of the oscillator, which are wired in such a way that the

electromagnetic force produces a torque about the axis of

the specimen. The large inertia mass of the oscillator

essentially provides a fixed reaction for the spring such

that the driving torque produces vibration of the central

mass and specimen top platen system.

The rigid mass with rotational inertia (Fig. A. 2)

represents the top platen including a porous stone, the

0-ring seal, part of the membrane, the four-arm central

mass, and an acceleration transducer, which monitors the

response of the system at the top of the specimen. All

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
500

these elements are connected to the top platen in such a


O fashion that they are considered part of the platen and

move as a rigid body for the full range of frequency of

interest. The coils and permanent magnets are designed

such that there is no contact between the moving and sta­

tionary parts.

In order to determine the boundary conditions at the

top of the specimen (x=L), the rigid mass can be con­

sidered as a free body as shown in Figure A.2b. The equa­

tion of motion for this rigid mass without apparatus damp­

ing becomes

•N 2Q
J. (L,t) + K. S(L,t) + T(L,t) = T sin(wt) (A.9)

C 31

where JA is the polar mass moment of inertia of the


°

equivalent rigid mass, is the torsional spring constant

corresponding to the oscillator spring, T(L,t) is the

torque transmitted to the specimen, and T q is the ampli­

tude of the sinusoidal applied torque. Substitution of

Eq. A.2 into E q . A.9 gives

2
JA^ - | (L ,t) + K A 0( L ,t) + GI-|| (L ,t ) = TQ sin(wt) (A.10)
31

where G and I are the snear modulus of the soil and the

polar area moment of inertia of the specimen, respec­

tively .
f
K
V

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
501

The solution governing the natural frequency of the


O system under torsional vibration can be obtained by solv­

ing Eq. A. 10 for the case of free vibrations. From E q . A .8

and after differentiation it follows that

2
2— j (L ,t) = -BjW2 sin ^ [AjSin(wt) + A 2 cos(wt)] (A.11)
dt S

0 (L,t) = B^sin ^ [A^sinCwt) + A^cosCwt)] (A.12)


s

30 w tjT,
Y ? (L,t) = Bj — cos — [AjSinCwt) + A2 cos(wt)] (A.13)
s s

Substituting Eqs. A . 11 to A . 13 into E q . A . 10 for the case

of free vibration yields

c
[JA w 2 - _
Ka ] sin —wL = GI— w cos —WL
r T . . _ - r /■ 14)
A 1 / \
(A.
s s s

2 2
By replacing KA by wA JA , G by Vg p, and I by J/pL, Eq.

A . 14 becomes

J y
A ,, , A N 2, wL . wL . /»ic\
J— [ 1 - (— ) ] Y ~ tan “ = (A.15)
s s

where J is the polar mass moment of inertia of the speci­

men, and wA is the circular natural frequency for tor­

sional vibration corresponding to the one-degree-of-

freedom system composed by the rigid mass of inertia JA

and the oscillator spring C^A ). Finally, by defining

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
502

o and
( A . 16)

Equation A . 15 can be simplified to

T F TAN F - 1.0 (A.18)

where T is termed active end stiffness/inertia factor and

F is a dimensionless frequency factor.

Equation A . 18 constitutes the characteristic equation

for free vibrations of the system. This equation would

yield the natural frequency of vibration provided all the

other terms are known. However, the shear modulus, G, and

hence the shear wave velocity, Vg , are in general unknown.

At resonance under forced vibrations, the applied

frequency, w, corresponds to the natural frequency of the

system for torsional vibrations, from which the shear

modulus of the specimen can be calculated. With a known

value of T (Eq. A . 16), the value of F can be determined

either graphically or by trial and error from E q . A . 18.

Finally, in accordance with E q . A . 17, once the system

resonant frequency, w, and the value of F are known, the

shear modulus of the soil can be obtained as

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
503

G « V2 p * (f^)2 P “ 4*2L 2 p (f/F)2 (A.19)


O
where f is the natural frequency of the- system in cycles

per second.

Measurements of System Vibrations

A piezoelectric accelerometer monitors the response

of the system at the top of the specimen. The accelerome­

ter is mounted horizontally oriented on one of the arms of

the vibrating central mass at a distance of 0.0349 meters

from the axis of rotation. This distance is termed the

transducer lever arm, TLA (37). The accelerometer pro­

duces an electrical charge which is proportional to

c acceleration. This accelerometer was provided by Columbia

Research Laboratories with the calibrations that are

traceable to the National Bureau of Standards. This cali­

bration is in the form of a charge sensitivity in peak

picocoulombs per peak "g" of acceleration. The value of

this calibration factor for this accelerometer (S/N 8 88)

is 27.0 pk -pc mb/ pk- g.

A charge amplifier is used to convert the electrical

charge to voltage and then it is read with conventional

voltmeters or oscilloscopes. The charge amplifier is also

a Columbia Research Laboratory (model 4102) . The

accelerometer charge sensitivity is set with a locking

dial on the unit and the output voltage of the charge

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
o amplifier becomes proportional to acceleration.

put is 2.5, 0.25 or 0.025 pk-Volt/pk-g depending on


The out­

whether the left, center or right button on the front

panel is depressed. The left button is used for readings

below 5 Volts. For larger voltages, the center and right

buttons should be used and the recorded voltages multi­

plied by 10 and 100, respectively, before using the cali­

bration factor given below.

Since the vibrations of the system are sinusoidal,

displacement is related to acceleration by

2
Displacement = Acceleration/(2vf) (A.20)

where f is the frequency of the applied vibration (Hz).


C
The output from the charge amplifier is usually read

on a voltmeter calibrated to give root-mean-square values

of voltage Thus the value of the displacement

calibration factor, DCF, is given by

DCF=\ j2 [9.80 m/sec 2 /g] / [ (2 .5 V/g)(2Trf)2 rad/sec2 ] (A.21a)

DCF = 0.1404/f2 (meters/Volt rms ) (A.21b)

Hence, the rotational calibration factor, RCF,* is given

by

1. Most of the terminology is compatible with ASTM D

r 4015-81.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
505

RCF - DCF/( 0 .0349meter) (A.22a)


( J

RCF - 4.02/f2 (Radians/Volt ) (A.22b)


r ms

Eq. A.22b indicates that for acceleration transducers the

vibration frequency should be included as a term in the

calibration factor. Consequently, the peak rotation

(radians) at the top of the specimen can be obtained as

6 (L ,t) - (RCF)(R T O ) (A.23)

where RTO is the rotational (acceleration) transducer out­

put (Volts ).
r rms

In the case of torsional vibrations, the shear strain

c in each cross section of a hollow cylinder specimen varies

from a minimum at the inner perimeter to a maximum at the

outer perimeter of the specimen. For small rotations, the

shear strain, Y, at a distance r from the axis of rotation

is given by

Y(r) = j- r (A.24)

where 0 is the rotation angle in the cross section at a

height 1 from the bottom of the specimen.

Accordingly, the average shear strain, Ya v » can be

expressed as

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
506

O „
av “
JA T(r)da
JdA
2 6 , ro ~ rl .
3 1 ^2 _ 1 (A.2a)
o ~ i

where r and r. are the outer and inner radii of the


o i
specimen, respectively. Thus, the average shear strain

for each cross section occurs at a radius equal to

r 3 - r3
rav - I i 4 7 4 l <A '2 6 >
o i

In the case of a solid cylinder specimen, r^ = 0, hence

the average shear strain occurs at a radius equal to

(2/3)rQ . In terms of the outer and inner diameters, OD

and ID, E q . A.25 becomes,

c
^avJ I I l°?Y ~ IP2 ] CA *27)
OD - OD

Substitution of E q . A.23 into E q . A.27 yields the expres­

sion for the average shear strain at the top of the speci-

ui 0 XX y

V. u,t) - 1 (RCF)(RTO) , OD3 - ID3 , (A>28)


aV OD 2 - ID2

For solid cylindrical specimens, E q . A.28 simplifies to

Yav (L ,t) = i (RCF)(RTO)(D/L) (A.29)

where D and L are the diameter and length of the specimen

respectively.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
507

Measurement of System Resonant Frequency


O
The frequency of excitation is adjusted to produce

resonance of the system composed of the vibration excita­

tion device, the specimen and its attached platens. The

system resonant frequency should correspond to the mode of

vibration of the specimen for which the only node isat

the fixed end of the specimen (57). This will be the

lowest system resonant frequency for the model in Figure

A . 2. At this resonant frequency, the mode shape of this

system is practically linear, that is the rotations

changes linearly from zero at the base to maximum at the

top of the specimen (36,192). The system will experience

resonance at various higher frequencies. However, as the

c damping of the system increases more of the higher reso­

nances are damped out. The tendency for the higher reso­

nances to disappear increases as the ratio increases.

This system approaches a single-degree-of-freedom system-

as JA /*J becomes very large (56).

The resonance condition is established by observing

the Lissajous figures formed on the screen of an X-Y

oscilloscope by the torque and acceleration signals. It

is well known that when two sinusoidal voltages are

applied to the vertical and horizontal (Y and X) channels

of an oscilloscope, the figure traced on the screen

depends on the frequency and phase difference of the two

f signals. Figure A. 3 shows the patterns obtained from two

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
508

/ .
\J SCREEN
Elliptical
V ERTIC AL INPUT

PATTERN

H O R I Z O N T A L INPUT

a) T h e tw o signals h a v e th e s a m e fr e q u e n c y

SCREEN-

c VERTIC AL INPUT
Butterfly
PATTERN

H O R IZO N TA L INPUT 12

b) T h e f r e q u e n c y o f t h e v e r t i c a l i n p u t is
t w ic e th a t o f th e h o riz o n ta l input

Figure A.3 Lissajous Figures from Two Sine Waves 90


Out of Phase.

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
509

sinusoidal waves 90° out of phase. When the two sine

waves are of the same frequency, an elliptical patterns

appears on the screen. For a perfectly circular pattern

the two signals should be of the same amp li tud e. When the

vertical frequency is twice that of the horizontal fre­

quency, a butterfly pattern, or a figure eight on its

side, appears on the screen.

The driving current and thus the input torque are

known by measuring the voltage drop across a 5 8 precision

(temperature and frequency stable) power resistor in

series with the driving coils. Consequently, this voltage

will be in phase with the forcing torque function applied

to the system. The voltage proportional to the driving

torque is applied to the horizontal amplifier of the

oscilloscope whereas the output from the accelerometer,

after being converted to voltage by the charge amplifier,

is applied to the vertical amplifier of the oscilloscope-.

As for the case of a damped single-degree— of-freedom sys­

tem, the sinusoidal excitation torque and the acceleration

signals are 90° out of phase when the applied frequency

corresponds to the undamped natural frequency of the sys­

tem (38,39,176). Resonance of the system can be esta­

blished by adjusting the frequency of excitation to pro­

duce an ellipse with vertical and horizontal axes

(38,40,57). The undamped natural frequency of the system

is determined by recording the frequency associated with a

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
510

90° phase difference between the driving torque and the


C /
rotational acceleration of the top mass. When the damping

of the system is negligible, this frequency is essentially

equal to the frequency that produces maximum amplitude of

displacement, velocity or acceleration of the vibration

end of the specimen (39,57).

Determination of Oscillator Constants

In order to calculate the shear modulus of the speci­

men using Eqs. A.16 to A. 19, in addition to the system

resonant frequency, it is necessary to know the natural

frequency of vibration of the torsional motion, wA , and

the polar mass moment of inertia, , corresponding to the

top rigid mass (Fig. A.2). This rigid mass and the oscil­

lator spring constitutes a single-degree-of-freedom sys­

tem. As mentioned earlier, during a test, the top platen,

porous stone, 0-rings, etc., are rigidly attached to the

moving parts of the oscillator and hence their polar mass

moment of inertial must be considered as part of the

apparatus inertia, J^.

The natural frequency of vibration of the oscillator

top platen system, f ^ , can be determined by operating the

Hardin oscillator at resonance with the top platen, 0-

rings, etc. attached to it as if a test were to be per­

formed but not having a specimen in place. In the new

apparatus the value of w^ varies according to the specimen

size and type (-T^.ble A.l).

