You are on page 1of 389

YIELDING AND FAILURE OF CEMENT TREATED SOIL

XIAO HUAWEN

NATIONAL UNIVERSITY OF SINGAPORE


2009
YIELDING AND FAILURE OF CEMENT TREATED SOIL

XIAO HUAWEN
(B.Eng., M.E., HHU)

A THESIS SUBMITTED

FOR THE DEGREE OF DOCTOR OF PHILOSOPHY

DEPARTMENT OF CIVIL ENGINEERING

NATIONAL UNIVERSITY OF SINGAPORE

2009
Dedicated to my wife and daughter

i
ACKNOWLEDGEMENTS

The author wishes to express his profound gratitude and sincere appreciation to

his supervisor, Professor Lee Fook Hou for the valued advice, constructive criticisms

and endless guidance throughout this research study. Without his help, this research

work could not have been accomplished.

Grateful acknowledgement is given to the technical staffs who have assisted the

author in the experimental studies. They are, Mdm. Jamilah Bte Mohd, Mr. Foo Hee

Ann, Mr. John Choy Moon Nien, and Mr. Tan Lye Heng. Sincere appreciation is also

expressed to Dr Chew Soon Hoe, the lab supervisor.

The author also deeply appreciates the financial assistance in the form of research

scholarship as well as facilities provided by the National University of Singapore to

perform his research study.

Acknowledgements are also due to:

(a) Dr. Chin Kheng Ghee who gave grate help at the beginning of the research.

(b) Fellow colleagues of NUS, in particular Dr Shen Ruifu, Dr. Xie Yi, Dr. Cheng

Yonggang, Dr. Zhou Xiaoxian, Mr. Sun Daojun, Dr. Sindhu Tjahyono, Dr. Yeo Chong

Hun, Dr. Subhadeep Banerjee, Dr. Liu Xuemei, Dr. Ma Kang, Dr. Wang Zhengrong,

Mr. Vincent, Mr. Harrish, Miss Charlene, Mr. Meas; Mr. Isaac.

Finally, the author would like to express special appreciation to his wife Ms Liu

Wenyan for her always care, support, and encouragement and selfless accompany

during these years. Without her help, the author could not come through his research.

ii
TABLE of CONTENTS

Dedication i
Acknowledgements ii
Table of Contents iii
Summary x
List of Tables xiii
List of Figures xiv
List of Symbols xxxii
1 Introduction 1

1.1 Background 1

1.2 Behavior of cement-treated soil 3

1.2.1 Behavior of natural and artificially lightly cemented soil 4

1.2.2 Behavior of cement-treated sand and clay 5

1.2.3 Behavior of cement-treated Singapore marine clay 6

1.3 Objectives and organization of the thesis 6

2 LITERATURE REVIEW 9

2.1 Introduction 9

2.2 Basic concepts and mechanism of cement stabilization 9

2.2.1 Mechanism of cement stabilization 9

2.2.2 Structure and microstructure of treated soil 11

2.3 Factors on the strength of cement-treated soil 13

2.3.1 Characteristics of stabilizing agents 14

iii
2.3.2 Characteristics of soil 15

2.3.3 Mixing conditions 16

2.3.4 Curing conditions 16

2.4 Experimental studies on cement-treated soil 17

2.4.1 Changes in basic property of cement-treated soil 17

2.4.2 Engineering behavior 18

2.4.2.1 Prime factors on strength and deformation of 18


cement-treated soil

2.4.2.2 Permeability 20

2.4.2.3 Compressibility 20

2.4.2.4 Stiffness 22

2.4.2.5 Shear strength parameters 22

2.4.3 Stress-strain behavior of cement-treated soil under 23


triaxial condition

2.4.4 Studies on cement-treated Singapore marine clay 24

2.4.5 Primary yielding and post-yield behavior of cemented 26


soil

2.4.6 Softening and post-peak shearing behavior of cemented 29


soil

2.5 Theoretical studies 32

2.5.1 Review of theoretical studies for natural or artificial 32


cemented soil

2.5.2 Constitutive frameworks for cement-treated soil – a 36


summary

2.6 Outstanding issues 37

2.6.1 Outstanding issues 37

2.6.2 Scope of work in the current Study 38

iv
3 EXPERIMENTAL METHODOLOGY AND SETUP 57

3.1 Materials 57

3.1.1 Untreated marine clay 57

3.1.2 Ordinary Portland cement 57

3.2 Variables investigated 58

3.3 Sample preparation procedure 60

3.4 Testing procedure and apparatus 62

3.4.1 Basic properties 63

3.4.2 Isotropic compression 63

3.4.3 Triaxial compression and strength 63

3.4.4 Microstructural properties 65

3.4.5 Tensile splitting strength test 66

3.4.6 Remoulded cement-treated marine clay 67

4 PREPEAK BEHAVIOR OF CEMENT-TREATED SINGAPORE 77


MARINE CLAY

4.1 Purpose of experiment 77

4.2 Definition of primary yielding and yield locus for 81


cement-treated soil

4.2.1 Definition of artificial structure 81

4.2.2 Definition of yielding and yield locus for cemented soils 82

4.2.3 Summary 89

4.3 Isotropic compression behavior of cement-treated soil 89

4.3.1 General behavior under isotropic compression 89

4.3.2 Cement content effect 90

v
4.3.3 Total water content effect 90

4.3.4 Curing stress effect 91

4.3.5 Curing period effect 92

4.3.6 Summary 92

4.4 Triaxial compression behavior of samples consolidated at 93


pressures below the primary isotropic yield stress p 'py

4.4.1 General behavior under triaxial compression 93

4.4.2 Cement content effect 95

4.4.3 Total water content effect 95

4.4.4 Curing stress effect 96

4.4.5 Curing period effect 96

4.4.6 Summary 97

4.5 Yielding behavior of cement-treated Singapore marine clay 97

4.5.1 Primary yielding and yield locus 98

4.5.2 Relation of primary yield locus to other parameters 99

4.5.2.1 Correlation of isotropic yield stress to 100


unconfined compressive strength

4.5.2.2 Correlation of isotropic yield stress to 101


post-curing void ratio

4.5.2.3 Direct correlation of isotropic yield stress to 102


cement and total water content

4.5.2.4 Correlation of unconfined compressive 103


strength to mix proportions

4.5.2.5 Summary 105

4.6 Triaxial compression behavior of samples consolidated at 105


effective pressure higher than the isotropic primary yield
stress p 'py

vi
4.6.1 General behavior under triaxial compression 106

4.6.2 Evolution of yield locus 108

4.7 Tensile splitting strength test and modification of primary yield 109
locus

4.7.1 Stresses along radial loading section 109

4.7.2 Tensile strength of cement-treated marine clay 110

4.7.3 Correlation of tensile strength to unconfined 112


compressive strength

4.7.4 Modification of primary yield locus of cement-treated 112


marine clay under triaxial loading condition

5 ULTIMATE STATE OF CEMENT-TREATED MARINE CLAY 193

5.1 Introduction 193

5.2 Post-peak behavior 195

5.2.1 Onset of strain softening 195

5.2.2 Special specimen with lubrication and radial excess 196

5.2.2.1 Effect of slenderness ratio 198

5.2.2.2 Effect of lubrication 199

5.2.2.3 Effect of combination of slenderness and 199


lubrication

5.2.2.4 Effect of enlarged low-friction end caps 200

5.2.2.5 Stress state at post-peak stage 200

5.2.2.6 Results of CID short specimens 202

5.2.3 Critical state and residual state 203

5.3 Constitutive behavior of remoulded cement-treated marine clay 205

5.3.1 Effects of remoulding on cement-treated marine clay – 205


previous works

vii
5.3.2 Isotropic compression behavior of remoulded 208
cement-treated marine clay

5.3.3 Triaxial compression behavior 211

5.3.3.1 Remoulded marine clay 211

5.3.3.2 Remoulded cement-treated marine clay with 212


100% pre-remoulding total water content

5.3.3.3 Remoulded treated marine clay with 50% 216


cement content –effect of pre-remoulding total
water content

5.3.3.4 Remoulded treated marine clay with 50% 217


cement content and 133% pre-remoulding total
water content –effects of curing stress

5.3.3.5 Summary of findings 218

5.4 Discussion 219

6 A CONSTITUTIVE MODEL FOR CEMENT-TREATED 269


MARINE CLAY

6.1 Introduction 269

6.1.1 Expanded structure surface approach 270

6.1.2 Superposition method 275

6.1.3 Other methods 278

6.1.4 Difficulties in using existing models for cement-treated 278


marine clays

6.2 An empirical constitutive framework for cement-treated marine 279


clay

6.2.1 Empirical yield function one 280

6.2.2 Empirical yield function two 281

6.3 A theoretical constitutive framework for cement-treated marine 282


clay

6.3.1 Basis of theoretical model 286

viii
6.3.2 Yield locus 288

6.3.2.1 Deduction of yield locus 288

6.3.2.2 Consistency checks 291

6.3.2.3 Comparison with test data 292

6.4 Parameters of theoretical yield function 294

6.4.1 Friction coefficient M 295

6.4.2 Preconsolidation stress p0' 295

6.4.3 Determination of cohesion 298

6.4.3.1 Determination of initial cohesion C0 299

6.4.3.2 Cohesion degradation 301

7 CONCLUSIONS 329

7.1 Conclusions 329

7.1.1 Pre-peak behavior of cement-treated Singapore marine 329


clay

7.1.2 Ultimate state of cement-treated marine clay 332

7.1.3 Constitutive frame work of cement-treated marine clay 334

7.2 Recommendations for future study 335

REFERENCES 337

ix
SUMMARY

This study deals with the development of a constitutive framework for cement-treated

Singapore marine clay. Element testing was conducted for a wide range of mix proportions and

different curing conditions to shed light on the constitutive behavior of the cement-treated clay.

Based on the experimental results, common trends were identified which forms the basis of a

framework of behavior of this material. Finally, a constitutive framework was postulated for

the primary yielding and early post-yield behavior of cement-treated marine clay.

Coop and Atkinson’s method was introduced to define the primary yield stress and the

primary yield locus. The results showed that a reasonably consistent primary yield locus of

cement-treated marine clay can be obtained.

The results of isotropic compression tests showed that the post-yield compression index

seems to be independent on cement content and curing period. The post-yield compression

index is also independent on total water content and curing stress. However, the isotropic

primary yielding stress increases with the increase in cement content, curing load and curing

time as well as decrease in total water content.

The experiment results showed that behaviors under triaxial compression are consistent

with Chin’s (2006) study. The results also showed that increasing the cement content, curing

stress and curing period or decreasing the total water content all has a similar effect in

increasing the peak strength of the specimens. The post-yield behavior of the cement-treated

soil appears to be influenced by densification effects as well as breakage of inter-aggregate

bonds. The yield locus of the cement-treated marine clay evolves into a shape which is

well-fitted by an ellipse.

x
Experiment results suggest that excess pore pressure and stress ratio may be better

indicators for shear band initiation than the deviator stress itself. The short specimens with

enlarged low-friction end caps showed significantly slower rate of strain softening and a more

uniform post-peak behavior than that of conventional specimens showing a single shear band.

The short specimens probably reached a critical state at a shear strain of about 20%, with a

friction coefficient much higher than that of conventional long specimen.

The experiment results showed that the isotropic compression curve of the remoulded

treated marine clay depends mainly on the cement content and lies between the compression

curves of the untreated marine clay and the corresponding intact treated marine clay. The

undrained stress path and deviator stress-strain of remoulded cement-treated marine clay is

similar to that of remoulded untreated marine clay. The excess pore pressure changes very

small after peak point. Base on the experiment study, the remoulded cement-treated marine

clay may be considered as a reference “remoulded” state towards which the microstructure of

the cement-treated clay will evolve with continual shearing. The ultimate state of remoulded

cement-treated specimen seems to be only dependent on the cement content, thereby making it

a convenient reference state.

Based on the observation that the cement-treated soils do have a structure surface (or

primary yield surface) but that this yield surface can change shape and size during shearing, a

new theoretical approach was attempted. In the theoretical framework, a new yield function

was derived by introducing true cohesion parameter into a modified form of the modified Cam

Clay energy equation. The proposed yield function can fit well the observed primary yield

locus and evolution of yield locus for cement-treated marine clay specimens. The parameters

xi
of the proposed model can be obtained from conventional test.

Keywords: cement-treated marine clay, primary yield, post-yield, cohesion, remoulded

state, constitutive framework

xii
LIST OF TABLES

1.1 Typical properties of Singapore marine clay 2

2.1 Factors affecting the strength increase (after Terashi, 1997) 14

3.1 Basic properties of Singapore upper marine clay 57

3.2 Chemical compositions and physical properties of Ordinary Portland 58


Cement

3.3 Mixture proportions and experiments in this study 60

3.4a Basic and engineering properties in this study 68

3.4b Basic and engineering properties in this study 69

4.1 Summary of tensile splitting strength test results 114

5.1 Summary of Friction Coefficient and Microstructure after Different 220


Processes

xiii
LIST OF FIGURES

2.1 Schematic illustrations of improved soil (after Saitoh et al., 1985) 40

2.2 SEM micrographs of lime improved soil (after Locat et al., 1990) 40

2.3 Microstructures on slip planes of several cement-treated specimens at the 41


end of triaxial shearing tests: CID50; CIU50; CIU500;
CIU1500; CID500; CID1500 (after Chin, 2006)

2.4 Effect of cement type on compressive strength of soil-cement for: (a) 42


Kanagawa; and (b) Saga soils (after Kawasaki et al., 1981)

2.5 Effect of different stabilizers on compressive strength of different soils in 42


Sweden (after Ahnberg et al., 1995)

2.6 Effect of grain size distribution on cement stabilization (after Niina et al., 43
1977)

2.7 Effect of organic content on the unconfined compressive strength of peat 43


soils (after Huat, 2006)

2.8 Effect of mixing time on cement stabilization (after Nakamura et al., 44


1982)

2.9 Effect of curing time on strength (after Kawasaki et al., 1981) 44

2.10 Effect of curing temperature on compressive strength of silt (after Enami 45


et al., 1985)

2.11 UCT results for specimens treated under drained isotropic curing stress 45
(after Chin, 2006)

2.12 Plasticity chart for cement treated clay under various stress states during 46
curing period (after Chin, 2006)

2.13 Effect of initial water content and cement content on e-logσ’v 46


relationship (after Kamruzzaman, 2002)

2.14 Isotropic compression behavior for cement treated clay under various 47
stress states during curing period (after Chin, 2006)

xiv
2.15 Peak strength envelope for treated specimens with OCR=1 in p’-q 47
stress plane (after Chin, 2006)

2.16 Triaxial behaviour of treated clay under (a) drained; and (b) undrained 48
conditions (after Endo, 1976)

2.17 Consolidated drained triaxial behaviour with different confining 48


pressures (after Tatsuoka and Kobayashi, 1983)

2.18 Undrained stress path of cement-treated clay for (a) 10%; (b) 30%; and 49
(c) 50% of cement contents (after Kamruzzman, 2002)

2.19 Definition of yield surfaces (after Smith, 1992) 50

2.20 Variation of incremental yield stress with curing void ratio and cement 50
content(after Rotta et al, 2003)

2.21 Relationship between initial curing void ratio, unconfined compressive 51


strength, incremental yield stress and initial bulk density (after Consoli et
al, 2006)

2.22 Undrained triaxial compression of Todi clay: stress-strain response, pore 52


water pressure difference and strain homogeneity indexes (after Viggiani
et al,1993)

2.23 Size effect on triaxial behavior of specimen with height-diameter ratio 52


1(after Abdulla & Panos, 1997)

2.24 Effect of slenderness ratio on drained triaxial behavior of specimen with 53


cement content 4% and confining stress 200kPa(after Abdulla & Panos,
1997)

2.25 Effect of the slenderness of the specimen on the results obtained from 53
constant p’ triaxial test (after Callisto & Calabresi, 1998)

2.26 Post-ruptured strength envelope of samples with mix proportion 54


s:c:w=5:1:6 (after Chin, 2006)

2.27 Successive yield surfaces for increasing degrees of bonding. Surface A 54


corresponds to unbonded material (after Gens and Nova, 1993)

2.28 Reference surface, structure surface and bubble (yield surface) for 55
destructuration model (after Rouaina & Wood, 2000)

xv
2.29 Representation of composite action of frictional and bond forces in 55
cemented soil (after Vatsala et al., 2001)

2.30 yield surface for cement treated clay (after Lee et al., 2004) 55

2.31 A new family of yield locus (after McDowell, 2000) 56

3.1 Soil-cement and water-cement ratios for some previous studies on deep 70
mixing and jet grouting (after Lee, 2005) and for this study

3.2 Working ranges of dried-pulverized and slurry clay cement mixes (after 70
Lee, 2005, mix proportion for this study were added into the same plot)

3.3 Distribution curve of air content (%) 71

3.4 A fully computer controlled triaxial stress path 71

3.5 Normal triaxial test apparatus (a) high pressure with GDS pressure 72
controller (b) lower pressure

3.6 Hitachi Tabletop microscope TM-1000 73

3.7 Tensile strength test setup (a) setup without sample (b) loading platform 74
with sample (c) loading platform with sample under water

3.8 Preparing remoulded cement-treated marine clay specimen (a) 75


cement-treated marine clay specimens, (b) cutting into pieces, (c) dried
pieces, (d) powder of remoulded specimen, (e) setup for preloading of
remoulded specimen, (f) preloading of remoulded specimen

4.1 Determination of initial yield stress 115

4.2 Determination of primary yield locus for cement-treated marine clay 117
specimen (mix proportion 10:3:13 cured under atmospheric pressure for
7days)

4.3 Isotropic compression curves for specimens with different cement 118
content and 100% total water content cured under atmospheric pressure
for 7 days

4.4 Effect of cement content on primary yielding stress under isotropic 118
compression for specimens with 100% total water content cured under
atmospheric pressure for 7 days

xvi
4.5 Isotropic compression curves for specimens with different total water 119
content and 50% cement content cured under atmospheric pressure for 7
days

4.6 Effect of total water content on primary yielding stress under isotropic 119
compression for specimens with 50% cement content cured under
atmospheric pressure for 7 days

4.7 Isotropic compression curves for specimens with mix proportion 10-1-11 120
cured under different curing stress for 7 days

4.8 Isotropic compression curves for specimens with mix proportion 20-3-23 120
cured under different curing stress for 7 days

4.9 Isotropic compression curves for specimens with mix proportion 5-1-6 121
cured under different curing stress for 7 days

4.10 Isotropic compression curves for specimens with mix proportion 2-1-4 121
cured under different curing stress for 7 days

4.11 Isotropic compression curves for specimens with mix proportion 2-1-5 122
cured under different curing stress for 7 days

4.12 Effect of curing stress on primary yielding stress under isotropic 122
compression for specimens with different mix proportion cured for 7
days

4.13 Isotropic compression curves for specimens with mix proportion 20-3-23 123
and different curing time under atmospheric pressure

4.14 Isotropic compression curves for specimens with mix proportion 5-1-6 123
and different curing time under atmospheric pressure

4.15 Isotropic compression curves for specimens with mix proportion 2-1-4 124
and different curing time under atmospheric pressure

4.16 Isotropic compression curves for specimens with mix proportion 2-1-4.5 124
and different curing time under atmospheric pressure

4.17 Isotropic compression curves for specimens with mix proportion 1-1-3 125
and different curing time under atmospheric pressure

4.18 Effect of curing time on primary yielding stress under isotropic 125
compression for specimens with different mix proportion cured under

xvii
atmospheric pressure

4.19 Undrained triaxial shearing behavior of specimens with mix proportion 126
10-1-11 cured under atmospheric pressure for 7 days (a) stress path (b)
deviator stress-strain curve

4.20 Drained triaxial shearing behavior of specimens with mix proportion 127
10-1-11 cured under atmospheric pressure for 7 days

4.21 Undrained triaxial shearing behavior of specimens with mix proportion 128
20-3-23 cured under atmospheric pressure for 7 days (a) stress path (b)
deviator stress-strain curve

4.22 Drained triaxial shearing behavior of specimens with mix proportion 129
20-3-23 cured under atmospheric pressure for 7 days

4.23 Undrained triaxial shearing behavior of specimens with mix proportion 130
5-1-6 cured under atmospheric pressure for 7 days (a) stress path (b)
deviator stress-strain curve

4.24 Drained triaxial shearing behavior of specimens with mix proportion 131
5-1-6 cured under atmospheric pressure for 7 days

4.25 Undrained triaxial shearing behavior of specimens with mix proportion 132
10-3-13 cured under atmospheric pressure for 7 days (a) stress path (b)
deviator stress-strain curve

4.26 Drained triaxial shearing behavior of specimens with mix proportion 133
10-3-13 cured under atmospheric pressure for 7 days

4.27 Undrained triaxial shearing behavior of specimens with mix proportion 134
2-1-3 cured under atmospheric pressure for 7 days (a) stress path (b)
deviator stress-strain curve

4.28 Drained triaxial shearing behavior of specimens with mix proportion 135
2-1-3 cured under atmospheric pressure for 7 days

4.29 Undrained triaxial shearing behavior of specimens with mix proportion 136
2-1-4 cured under atmospheric pressure for 7 days (a) stress path (b)
deviator stress-strain curve

4.30 Drained triaxial behaviors of specimens with mix proportion 2-1-4 cured 137
under atmospheric pressure for 7 days

xviii
4.31 Undrained triaxial shearing behavior of specimens with mix proportion 138
2-1-5 cured under atmospheric pressure for 7 days (a) stress path (b)
deviator stress-strain curve

4.32 Drained triaxial shearing behavior of specimens with mix proportion 139
2-1-5 cured under atmospheric pressure for 7 days

4.33 Undrained triaxial shearing behavior of specimens with mix proportion 140
2-1-5.5 cured under atmospheric pressure for 7 days (a) stress path (b)
deviator stress-strain curve

4.34 Drained triaxial shearing behavior of specimens with mix proportion 141
2-1-5.5 cured under atmospheric pressure for 7 days

4.35 Undrained triaxial shearing behavior of specimens with mix proportion 142
2-1-4 cured under 50kPa effective confining pressure for 7 days (a) stress
path (b) deviator stress-strain curve