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
511

Determination of the rotational inertia, J^, is com­

plicated by the fact that all the components do not have

simple geometry. Consequently, it is very difficult to

determine the value of precisely from physical measure­

ments and from values of the density of the mass. Thus,

indirect procedures have to be used for determining J^.

Some of these procedures are discussed below.

According to Drnevich et al. (40) a calibration rod

of known torsional stiffness (K ,) can be coupled to the


r od
platens in place of the specimen as shown in Figure A.4a.

With the calibration rod in place, the resonant frequency,

w ., of the system composed of oscillator, calibration


r od
rod and top platen components can be determined. Neglect­
r ing the rotational inertia of the calibration rod, the

natural frequency of the system with two springs in series

4e
— ~ ft-w’o n Kxr
~J

K J + Ka
W rod A (A.30)
rod J.
A

2
Replacing by w^ in E q . A.30, the polar mass moment

of inertia of the top platen system, J^, can be obtained

as

ja 2 V * r (A*31)
tfrod - fI ]

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
512

Top platen system


O Ka
JA

Calibration rod
with Known
P, G, I, Krod

a) Drnevich’s Method

Ka Ka

J2

Calibration rod
with unknown

b) Tatsuoka and Silver's Method


Figure A.4 Methods for Determining Top Platen System
Rotational Inertia.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
513

, A second alternative procedure Is to couple the metal

calibration rod to the platens in place of the specimen

and then use Eqs. A . 16 to A . 19 to back figure the top pla­

ten system inertia, , from the known shear modulus of

the rod (40).

It is clear that in the two procedures just

described, the accuracy of the value of is based on the

accuracy of the values of J and 6 determined for the metal

calibration rod. Any determination of a value of J is

very sensitive to minor error in measuring the small diam­

eter of the calibration rod. Furthermore, in some cases

it is not easy to determine a shear modulus value for the

calibration rod material.

c Tatsuoka and Silver (176) have presented a new method

in which the value of JA can be determined without having

to know the material properties of the calibration rod.

As shown in Figure A.4b, two calibration rods are required

which should be essentially identical except for the

dimensions of a mass, which is added at the top. The two

rods should preferably be machined from the same piece of

metal stock. However, the top portions of the two cali­

bration rods should have different sizes but simple

geometry. Consequently, it is easy to determine the

values of and for the top portions of the calibra­

tion rods from their dimensions and densities. After

determining the resonant frequency of the system with each


C
\

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
514

one of the calibration rods in place, and f ^ > the rota­


c tional inertia of the top platen system can be obtained as

(176)

J2 - 4 f > 2 J 1
J. - s ------ (A.32)
12
(T1 )2 - 1
2

It should be noted that in this calibration method it is

not necessary to know the values of J, L, and G for the

thin calibration rod. In using this calibration tech­

nique, the ratio Jj/J ^ should be chosen so that values of

f^ and f2 are not close together (176).

c Once w^ and

ten systems
are known for each one of the top pla­

(hollow and solid cylinder specimens), the

torsional oscillator spring constant, K A , may be calcu-


A
lated from

kA = (2irfa )2 JA (A.33)

where is expressed in N-m/rad. Calculation of the

value of K.a> which should be unique for different top pla­

ten systems, provides a check on the accuracy of the cali­

bration procedures. The corresponding values of

, f^, and obtained for the new apparatus are summar­

ized in Table A.l according to the specimen size. For

determining , a steel calibration rod with a diameter of


c

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
515

12.7 mm and 190.5 mm long was used.

Two additional calibration factors for the oscillator

are also required. They are the torque calibration fac­

tor, T C F , and the oscillator (apparatus) damping coeffi­

cient, A D C ^ . The purpose of these two factors is illus­

trated in the next section.

The torque calibration factor, TCF, can be determined

by operating the oscillator at frequencies usually between

0.5 and 2 times the undamped natural frequency. For exam­

ple, frequencies equal to (\|2/ 2 ) f^ and \ [2 f^ are typi­

cally used. With the oscillator operating at these fre­

quencies, the corresponding acceleration reading (RTO),

and the torque reading (CR) are obtained. The torque

reading, CR, corresponds to voltage drop across the 5fl

resistor associated with the current applied to the tor­

sional coils.

It is well known (28,133) that for a single-degree-

of-freedom system with no viscous damping, the dynamic

amplification factor under forced vibrations is given by

6 = ---- 9 (A.34)
i - (-^r
n

where f and f are the frequency of the driving force and

the natural frequency of the system, respectively. It is

seen that for f/f = \ 12/ 2 , 8 equal 2.0 whereas for

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
516

c f/fA “ \ !2 the magnification factor

becomes equal to 1.0 .


(absolute value)

If the torque is applied statically to the vibrating

mass of the oscillator, the angular displacement, 6


s
corresponds to

Q Dynamic Rotation (RCF)(RT0)


s " -------- 3--------- =------3----- (A.35)

Accordingly, a relationship between the corresponding

static rotation and the voltage reading associated with

the current applied to the coils can be established. For

the case f = \ 12/2 fA , a constant is defined as

* (R C F ) (R T O ) -

c ° 1 ----- 2 CR. <A '3 6 >

whereas if the applied frequency is \|2 f^, the

corresponding constant becomes

* (RCF)(RTO)_
C2 STT <A ‘3 7 >

The values of Cj and should in general agree

within 10%.

Consequently, the torque calibration factor, TCF, can

be expressed as

* Consideration-of the apparatus damping can be given in

r the evaluation of the dynamic amplification factor.

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
517

o where TCF is expressed in (N-m)/volt as indicated in


( A . 38)

rms
Table A.I.

It follows that the torque applied by the oscillator

to the vibrating central mass at any stage of vibration

can finally be obtained as

T - (TCF)(CR) (A.39)

where the torque T is obtained in (N-m).

Although very low, there is some quantity of damping

associated with the spool shaped spring in the oscillator,

which is made of high strength aluminum. This damping is

referred to as apparatus damping. The corresponding

apparatus damping coefficient for torsional motion ADC^

can be determined by recording the amplitude decay curve

under free vibration of the apparatus. The apparatus

damping coefficient relates to the apparatus damping

ratio, Da , by the expression.

ADC.
A
D (A.40)
A 2(K a /Ja >

With the oscillator and the top platen attached and

the system vibrating at resonance, the power to the exci­

tation device is cut off and the resulting free vibration


r

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
Reproduced with permission of the copyright owner. Further reproduction

Table A,1 Calibration Factors for Hardin Oscillator.

Symbol Units Specimen


Solid Hollow Hollow

Outer diameter mm 71.1 71.1 101.6

Inner diameter mm 38.1 71.1

Accelerometer
Lever Arm m 0.0349 0.0349 0.0349

RCF1 Pk-Rad/V 4.02/f2 4.02/f2 4.02/f2


rms
N-m/Rad 420 420 420
ka
prohibited without perm ission.

Hz 90.5 90.8 62.0


fA
L.
kg-m 2 1 .3xlO~ 3 1 .29xl0"3 2 .77x10**3
JA
2
ADC. (kg-m )/sec 4.71x10-3 5.37xl0" 3 4.69xl0**3
A
TCF (N-m )
'
/Vr m s 1.498xl0-2 1 .498xl0“ 2 1 ,498x10-2

1. Based on use of a charge amplifier with output of 2.5 peak


Volts per peak "g" of acceleration; f = frequency of oscillation
in Hz
519

decay curve Is stored on the screen of the oscilloscope.

A photograph of the stored trace can be taken and the log­

arithmic decrement, 5^, calculated from

6A " (1/n) ln (A1/An + 1 ) (A.41)

where n is the number of decay cycles, A^ is the peak to

peak amplitude for the first cycle after power is cut off,

and Afl+j is the peak to peak amplitude for the n+1 cycle

after the power is cut off. The value of the apparatus

damping coefficient can be obtained from

ADC a = 2 fA JA SA (A.42)

c All the information on the calibration factors for

the new apparatus is summarized in Table A.l according to

the specimen size. It is noted that the apparatus damping

coefficient changes slightly as a function of the top pla­

ten system size but it is indeed very low. The calibra­

tion information for the three different specimen sizes

confirms the accuracy of the calibration procedures.

Analytical Solution for Large Amplitude Vibrations

As the amplitude of vibration increases, the stress-

strain characteristics of sands becomes nonlinear and ine­

lastic, yielding a histeretic closed loop in the shear

stress-shear strain space when subjected to steady-state

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
520

^ ^ vibrations (40). Consequently, the energy dissipating

properties of the soil specimen should be accounted for in

the equation for torsional motion of the system shown in

Figure A.2. It is also appropriate to consider the con­

tribution of the apparatus damping in the response of the

system.

The general steady-state solution for forced tor­

sional vibrations of cylindrical sand specimens has been

obtained by using viscous damping theory, that is, by

assuming that a Kelvin-Voigt model represents the response

of the material (56,62). Equivalent shear modulus, G, and

damping ratio, D, for the soil specimen are defined as the

elastic modulus and the viscous damping ratio of a uni-

^ form, linear visco-elastic (Kelvin-Voigt model) specimen

of the same mass density and dimensions as the soil speci­

men necessary to produce a resonant column having the

measured system resonant frequency and response due to a.

vibratory torque input (40,52). In other words, a model

for the system is assumed which includes parameters to

describe the end platen conditions, the specimen and the

excitation device. By using the analytical solution for

the system model, the unknown parameters (G and D) are

assumed and varied until the motion of the model matches

the motion measured in the test under a given torsional

vibration (39).

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
521

It is important to point out that the energy dissi­


O pating characteristics of sands are not viscous in nature

but they are due to the nonlinear and inelastic stress-

straln characteristics of sands. Some of the applied

energy is consumed in producing permanent deformation and

displacement of the soil grains, or transformed into

energy of sound, heat, etc. Histeretic damping in soils

apparently is strain rate independent. The effects of

viscous damping, which is strain rate dependent seem to be

mino r for granular materials (171). Nevertheless, obser­

vation of the decay of free vibrations on specimens of

granular materials suggest that the damping capacity of

sands behaves like viscous damping (54,56). Consequently,

the Kelvin-Voigt model has been considered to be a useful

tool for describing the behavior of sands subjected to

forced vibrations over a wide frequency range (39,56,62).

Since the resistance to torsional deformation is the

sum of "elastic" and "viscous" components, the shear

stress in any cross section of the cylindrical specimen

can be obtained as (56,62)

T (A.43)

where x^ and x^ are the elastic and viscous components,

respectively.

r (A.44)

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
2
where p is the shear coefficient of viscosity (N - s/s ).

Hence,

2
x(r,x,t) = Gr-|| + pr -^"at (A.46)

The torque T(x,t), acting on a section of the speci­

m en at time, t, can be obtained by integrating E q . A.46.

Thus,

c T( x ,t ) = /A rxdA = GI i | + Pi-glgl

Substituting E q . A . 47 into E q . A.lb and differentiating


(A.47)

yields E q . A.48, which is the differential equation

governing torsional vibrations of the system shown in Fig­

ure A.2,

iff , £ l!| + £ -41- (A.48)


3t P 3x P 3x 3t

The shear coefficient of viscosity, p, is related to the

shear damping ratio, D, by the expression (40)

D = I (l ^ ) (A.49)

in which w is the circular frequency of the driving


r

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
523

torque. The shear damping ratio, D, will also correspond

to the area of the stress-strain histeresis loop divided

by 4 it times the elastic energy stored in the specimen at

maximum strain amplitude (40,57,171).

The steady-state solution for E q . A . 48 takes the form

0(x,t) ■■ u(x)cos(wt) + v(x)sin(wt) (A.50)

where u(x) and v(x) are also functions of the specimen

properties, which include four arbitrary constants. The

expressions for u(x) and v(x) are presented elsewhere

(39,56,62). The four arbitrary constants are determined

according to the boundary conditions of the specimen-

oscillator system (Fig. A.2) as illustrated in the analyt­

ical solution for small amplitude vibrations.