4.36 Drained triaxial shearing behavior of specimens with mix proportion 143
2-1-4 cured under 50KP effective confining stress for 7 days

4.37 Undrained triaxial shearing behavior of specimens with mix proportion 144
2-1-4 cured under 100kPa effective confining pressure for 7 days (a)
stress path (b) deviator stress-strain curve

4.38 Drained triaxial shearing behavior of specimens with mix proportion 145
2-1-4 cured under 100KP effective confining stress for 7 days

4.39 Undrained triaxial shearing behaviors of specimens with mix proportion 146
2-1-4 cured under 250kPa effective confining pressure for 7 days (a)
stress path (b) deviator stress-strain curve

4.40 Drained triaxial shearing behavior of specimens with mix proportion 147
2-1-4 cured under 250KP effective pressure for 7 days

4.41 Cement content effect on undrained stress-strain behavior of specimens 148


cured under atmospheric pressure for 7 days of curing period (a) stress
path (b) deviator stress-strain curve

4.42 Cement content effect on drained stress-strain behavior of specimens 149


cured under atmospheric pressure for 7 days of curing period

4.43 Total water content effect on undrained stress-strain behavior of 150


specimens cured under atmospheric pressure for 7 days of curing period

xix
(a) stress path (b) deviator stress-strain curve

4.44 Total water content effect on drained stress-strain behavior of specimens 151
cured under atmospheric pressure for 7 days of curing period

4.45 Curing stress effect on stress-strain behavior of specimens with mix 152
proportion 2-1-4 cured for 7 days (a) stress path (b) deviator stress-strain
curve

4.46 Curing stress effect on drained stress-strain behavior of specimens with 153
mix proportion 2-1-4 cured for 7 days

4.47 Curing time effect on undrained triaxial behavior of specimens with mix 154
proportion 2-1-4 cured under atmospheric pressure (a) stress path (b)
deviator stress-strain curve

4.48 Curing time effect on drained triaxial behavior of specimens with mix 155
proportion 2-1-4 cured under atmospheric pressure

4.49 Curing time effect on undrained triaxial behavior of specimens with mix 156
proportion 51-6 cured under atmospheric pressure (a) stress path (b)
deviator stress-strain curve

4.50 Curing time effect on drained triaxial behavior of specimens with mix 157
proportion 5-1-6 cured under atmospheric pressure

4.51 Primary yield loci normalized by isotropic primary yield stress for 158
specimens with different mix proportion and curing stress and cured for
7 days(a) fitted by function (4-1) (b) fitted by function (4-2)

4.52 Primary yield loci normalized by isotropic primary yield stress with 159
different curing period (a) fitted by function (4-1) (b) fitted by function
(4-2)

4.53 Unconfined compressive strength tests for specimens with different 160
curing stress and 7 days of curing time period (a-d)

4.54 Unconfined compressive strength tests for specimens with different 161
curing stress and 7 days of curing time period (e-h)

4.55 Unconfined compressive strength tests for specimens with different 162
curing time and zero curing stress (a-d)

4.56 Unconfined compressive strength tests for specimens with different 163

xx
curing time and zero curing stress (e-g)

4.57 Relationship between Isotropic primary yield stress and unconfined 164
compressive strength for specimens (a) with 7 days of curing and (b)
different curing time period

4.58 Relationship between incremental Isotropic primary yield stress and 165
post-curing void ratio for cement-treated marine clay specimens cured
for 7 days

4.59 Relationship between parameters and ratio of pre-curing total water 165
content to the cement content Cw/Aw (a) A1-Cw/Aw, (b) B1-Cw/Aw

4.60 Relationship between Isotropic primary yield stress and cement content 167
and total water content for cement treated specimens cured under
atmospheric pressure for 7days (a) p 'py versus cement content (b)

parameter A2 versus total water content (c) parameter B2 versus total


water content (d) comparison between experimental data and simulation
results by using formula (4-11)

4.61 Comparison between unconfined compressive strength obtained from 168


this study and Lee’s study (1999)

4.62 Simulated unconfined compressive strength by using Lee et al.’s (2005)’s 169
formula and this study’s formula

4.63a Simulated unconfined compressive strength by using modified formula 170


of this study and experimental unconfined compressive strength from
Lee’s (1999) study

4.63b Simulated unconfined compressive strength by using Lee’s (2005) 170


formula with different q0 and experimental unconfined compressive
strength from this study

4.64a Post-yield stress-strain behavior of CIU specimen with mix proportion 171
10:1:11 cured under atmospheric pressure for 7days (a) stress path, (b)
deviator stress-strain curve

4.64b Post-yield stress-strain behavior of CID specimen with mix proportion 172
10:1:11 cured under atmospheric pressure for 7days

4.65a Post-yield stress-strain behavior of CIU specimen with mix proportion 173
5:1:6 cured under atmospheric pressure for 7days (a) stress path, (b)

xxi
deviator stress-strain curve

4.65b Post-yield stress-strain behavior of CID specimen with mix proportion 174
5:1:6 cured under atmospheric pressure for 7days

4.66a Post-yield stress-strain behavior of specimen with mix proportion 175


10:3:13 cured under atmospheric pressure for 7days (a) stress path, (b)
deviator stress-strain curve

4.66b Post-yield stress-strain behavior of specimen with mix proportion 176


10:3:13 cured under atmospheric pressure for 7days

4.67a Post-yield stress-strain behavior of specimen with mix proportion 2:1:3 177
cured under atmospheric pressure for 7days (a) stress path, (b) deviator
stress-strain curve

4.67b Post-yield stress-strain behavior of specimen with mix proportion 2:1:3 178
cured under atmospheric pressure for 7days

4.68a Post-yield stress-strain behavior of CIU specimen with mix proportion 179
2:1:4 cured under atmospheric pressure for 7days (a) stress path, (b)
deviator stress-strain curve

4.68b Post-yield stress-strain behavior of CID specimen with mix proportion 180
2:1:4 cured under atmospheric pressure for 7days

4.69 Evolution of yield locus for cement treated marine clay specimens on 183
normalized stress space (a) mix proportion 10:1:11 (b) 5:1:6 (c) 10:3:13
(d) 2:1:3 (e) 2:1:4

4.70 Comparison between evolutions of yield locus for cement treated marine 184
clay specimens on normalized stress space (a) different cement content
(b) different total water content

4.71 Microstructure of cement treated marine clay specimens with mix 185
proportion 5:1:6 under IPC pressure 2500kPa and drained shearing
(CID2500-1250) (a) outside of slip band (b) inside of slip band

4.72 Microstructure of cement treated marine clay specimens with mix 186
proportion 5:1:6 under IPC pressure 2500kPa and undrained shearing
(2500-1250CU) (a) outside of slip band (b) inside of slip band

4.73 Theoretical solutions to the stress distribution of cylinder specimen with 187
thickness t and diameter D under two radial and opposite line loading

xxii
(after Timoshenko and Goodier, 1970)

4.74 The failure pattern of cement-treated marine clay cylinder specimen 187
(a)split specimen (b) top loading face after Brazilian test (c) specimen
with central crack

4.75 The tensile splitting strength of cement-treated marine clay cylinder 188
specimen (a) tensile strength versus cement content (b) tensile strength
versus total water content (c) tensile strength versus curing stress

4.76 Comparisons between very fast and vey slow compression rate 189

4.77 Comparisons between compression under dry condition and water 189

4.78 Prolonged submerged test (a) remoulded marine clay sample (b) 190
cement-treated marine clay sample

4.79 Relationship between tensile strength and unconfined compressive 190


strength

4.80 Modified primary yield locus for cement-treated marine clay specimens 191
with different cement content and under triaxial loading

4.81 Modified primary yield locus for cement-treated marine clay specimens 191
with different total water content and under triaxial loading

4.82 Modified primary yield locus for cement-treated marine clay specimens 192
with different curing stress and under triaxial loading

4.83 Modified primary yield locus for cement-treated marine clay specimens 192
with different curing time period and under triaxial loading

5.1 Undrained triaxial compression of Todi clay: stress-strain response, pore 222
water pressure difference and strain homogeneity indexes(after Viggiani
et al,1993a)

5.2 Localization of cement-treated marine clay specimens with mix 223


proportion 2-1-4 in CIU test (a) 50CIU, (b) 100CIU, (c) 250CIU

5.3 Localization of cement-treated marine clay specimens (mix proportion 224


2-1-4) in CIU test with maximum consolidation stress 500kPa (a)
500-50CIU, (b) 500-250CIU

5.4 Localization of cement-treated marine clay specimens (mix proportion 225

xxiii
2-1-4) in CIU test with maximum consolidation stress 2000kPa (a)
2000-50CIU, (b) 2000-400CIU

5.5 Localization of cement-treated marine clay specimens (mix proportion 226


5-1-6) in CIU test with maximum consolidation stress 50kPa, 1000kPa
and 2000kPa respectively (a) 50CIU, (b) 1000-250CIU, (c) 2000-50CIU

5.6 Effect of slenderness ratio (ratio of height to diameter H/D) on 227


specimens in CIU test (mix proportion 2-1-4, with confining stress
1500kPa, lateral space Ls=0, sample diameter is the same as that of caps)

5.7 Samples after shearing in CIU test (mix proportion 2-1-4) 227

5.8 Effect of end restraint (lubrication) on specimens in CIU test (mix 228
proportion 2-1-4, with confining stress 1000kPa, Ls=0, sample diameter
is the same as that of caps), slenderness ratio = 2.