However, it is important to note that in order to

take apparatus damping into account, a new term has to be

included in the equation of motion for the rigid mass at

the top of the specimen. Thus, E q . A.9 becomes

•(L ,t ) + ADC,||(L,t) + (A.51)

K . 6( L ,t ) + T(L,t) = T sin(wt)
A O

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
5 24

o Substituting E q . A.47 into E q . 51 results in

(A.52)

Gl||(L,t) + pi

Because the response of the system is measured at the

top of the specimen, the analytical solution is usually

given as 0(L,t) « A^sin(wt — > in which A^ is the ampli­

tude of rotation at the top of the specimen and <j>f is the

phase difference between the applied torque and the

resulting angular displacement at the top of the specimen.

Finally, to render the response in a dimensionless

form, a resonance factor can be defined as

A
f
R (A.53)
•f 0
s

in which 0g is the static angular displacement which is

given by

Figure A.5a shows a typical plot which illustrates the

relationship between the resonant factor, R^ , and the fre­

quency of the driving torque, w, for a case in which

= 1.0 and no spring is attached to the top platen system

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
525

o
ka = o
o 0.20 V a iu e of
co
**— 0.40
CD 0.01
O
c 0.02
2 0.04
o
(0
ffi
cc
0.02

2 3 4
1T 0jL
V a i u e of F * = ----------—
2 Vs

a) Theoretical R e s p o n s e
100
Theoretical response for « 0.025 - 2D
29.13 psi
7.40 psi > Experimental response, (pressure)

c o
CO
«*-*

<r
10
J * / J = 0 .3 2 2

o ka =0
c
cc
c
o
CO
CD
CC
0.1

0.01
1 2 3 4 5
, IT C liL
V a lu e of F = — ------
2 Vs

b) M e a s u r e d R e s p o n s e of a S p e c i m e n
of D r y O t t a w a 2 0 - 3 0 s a n d

Figure A. 5 Steady-State Forced Vibrations of Fixed Base


Cylindrical Specimen (After Ref. 56).

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
526

(K a «0). The applied frequency is expressed in terms of a


O dimensionless frequency factor, F* , which is given by

The occurrence of various resonances is clearly evi­

dent. However, as the damping increases, more of the

higher resonances are damped out. Figure A.5b compares

the theoretical and experimental steady— state response of

a specimen of dry Ottawa 20-30 sand, with a void ratio of

0.54 subjected to two different confining pressures (56).

The value of J./J was 0.322. The values of F* and


A
pw/G(=2D) for the theoretical solution were chosen such

c that there was good agreement between the theoretical and

experimental results at the first resonance. The

correspondence between the curves is good given the errors

which may exist in the measurements, confirming the vali­

dity of the Kelvin-Voigt model. Consequently, it is con­

sidered that the Kelvin-Voigt constitutive model is ade­

quate for obtaining modulus and damping values under

sinusoidal excitation irrespective of the shear strain

amplitude and damping (39). It is noted that the theoret­

ical curve with pw/G constant agrees well with the

results. This means that the shear coefficient of viscos­

ity, y , decreases as the frequency increases such that the

ratio pw/G remains constant with frequency (56).

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
527

The general solution for the system in Figure A.2 is


o included in the computer program contained in ASTM D

4015-81 (3). Therefore, this program is useful for reduc­

ing data for the Hardin-type oscillator which employs a

spring attached to the top platen system and with some

quantity of damping associated with it. Consequently, the

program uses the apparatus calibration factors, specimen

measurements and measured vibration data (that is, system

resonant frequency, f, amplitude of rotation, RTO, and

torque reading, CR) to calculate iteratively all pertinent

information about the specimen including shear modulus, G,

shear damping ratio, D, and average shear strain ampli­

tude, y • The program is written in ANSI FORTRAN and all


av

c units in

sionless
the program are SI.

system frequency factor,


The unknowns are the dimen-

F= (wL/V ),
s
and the

specimen damping ratio, D. The program uses an Iterative

scheme to establish the value of F associated with the

undamped natural frequency of the system by setting the

phase angle equal to 90° at the active end of the speci­

men. At the same time a second iterative scheme is used

to adjust the specimen damping ratio, D, until the calcu­

lated displacements of the model match the displacements

measured during the test. It is important to note that

the data reduction procedure in the program is based on a

phase difference of 90° and not on maximum amplitude dis­

placements. Accordingly, the "resonant" frequency of the

f system should correspond to the frequency which produces a

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
528

vertical ellipse on the screen of the oscilloscope,


O irrespective of the shear strain amplitude or damping

ratio. It is more efficient from the calculation point of

view to equal the phase difference equation to 90° than to

establish the F and D values corresponding to the maximum

points of the frequency response curves (39). As the

damping of the system increases, the difference between

the frequency producing maximum amplitude of rotation and

the frequency corresponding to a phase difference of 90°

between the torque and acceleration signals increases.

The latter corresponds to the undamped natural frequency

of the system.

The analytical solution for the Hardin oscillator—

c specimen system indicates that the system dimensionless

frequency factor, F, is relatively independent of damping

in either the apparatus or the specimen for small damping

values (176). With the new Hardin oscillator, dynamic

properties of sand specimens are determined at low shear

strain amplitudes usually ranging from 10 % to


-2
1.5 x 10 %. Consequently, for most cases the damping of

the system is low enough and close approximation may be

obtained by using the frequency equation corresponding to

the "elastic" solution of the system (Eq. A . 18), which

assumes that there is no damping any where in the system

and that one end of the specimen is perfectly fixed.

Accordingly, the shear modulus values can be calculated

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
529

from E q . A . 19. Thus

G = p (2irL)2 (f/F)2 (A.56)

where the values of the dimenslonless frequency factors

are obtained by use of Figure A. 6 . This figure gives

values of F as a function of the active end

inertia/stiffness factor, T,j, (Eq. A . 16), for cases in

which the apparatus damping coefficient is essentially

zero and the specimen damping ratio, D, is estimated to be

less than 10%.

Hardin (57) considers that the assumption of small

damping can be assumed to be valid when the damping of the

system including apparatus and soil damping is such that

the energy dissipated during one cycle of vibration is

less than 50% of the maximum strain energy stored in the

system at maximum displacement. This provision shall be

considered satisfied if the logarithmic decrement of the

system, as determined in the next section, is less than

0.4, or if the ratio of the displacement amplitude of

vibration at the system resonant frequency is five times

the displacement at a frequency equal to one half of the

system resonant frequency for a fixed excitation current.

From the foregoing discussion, it can be seen that

the damping characteristics of the soil specimen can be

indirectly determined from the measured steady-state

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
530

0.5

F
0.2

0.1

D < 0.1
0.05 ADC - 0
0.03
10 . 20. 50. 100 . 20 0 . 500. 1000.
Tt

c a)

3.
2.

1.

0.5 ADC » 0
0.3
0.1 0.2 0.5 1. 2. 5. 10 .

b)

Figure A .6 Dimensionless Frequency Factor for Resonant


Column Apparatus with Base of Specimen
Fixed (After Ref. 3).

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
531

response with due consideration to damping and strain


O energy stored in the apparatus. For a given system (J^/J),

the resonance factor is essentially a function of the

specimen damping. Since the oscillator is calibrated so

that the value of the forcing function, T(L,t), is known,

the dynamic magnification factor may be determined from

the measured amplitude of vibration at resonance as fol­

lows

HMF - ifza - l 1 (A.57)


s

Substituting E q . A.22b into E q . A . 57 results in the sim­

plified expression,

C MMF = [ ( 1 5 8 . 7 0 3 X RTO) / (TCF) ( CR) ] J (A.58)

Shear damping ratio, D, is given as part of the output if

the ASTM D 4015-81 computer program is used. Otherwise ,-

shear damping ratio values may be calculated from the

equation (3,40)

D(Z) = ^ (^r) (A.59)

In which A is an amplification coefficient. For the new

apparatus, approximate values of A are given as a function

of TT , in Figure A.7. N o t e sthat if the values of are

greater than about 2, the value of A is approximately

twice Tx .

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
2000.

1000.

500. ADC

200.

a d 100.

50.

G 20.

10.

5.

2.

1 .
0.1 0.2 0.5 1. 2. 5. 10. 20. 50. 100.200.500.1000.
Ty

Figure A.7 Damping Amplification Coefficient for


Resonant Column Apparatus with Base of
Specimen Fixed (After Ref. 3).

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
533

Determination of Specimen Damping Ratio from Free Vibrations

The specimen damping characteristics can also be

determined from observation of the amplitude decay of free

vibrations of the system, as was indicated in determining

the apparatus damping. With the system, vibrating at

resonance, the power is shut off and the logarithmic

decrement of the system, 6 , determined from the resulting


s
amplitude decay curve. The observed 6 is a property of
s
the system, including the specimen and its attached

apparatus. This procedure is theoretically exact only if

the bottom end (passive end) can be assumed to be rigidly

fixed (40).

Since in the Hardin oscillator the active end is res­

trained by a spring with some damping, a system energy

ratio, S, must be calculated as follows (40)

(A.60)

Finally, the specimen damping ratio, D, can be calculated

from (40),

D(%) = 100 [6g (l+s) - S 6a ] / 2 * (A.61)

where 6^ is the apparatus logarithmic decrement.

Nevertheless, it is considered that the method of

free vibration decay for measuring damping is less


r“>\

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
534

desirable than the steady-state vibration method (3 9 ).


C
Besides being an added step in the testing procedure, it

does not allow full consideration of apparatus stiffness

and damping. Also for large damping, there are only two

or three measurable peaks, which impedes an accurate

determination of the system logarithmic decrement.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
Appendix B

Data Reduction Procedures for Torsional Shear Tests

In the new resonant column/torsional shear test

apparatus a solid or hollow cylindrical specimen can be

subjected to a simultaneous application of a confining

cell pressure, a vertical compressional force and a tor­

sional force. The axial force is applied by using a dou­

ble acting air cylinder mounted on a frame on top of the

chamber lid whereas the torsional force is applied to the

specimen by a torque motor system which consists of a

stepper motor and a gear reduction box with reduction

ratios of 200 to 1 and 3200 to 1. Gears on the reduction

box engage with gears mounted on the shaft of the Hardin

oscillator, causing the entire oscillator and the upper

platen to rotate with it.

Stress/strain components and distributions, measure­

ment system characteristics and calibrations as well as

the corresponding data reduction procedures for torsional

shear tests are discussed below.

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
536

Stress and Strain Components In Torsional Shear Test Specimens


(
Stress Components

Figure B.l shows the stress components on hollow tor­

sional shear test specimens in terms of cylindrical coor­

dinates. It is evident that the stress condition in tor­

sional shear test specimens is axially symmetric and that

six independent stresses must be specified to completely

define the state of stress within hollow (or solid)

cylindrical specimens and consequently to analyze the

results of torsional shear tests. o and o correspond to


Z IT
the stresses associated with the vertical compressional

force and the confining cell pressure, respectively. The

applied torque develops shear stresses on horizontal


c planes, Tz 0> and complementary shear stresses develop

automatically on vertical radial planes, T 0Z » Since the

confining pressure acts through flexible membranes there

are no shear stresses on vertical circumferential sur­

faces. Accordingly, r and t equal zero and the radial


rz zr
stress, , is a principal stress, provided there are no

end effects. It is noted that the only unknown stress is

the circumferential stress, O q , which is the normal stress

in the direction of particle shear deformation. This

stress must be estimated for analyzing the test results.

The relationship between the circumferential stress,

o Q , and the applied boundary stresses can be established

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
537

T ze
f
z

Oz
Tzi

Tez

O 'r

a) A pplied B o u n d a r y b) S t r e s s e s on an

S tresses E le m e n t in th e Wall

c
CE- a n g le b e tw e e n
CT\ d ire c tio n
T Pole
an d z a x is
ze

OS

9Z

c) M o h r 's C ircle D ia g ra m

Figure B -1 Stress Components in Hollow Torsional Shear


Test Specimens.

R eproduced w ith perm ission o f the copyright owner. F urth er reproduction prohibited w itho ut perm ission.
538

o by considering force equilibrium along the radial direc­

tion, as illustrated in Figure B.2. The corresponding

equation for radial equilibrium becomes

°e - ° r = r -sir (B-2)
Since inthe new resonant column/torsional shear apparatus

there is no gradient of radial stress, , across the

cylinder wall (inner and outer pressures are the same),

the circumferential stress, o 0, will be equal to o^. The

same conclusion is reached from plasticity theory (198).