5.9 Samples after shearing in CIU test (mix proportion 2-1-4) 228

5.10 Effect of slenderness and end restraint on specimens in CIU test (mix 229
proportion 2-1-4, with confining stress 1000kPa)

5.11 Samples after shearing in CIU test (a) Specimen H/D=1, porous stone, 229
Ls=0, (b) Specimen H/D=1, Teflon, Ls=0 (c) Specimen -H/D=1, Teflon,
Ls=6mm

5.12 Effect of slenderness, end restraint and lateral space on specimens in 230
CIU test (mix proportion 2-1-4, specimen 1 and 2 with confining stress
1000kPa, specimen 3 and 4 with confining stress 500kPa)

5.13 Post-peak stress statuses at different range of axial strain in CIU test 231
(mix proportion 2-1-4, with confining stress 1000kPa)

5.14 Post-peak stress statuses at very larger axial strain in CIU test (mix 232
proportion 2-1-4, with confining stress 1000kPa)

5.15 Short samples with enlarged low friction platens after shearing to axial 233
strain of about 55% in CIU test (a) confining stress 1000kPa (b)
confining stress 500kPa

5.16 Comparison between conventional specimen and short specimen in CID 233
test (mix proportion 2-1-4, effective confining stress 250kPa)

5.17 Comparison between conventional specimen and short specimen in CID 234

xxiv
test (mix proportion 2-1-4, effective confining stress 500kPa)

5.18 Critical state of conventional specimen and short specimen 235

5.19 Microstructure of cement treated marine clay short specimens with mix 236
proportion 2:1:4 under effective confining pressure 500kPa and drained
shearing (500CID)

5.20 Isotropic compression curves for remoulded cement-treated marine clay 237
specimens with different cement content and 100% total water content
cured under atmospheric pressure for 7 days before remoulding

5.21 Isotropic compression curves for remoulded and unremoulded 237


cement-treated marine clay specimens and remoulded marine clay

5.22 Isotropic compression curves for remoulded cement-treated marine clay 238
specimens with different total water content and 50% cement content
cured under atmospheric pressure for 7 days before remoulding

5.23 Isotropic compression curves for remoulded cement-treated marine clay 238
specimens with different curing stress and 50% cement content and
133% total water content before remoulding

5.24 Isotropic compression curves for remoulded cement-treated marine clay 239
specimens with different cement content and curing time before
remoulding

5.25 Undrained triaxial shearing behaviors of remoulded marine clay 240


specimens with 70 kPa preconsolidation pressure

5.26 Undrained triaxial shearing behavior of remoulded cement-treated 241


marine clay specimens with 10% cement content and 70 kPa
preconsolidation pressure

5.27 Undrained triaxial shearing behavior of remoulded cement-treated 242


marine clay specimens with 10% cement content and different
preconsolidation pressure

5.28 Undrained triaxial shearing behavior of remoulded cement-treated 243


marine clay specimens with 10% cement content and 70kPa
preconsolidation pressure and different curing period

5.29 Undrained triaxial shearing behavior of remoulded cement-treated 244


marine clay specimens with 10% cement content and 44kPa

xxv
preconsolidation pressure and different curing period

5.30 Undrained triaxial shearing behavior of remoulded cement-treated 245


marine clay specimens with 20% cement content and 70 kPa
preconsolidation pressure

5.31 Undrained triaxial shearing behavior of remoulded cement-treated 246


marine clay specimens with 20% cement content and different
preconsolidation pressure

5.32 Undrained triaxial shearing behavior of remoulded cement-treated 247


marine clay specimens with 20% cement content and 70kPa
preconsolidation pressure and different curing periods

5.33 Undrained triaxial shearing behavior of remoulded cement-treated 248


marine clay specimens with 20% cement content and 44kPa
preconsolidation pressure and different curing periods

5.34 Undrained triaxial shearing behavior of remoulded cement-treated 249


marine clay specimens with 30% cement content and 70 kPa
preconsolidation pressure

5.35 Undrained triaxial shearing behavior of remoulded cement-treated 250


marine clay specimens with 30% cement content and different
preconsolidation pressure

5.36 Undrained triaxial shearing behavior of remoulded cement-treated 251


marine clay specimens with 30% cement content and 70kPa
preconsolidation pressure and different curing periods

5.37 Undrained triaxial shearing behavior of remoulded cement-treated 252


marine clay specimens with 30% cement content and 44kPa
preconsolidation pressure and different curing periods

5.38 Undrained triaxial shearing behavior of remoulded cement-treated 253


marine clay specimens with 50% cement content and 70 kPa
preconsolidation pressure

5.39 Undrained triaxial shearing behavior of remoulded cement-treated 254


marine clay specimens with different cement content and 70 kPa
preconsolidation pressure

5.40 Simulation of stress path of remoulded cement-treated marine clay 255


specimens with 50% cement content by modified Cam clay model

xxvi
5.41 Undrained triaxial shearing behavior of remoulded cement-treated 256
marine clay specimens with 50% cement content and different
preconsolidation pressure

5.42 Undrained triaxial shearing behavior of remoulded cement-treated 257


marine clay specimens with 50% cement content and 70kPa
preconsolidation pressure and different curing period

5.43 Undrained triaxial shearing behavior of remoulded cement-treated 258


marine clay specimens with 50% cement content and 44kPa
preconsolidation pressure and different curing period

5.44 Experiment and simulated Stress paths of remoulded cement-treated 259


marine clay specimens with preloading 44kPa (from cement-treated
marine clay with 50% cement content and different total water content,
cured under atmospheric pressure for 7 days)

5.45 Stress-strain curve of remoulded cement-treated marine clay specimens 260


with preloading 44kPa (from cement-treated marine clay with 50%
cement content and different pre-remoulding total water content, cured
under atmospheric pressure for 7 days)

5.46 Excess pore pressure-strain curve of remoulded cement-treated marine 261


clay specimens with preloading 44kPa (from cement-treated marine clay
with 50% cement content and different pre-remoulding total water
content, cured under atmospheric pressure for 7 days)

5.47 Undrained triaxial shearing behavior of remoulded cement-treated 262


marine clay specimens with preconsolidation pressure 70kPa (remoulded
from cement-treated marine clay with 50% cement content and 133%
total water content and cured under atmospheric pressure for 7 days)

5.48 Undrained triaxial shearing behavior of remoulded cement-treated 263


marine clay specimens with preconsolidation pressure 70kPa (remoulded
from cement-treated marine clay with 50% cement content and 133%
total water content and 50kPa curing stress)

5.49 Undrained triaxial shearing behavior of remoulded cement-treated 264


marine clay specimens with preconsolidation pressure 70kPa (remoulded
from cement-treated marine clay with 50% cement content and 133%
total water content and 100kPa curing stress)

5.50 Uundrained triaxial shearing behavior of remoulded cement-treated 265

xxvii
marine clay specimens with preconsolidation pressure 70kPa (remoulded
from cement-treated marine clay with 50% cement content and 133%
total water content and 250kPa curing stress)

5.51 Microstructure at post-rupture states, as shown in the mechanical 266


behaviour (after Chin (2006))

5.52 Comparison between SEM of remoulded cement-treated marine clay and 267
cement-treated marine clay after drained shearing test (a)remoulded
cement-treated marine clay specimen, (b)rupture plane of cement-treated
marine clay (CID2500-1250)

6.1 Relationship between stress ratio at the top of the yield locus(empirical 304
method one) and plastic volumetric strain for different mix proportion (a)
10-1-11, (b) 5-1-6, (3) 10-3-13, (d) 2-1-3, (e) 2-1-4, (f) for all mix
proportions

6.2 Simulated yield loci through empiric method one with experimental data 307
for different mix proportion (a) 10-1-11, (b) 5-1-6, (3) 10-3-13, (d) 2-1-3,
(e) 2-1-4

6.3 Relationship between stress ratio at the top of the yield locus(empirical 308
method two) and plastic volumetric strain for different mix proportion
(a) 10-1-11, (b) 5-1-6, (3) 10-3-13, (d) 2-1-3, (e) 2-1-4, (f) for all mix
proportions

6.4 Simulated yield loci through empiric method two with experimental data 311
for different mix proportion (a) 10-1-11, (b) 5-1-6, (3) 10-3-13, (d) 2-1-3,
(e) 2-1-4

6.5 Simulated yield loci through empiric method one with tensile strength (a) 312
mix proportion 10-1-11, (b) mix proportion 10-3-13

6.6 313
Demonstration of proposed yield locus (a) p '− q plane (b) p '/ p0' − q / p0'

plane

6.7 Simulated primary yield loci for cement-treated marine clay specimens 314
with different cement content and cured under atmospheric pressure for 7
days

6.8 Simulated primary yield loci for cement-treated marine clay specimens 314
with different total water content and cured under atmospheric pressure
for 7 days

xxviii
6.9 Simulated primary yield loci for cement-treated marine clay specimens 315
with 50% cement content and 133% total water content and cured under
different confining pressure for 7 days

6.10 Simulated primary yield loci for cement-treated marine clay specimens 315
with 50% cement content and 133% total water content and cured under
atmospheric pressure for different curing time periods

6.11 Simulated evolution of yield locus for cement-treated marine clay 316
specimens with 10% cement content and 100% total water content and
cured under atmospheric pressure for 7days

6.12 Simulated evolution of yield locus for cement-treated marine clay 316
specimens with 20% cement content and 100% total water content and
cured under atmospheric pressure for 7days

6.13 Simulated evolution of yield locus for cement-treated marine clay 317
specimens with 30% cement content and 100% total water content and
cured under atmospheric pressure for 7days

6.14 Simulated evolution of yield locus for cement-treated marine clay 317
specimens with 50% cement content and 100% total water content and
cured under atmospheric pressure for 7days

6.15 Simulated evolution of yield locus for cement-treated marine clay 318
specimens with 50% cement content and 133% total water content and
cured under atmospheric pressure for 7days

6.16 Isotropic compression curves of remoulded and unremoulded 319


cement-treated marine clay specimens with 10% cement content

6.17 Isotropic compression curves of remoulded and unremoulded 319


cement-treated marine clay specimens with 20% cement content

6.18 Isotropic compression curves of remoulded and unremoulded 320


cement-treated marine clay specimens with 30% cement content

6.19 Isotropic compression curves of remoulded and unremoulded 320


cement-treated marine clay specimens with 50% cement content

6.20 Isotropic compression curves of remoulded and unremoulded 321


cement-treated marine clay specimens with 50% cement content and
133% total water content before remoulding

xxix
6.21 Idealization of the isotropic compression behavior of reconstituted and 321
structured soils

6.22 Relationships between q1 and qu 322

6.23 Comparison between Cot and Cos (Cot was obtained from experiment, 322
Cos was obtained by using qu and yield function)

6.24 Relationship between C0 and qu 323

6.25 Degradation of C with plastic work on normalized plane for 323


cement-treated marine clay (with mix proportion 10:1:11)
6.26 Degradation of inherent cohesion C for cement-treated specimen with 324
10% cement content and 100% total water content

6.27 Degradation of inherent cohesion C for cement-treated specimen with 324


20% cement content and 100% total water content

6.28 Degradation of inherent cohesion C for cement-treated specimen with 325


30% cement content and 100% total water content

6.29 Degradation of inherent cohesion C for cement-treated specimen with 325


50% cement content and 100% total water content

6.30 Degradation of inherent cohesion C for cement-treated specimen with 326


50% cement content and 133% total water content

6.31 Cohesion degradation with evolution of yield locus and plastic strain for 326
cement-treated marine clay specimens with 10% cement content and
100% total water content and cured under atmospheric pressure for
7days
6.32 Cohesion degradation with evolution of yield locus and plastic strain for 327
cement-treated marine clay specimens with 20% cement content and
100% total water content and cured under atmospheric pressure for
7days

6.33 Cohesion degradation with evolution of yield locus and plastic strain for 327
cement-treated marine clay specimens with 30% cement content and
100% total water content and cured under atmospheric pressure for
7days
6.34 Cohesion degradation with evolution of yield locus and plastic strain for 328
cement-treated marine clay specimens with 50% cement content and
100% total water content and cured under atmospheric pressure for

xxx
7days
6.35 Cohesion degradation with evolution of yield locus and plastic strain for 328
cement-treated marine clay specimens with 50% cement content and
133% total water content and cured under atmospheric pressure for
7days

xxxi
LIST OF SYMBOLS

Ac Cement content

Aw Total water cotent

Al O Aluminium oxide
2 3

C True or inherent cohesion

C0 Initial true or inherent cohesion

C0s Simulated value of initial true cohesion

C0t Experimental value of initial true cohesion

C´ Cohesion intercept

CS Di-calcium silicate
2

CA Tri-calcium aluminate
3

CS Tri-calcium silicate
3

C AF Tetra-calcium alumino-ferrite
4

Ca(OH) Calcium hydroxide, or lime


2

CAH Calcium aluminate hydrates

CaO Calcium Oxide

CASH Calcium Aluminate Silicate Hydrate

Cc compression index

CDM Cement Deep Mixing

CID Isotropically consolidated drained triaxial test

CIU Isotropically consolidated undrained triaxial test

CSH Calcium Silicate Hydrate

Cu Undrained shear strength

xxxii
e Void ratio

ecur Post-curing void ratio

e0 Initial void ratio

E Tangential elastic Young’s modulus

Fe O Iron oxide
2 3

G Tangent shear modulus

Gs Specific gravity

HO Water
2

ICL Intrinsic compression line, Burland (1990)

ICT Isotropic compression test

I Liquidity index
L

IOCR Isotropic overconsolidation ratio

K0 Coefficient of earth pressure at rest

K 2O Potassium Oxide

L/D Slenderness ratio, ratio of length to diameter

LL Liquid limit
'
M Stress ratio, q/p at critical state

MgO Magnesia

N Stress ratio at the top of yield locus

NaOH Sodium hydroxide

Na2O Sodium Oxide

NCL Normally Consolidation Line

P line loading along central vertical section

xxxiii
OPC Ordinary Portland Cement

PI Plasticity index

PL Plastic limit

p' Mean normal effective stress

p'1 The mean effective stress at the intersection between yield locus and apparent
tensile cut-off line

Pap atmospheric pressure

p'cur Effective isotropic curing stress

p'c Pre-shear effective confining pressure

p'0 Precompression pressure

p'0u Precompression pressure borne by the remoulded soil skeleton

p'u mean effective stress carried by the remoulded or unstructured cement-treated

q Deviator stress

q1 The deviator stress at the intersection between yield locus and apparent tensile
cut-off line

q0 Experimentally fitted value, Lee et al. (2005)’s framework

qu Unconfined compressive strength

S/C Soil-cement ratio

SCL sedimentation compression line

S:C:W Soil-cement-water ratio

SEM Scanning Electron Microscopy

SiO Silica
2

SO3 Sulphuric Anhydride

TST Tensile splitting strength test

xxxiv
u Excess pore water pressure

UCT Unconfined compressive test

V Specific volume, 1+e

VIU initial specific volume of unstructured soil when p0u


'
=1

Vpy specific volume of structured soil at primary yield stress

W plastic work

w Moisture content

Wvp plastic work due to plastic volumetric strain ε vp

Wsp plastic work due to plastic shear strain ε sp

w/c Water-cement ratio

γb Bulk density

γd Dry density

εa Axial strain

εs Deviatoric strain

εv Volumetric strain
'
η Stress ratio, q/p

κ Slope of unloading-reloading line in v-lnp' plane after initial-yield

κU Slope of unloading-reloading line in v-lnp' plane of unstructured soil

λ compression index in natural log scale ( λ )

λU isotropic compression index of the unstructured soil

σ'1,σ'3 Principle effective stresses

σt Tensile strength

σx stress along horizontal direction(x)

xxxv
σy stress along vertical direction(y)

∆ p'0s additional preconsolidation stress that can be resisted due to structure

∆ p's additional mean effective stress that can be resisted due to structure

∆ p'0si The initial value of additional preconsolidation stress that can be resisted due
to structure

ϕ' frictional angle

xxxvi
CHAPTER 1
INTRODUCTION

1.1 Background

About one quarter of Singapore Island is underlain by marine clays of the Kallang

Formation (Pitts, 1992). As noted by Tan et al. (2002), Singapore marine clay is a lightly over-

consolidated, structured soil. Details of the Singapore marine clay have been presented by Tan

(1983) and Yong et al. (1990), amongst others. Table 1.1 shows some of the engineering

properties of Singapore marine clay. As this Table shows, Singapore marine clay is

characterized by low undrained shear strength and high compressibility; these two properties

often give rise to a variety of geotechnical problems in construction, such as ground heave,

large soil settlement and collapse in foundation and sub-structure construction, which can

cause superstructural damage such as crack, tilt and collapse. Ground improvement techniques

such as sand and vertical drain and chemical stabilization have been used to improve the

properties of soft soils in foundation and sub-structure construction. In Singapore, one of the

most commonly used ground improvement schemes is chemical stabilization using cement

because cement is relatively abundant (compared to other chemicals), cheap and efficient

(Broms, 1984). The improvement process is also relatively fast compared to methods

involving consolidation and incurs little or no settlement to the surrounding ground. Dynamic

compaction cannot be used as the Singapore marine clay is highly compressible with low

permeability and the vibration due to dynamic compaction would have been unacceptable in a

densely built urban environment.

There are two main approaches to introduce cement into the soil matrix. They are cement

deep mixing (CDM) and jet grouting pile (JGP). The former introduces and mixes cement

slurry or powder into the soil matrix by a rotating mixing tool (e.g., Babasaki et al. 1991;

Bruce et al. 1998) whereas the latter involves breaking up the soil matrix by a high velocity

1
CHAPTER 1

grout or water jet with concurrent introduction of cement grout (e.g., Gallavresi 1992; Chia

and Tan 1993; Yong et al. 1996).

Table 1.1 Typical properties of Singapore marine clay


Geotechnical properties Tan(1983) Yong et al.(1990)
Upper Lower Upper Lower
member member member member
Unit weight(kN/m3) 14.0-16.0 15.0-17.5 15.0-16.0 15.0-17.5
Water content (%) 60-110 47-70 60-90 50-70
Atterberg limits Liquid limit (%) 75-115 63-80 80-120 60-90
Plasticity index 50-77 39-55 50-80 40-55
(%)
Undrained shear strength Undrained 5-20 8-50 … …
Su(kPa) unconsolidated
triaxial test
Field vane 8-40 35-50 … …
Su/p0’ 0.18-0.41 0.25-0.41 0.18-0.30 0.25-0.30
Sensitivity 1.5-6 3-5 1.5-6 3-5
Compression index Cc 0.70-1.30 0.45-0.95 0.60-1.20 0.40-1.00
In situ void ratio 1.72-2.50 1.40-1.85 1.70-2.50 1.40-1.85
Coefficient of 0.01-0.20 … 0.10-1 1
permeability k(x10-9m/s)
Coefficient of lateral 0.52-0.72 0.52-0.72 0.58-0.70 0.58-0.70
Earth pressure at rest(ko)
Note: p0' denotes effective overburden stress

It is well known that the introduction of cement will increase in the strength of soft clay.

The short-term gain in strength is the result of primary hydration reaction, which also leads to

a reduction in moisture content during the chemical reaction. The long-term gain in strength is

largely a result of secondary pozzolanic reaction between the lime and the clay minerals (e.g.,

Kezdi, 1979; Bergado et al., 1996). Besides the increase in strength, the cement stabilization

also causes significant improvement in other engineering properties such as compressibility,

stiffness, permeability and stress-strain behavior due to the formation of structure within the

soil grain assemblage induced by cementation.

Because of the rapidly rising demand for cement stabilization in soft soils for urban

development, the engineering behavior of cement-treated soil has attracted worldwide interest.

However, constitutive framework of improved marine clay is still not available due to

insufficient research. In fact, the cement-treated marine clay is often modeled as a Mohr-

Coulomb linear elastic-perfectly plastic material in numerical modeling (e.g. COI 2005; Lee

2
CHAPTER 1

2008), which may not be consistent with the real constitutive behavior of cement-treated

marine clay.

1.2 Behavior of cement-treated soil

The constitutive behaviors of natural soils have been extensively investigated (e.g. Al-

Tabbaa and Muir Wood, 1989; Smith et al., 1992; Shen, 1993a, 1993b; Wheeler, 1997;

Rouania and Muir Wood, 2000; Kavvadas and Amorosi, 2000; Asaoka, 2000; Gajo et at., 2001;

Liu and Carter, 2002; Koskinen and Karstunen, 2002; Baudet and Stallebrass, 2004;

McDowell and Hau, 2004). The constitutive behaviors of natural, weakly cemented soils or

soft rocks have also been studied by various researchers (e.g. Wong et al., 1975; Oka et al.,

1985; Nova, 1988; Gens and Nova, 1993; Airey, 1993; Lagioia and Nova, 1995; Malandraki

and Toll, 1995, 2000; Callisto et al. 2004). The engineering properties especially the

unconfined compressive strength, modulus and permeability of cement-treated soil have been

studied since 1960’s (e.g. Wissa et al., 1965; Enami et al., 1980; Nakamura et al., 1982;

Kauschinger et al., 1992; Terashi, 1997; Uddin et al., 1997; Balasubramanian et al., 1998; Yin

and Lai, 1998; Huang and Airey, 1998; Miura et al., 2001; Kamruzzaman, 2002; Hopribulsuk

et al, 2003; Lorenzo and Bergado, 2004; Lee et al., 2005; Chin, 2006). However, the

constitutive behaviors of cement-treated soils have been studied relatively late. Most of the

studies were mainly based on isotropic compression and undrained triaxial shear test (Hirai et

al., 1989; Matsuoka et al., 1995; Uddin et al., 1997; Yu et al., 1998; Kasama et al., 2000; Rotta

et al., 2003; Lee et al., 2004; Horpibulsuk et al., 2004; Namikawa et al., 2006; Namikawa et al.,

2007; Chiu et al., 2008; Horpibulsuk et al., 2009). The constitutive model for cement-treated

soil especially for cement-treated clay is very scarce (Lee et al., 2004; Horpibulsuk et al.,

2009). As such, the stress-strain behavior of cement-treated clayey soil is still largely unclear

and research on this area remains scanty. As mentioned earlier, up to now, cement-treated

marine clay is often modeled as a Mohr-Coulomb material in numerical modeling (e.g. COI

2005; Lee 2008).

3
CHAPTER 1

1.2.1 Behavior of natural and artificially lightly cemented soil

Numerous studies on the mechanical behavior of natural cemented soil have been

conducted over the last 30 years (e.g. Wong et al., 1975; Oka et al., 1985; Leroueil et al., 1990;

Airey, 1993; Coop and Atkinson,1993; Cuccovillos and Coop, 1999; Callisto et al. 2004).

Based on the stress-strain behavior under triaxial loading condition, many constitutive models

have been developed for the behavior of natural cemented soil for a long time, which considers

the ‘small strain’ response, the material’s memory of its stress history and soil structuration

effects (e.g. Jardine et al, 1986; Gens and Nova, 1993; Matsuoka and Sun, 1993; Kavvadas,

1995; Muir Wood, 1995b; Liu et al., 1997).

For cemented soils, Leroueil and Vaughan (1990) stated that the patterns of behavior

observed in all cemented soils are similar even though the cementation may result from

different causes. More recently, researchers used artificially created lightly cemented soil to

study the naturally cemented soil by re-creating cementation under controlled condition (e.g.