It is important to note that if end effects develop a new

term should be included in the equation for radial equili­

brium as discussed in a later section. On the other hand,

c for an apparatus in which different pressures can be

applied to the inner and outer cells of a hollow cylinder,

a gradient of radial stress, o ^ , across the cylinder wall

will be established and the circumferential stress will

then be different from o ^ , as shown by equation B.l.

Once oz , Oq, T2 0 , and r Qz are known they can be

resolved to determine the magnitude and orientation of the

principal stresses, as illustrated in Figure B.l. It is

noted that the radial stress, o^ , corresponds to the

intermediate principal stress, o^ . The remaining princi­

pal stresses, and , which act in the circumferential

plane, are given by


A
{

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
539

dr

c
Note: a ® is in de p e n d en t of ■&

(err) (rde ) + 2 c- 9- (dr d<t) = (^ r + Ser r ^ ) (r+ dr) d-e-

cr$ _ o*r .= r 6o~r


d r

Figure B.2 Radial Equilibrium in Torsional Shear


Test Specimens.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
540

o /
°z + °e .
° i ------- 2 .

°z " °e>2 . _2
+ \ (-=“ 2--- > + T20 ( B .2a)

°z - °0 °2 " ° 6,2 . _2
a
\ ( 2--- > + T0z (B ,2b)

3 2

and the orientation of major principal stress, 0j, with

respect to the vertical axis can be obtained as

2 To
tan 2a = --- =-2- (B.3)
°z " 0

Equation B.3 reflects one of the main features of the tor­

sional shear test. That is, the radial, axial and tor­

sional shear stresses can be changed as desired to obtain

any orientation of the planes of principal stress. For

c example, if the radial stress, o ^ , is kept constant and

the axial and torsional stresses are changed so that their

ratio, t q/Ao remains constant, it follows that


* z0 z

“ ■* <2“> - , + I / 6- o - 2 - K < B -4 >


r z r z

indicating that a remains constant and that the cell pres­

sure acting equally in all directions has no effect on the

direction of the major principal stress. Alternatively, a

test may be desired in which the mean confining stress,


*
a , is to remain constant. In this case, the cell pres-
hi

sure is decreased and the axial stress increased in the

ratio of 2 to 1 such that

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
541

2T>fl 4T, fl
( i tan ( 2 0 -------------------------------------------(b .5)
V / . , Z, 2
O + 4 0 - ( o r— )
r z ' r 2 '

Note that in this case Ac. - — Ac /2. However, the orien-


2 z *
tation of the principal stresses can be chosen to be the

same as in the test at constant radial stress, c^.

Figure B.3 shows some of the multiple stress paths

and the corresponding orientations of o^ that can be

applied on torsional shear stress specimens and depicts

the enormous capabilities of this test. As indicated in

Figure B.3d, even a continuous rotation of the principal

stresses can be produced by changing the applied boundary

stresses in such a way that the deviator stress ( 0^ - c^)

c remains constant. From Eqs. B.2a and B.2b it follows that

<°1 - °3> J °z * V , 2 . 2
2 \ ^ 2---- > + z9 <B -6 >

Continuous rotation of principal stress directions appears

to have an important role in properly simulating the

stress paths which are followed e.g. by elements beneath

ocean structures (83,181,182).

Equation B.6 implies that two components of the max­

imum shear stress may exist within the specimen:

1. (oz - C q )/2 is the shear stress component associated

with the triaxial mode and

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
54 2

o a = angle b e t w e e n
a -.j direction a n d
z axis
zB
tan 2 a = 2 T ZS
-°e

2a

c a) Torsional loading
Isotropic consolidation
b) Torsional loading
Anisotropic consolidation

zB zB

c) C o n s t a n t a d) C o n s t a n t
deviator stress,

Figure B.3 Potential Loading Schemes for Cyclic


Torsional Shear Tests.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
543

2. t Q is Che shear stress related to the torsional mode


Z o
of deformation.

It is obvious that automatic controls are needed to

change the axial, torsional and cell pressures simultane­

ously with fixed ratios. This capability is not available

yet in the new resonant column/torsional shear test

apparatus. Only the torsional loading is automatically

controlled. Accordingly, only stress paths a_ and b^ in

Figure 6.3 can be applied. However, computer controlled

solenoid values can be readily used to enhance the capa­

bilities .

It is of interest to consider the effect of the

intermediate principal stress, in the light of tor­

sional shear tests. Conventionally, the contribution of

the intermediate principal stress to the strength-

deformation characteristics of soils has been analyzed in

terms of the parameter b which is given by the expression

(9)

b (B .7 )

Accordingly, in the axial compression triaxial test b=0

whereas in the axial extension triaxial test b=1.0. This

fact has been used to explain differences in behavior

between the two tests. From Eqs. B.2a and B.2b and

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
544

^ considering it follows that the parameter b

corresponding to the torsional shear tests is given by the

expression

(B.8)

M < °Z 2 °r > 2 + Tz 9

On the other hand, from Figure B.l it is observed that

a - a
z r
cos 2ct = — ■■ (B.9)

Substituting E q . B.9 into E q . B.8 yields

c
b - 1 - £?■?— 2-° (B .10)

2
Since cos 2a = 1 - sin a, the parameter b can be obtained

as

b = sin2 a (B.ll)

Consequently, there is a unique relationship between the

parameter b and the orientation of the principal stresses.

Changes in , and q may imply a change in b (and a)

and hence, on the magnitude of the intermediate principal

stress, <?2 , relative to and . It is to be noted that

equation B.ll holds for the case in which the pressure in

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
545

f , the inner and outer cells of a hollow cylinder specimen

are the same. When this is not the case, the parameter b

is a function of both the orientation of the principal

stress, a, and the pressure gradient across the wall of

the cylinder. This fact has been considered by some

investigators hoping to separate the effects of o^ from

the effects of a (65,66). In principle, by changing a and

the pressure gradient across the wall it is possible to

keep the parameter b constant, and hence the relative mag­

nitude of . Unfortunately, a gradient of radial stress

across the wall implies additional non-uniformities in the

stress distributions. These non-uniformities are dis­

cussed in a later section. On the other hand, it is not

clear whether the parameter b is enough to account for the


c effect of the intermediate principal stress (140). If the

two hypothetical tests described earlier are considered,

namely; torsional shear tests with a) constant radial

stress, o ,and b) constant mean confining stress, a , it


r m
is possible to change , a and q in such a way that

the two tests have the same a (and hence the same b).

However, in the first case o. = c remains constant


* 2 r
whereas in the latter case a^ is continuously decreased

(Ac, = -Ac /2). Consequently, in the author's opinion,


6 Z
torsional shear tests on hollow specimens under different

inner and outer cell pressures may not be as much of an

improvement as claimed elsewhere (65,66).

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
546

Strain Components

It is seen that if any circumferential sheet within

the cylinder (see Figure B.4) were unwound to a plane, the

deformation imposed on such an elemental cylinder by

twisting the top of the specimen which is rigidly fixed at

its base, would be similar to simple shear but without the

problems associated with the lack of complementary

stresses ever present in the simple shear test (5,140).

Consequently, the strain components in hollow torsional

shear specimens in terms of cylindrical coordinates are

the shear strain in the circumferential plane T2 q » the

axial strain e , the radial strain e and the circumferen-


Z IT

tial strain eQ . y and y Q are equal to zero,


o zr ro

c It is evident again that the strain (as with the

stress) condition is axially symmetric, that is, indepen­

dent of 0 (198). In the absence of end restrain effects,

it is also independent of z. It is noted that in any case

the shear strain, Yz g^ is theoretically the same at all

cross sections along the height of the specimen. Minor

deviations from these "uniform" patterns have been attri­

buted to the effect of the weight of the specimen under

low levels of confining stresses (174).

The radial strain, , and the circumferential

strain, e Q , are given by (65,69)

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
547

o
1.

:jjj_

ir

a) Distorsion of a s p e c i m e n d u e to torsion

li
c
A = dime n s i o n

b) Distorsion o n a circunferencial piane


at a distance r f r om axis.

Figure B.4 Strain Components in Hollow Torsional


Shear Test Specimens.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
548

3u
(B.12)
O 3r

and

_u
■0 r

where u is the displacement of a point in the radial

direction. It is noted that if the behavior of the soil

were linear elastic the radial and circumferential strains

would be equal. Otherwise, eand 6g are different even

if o^ and Og are uniformly distributed across the wall of

the cylinder.

In addition to principal stress components, principal

strain components are defined as follow (181,182)

c
ez + e 0 1 (B .14a)
2 \ < £z - V 2 + Yz0

(B.14b)

ez + £0 1
2 \
(e - ea) + r2z 0 (B .14c)

The torsional shear tests has been successfully used

to simulate plane strain conditions (75,81,173,175). From

Eq. B.12 it follows that the requirement for a plane

strain condition is satisfied by the radial displacement,

u, being independent of r, so that uniform radial transla­

tion is the required condition for plane strain (65). In

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
54 9

o other words, a plane strain condition develops if the

thickness of the cylinder wall is kept constant. This can

be achieved by adjusting the pressure in the inner and

outer cells of a hollow cylinder. However, this procedure

is rather complicated and as mentioned earlier, different

inner and outer cell pressures increase non-uniformities

of stresses across the cylinder wall.. On the other hand,

the need to apply shear stresses at the ends of the

cylinder is incompatible with the freedom for the cylinder

to change radius at the ends (radial translation). This

restraint will inevitably lead to non-homogeneous strain

if the whole cylinder is considered and the radius of the

cylinder changes away from the end platens. End restrain

c effects are discussed in detail in a later section.

should be noted that under uniform radial displacement


It

( er = 0 ) the circumferential strain, S q , does not equal

zero. It may vary across the wall, which contributes to

non-uniformities in the internal stresses.

Consequently, plane strain conditions in torsional

shear test specimens are more usually achieved by means of

tests in which no radial displacement is permitted. If

the cross-sectional area of the vertical loading piston is

the same as that of the hollow cylinder specimen, it is

possible to impose no lateral strain by preventing any

change in the volume of cell water during torsional load­

ing (75,81). Under these conditions, the cell pressure


r

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
550

continuously changes without permitting any lateral defor­

mation of the specimen. Continuous measurement of the

cell pressure permits the changes in lateral pressure

under plane strain conditions to be known throughout the

tests duration. Consequently, when a specimen is K -

consolidated and sheared under plain strain conditions,

where the cross area is kept constant, the mean confining

stress, , the direction of o^ , angle a, and hence b

change during torsional shearing (52). Since the specimen

is twisted without allowing lateral displacement, the test

is considered to simulate the stress conditions to which

an element of soil within deposits under level ground is

subjected during earthquakes (75,81). Under undrained

shearing, a condition of plane strain can be alternatively

achieved by fixing the length of the specimen (173). In

such a case, the axial stress and not the cell pressure

will change during torsional loading. In the case of

plane strain tests, without allowing radial displacement,

the cross section of the specimen is kept constant, there­

fore

earea ' 0

However, the radial and circumferential strains may be

still different than zero such that

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
551

o It is difficult to know if the strains

torsional shear tests specimens under "plane strain" con­


e and
r
e Q exist in
o

ditions. If there are not such strains there may be no

end restraints and hence, shorter specimens may be used

for plane strain tests (52).

The new resonant column/torsional shear test

apparatus does not yet have the capability to perform

plane strain tests. Change in the volume of the cell

water cannot be controlled since the Hardin oscillator

cannot be submerged. Besides, the cross-sectional areas

of the vertical loading piston and of the specimen are not

the same. One alternative would be to use a mechanism to

fix the height of the specimen during undrained torsional

c shear tests. Alternatively, proximity transducers

be used to measure radial displacements of the cylinder


could

walls. Proximity transducers sense the gap between the

transducer force and a target such as an aluminum foil

target attached to the membrane. Computer controlled

solenoid valves would then be used to automatically change

the cell pressure to prevent any radial displacement.

This feature would greatly enhance the capabilities of the

new resonant column/torsional shear test apparatus.

Stress/Strain Distributions

On a cylindrical column of soil subjected to tor­

r sional loading, the shear strains range from a minimum at

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
552

Che inner radius to a maximum at the outer radius; the


O resulting variation in shear stresses will depend upon the

stress-strain properties of the tested specimen. For

small rotations, the shear strain, y at a distance r


Z o
from the axis of rotation is given by

Tz e (r) " I r (B.17)

where 6 is the rotation angle in the cross section at a

height 1 from the bottom of the specimen. Since 6/1

remains constant along the height of the specimen, y 0, is


zo
independent of z. Accordingly, the shear strain, Yz g> can

be defined at any cross section. For example, at the top

of the specimen, 6 equals the maximum twist angle and 1

c corresponds to the length of the specimen, L.

The average shear strain, y q , can be expressed as


av

^ATzfl(r)dA 2 6 r ro “ ri ,
Tz0 JTdA 3 L 2 2 (B.18)
av ■'A r - r.
o i

where r and r. are the outer and inner radii, respec-


o i
tively. Thus, the average shear strain for each cross

section occurs at a radius equal to

3 - r3
rav = J 1 r2 _ r 2 5 (B . 19)
o ~ i

( In the case of a solid cylinder specimen, rQ = 0, and

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
553

c ^
hence the average shear strain occurs at a radius equal to

(2/3)r
o
. In terms of the outer and inner diameters, OD

and ID, E q . B.18 becomes

Y .e 0P* ~ l p -' 1 (*.2 0 )


av 3 L OD2 - OD2

The non-uniform distribution of shear strains induced

by twisting the top of the specimen leads to variations in

the shear stress, Tz g > across the cylinder wall. In gen­

eral, the stress and strain distributions depend on the

materials constitutive law. For small shear distorsions,

the behavior of soils is considered to be approximately

"elastic." Consequently, in a hollow cylinder specimen

subjected to a torque T which produces small shear

strains, the shear stress, Tz g> can be assumed to vary

linearly with radius. From elasticity theory, the shear

stress, Tz g» is related to the shear strain, Yz g> by the

ex pre ssi on,

Tz e (r) = G Yz g(r) (B.21)

where G is the equivalent "elastic" shear modulus of the

soil specimen. Substituting E q . B.17 into E q . B.21 yields

Tz 6 (r) = G | r ( B. 22 )

On the other hand, the torque T can be expressed as


r~
K

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
554

x rn ro 1
(^} T - Jt° Tz6(r) [2irr]r dr » 2ir Jr° T20 (r)[rz ]dr (B.23)

By substituting E q . B.22 into B.23 and integrating, it

follows that

T - G ^ I (B.24)
li

4 4
where I * “ r^)/2 is the area polar moment of inertia

of the specimen cross section. By combining Eqs. B.22 and

B.24, the shear stress, t can be obtained as


zo

x2 0 (r) = I r (B .25)

The average shear stress t q can be expressed as


av

c
“ T rav <B '2 6 >
av

By substituting E q . B.19 into E q . B.26 the expression for

the average shear stress becomes

r3 - r3
Tz0 = Ifi? 7"2 2W 4 47 T (B .27)
av (ro - ri )(ro - r± )

and in terms of the outer and inner diameters E q . B.27

becomes

32 [OD3 - ID3 ]
z6 3 ir
(B.28 )
av [OD2 - ID2 ][OD4 - ID4 ]

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
555

At large shear strains, the assumption of linear dis­


c tribution of shear stresses is incorrect. On the con­

trary, as the shear strains increase the plastic nature of

the behavior gradually becomes dominant and the shear

stress tz g approaches a limiting value irrespective of the

radius, r. Consequently, it is more appropriate to assume

a uniform distribution of shear stresses across the

cylinder wall. In this case, the relationship between the

average shear stress, Tz g» ai*d the torque can be esta­

blished from E q . B.23 as

T- *= ------^ --- =- T (B .29 )


9Z [rI - r»]

c and in terms of the outer and inner diameters,

Eq. B.29 becomes


OD and ID,

Tfw " -------- =5“ T (B .30)


02 * [OD3 - ID3 ]

In summary, there are two extreme behaviors regarding

deformation properties of soils: a perfectly linear elas­

tic behavior and a perfectly plastic behavior. For these

two extreme cases, the relationship between the applied

torque and the average shear stress, Tz g» is given by

equations B.28 and B.30. The most correct relation

between torque and shear stress is non-linear shear strain

dependent and lies between these two extremes. However,

it has been found that the assumed stress distribution


c

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
556

does not s i g n i f i c a n t l y affect the relationship between the


O average shear stress and the applied torque (38,75,88).

Table B.l presents the values of the ratio Xgz /T (Eqs.


av
B. 28 and B.30) for the two extreme stress distributions,

according to the specimen sizes in the new resonant

column/torsional shear test specimens. It is evident that

the relationship between the average shear stress and the

applied torque does not differ appreciably (about 11% for

the solid cylinder) regardless of whether a linear varia­

tion of shear stresses (elastic model) or a uniform shear

stress distribution (rigid plastic model) is assumed. The

difference is further reduced if the hollow cylinder is

used (about 1 to 3%). Consequently, either E q . B.28 or

E q . B.30 can be used without significant error in the cal­


c
culated average shear stress, especially in the case of

hollow cylinders.

Although the average shear stress may not be affected

significantly by the actual stress-strain distributions,

the variation in T q z and in y 0 across the cylinder wall

is an undesirable feature of the torsional shear test,

especially if strain measurements are to be made in order

to generate constitutive equations. On the other hand,

since soils undergo volume changes when subjected to

shear, the non-uniform distribution of shear stresses,

x Q , causes different tendencies for volumetric strains


ZO
across the wall of the specimen, which in turn may cause

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
Reproduced with permission of the copyright owner. Further reproduction O

Table B.l Relationship Between Average Shear Stress


and Applied Torque for New Apparatus

S pecimen Solid Hollow Hollow

Outer diameter (mm) 71.1 71.1 101.6

Inner diameter (mm) 38.1 71.1

A = T /T 9446.5 12220.3 5485.8


(linearvdistribution)
prohibited without perm ission.

B = TQV /T 10627.3 12559.9 5541.1


(uniform distribution)

A/B 0.889 0.973 0.990

* From Eqs. B.26 and B.28 with 0D and ID in meters

557
558

non— uniformities in the distribution of axial, radial and


O circumferential stresses, even in the case of equal inner

and outer pressures. Accordingly, the radial stress, o^,

need not be constant and equal to the circumferential

stress, Og, when a torque is applied, even to an isotropic

material. Only the average values of these stresses,

and U g , are equal. The actual distributions of these

stresses depend on the specimen geometry and the stress-

strain properties of the soil. For anisotropic materials

stress non-uniformities may arise, even without an applied

torque (65,66).

It is clear that with hollow cylinder specimens the

difference between the maximum and minimum shear stresses

c can be substantially reduced and with them the correspond­

ing non-uniformities in the distributions of normal

stresses across the cylinder wall. Limiting values of the

deviation from a uniform shear stress distribution can be

established by comparing the maximum and minimum shear

stress values for a linear distribution with the average

shear stress defined by assuming a uniform distribution of

shear stresses. From Eqs. B.25 and B.29, it follows that

= 4 (! - - ° 0 (B .3la)
(Tz 0Jav 3 1 - n4”

and

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
559

O (B.31b)

where n corresponds to the ratio, r./r . Equations B.31a


1 o
and B.31b are shown graphically in Figure B.5, together

with the corresponding values for the different specimen

sizes in the new resonant column/torsional shear test

apparatus. The values for the hollow cylinder used by

Hight et al. (65) are also given for comparison. It is

evident from this figure that by adopting a specimen

geometry with a large outer radius and a thin wall, the

non-uniformities produced by the applied torque can be

minimized. For example, a ratio n = 0.70 leads to a max­

imum relative difference in the distribution of the tor­

sional shear stress of about 15 to 20% with respect to a

uniform distribution, which seems quite acceptable (140).

It is to be noted that the rate of improvement in stress

uniformity decreases as the radius ratio increases.

Accordingly, unpractical specimens sizes may be required

to obtain a further reduction in stress non-uniformities

for radius ratios above 0.7. It is important to note that

the shear stress differences shown in Figure B.5 decrease

significantly as one moves towards the mean radius and/or

as failure is approached.

There is yet another aspect related to stress distri­

butions in hollow torsional shear test specimens which

requires some discussion. In common with all testing

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
Reproduced with permission of the copyright owner. Further reproduction

1.4-
(linear)

1.2
Uniform
1. 0 -
T max
ez 0.8
n'0z)av
0.6 Linear Uniform

I.
prohibited without perm ission.

0.4
w
io o
CM CM
0.2 — (r z0) min
(linear) o o
o -

0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
n = ri/ro

F i g u r e B, 5 Shear S t r e s s Distributions i n Ho l l ow T o r s i o n a l Shear Te s t

5 60
Specimens.
561

o apparatus utilizing rigid loading platens, radial forces

due to frictional restraint and stiffness of the platens

are developed at the end of the specimen, if it has a ten­

dency to expand or contract during torsional loading

(52,65,97,140). This tendency is always present when

there is large volume changes or a change in length at

constant volume. Lubrication of the platens is not possi­

ble because in the torsional shear test apparatus friction

of the platens in necessary for transmission of torque to

the specimen (65). Restraint on radial displacement at

the specimen ends primarily results in the development of

radial shear stresses, Tz r » and then complementary

stresses, Tr z > which decay with distance from the platens

c (St. Venant's principle).

radial shear stresses,


It is clear that due to these

the intermediate principal stress,

, rotates out of the plane of the cylinder wall. The

radial frictional forces, which are self-equilibrating,

and the associated shear stresses also result in addi­

tional circumferential (hoop) stresses, a Q, and in bending

moments which affect the distribution of the axial normal

stress, Oz . The stress components associated with end

restrain effects are illustrated in Figure B.6. Since the

shear stresses, Tz r » cannot be balanced by shearing

stresses on the inner and outer surfaces of the specimen,

a moment is developed which must be balanced by changes in

the axial stress (192). These stresses decrease the

applied axial stresses on the outer half of the specimen


fV

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
562

C Tez

O ’r

c
o ’z d u e to M
<,

Figure B.6 Stress Components in Hollow Torsional Shear


Test Specimens Including End Effects
(After Ref. 140).

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
563

and increase them on the inner half or vice versa depend­


u ing on the direction of the radial frictional forces.

This can have important effect on e and e a for a fric-


r o
tional material (140).

It is important to note that once radial shear

stresses ( t z r ) develop, a new term should be included in

the equation for radial equilibrium (Eq. B.l). Conse­

quently, O q cannot be assumed to be equal to , even if

the radial stress gradient, 3cr /3r, is zero. The average

change in <Jg can be expressed by (65)

r + r. 3t
A oe - ( ° 2 A ) (B.32)

C It is clear that stress distributions due to end restraint

effects are dependent on specimen geometry, the stress-

strain characteristics of the specimen and the applied

pressure and load combinations. These effects have been

comprehensively studied by Saada et a l . (140) and Hight et

al. (65). Saada et al. (140) used the elasticity theory

solution for a hollow cylinder subjected to a tangential

force F per unit length at both of its ends (see Fig. 3.6)

in order to estimate the distance beyond which the various

stresses caused by end restraint effects, which decay with

distance from the platens, become negligible. Based on

this analysis, Saada et a l . (140) concluded that the

stresses a , o_, and t due to end effects were all of


z 9 zr
the same order of magnitude and that in order to have a
V

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
564

c central zone in the specimen free from end effects with a

length at least equal to the length of the zone influenced

by a platen, the total length should be

L > 5.44 ^ - r^ (B.33)


o i

This requirement is satisfied for all specimen sizes in

the new resonant column/torsional shear test apparatus.