Abdulla & Panos, 1997; Malandraki and Toll, 2000; Consoli et al, 2000; Kasama et al., 2000;

Rotta et al, 2003; Consoli et al, 2006). However, the cement content in these studies is often

too low to be representative of cement-treated soil and the results of these studies may

therefore not be applicable for cement-treated soils due to the following reasons. For instance,

in Rotta et al.’s (2003) experiments, the cement content used ranges from 1% to 3%, which is

well below what is usually used in cement treatment of soft clays. In such models, the effect of

cementation is often modeled by an enlarged yield surface with the same shape as that of the

remoulded soil (e.g. Gens and Nova, 1993; Rouainia & Wood, 2000). During undrained

loading, the loss of shear strength is often associated with a loss of effective stress arising from

volumetric collapse, rather than a direct loss of bonding. In drained conditions, loss of strength

is usually not modeled as volumetric densification offsets the loss of bonding. Chin (2006)

noted that, in cement-treated marine clay, strain softening can occur after an episode of

volumetric densification and strain hardening. Furthermore, during the strain softening, there is

a significant decrease in the stress ratio, which is suggestive of a “flattening” of the yield locus.

4
CHAPTER 1

1.2.2 Behavior of cement-treated sand and clay

In general, the behavior of cement-treated soil is very different from that of untreated soil

due to the structure effect. As such, the significance of structure to cement-treated soil is

obvious. The structure and its effect on the behavior of cement-treated soil attracted wide

interest (e.g. Kezdi, 1979; Saitoh et al., 1985; Chew et al., 2004; Chin, 2006).

Changes in physical properties such as water content, liquid limit, plastic limit with

cement content, curing time and curing stress has been investigated widely (e.g. Locat, 1990;

Uddin et al., 1997; Petchgate et al., 2001; Chew et al., 2004; Lee et al., 2005; and Chin, 2006).

The change in particle size, particle distribution and some aspects of micro-structure due to

cementation has been also studied (e.g. Suzuki et al. 1981, Shen 1998, Chew et al., 2004; Chin

2006).

The early research works carried out to understand the effects of the various factors on the

strength of cement-treated soil were based on unconfined compressive strength, which is

widely used as an engineering property to represent the effectiveness of the stabilization

method. Factors which have been found to affect the unconfined compressive strength

included stabilization agents, characteristics of soil, mixing condition, and curing condition

(e.g. Niina et al., 1977; Enami et al., 1980; Kawasaki et al., 1981; Nakamura et al., 1982;

Terashi, 1997).

The engineering behavior of cement-treated soil such as permeability, compressibility,

stiffness, shear strength, and stress-strain behavior was also investigated widely. In general, the

results obtained by different researchers showed the same trend for compressibility, stress-

strain behavior (e.g. Endo, 1976; Tatsuoka and Kobayashi, 1983; Shibuya et al., 1992; Uddin

et al., 1997; Balasubramanian et al., 1998; Miura et al., 2001; Chin, 2006) whereas the results

for permeability and shear strength by different researchers showed more variation (e.g.

Suzuki et al.,1981; Kauschinger et al., 1992; Uddin et al., 1997;Yin & Lai, 1998).

However, to date, the constitutive behavior of cement-treated soft clay has not been

investigated sufficiently to enable development of constitutive models. Consequently only a

5
CHAPTER 1

few of constitutive models were proposed for cement-treated soil and these are mainly for

cement-treated sand or sandy soil (e.g. Hirai et al., 1989; Yu et al., 1998; Namikawa et al.,

2007).

1.2.3 Behavior of cement-treated Singapore marine clay

The behavior of cement-treated Singapore marine clay has been investigated since 1990’s.

Lee (1999) and Lee et al.(2005) studied the strength characteristics of cement-treated

Singapore marine clay specifically in the context of jet grouting. Tan et al. (2002) investigated

the properties of cement-treated Singapore marine clay through unconfined strength test with

local strain measurement method. Kamruzzaman (2002) and Chew et al. (2004) studied the

physicochemical and micro-structural properties of cement-treated marine clay, while Chin

(2006) focused on the micro-structural changes and macroscopic behavior of cement-treated

marine clay under triaxial loadings at low and high stresses. So, however, a complete

constitutive framework for cement-treated Singapore marine clay is still not available.

1.3 Objectives and organization of the thesis

This study follows upon the work of Chin’s (2006) initial study on the constitutive

behavior of cement-treated Singapore marine clay. This study can be sub-divided into several

parts. Firstly, element testing was conducted for a wide range of mix proportions and different

curing conditions to shed light on the constitutive behavior of the cement-treated clay. Next,

based on the experimental results, common trends were identified which forms the basis of a

framework of behavior of this material. Finally, a constitutive framework was postulated for

the primary yielding and early post-yield behavior of cement-treated marine clay. The detailed

issues to be studied will be discussed in the next chapter. The main objectives of the thesis are

summarized as below.

(a) To investigate the primary yielding behavior of cement-treated marine clay, including the

proper method to determine the primary yield locus.

6
CHAPTER 1

(b) To investigate the early post-yield behavior of cement-treated marine clay, especially the

evolution of yield locus during hardening phase.

(c) To develop a constitutive model for the primary yielding and early post-yield behavior of

cement-treated marine clay.

The layout of the thesis following this chapter is introduced briefly as below.

ƒ Chapter 2 presents a comprehensive literature review on experimental and theoretical study

as well as the outstanding issues relevant to cemented soil. The scope of the current study is

also highlighted.

ƒ Chapter 3 introduces experiment set-up and methodology, including the experiment work

done in this research.

ƒ Chapter 4 and Chapter 5 present experimental results and analysis, mainly on primary

yielding and post-yield behavior of cement- treated clay. Instead of trying to investigate the

critical state of cement treated marine clay, a reference state to study the structure effect on

constitutive behavior of cement treated marine clay is also introduced. The ultimate state of

cement-treated soil is also discussed.

ƒ Chapter 6 introduces the proposed empirical constitutive framework as well as theoretical

constitutive framework based on experimental study. Parameters of the proposed constitutive

model are also discussed.

ƒ Chapter 7 summarizes the main findings of the current study and discussions on the

proposed constitutive model. Besides, recommendations are made for future studies.

7
CHAPTER 2
LITERATURE REVIEW

2.1 Introduction

This chapter provides a review on experimental and theoretical studies on the behavior of

cement-treated soil. The fundamental concepts and mechanism of cement stabilization is first

discussed, followed by the factors on the hardening characteristic of cement-treated clay. In

experimental study, current understanding on the changes of soil properties and behavior due

to the cement stabilization is presented. The limitations of those studies are reviewed. In

theoretical modelling, current approaches to model the constitutive behavior of cemented soil

are discussed. Some outstanding issues relating to yielding and softening are then presented.

Finally the issues which will be examined in the current study are presented.

2.2 Basic concepts and mechanism of cement stabilization

2.2.1 Mechanism of cement stabilization

Ordinary Portland Cement (OPC) is the most commonly used cement in cement-soil

stabilization. Its main components are tricalcium and dicalcium silicates (C3S and C2S),

tricalcium aluminate (C3A) and tetracalcium alumino-ferrite (C4AF). In accordance with the

notation commonly used in cement chemistry, C herein represents CaO, S represents SiO2, A

represents Al2O3 and F represents Fe2O3. In the presence of water, hydrations of these

compounds form colloidal hydrated products of very low solubility, which are calcium silicate

hydrates [CSH]; calcium aluminate hydrates [CAH]; and calcium aluminate silicate hydrates

[CASH]. The hydrations of aluminates are mainly responsible for setting, i.e. solidification of

the cement paste whereas the hydrations of silicates lead to hardening of cement paste. Also,

the hydration of calcium silicates produces calcium hydroxide [Ca(OH)2], or lime. The

9
CHAPTER 2

Ca(OH)2, together with NaOH and KOH that are present in small amounts, causes a rise in pH

in the pore liquid.

For simplicity, tricalcium silicate is chosen to illustrate the cement stabilization

mechanism in the discussion below. There are two major chemical reactions which govern the

soil cement stabilization mechanism: the primary hydration and the secondary pozzolanic

reactions. The former is represented by Equation (2.1) below and occurs between cement and

water (from soil or cement slurry), resulting in rapid strength gain due to the formation of

primary cementitious products such as CSH. This reaction also leads to the short-term

hardening of cement-treated soil. In addition, lime is produced and the concentrations of Ca2+

and OH- ions in the pore water increases [Equation (2.2)] through the hydrolysis of the lime.

C3S + H2O C3S2HX (hydrated gel) + Ca(OH)2


(2.1)
(primary cementitious products)

Ca(OH)2 Ca2+ + 2(OH)- (2.2)


(hydrolysis of lime)

The secondary pozzolanic reaction, also termed as solidification, occurs once the pore

chemistry in soil system achieves a sufficiently alkaline condition when sufficient

concentration of OH- ions is present in the pore water. The resulting alkalinity of the pore

water promotes dissolution of silica and alumina from the clays, which then react with the Ca2+

ions, forming calcium silicate hydrate (CSH) and calcium aluminate hydrate (CAH), which are

the secondary cementitious products [Equations (2.3) and (2.4)]. These compounds crystallize

and harden with time, thereby enhancing the strength of the soil cement mixes.

Ca2+ + 2(OH)- C-S-H


(2.3)
+ SiO2 (soil silica) (secondary cementitious product)

Ca2+ + 2(OH)- C-A-H


(2.4)
+ Al2O3 (soil alumina) (secondary cementitious product)

It should be noted that the Eqs. (2.1) – (2.4) only apply to tricalcium silicate (C3S), which

is the main constituent of OPC. The others main constituents of cement such as dicalcium

silicates (C2S), tricalcium aluminate (C3A) and tetracalcium alumino-ferrite (C4AF) are also

involved in both the hydration and pozollanic reactions to produce calcium silicate hydrate

(CSH), calcium aluminate hydrate (CAH) and calcium aluminate silicate hydrates (CASH). A

10
CHAPTER 2

complete set of chemical equations involving these reactions as well as lime treated clay have

been presented by Kezdi (1979).

2.2.2 Structure and microstructure of treated soil

The term “structure” was first introduced for clays (e.g. Mitchell, 1976), and has been

described by Burland (1990) as being the combination of bonding (i.e. inter-particle cementing)

and fabric (i.e. the arrangement and distribution of the particles comprising the soil). A

common effect of structure is to give the soil a higher strength and stiffness and to allow the

soil to exist at a higher volumetric state than that of the same material in a reconstituted state.

For sands, structure has often been equated solely with interparticle bonds although the fabric

of geologically aged sands may also influence the peak strength (e.g., Dusseault &

Morgenstern, 1979; Cuccovillo & Coop, 1999). The effect of structure has been observed on a

wide range of natural soils and weak rocks of both sedimentary and residual soils, and also for

artificially cemented soils. The structure of natural soils has been examined by numerous

researchers (Lerouiel and Vaughan, 1990; Burland, 1990; Nagaraj et al., 1998; Liu and Carter,

1999; Asaoka et al., 2000; Baudet and Stallebrass, 2004; Low et al., 2008), which are often

formed by solute deposition at inter-particle contacts, charge deficiencies, and van der Waal

forces (Mitchell 1993), rather than by hydration and pozzolanic reactions. The structure of

lime-treated clays has been studied by Locat et al. (1990, 1996) and Rao and

Rajasekaran(1996), amongst others. However, the absence of the primary hydration reaction in

lime treatment may lead to a different microstructure from that produced by cement treatment.

The structure of artificially cemented soils has been investigated by various researchers (e.g.