Based on linear elastic and plastic finite element

analyses, Hight et al. (65) concluded that

the distribution of o^ is not significantly affected

by end restraint

c b. changes in the distribution of both and a Q are

sensitive to geometry (more to height than to

ra di us) , to load/pressure combinations and to

Poisson's ratio,

c. the distribution of stresses is basically similar for

elastic and plastic materials,

d. stress non-uniformities due to end restrains obvi­

ously increase as failure is approached, and

e. a reasonable uniform distribution of stresses can be

obtained with a suitably "high" specimen, at least

over a central position.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
565

o These conclusions confirmed the previous ones drawn

by others (97,140). If end restraint effects cannot be

totally eliminated, their influence needs to be con­

sidered, especially in the interpretation of strength.

For example, uncertainty in the value of the mobilized

friction angle, $, can arise when failure is approached

and the sample shows a tendency to bulge (169). The error

in friction angle would be mainly determined by the real

value of and by the unknown change in Og, which as

indicated is given by E q . B.32. Unfortunately, the value

of the end restrain force (and hence of XZ T ) depends on

the tendency to expand radially and there is no direct way

of measuring it or evaluating its magnitude (140).

c Measurement of Applied Boundary Stresses

The applied boundary stresses are measured by means

of a differential pressure transducer, a load cell and a

torque transducer. The characteristics of this measure­

ment system and the corresponding calibration factors and

procedures are discussed below.

Differential Pressure Transducer

The differential pressure transducer is located

beneath the bottom of the chamber. The positive side of

the transducer is connected to the chamber while the nega­

c tive side is connected to one of the drainage lines from

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
566

^ , the bottom platen. Accordingly, the transducer outputs a

voltage which is proportional to the effective confining

stress. This transducer is a four channel strain gauge

and was provided by Sensotec, Inc. (Model No. P30P). The

full scale capacity of the transducer is 344.7 kPa (50

psi) differential but pressures of 689.4 kPa (100 kPa) are

not supposed to harm the unit (166). The calibration fac­

tor provided by the manufacturer is given in the form of

transducer output (millivolts) per volt of excitation for

a differential pressure corresponding to the full scale

capacity of the transducer. This calibration factor is

1.991 MV/V. With an excitation voltage of 10 Volts DC,

the pressure transducer calibration factor, PTCF, becomes

c equal to 17315.42 kPa per Volt output.

In the new apparatus, excitation to the transducer is

provided with an ADIO (Analog/Digital Input/Output) card

(AIM05), which plugs into the MACSYM 200 I/O back plane

and provides bridge completion, amplification, filtering

and excitation for a wide range of sensors, transducers

and actuators. The excitation voltage was set at 10 Volts

DC, the excitation voltage recommended by the manufac­

turer. Under these conditions, the transducer was recali­

brated by applying vacuum to the negative side of the

transducer (bottom platen connection) and using a 30 Hg-

inches vacuum gauge with an accuracy of about 3 kPa. The

results of this recalibration are presented in Tables B.2a

r
V

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
Reproduced with permission of the copyright owner. Further reproduction

Table B.2 Results of Pressure Transducer Calibration.

CALIBRAT.RDR - VERSION OF 9/16/03


a)

Differential pressure transducer calibration

02/15/85 Calibration applying vacuum

LEAST SQUARES FIT

EQUATION OF LINE: Y *=.000056302 (X) +.00017246

pressure output CAL. output; DIFFERENCE ERROR 'A 1


prohibited without perm ission.

kpa volts volts volts

33.0773 .002054 .00200253 -.0000285311 .50912


67.7546 .003995 .0039926 2.39769E-6 .0427853
95.8720 .005575 .00557796 -2.9644E-6 .0528979
96.2115 .005604 .00559706 6.93920E-6 .123827
67.7546 .00390 .0039926 -.00001260219 .224879
33.0773 .002075 .00200253 -7.53114E-6 .134389
33.8773 .002090 .00200253 .0000154680 .276031
37.4092 .002313 .00220610 .0000260226 .478632

MAX. ERROR (PERCENT FULL SCALE OUTPUT) = .509118


COEFFICIENT OF CORRELATION ’ R’ ■"** a999939

567
STANDARD ERROR OF ESTIMATE 0000104678
PTCF = 17736.2 kPa/Volt
Reproduced with permission of the copyright owner. Further reproduction

Table B.2, continued.

CALIBRAT.RDR - VERSION OF S/16/83


b)

Pressure transducer calibration with vacuum

02/15/85 Including only readings for 10 and 20 Hg inches

LEAST SQUARES FIT

EQUATION OF LINE: Y =.0000564144 (X) +.000165167


prohibited without perm ission.

pressure output CAL.output DIFFERENCE ERROR •/. FSO


Kpa volts volts volts

33.8773 .002056 .00207633 -.0000203331 .508964


67.7546 .003SS5 .0039875 7.50041E-6 .187745
67.7546 .00398 .0039875 -7.49948E-6 .187722
33.8773 .002075 .00207633 -1.3331B8E-6 .0333714
33.8773 .002098 .00207633 .0000216668 .542347

MAX. ERROR (PERCENT FULL SCALE OUTPUT) .542346


COEFFICIENT OF CORRELATION ’ R’ .999885
STANDARD ERROR OF ESTIMATE .0000183236
PTCF = 17726.1 kPa/Volt Ln
On
00
569

and B.2b, together with corresponding least square fit­


G tings. It is noted that the transducer calibration factor

determined by this procedure differs at most by 2.4% from

the value provided by the manufacturer, confirming its

accuracy. The main characteristics of the pressure trans­

ducer and all the other transducers in the new apparatus


are summarized in Table B.5. The effective radial stress,

a , can be obtained as
r*

ar = (PTCF)(PTO) (B.34)

where is obtained in (kPa) and PTO is the pressure

transducer output (Volts) after subtracting the reading

under zero differential pressure. It is to be noted that

C in the new resonant column torsional shear apparatus

independent measurements of the cell pressure and pore

water pressure cannot be made. Consequently, the pore

water pressure, u, at any time during an undrained test

must be calculated in terms of the consolidation effective


*
radial stress, a . It follows that
c

u = o (B .35)
c

Although differential pressure transducers avoid

errors due to differences in calibration of two different

transducers, an independent measurement of the cell pres­

sure is required, especially for B— parameter (pore

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
570

pressure) measurements. At the present time the change in

the total confining pressure, , is measured with a 700

kPa pressure gauge, which does not have enough accuracy

for precise measurements.

Load Cell

Any additional axial load applied by the double act­

ing air cylinder on the specimen is measured by means of a

load cell located within the chamber, underneath the bot­

tom platen. In this way friction related problems are

avoided and the actual axial load applied to the specimen

is effectively measured. This load cell is also a four

arm strain gauge bridge and was provided by LEBOW Associ­

ates, Inc. (S/N 2628), with calibrations that are trace­

able to the U.S. National Bureau of Standards. This load

cell is designed in such a way that the readings are not

affected by changes in the external cell pressure. The .

nominal full scale capacity of the transducer is 3.336 KN

(750 pounds) with a static overload capacity equal to 50%

of nominal capacity. The calibration factor provided by

the manufacturer corresponds to 1.965 millivolts (output)

per volt of excitation under full load capacity in


«

compression. Consequently, for an excitation voltage of

10 Volts DC, the load cell calibration factor, LCCF, in

terms of the data provided by the manufacturer becomes

equal to 169.77 KN per volt output.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
571

The excitation and signal conditioning for the load


G
cell is the same as described for the differential pres­

sure transducer. With an excitation power of 10 DC volts,

the load cell was recalibrated by placing dead weights on

top of the bottom platen with a maximum load corresponding

to three different percentages of the rated capacity. The

results of this recalibration are presented in Table B.3a,

B.3b and B.3c. The difference between the load cell cali­

bration factors calculated using least square fittings and

the nominal value provided by the manufacturer is at most

0.075%. A summary of the characteristics of the load cell

is also included in Table B.5. The axial stress incre­

ment, Ao , can be obtained as


z

C 4C . ( L C C F ) d C O )
z A

where LCO is the load cell output (Volts) after subtract­

ing the reading under zero axial load, A is the area of -


2
the cross section of the specimen (meters ) and A i s

obtained in k P a . It should be noted that the bottom pla­

ten, the rotation transducer platform and the filling of

the cell with water contribute to changes in the load cell

readings. As we are concerned with the additional axial

load applied to the specimen, the "zero” load cell reading

ic taken after specimen preparation is completed.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
572

/ v

Table B.3 Results of Load Cell Calibration.

CflLIBRftT.RDR - VERSION OF 9/16/83


a)

Calibration of load cell for measuring vertical loads

07/02/64 First calibration

LEAST SQUARES FIT

EQUATION OF LINE: Y =5.8859E-6 (X) +. 000187906

load out put CAL.output DIFFERENCE ERROR % FSO


newtons volts volts volts

0 .000187755 .000187906 — 1.50656E—7 .0139569


4.13855 .000212789 .000212265 5.24276E-7 .0485693
13.9456 .000270605 .000269988 6.17001E-7 .0571594
S3.7526 .000328422 .000327711 7.11007E-7 .0656681
33.5596 .000385642 .000385434 2. 0S005E-7 .0192697
.000442863 .000443157 -2.93978E—7 .0272343

c
43.3666
53.1736 .000501275 .00050088 3.95055E-7 .0365982
62.9904 .000559688 .000558661 1.02737E-6 .0951758
72.B072 .000616312 .000616441 -1.29279E—7 .0119765
82.624 .000673532 .000674222 -6.89994E-7 .0639214
92.440B .1000731349 .000732003 -6.53672E—7 .0605566
102.258 .000790954 .000789786 1.16829E-6 .108231
112.074 .000847578 .000847562 1.63564E-8 .00151526
121.891 .000905395 .000905344 5.1513BE-8 .00477227
131.71B .000962615 .000963184 -5.69213E-7 .0527322
141.554 .00102103 .00102108 -4.78467E-8 .00443255
151.4 .00107944 .00107903 4.09549E-7 .0379409
141.554 .00102162 .00102108 5.42146E-7 .0502248
131.71B .000963807 .000963184 6.22764E-7 .0576932
121.891 .000905395 .000905344 5.1513BE-8 .00477227
112.074 .000846982 .000847562 -5.79632E—7 .0536975
102.258 .000788569 .000789786 — 1.216715E—6 .112717
92.4408 .000731349 .000732003 -6.53672E-7 .0605566
82. 624 .000673532 .000674222 -6.89994E—7 .0639214
72.8072 .000616312 .000616441 — 1.29279E-7 .0119765
62.9904 .000559092 .000558661 4.31319E-7 .0399576
53. 1736 .000501275 .00050088 3.95055E-7 .0365982
43.3666 .000443459 .000443157 3.020IE—7 .0279784
33.5596 .000385642 .000385434 2. 08005E-7 .0192697
23.7526 .000327229 .000327711 -4.81989E—7 .0446517
13.9456 .000270605 .000269988 6.17001E-7 .0571594
4.13855 .000211596 .000212265 -6.68733E-7 .0619519
0 .000186563 .000187906 -1.342647E-6 .124384

MAX. ERROR (PERCENT FULL SCALE OUTPUT) = .124384


COEFFICIIENT OF CORRELATION ’R’ = .999998
STANDARD ERROR OF ESTIMATE = 5. 80856E-7
LCCF = 169.897 KN/Volt

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
573

O
Table B.3, continued.

CALIBRAT. RDR - VERSION OF 9/16/83


b)

Cali brat iori of load cell for vertical loads

02/16/B5 Second Calibration on June 13/84

LEAST SQUARES FIT

EQUATION OF LINE: Y =5.88984E-S (X) +.000187086

load output CAL. output DIFFERENCE ERROR ‘A FSO


newtons volts volts volts

0 .000185966 .000187086 -1.119814E-6 .0361226


4.13855 .000211596 .000211461 1.34765E-7 .00434721
43.4744 .000442663 .000443143 -2.8027E-7 .00904085
82.722 .000673532 .000674306 -7.73522E—7 .024952
122.048 .000905991 .00090593 6.1409IE—8 .00198091
161.237 .00113666 .00113675 -8.67294E—8 .00279769
200.435 .00136733 .00136762 -2.85847E—7 .00925302
c 298.505
396.575
.0019449
.00252306
.00194523
.00252285
—3.33996E—7
2.C3949E-7
.0107739
.00673698
494.645 •00310004 .00310047 -4.28176E-7 .0138119
396.575 .00252247 .00252285 -3.81144E-7 .0122948
298. 505 .0019455 .00194523 2.66009E-7 .00858083
200.435 .00136793 .00136762 3.13157E-7 .0101017
161.237 .00113845 .00113675 1.70327E-6 .0549436
122.048 .000907183 .00090593 1. 25339E-6 .0404313
82.722 .000673532 .000674306 —7.73522E-7 .024952
43.4744 .000443459 .000443143 3.15718E-7 .0101843
4.13855 .000211596 .000211461 1.34765E—7 .00434721
0 .000187159 .000187086 7.31961E-S .00236114

MAX. ERROR (PERCENT FULL SCALE OUTPUT) = .0549431


COEFFICIENT OF CORRELATION ’R’ = .999999
STANDARD ERROR OF ESTIMATE = 9. S1323E-7
LCCF = 169.784 KN/Volt

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
574

o
Table B.3, continued,

CRLIBRAT. RDR - VERSION OF 9/16/83


c)

Calibration of load cell for vertical loads

02/16/85 Third calibration on June 13/84

LEAST SQUPRES FIT

EQUATION OF LINE: Y =5.88882E-6 (X) +.000188023

load output CAL.output DIFFERENCE ERROR '/. FSO


newtons volts volts volts

0 .000187159 .000188023 -8.63714E—7 .017058


4. 13855 .000211 .000212394 — 1.393899E-6 .0275289
102.209 .000788569 .000789913 —1.34448IE-6 .0265529
200.279 .00136673 .00136743 -7. 00355E-7 .0138317
298.349 .00194311 .00194495 —1.B3715E—6 .0362829
396.419 .00252187 .00252246 -5.94184E-7 .0117349
494.489 .00309885 .00309998 -1.131091E—6 •0223385
c 592.559
631.885
.00367761
.00390887
.0036775
.00390908
1.11992E-7
-2.11B76E-7
.00221178
.00418445
671.221 .00414073 .00414072 5.58794E-9 .000110359
710.409 .0043714 .0043715 -9.54606E-8 .0018853
749.60S .00460267 .00460233 3.3B536E—7 .00668592
788.855 .00483394 .00483345 4.89B76E—7 .00967482
828.083 .00506341 .00506446 —1.047272E—6 .0206831
788.855 .00483453 .00483345 1.07987E—6 .0213269
749.608 .00460207 .00460233 -2.61702E-7 .00516849
710.409 .0043714 .0043715 -9.54606E—8 .0018853
671.221 .00414133 .00414072 6.0536E-7 .0119556
631.885 .00390887 .00390908 -2.11876E-7 .00418445
592.559 .S3367761 .0036775 1.11992E—7 .00221178
494.489 .00310063 .00309998 6.48B99E—7 .0128155
396.419 .00252306 .00252246 5.95814E-7 .011767
298.349 .0019455 .00194495 5.52B55E—7 .0109187
200.279 .00136793 .00136743 4.99655E-7 .00986795
102.209 .000792146 .000789913 2.23255E—6 .0440919
4.13855 .000213981 .000212394 1.587C9E-6 .0313443
0 .000188947 .000188023 9.24279E-7 .0182541

MAX. ERROR (PERCENT FULL SCALE OUTPUT) = .0440924


COEFFICIENT OF CORRELATION ’R’ =1
STANDARD ERROR OF ESTIMATE = 1.20631E-6
LCCF = 169.6 KN/Volt

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
575

C Finally, the total axial stress, oz , is obtained as

a * o + Ac (B.37)
z r z
c

Torque Transducer

The torque transducer is located in the base plate of

the chamber with the sensing head within the chamber and

the transducer body outside the chamber. The bottom pla­

ten rests on top of the load cell which in turn is

attached to the head of the torque transducer. This sens­

ing head is not in contact with the base plate and hence,

the actual torque applied to the specimen is effectively

measured. This transducer is also a four arm strain gauge

bridge and was provided by LEBOW Associates, Inc. (S/N

555), with calibrations that are traceable to the U.S.

National Bureau of Standards. The rated capacity of the

torque transducer is 56.5 N-m (500 In.Lbs) with an over­

load capacity of 50%. The calibration factor provided by

the manufacturer corresponds to 1.968 millivolts (output)

per volt of excitation for clockwise rotation and 1.967

mV/V for counter-clockwise rotation. Accordingly, for an

excitation voltage of 10 volts DC, the torque transducer

calibration factor, TTCF, becomes equal to 2870.4 N-m per

volt output for clockwise rotation and 2871.86 N-m per

volt output for counter-clockwise rotation.

r
v

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
576

The excitation and signal conditioning for the torque


c transducer is the same as described for the pressure

transducer and the load cell. The torque transducer was

recalibrated by applying torque to the bottom platen by

means of a steel bar and two pulleys clamped to a fixed

support. Dead weights were suspended from the pulleys

which transmitted the load horizontally to the steel bar

with a lever arm of 0.508 meters. The results of this

recalibration are presented in Table B.4a and B.4b. Least

square fittings were performed considering all the data

points (Table B.4a) and considering independently the data

for clockwise rotation and counter-clockwise rotation

(Table B.4b). The calculated calibration factors differ

c from the nominal values provided by the manufacturer,

at most 0.8%. The torque applied to the top of the speci­


by

men, T, can be obtained as

T = (TTCF)(TTO)/1000 (B.38)

where TTO is the torque transducer output after subtract­

ing the reading under zero applied torque and T is

obtained in KN-m. By combining Eqs. B.28 and B.38 the

average shear stress, x ^ , acting on any cross section


av
of the specimen is given by

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
577

c
Table B.4 Results of Torque Transducer Calibration.

CALIBRAT. RDR - VERSION OF 9/16/83


a)

Data for- calibration of torque tranducer

07/03/84 Second calibration(June)

BI-POLARITY DATA
PQS. & NE6. BRANCHES INDEPENDENT.
LEAST SQUARES FIT

EQUATION OF LINE (NEG. BRANCH): Y =.00035147 (X)

torque output CAL. out put DIFFERENCE ERROR ’A FSO


newt on-met er volts volts volts

0 -.000171065 0 -.000171065 1.0466


-2. 10239 -.0008744 -.000738927 -.000135473 .828842
-7.08434 -. 0025773 -.00248993 -.0000873657 .534517
-12.0663 -.00427723 -.00424094 -.0000362853 .221999
c -17.0532
-22.0402
-. 00599504
-.00770986
-.00599369
-.00774647
-1.34902IE-6
.0000366126
.00825352
.224002
-27.0271 -.00944555 -.00949922 .0000536693 .328357
-32.024 -.011186 -.01125548 .0000694804 .425092
-37.0259 -.01291573 -.0130135 .0000977702 .598173
-42. 0328 -. 01465082 -.01477327 .000122455 .749198
-47. 0396 -.0163448 -.016533 .000188215 1.15153
-42. 0328 -.01480877 -.01477327 -.0000354955 .217167
-37.0259 -.01308262 -.0130135 -.00006912 .422887
-32. 024 -. 0113538 -.01125548 -.0000983197 .601535
-27.0271 -.00961006 -.00949922 -.0001108404 .678139
-22.0402 -.00789881 -.00774647 -.000152338 .932026
-17.0532 -. 006173B5 -.00599369 -.000180159 1.10224
-12. 0663 00442446 -.00424094 -.000183516 1.12278
-7.08434 -.00269353 -.00248993 -.000203596 1.24563
-2.10239 -.000956059 -.000738927 -.000217132 1.32844
0 -.000191927 0 -.000191927 1. 17424
0 -.000151992 0 -.000151992 .92991
0 -.000131726 0 -.000131726 .80592

MAX. ERROR (PERCENT FULL SCALE OUTPUT) = 1.32844


COEFFICIENT OF CORRELATION ’R’ = .991956
STANDARD ERROR OF ESTIMATE = .000706588
TTCF = 284 5 .2 iv -iu /V o lt

r
v

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
578

Table B.4, continued,

a)

EQUATION OF LINE (POS. BRANCH): Y =.000345445 (X)

torque output CAL.output DIFFERENCE ERROR % FSO


newt on-met er volts volts volts

0 -.000171065 0 -.000171065 1.04955


0 -.000191927 0 -.000191927 1.17755
0 -.000151992 0 -.000151992 .932529
S. 10239 .000568032 .000726259 -. 000158227 .970785
7.08434 .00231326 .00244725 -.000133987 .822061
12.0863 .00406325 .00416824 -.000104988 .644142
17.0532 .00580609 .00589094 -. 0000848453 .520559
22.0402 .00755191 .00761367 -. 0000617574 .378905
27.0271 .00929773 •00933636 -.0000386341 .237035
32.024 .0110507 .0110625 -.00001181569 .0724938
37.0259 .0127912 .0127904 8. 028E-7 .00492549
42. 0328 .0145441 .01452 .000024038 .14785

c 47.0396
42.0328
37. 0259
32. 024
27.0271
.0162989
.0145829
.0128454
.0110348
.00934124
.0162496
.01452
.0127904
.0110625
.00933636
.0000493247
.0000628987
.000055003
.0000322843
4.87547E—6
.302626
.385908
.337464
.198077
.0299129
22.0402 .00760078 .00761367 -.00001288718 .0790678
17. 0532 .00586271 .00589094 -. 0008282251 .173172
12. 0663 .0041008 .00416824 -.000067438 .413758
7.08434 .00236034 .00244725 -.0000B69068 .533207
2.10239 .000614524 .000726253 -.0001117352 .685538
0 -.000131726 0 -.000131726 .80819

MAX. ERROR (PERCENT FULL SCALE OUTPUT) = 1.17755


COEFFICIENT OF CORRELATION ’R’ = 993216 .
STANDARD ERROR OF ESTIMATE = 00065203 .
TTCF = 2894 .6 N-m/Volt

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
579

c Table B.4, continued.

CALIBRAT.RDR - VERSION OF 9/16/83


b)

Data for calibration of torque tranducer

07/03/84 Second calibration(June)

BI-POLARITY DATA
LEAST SQUARES FIT

EQUATION OF LINE: Y =.000348457 (X) -.0001175415

torque output CAL.output DIFFERENCE ERROR FSD


newt on—met er volts volts volts

0 -.000171065 -.0001175415 -.0000535235 .327465


-2.10239 -. 0008744 -.000850135 -.0000242653 .148459
-7. 08434 -. 0025773 -.00258613 8.83196E—6 .0540353
-12.0663 -.00427723 -.00432213 .0000449028 .274722
-17.0532 -.00599504 -.00605985 .0000648145 .396545
-22.0402 -.00770986 -.00779761 .0000877506 •536872
-27.0271 -.00944555 -.00953533 .0000897832 .549308
-32.024 -.011186 -.01127654 .0000905385 .553929
-37.0259 -.01291573 -.01301949 .00010376 .534817
-42.0328 -.01465082 -.01476418 .000113359 .693546
-47.0396 -. 0163448 -.0165086 .000164036 1.0035
-42.0328 -.01480877 -.01476418 -. 0000445917 .272e19
.386243

c
-37.0259 —•01308262 -.01301949 -.0000631306
-32.024. -.0113538 -.01127654 -.0000772616 .472698
-27.0271 -.00961006 -.00953533 -.0000747265 .457188
-22.0402 -.00789881 -.00779761 -.0001011998 .619156
-17.0532 -.00617385 -.00605985 -.0001139957 .697443
-12. 0663 -.00442446 -.00432213 -.0001023277 .626056
-7.08434 -.00269353 -.00258613 -.000107398 .657078
-2.10239 -.000956059 -.000850135 -.0001059243 .648061
0 -.000191927 -.0001175415 -. 0000743855 •455102
0 -.000151992 -.0001175415 -.0000344505 .210773
2.10239 .000568032 .800615052 -.0000470197 .287674
7.08434 .00231326 .00235105 -.0000377889 .231198
12.0663 .00406325 .80408705 -.0000237995 .145609
17.0532 .00580609 .00582477 -.0000186814 .114296
22.0402 .00755191 .00756253 -.00001061801 .0549626
27.0271 .00929773 .00930025 -2.51923E—6‘ .015413
32. 024 .0110507 .0110415 9. 2443IE—6 .0565581
37. 0259 .0127912 .0127844 6.794E-6 .0415667
42. 0328 .0145441 .0145291 .0000150036 .0917944
47.0396 .0162989 .0162738 .0000251476 .153857
42.0328 .0145829 .0145291 .000053B044 .329183
37.0259 .0128454 .0127844 .0000609942 .373172
32. 024 .0110948 .0110415 .0000533443 .326369
27.0271 .00934124 .00930025 .0000409903 .250785
22. 0402 .00760078 .00756253 .0000382522 .234033
17. 0532 .00586271 .00582477 .0000379388 .232116
12.0663 .0041008 .00408705 .0000137505 .0841277
7. 08434 .00236034 .00235105 9.29111E-6 .0568444
2. 10239 .000614524 .000615052 -5.2771IE—7 .00322861
0 -.000131726 -.0001175415 -.00001418449 .0867829

MAX. ERROR (PERCENT FULL SCALE OUTPUT) = 1.00359


COEFFICIENT OF CORRELATION ’R’ = .999973
STANDARD ERROR OF ESTIMATE = .0000681967

c TTCF = 2 6 6 9 .8 N -m /V o lt

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
580

/ 32 [0D3 - I D 3 ] (TTCF)(TTO)
( T
1000 (B.39)
3u
av [OD2 - D02 ][0D4 - ID4 ]

where tz a is obtained in kPa.


o
av

Table B.5 summarizes the main characteristics of the

various transducers used for measuring the applied boun­

dary stresses. It should be noted that only static load­

ing was used for calibrating the various transducers.