Hirai et al. ,1989; Coop & Atkinson, 1993; Huang &Airey, 1998; Kasama et al., 2000;

Malandraki and Toll, 2000; Consoli et al., 2000; Schnaid et al., 2001; Kamruzzaman, 2002;

Rotta et al., 2003; Chew, 2004; Chin, 2006). However, most of them focused on lightly

cemented soils (usually less than 5% cement content), which may not have the same behavior

as heavily cement-treated soil created by ground improvement (usually higher than 10%

cement content).

11
CHAPTER 2

The microstructure of cement-treated clay is often significantly different from that of the

untreated clay. Kezdi (1979) suggested that a soil-cement skeleton matrix may be formed due

to the inclusion of cement with each skeletal unit consisting of a core of hydrated cement gel

(tobermorite gel) and secondary cementitious product (CSH and CAH) connecting the adjacent

clay particles. In addition, the inter-particle bond strength also increases due to reduction of

diffused double (absorbed) layer and flocculation of the secondary cementitious materials.

Saitoh et al. (1985) proposed schematic diagrams to illustrate the change in structure of

soil-cement mixtures during hardening, as shown in Fig 2.1. They postulated that the initial

condition immediately after mixing consists of clusters of clay particles, surrounded by cement

slurry. The primary hydration reaction involves only the shell of cement slurry, which forms

hardened cement bodies. The secondary pozzolanic reaction involves the inner clay particles,

leading to the formation of hardened soil bodies. They also postulated that the strength of the

improved soil will depend upon the strength characteristics of both types of hardened bodies.

Locat et al. (1990) presented micrographs on the microstructure of lime-treated sensitive

clay (see Fig 2.2). As Fig 2.2(A) shows, before addition of lime, the clay has an open

microfabric, and individual particles and aggregates could be seen. As Fig 2.2(B) shows, after

10 days of curing with quicklime, the soil has been flocculated into larger lumps; Fig 2.2(C-F)

show the lumps cemented together by the subsequent pozzolanic reaction products.

For cement-treated clay, Chew et al. (2004) noted that, as the cement content increases

from 10% to 50%, the flocculated nature of the fabric becomes more evident, with soil particle

clusters interspersed by large opening. They attributed this to the dissolution of silica and

alumina from the clay minerals and their subsequent reaction with the Ca2+ ions to form CSH

and CASH, which are then deposited onto the particle surface. They also noted a significant

amount of entrapped water within the flocculated particle clusters, similar to that observed by

Locat el al. (1996) for lime stabilized clay.

More recently, Kamruzzaman (2009) and Chin (2006) also explore the evolution of

cement-treated clay microstructure under different loading conditions and its effect on

macroscopic properties. Kamruzzaman(2009) found that during Ko consolidation(oedometer

12
CHAPTER 2

consolidation), the largest intercluster voids are collapsed at the early stage of consolidation

and small intracluster pores are compressed only after the pressure exceeded the

preconsolidation pressure. A significant reduction in apparent compression index of the treated

clay is observed after a certain level of stress, which indicates that a complete destructuration

is possible by applying a very high stress. By investigating specimens sheared under undrained

condition, Kamruzzaman (2009) stated that at high confining stress (beyond the yield stress),

the specimens has apparently destructured during isotropic consolidation stage but still

appeared to have some cementation effect during shearing. At low confining stress, the

destructuration of treated clay only takes place during shearing stage. However, Kamruzzaman

(2009) didn’t conduct drained triaxial test and generalize destructuration for both undrained

and drained stress paths under different loading conditions. Chin (2006) found that loss of

structure in the cement-treated soil under triaxial loading depends on the stress status. For

specimens sheared at effective confining stresses much lower than their isotropic yield stress,

large aggregates remained visible even within the rupture band. On the other hand, for

specimens sheared at higher effective confining stresses(i.e. at stresses close to and higher than

isotropic yield stress), the dominant microstructure change appears to be aggregate squashing

and void closure at pre-peak stage and massive aggregate and particle break-up at post-peak

stage, which increases in severity with compression pressure and volumetric

compression(Figure 2.3). However, Chin’s (2006) study was limited to only one mix

proportion. As such, whether there is similarity among different mix proportions is not clear.

2.3 Factors on the strength of cement-treated soil

The strength of the cement-treated soil can be affected by a number of factors. The early

research works carried out to understand the effects of the various factors on the strength of

cement-treated soil were based on unconfined compressive strength which is widely used as an

engineering index to represent the effectiveness of the stabilization method.

13
CHAPTER 2

Terashi (1997) summarized the factors that influence the strength of the improved soil

into four categories: characteristics of stabilizing agent; characteristics and condition of soils;

mixing conditions; and curing conditions (see Table 2.1).

Table 2.1: Factors affecting the strength increase (after Terashi, 1997)

I. Characteristic of stabilizing agent 1. Type of stabilizing agent


2. Quality
3. Mixing water and additives
II. Characteristics and conditions of soil 1. Physical chemical and mineralogical
(especially important for clays) properties of soil
2. Organic content
3. pH of pore water
4. Water content
III. Mixing conditions 1. Degree of mixing
2. Timing of mixing/re-mixing
3. Quantity of stabilizing agent
IV. Curing conditions 1. Temperature
2. curing time
3. Humidity
4. Wetting and drying/freezing and thawing,
etc.

2.3.1 Characteristics of stabilizing agents

In general, the strength of an improved soil increases with the amount of stabilizing agent.

However, the rate of increment is not proportional to cement contents (Kawasaki et al. 1981).

For Bangkok Clay, Uddin et al. (1997) observed that the greatest increment rate lies in cement

contents ranging from 10% to 25%. For Singapore marine clay, Kamruzzaman (2002) noted

significant increase in the strength of the treated soil within the cement content range of 5-40%.

On the other hand, Miura et al. (2001) and Horpibulsuk et al. (2003) noted that the ratio of the

initial water content of the clay to the cement content (clay-water/cement ratio) is a more

appropriate parameter for quantifying the strength development of the cement treated soft

clays, instead of cement content only. Similarly, Lee et al. (2005) showed that the unconfined

compressive strength of the improved clay is dependent on both soil/cement and water/cement

ratios.

Differences of improvement arising from the use of different types of cement have also

been investigated. Kawasaki et al. (1981) compared the effect of slag cement and ordinary

14
CHAPTER 2

Portland cement for two different types of soils in Japan, Kanagawa and Saga soils. The result

in Figure 2.4 shows that the improvement obtained is dependent on the type of stabilizing

agent and soil, or more precisely the chemical reactions that occur between the stabilizing

agents and the soils. Similarly, Ahnberg et al. (1995) compared the effect of cement, lime and

mixture of cement and lime mixed with different soils in Sweden, as shown in Figure 2.5.

Based on field results from sandy ground with 5% of fines, Saitoh et al. (1990) noted that

blast-furnace cement produces higher compressive strength than ordinary Portland cement.

Besides cement or lime, a mixture of different stabilizing agents such as fly ash-cement

mixture has also been studied (Balasubramaniam et. al, 1998).

2.3.2 Characteristics of soil

It is well-recognised that different types of soil (e.g. peat, clay, silt, sand, etc.) affect the

chemical reactions between the soils and stabilizing agents and thus the properties of the

treated soil. Niina et al. (1977) noted the influence of grain size distribution on the unconfined

compressive strength of the cement-treated soil. As shown in Figure 2.6, the highest

improvement effect was obtained when the sand fraction is about 60%, irrespective of the

amount of cement content. Taki and Yang (1991) also revealed that coarse grained soil shows

a larger strength increase for given cement content. Stavridakis (2006) stated that the decrease

of bentonite and increase of sand induced the increases of strength of cement stabilized clay-

sand mixtures.

Among clay types, the effect of different mineralogy was studied by Wissa et al. (1965).

They noted that, generally, soils with higher pozzolanic reactivity show greater strength

increase upon cement treatment. For instance, montmorillonitic and kaolinitic clayey soils

were effective pozzolanic agents, compared to clays which contain mainly illite, chlorite or

vermiculite. They explained that the pozzolanic reaction between clay particles and hydrated

lime is dependent on mineral composition, especially the amorphous silica and alumina that

present in the soil. Saitoh et al. (1985) also highlighted the importance of pozzolanic reactivity

in the effectiveness of the cement-clay improvement.

15
CHAPTER 2

Endo (1976) showed that the increase in initial water content in the soil significantly

reduces the compressive strength of the mixture at any particular cement content. However,

Terashi et al. (1980) stated, at very low moisture content, i.e. near to plastic limit, the degree of

improvement is also not significant.

Huat (2006) investigated the effect of the organic content on the strength of cement

stabilized tropical peat soils. As shown in Figure 2.7, higher organic content of the soil

induced lower strength of cement stabilized soil.

2.3.3 Mixing conditions

The mixing conditions include factors such as the types of mixer, installation method,

timing and degree of mixing. The importance of mixing is to create a treated soil mass with a

high degree of uniformity. For cement-stabilized clay, Nakamura et al. (1982) investigated the

relationship between the unconfined compressive strength and mixing period for laboratory

prepared samples mixed under both cement powder and cement slurry (see Figure 2.8). The

figure shows that the decrease in mixing time caused the decrease in unconfined compressive

strength. In laboratory, the standardization procedure by the Japanese Geotechnical society

(JGS, 2000) arguably suggests a mixing period of 10 minutes as appropriate time to obtain a

“sufficient mixing” in the laboratory, using a Hobart mixer.

For in-situ soil-cement mixing, Yoshizawa et al. (1996) showed that the number of

mixing shafts, mixing blades and rotational speed may affect the strength of improved soil.

They stated that better improvement could be obtained when using four mixing shafts and

higher rotational speed as compared to single shaft with low speed. Other configurations which

may also influence the degree of mixing during installation of soil-cement mixing such as the

number of shafts, the configuration of mixing blades, the penetration/withdrawal speed, the

rotational speed of the shafts and the injection methods were also studied(e.g. Nishibayashi et

al., 1985; Enami et al., 1986; Nishibayashi, 1988; Saitoh et al., 1990)

2.3.4 Curing conditions

16
CHAPTER 2

During treatment period, the curing temperature, stresses, time and humidity also affect

the strength development of the treated soil. In general, the longer the curing period, the better

is the strength development, due to the progress of the pozzolanic reaction (Kezdi, 1979). As

shown in Fig. 2.9, the strength increases with curing time irrespective of soil types (Kawasaki

et al., 1981). A similar test results were obtained with Portland cement or fly ash cement

(Saitoh, 1988; Uddin et al., 1997; Lorenzo et al., 2004; Stavridakis, 2006; Kamruzzaman et al.,

2009)).

The influence of curing temperature on the unconfined compressive strength of the

laboratory treated soil was also studied by Enami et al. (1980), as shown in Figure 2.10. In

general, it shows that a higher strength could be obtained under a higher curing temperature.

The increase in unconfined compressive strength is almost linear with curing temperature

ranging from 0°C to 30°C, for samples of different ages up to 28 days.

Chin (2006) showed that unconfined compressive strength increases with the confining

stress at which cement-treated specimens were cured under drained condition whereas

increasing the confining stress in undrained curing condition does not lead to a significant

increase in the unconfined compressive strength (Figure 2.11).