Consequently, the calculated calibration factors hold only

for low frequency tests. For high frequency tests,

dynamic effects should be evaluated as explained in Appen-

dix A.

Measurement of Specimen Deformations

In the new resonant column/torsional shear test

apparatus, it is possible to measure the average axial

strain and the rotation (and hence the shear strain) at

the top of the specimen. The apparatus has no capabili­

ties for measuring radial displacements of the specimen.

The characteristics of the deformation measurement system

are described below.

Axial Strains

Changes in the length of the specimen during any

stage of the test are measured by a linear variable dif­

ferential transformer (LVDT) located on a yoke just above

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
Reproduced with permission of the copyright owner. Further reproduction

Table B.5 Characteristics of Transducers for


Measuring Applied Boundary Stresses

Transducer Pressure Load Torque

Manufacturer Sensotec, Inc, Lebow, Inc, Lebow, Inc.

Rated Capacity 344.7 kPa 3.336 kN 56.5 N-m

Overload capacity (%) 100 50 50

Output at rated (mV/V) 1.991 1.965 +1.968 cw


capacity -1.967 ccw

Nonlinearity^ ^ ± 0.04 cw
(% of rated output) ± 0.03 ± 0.02 ccw
(2 )
Hysteresis ± 0.09 cw
(% of rated output) ± 0.03 ± 0.00 ccw
prohibited without perm ission.

Max. excitation voltage (Volts) 10.0 20 20

Calibration factor 17.3 169.8 2.87


(Excitation - 10 Volts) (MPa/Volt) (kN/Volt) (kN-m/Volt)

Linearity of a transducer refers to the maximum deviation of


any calibration point from a straight line during any one
calibration cycle over the specified linear range.
(2 ) Hysteresis of a transducer refers to the maximum difference
in output at any given input value, within the specified range
when the value is approached first with increasing and then with
decreasing inputs.
582

o the bearing connection between the air cylinder piston and

the rotating shaft. This LVDT was provided by Schaevitz

Engineering (S/N 521). The full scale displacement of the

transducer is 25.4 mm (1 inch). A transducer exciter

demodulator (Daytrenic, model 201C) is used to power the

transducer, which produces an output that is dependent

upon AC induction variations as a function of displacement

position of the core. This device is a signal condition­

ing instrument which allows convenient adaptation of dif­

ferential transformer transducers to strip chart record­

ers, X - Y recorders, digital voltmeters and similar dev­

ices requiring DC input signals. The transducer exciter

demodulator supplies regulated (3 Volts at 3 kHz) excita­

tion to the transducer, demodulates the transducer output


c signal, acd produces a filter DC output signal precisely

proportional to the displacement input. The axial dis­

placement transducer was calibrated by using a caliper

with an accuracy of 0.0005 inches (0.0127 mm). The

results of this calibration is presented in Figure B.7.

Within the linear range, the LVDT calibration factor,

LVDTCF, is 84.615 mm per volt output. Accordingly, the

axial displacement of the top of the specimen is given by

4 = (LVDTCF)(LVDTO) (B.40)

where LVDTO is the LVDT output (Volts) and A is obtained

in mm. The corresponding axial strain, , can be

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
583

( forthwards displacement
b a c k w a r d s displacement
120 --
E
E
i-
z
o 110 --
Q.
LLI
O
Z"
UJ
cc
UJ
LL
UJ
DC
2
O 90
DC
U_
UJ
O
z 80
<
I—
co
Q

70 — Linear range —

0.35 0.40 0.45 0.50 0.55 0.60 0.65

V O L T A G E O U T P U T (Volts)

m 1 = _ = 8 4 . 6 1 5 mm/volt
0.205Volt
m 2 =
8mm = 106. 6 6 7 mm/volt
0.075 volt

Figure B.7 Results of LVDT Calibration.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
obtained as

e » i x 100 (B.41)
z L

where L is the length of the specimen in millimeters and

&z is obtained as a percentage.

Shear Strains

The relative rotation between the bottom and the top

of the specimen is measured by two non-contacting dis­

placement transducers located on a cylindrical platform

which is fastened to the lower platen of the specimen.

These transducers were provided by Kaman Sciences Measur­

ing Systems (Model KD-2310). This model allows for preci­

sion non— contact displacement measurements of metal tar­

gets and thickness measurements of non-conductive material

backed by metal. The system includes a sensor, a coaxial

cable and a signal conditioning electronic package, which

has been incorporated within the analog control unit. The

output voltage of the system is proportional to the dis­

tance between the face of the sensor and any metallic

(conductive) target. In the new apparatus, the two trans­

ducers focus on non-concentric circular cams attached to

the top platen and hence measure the distance between the

cams and the transducer face. As the top platen rotates

relative to the bottom, the distance changes, which causes

c the transducer output to change, and hence allows

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
585

measurement of the angle of rotation from the at rest


C position of the specimen. The analog control unit sums

the signals from the two transducers and provides an out­

put signal for rotation control, the digital voltmeter and

the recorder system.

The rotation measurement system was calibrated by

Soil Dynamics Instruments (166), the builders of the main

portion of the new apparatus. The calibration procedure

was as follows. A machinist's scale graduated to 0.01

inch was fastened to the exterior of the Hardin oscillator

which has an 0D of 0.102 meters. An optical microscope,

attached to the specimen base by means of a magnetic base

and a metal plate held by "C" clamps, was used to read the

^ graduated scale and consequently, to determine the angle

of rotation. The chamber lid was held in position by

placing four brass tubes over the chamber rods. The upper

locking clamp on the shaft was locked to prevent the pis­

ton from falling into the chamber but the lower damp part

was loose and thus could allow rotation. The top platen

for solid specimens was attached to the oscillator and

rotation was provided by use of the stepper motor with a

gear ratio of 3200 to 1. The resulting calibration was

provided in the form of equations obtained from least

square fittings. The rotation, 6, can be obtained as

RTO
6 = RTCF (B.42)
c

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
586

where RTOa is the rotation transducer output after adjust­


0 ing it as follows: If RTO < 0, then

RTO ** RTO (B .43)


a

and if RTO > 0, then

2.876
RTO - RTO + 0.01637 (RTO) (B .44)
a

where RTO is the measured transducer output. The rotation

transducer calibration factor, RTCF, is 6.482 Volts/Rad.

for 6 values larger than 0.05 radians and 6.41 Volts/Rad.

when 6 is smaller than 0.05 radians. The corresponding

average shear strain, can be obtained from E q . B.20.

c Determination of Shear Modulus and


Damping Ratio from Torsional Shear Tests

In an elastic material, the relationship between

stresses and strains is linear. Consequently, the shear

modulus, which is the slope of the shear stress versus

shear strain curve, is constant irrespective of the shear

strain level. At very low shear strains the deformations

exhibited by most sands could be considered essentially

"elastic," that is recoverable. However, as shearing

takes place under increasing shear strains, the charac­

teristics of sands become non-linear and inelastic.

Accordingly, the stress-strain characteristics cannot be

r
v

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
587

represented by a single shear modulus value but it changes

with the level of shear-strain. Hence, it is possible to

define different shear moduli such as the initial tangent

shear modulus, the tangent shear modulus, the secant shear

modulus, etc. The response of sands to seismic loading is

usually investigated by means of cyclic loading, unload­

ing, and reloading, yielding an approximately closed loop

in the shear stress-shear strain space. The loop reflects

the nonlinear an inelastic stress-strain characteristics

of sands. During cyclic loading there is generally a com­

plete stress reversal, i.e. approximately equal excursions

in both the positive and negative directions. Hence, a

simple representation of the behavior is often made in

c terms of an equivalent secant shear modulus,

nal damping ratio, D, as illustrated in Figure B.8*


G, and inter­

The

shear modulus is defined as the slope of a straight line

connecting the end points of the shear stress versus shear

strain curve. Hence it is calculated as

G = f*? I (B.45)
max min

T - maximum positive shear stress


max
T . = maximum negative shear stress
min
Ymax = maximum positive shear strain

= maximum negative shear strain


Ymin
will have the same units as the shear stress

r
V

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.
588

t max

Xmin

Ymax

W/2
AW

AW

min

Figure B .8 D e f i n i t i o n of E q u i v a l e n t H y s t e r e t i c S h e ar
M o d u l u s and D amp ing Ratio.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
589

The damping ratio, D, is defined as the ratio of the

energy dissipated per loading cycle (area inside shear

stress-shear strain loop) , Aw, to 2x times the maximum

elastic strain energy stored in the soil element, w.

Hence,

1 Area of hysteris loop /T>


D = -r— -------------- v- — (B.46)
Area of triangles OAB & OA B

The area inside the shear stress-shear strain loop may be

obtained by numerical integration or by the use of a plan-

imeter. It should be noted that the maximum elastic

strain energy stored, w, can also be obtained as

w = G (y * + Y* >/2 (B.47)
mi n max

c
—!
_

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
o

VITA
c

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
590

VITA

Adolfo AlarcSn— GuzmSn was born on January 4, 1953* In

Dolores, Department of Tolima, Colombia. He attended

school In Bogota, Department of Cundinamarca, Colombia.

In 1976, he graduated from the National University of

Colombia with a diploma in Civil Engineering. At gradua­

tion, he received the "Grado de Honor" award for the

highest standing in the School of Engineering. He also

received the "Manuel Ponce de Leon" award granted by the

Colombian Society of Engineers for his high graduation


C point average.

In 1977, he joined the faculty at the School of

Engineering of the National University of Colombia, as an

instructor in Soil Mechanics and Foundation Engineering.

At the same time, he enrolled in the graduate program in

the Department of Geotechnical Engineering. Concurrent

with his work and studies at the National University of

Colombia, Mr. AlarcGn worked for G6mez Cajiao Asociados,

Ltd., from 1980 to 1982. During this time, he directed

laboratory investigations of soil properties for a variety

of projects, including two major dams and a 150 km. long

railroad.

R eproduced w ith perm ission o f the copyright owner. Further reproduction prohibited w itho ut perm ission.
591

S In the fall of 1982, he commenced graduate studies at

Purdue University and received a Master of Science in

Civil Engineering degree in 1983. Since then, he has been

working on his doctoral thesis.

Mr. AlarcSn is a member of the Colombian Society of

Engineers and the Colombian Geotechnical Society. He is

also a student member of the American Society of Civil

Engineers and a member of the American Society for Testing

and Materials.

fv

R eproduced with perm ission o f the copyright owner. F urther reproduction prohibited w itho ut perm ission.

Вам также может понравиться