Harleston et al. (2009) studied the humidity effect on the strength of cement stabilized

pavement layers. They cured the specimens under 100% humidity and room humidity (50%-

80%) and found that as humidity increases, strength increases.

2.4 Experimental studies on cement-treated soil

2.4.1 Changes in basic property of cement-treated soil

Many studies have shown that cement treatment leads to an increase in plastic limit as

well as liquid limit (e.g. Locat, 1990; Uddin et al., 1997; Petchgate et al., 2001; Chew et al.,

2004; Lee et al., 2005; and Chin, 2006). Both Locat et al. (1996) and Chew et al. (2004)

attributed the increase in the liquid limit to the presence of trapped water within the largely

hollow cemented-soil clusters. Chew et al. (2004) also attributed the increase in plastic limit to

17
CHAPTER 2

the clustering of the soil particles to larger clusters of particles. As shown in Fig. 2.12, Chin

(2006) also found that liquid limit increases after cement treatment but decreases with curing

load. Chin (2006) attributed this decrease to the squashing of the clusters which reduces the

amount of trapped water within the clusters.

Studies have also shown that cement treatment leads to decrease in water content and

specific gravity (Uddin et al., 1997; Chew et al., 2004; Lorenzo et al., 2004, Chin, 2006).

Lorezo et al. (2004) attributed the decrease in water content with increasing cement content

and curing time to hydration and pozzolanic reaction.

Assarson et al. (1974) and Chew et al. (2004) noted that the soluble products of cement

hydration lead to an increase in the electrolytic concentration of the pore fluids and hence

increases the pH value. The dissolved bivalent calcium ions (Ca2+) replace the monovalent ions

(e.g. Na+, K+), which are normally attracted to the surface of the negatively charged clay

particles. The crowding of Ca2+ ions on the surface of the clay particles brings about the

flocculation of the clay particles (Suzuki et al., 1981; Shen 1999). Therefore, the particle

(actually clusters of cemented soil particles) size increases with increase of cement content.

2.4.2 Engineering behavior

2.4.2.1 Prime factors on strength and deformation of cement-treated soil

Cement content and curing time are generally used as controlled parameters in previous

studies on the behaviour of cement-treated clay (e.g. Kamon and Bergado, 1991; Bergado et al.,

1999). Miura et al. (2001) suggested that the ratio of the initial water content of the clay to the

cement content (water-clay/cement ratio) is a prime parameter influencing the strength and

deformation behavior of cement-treated clay at high water content. Based on unconfined

compression tests and oedometer tests on cement-treated remoulded soft Bangkok clay,

Lorenzo et al. (2004) found that the after-curing void ratio and cement content are sufficient to

characterize the strength and compressibility of cement-treated clay at high water contents. For

naturally cemented soils, Leroueil & Vaughan (1990) suggested that the density and bond

18
CHAPTER 2

strength are the two most important parameters governing the behavior of cemented soil. This

is consistent with the notion that void ratio (which is a measure of density) and cement content

(which is related to bond strength) are the two most important parameters governing the

behavior of cemented soil.

Huang & Airey (1998) found that the strength and stiffness increase with increasing

density and cement content, but the relative influence of cementation decreases as the density

increases, which is the coupled effect between cementation and density of treated soil. Huang

& Airey (1998) stated that the cementation bonding was dominant for specimen with low

density and the effect mitigates with increase in packing density of the soil. They attributed the

reducing influence of cementation to the greater contribution of friction and particle

interlocking to the strength as the density increases.

Tatsuoka and Kobayashi (1983) among others investigated the effect of confining stress

on behavior of cement-treated soil. They noted that the behavior of cement-treated soil under

drained triaxial compression changed from strain softening to strain hardening as the confining

stress increases. Similarly, Yu et al. (1997), Uddin et al. (1997) and Chin (2006) also observed

similar effects of confining pressure on the drained triaxial behaviour of cement-treated soil.

The importance of ambient effective stress (curing stress) during the formation of

cementation bonds on the behavior of naturally and artificially cemented soils was also well-

recognized (Consoli et al., 2000; Rotta et al., 2003; Suzuki et al., 2004; Chin et al., 2004;

Bergado, 2004; Consoli et al., 2006; Chin, 2006). Chin (2006) found that samples cured under

loading show increases in unconfined compressive strength and gross yield, which appear to

be well-correlated to their post-cured void ratio. He also stated that the post-yield compression

curves of samples with different curing stress converge towards a single line. Similar result

were reported by Rotta et al.(2003) and Consoli et al.(2006). Consoli et al. (2006) noted that

incremental yield stress in isotropic compression, unconfined compressive strength, and the

initial bulk modulus can be related with curing void ratio and cement content. They also

pointed out that the effects of increasing cement contents on soil compressibility, yielding and

strength were consistently more pronounced at lower void ratios, clearly demonstrating the

19
CHAPTER 2

coupled effect of density and cementation on the overall response of bonded soils, which was

postulated by Huang & Airey (1998).

2.4.2.2 Permeability

The permeability or the hydraulic conductivity of cement-treated soil has also been

investigated (e.g. Jefferies, 1981; Kilpatrick and Garner, 1992; Deschens et al., 1995; Yu et al,

1999; Porbaha et al 2000; Kamruzzaman, 2002; Chin, 2006). This property is important for the

design of cut-off walls where seepage needs to be suppressed. Porbaha at al (2000) noted that

the distribution of pore sizes inside the soil cement mix influences the coefficient of

permeability of the treated soil. Kauschinger et al (1992) found that the permeability of the

cement-treated clay reduces with the increase of cement content and curing time. Broderic and

Daniel (1990) suggested that this reduction could be due to the pozzolanic cementitious

products, which block the pores in the soil cement matrix. In contrast, Suzuki et al (1981)

observed an increase in permeability due to flocculation of cemented soil. Chin (2006)

reported that the permeability of cement-treated marine clay reduces with consolidation and is

much lower than that of untreated clay at a given void ratio.

2.4.2.3 Compressibility

The reduction in compressibility of soft clay arising from inclusion of cement is well

established. Uddin et al. (1997) postulated that a certain amount of cement (>5%) is required to

improve the compressibility of the untreated clay. Miura et al. (2001) suggested that the

resistance to compression of the treated clay is markedly enhanced until the consolidation

pressure reaches the apparent pre-consolidation pressure (yield stress on one dimensional

compression curve), after which large volumetric compression occurs. Huang &Airey (1998)

also suggested that the effects of the bonding are only significant for stresses below an

apparent pre-consolidation stress. Kamruzzaman (2002) found that in odometer consolidation

test, both of cement content and initial water content affects the position of the post-yield

20
CHAPTER 2

compression line and the magnitude of the one-dimensional vertical apparent preconsolidation

stress(see Fig.2.13). Similarly, Chin (2006) found that isotropic yield stress increases with

curing load (see Fig. 2.14).

Balasubramanian et al (1998) reported that, at stresses much higher than the apparent pre-

consolidation pressure, the e − log p ' relationship of the treated clay is almost parallel to the

virgin compression line of the untreated clay. Huang & Airey (1998) stated that the slopes of

the compression lines are practically independent of the amount of cement added, basing on

the results of isotropic compression with a range of cell pressure capacity up to 70MPa. As the

cement content increases, the normal consolidation line shifts to the right in a v : ln p ' plot,

which was attributed to the change in grading of the soil. They also found that at confining

pressures greater than 20MPa, departures from the linear v : ln p ' responses occurred and the

compression index in natural log scale ( λ ) decreases for all cement content. Rotta et al. (2003)

and Chin (2006) reported similar results for slightly artificially cemented soil and heavily

artificially cemented clay cured under stress. In Rotta et al.’s (2003) study, the treated soil was

cured under confined pressure up to 2MPa and isotropically consolidated up to 6MPa. It was

found that after primary yield, the paths of the specimens with different initial void ratios all

follow a post-yield compression line that is unique for each degree of cementation and

converges with the intrinsic compression line of the untreated soil as the isotropic stress

increases. Cuccovillo and Coop (1999) also observed similar feature for natural calcarenites

with different void ratio. However, Rotta et al’s (2003) study only used cement content up to

3%. Chin’s (2006) study used higher cement content up to 20%. In Chin’s study, the treated

soil was cured under confined pressure up to 250kPa and tested under isotropic compression. It

was observed that the post-yield compression line tends to converge to a unique compression

line as the isotropic compression stress increases.

Uddin et al. (1997) found that compression index (Cc) reduces with the increase in

cement content when cement content is lower than 15%. Kamruzzaman (2002) also reported

that compression index reduces with the increase in cement content and curing time. In

21
CHAPTER 2

contrast, Rotta et al. (2003) and Lorenzo and Bergado (2004) found that higher cement content

resulted in higher compression index at the post-yield state.

2.4.2.4 Stiffness

The small-strain elastic modulus of cement-treated soil has been measured using local

strain transducer (Goto et al., 1991; Shibuya et al., 1992; Tatsuoka et al., 1997, Tan et al, 2002;

Kamruzzaman, 2002). Tatsuoka et al. (1997) used Local Displacement Transducer (LDT) in

consolidated undrained triaxial test to measure the stiffness of cement-treated sandy soil in the

linear elastic range of less than 0.01% of strains. They reported that the strain measured from

local transducer showed much higher stiffness than that of the conventional method. In

addition, the stiffness is almost constant and independent of initial loading, unloading and

reloading cycle with the elastic range of strain of about 0.01%. Kamruzzaman (2002) used

Hall’s effect transducer in unconfined compression and undrained triaxial test. It was found

that the stiffness measured by Hall’s effect local transducer is about 3.0 and 2.16 times of that

by conventional method for unconfined compression and undrained triaxial test respectively.

Yin and Lai (1998) reported that the secant modulus of elasticity (conventional strain

measurement) increases with the increase in cement content and confining pressure, but

decreases with the increase of initial water content. They also suggested that at very high

confining stresses, the cemented structure may have been partially destroyed, thereby reducing

the stiffness.

2.4.2.5 Shear strength parameters

Broms (1986) postulated that the two components of the strength, namely frictional

resistance ( ϕ ' ) and cohesion intercept ( c ' ) increase via two processes. The frictional

resistance increases due to the formation of significant amounts of particle interlocking in the

clay-cement skeleton while the cohesion component increases due to the reduction of the

thickness of the diffused doubled-layer of adsorbed water. Yin & Lai (1998) reported that, for

22
CHAPTER 2

Hong Kong marine clay deposit, the cohesion of the treated soil increases with increase in

cement content and decrease in initial water content of the untreated soil. However, the

internal friction angle decreases with the increase of cement content and decrease of initial

water content. In contrast, Uddin et al. (1997) reported that, for Bangkok clay, both the

cohesion and friction angle increase with the increase of cement content and curing time and

reached asymptotic values at about 15% cement content. Horpibulsuk et al. (2004) also

reported similar results.

Azman et al. (1994) studied the effect of cement content and consolidation pressure on

shear strength parameters of black soil in Malaysia. They found that, as the cement content

increases, the treated soil showed higher friction angle and lower cohesion; this is clearly the

opposite of what Yin & Lai (1998) has reported. On the other hand, higher cohesion and lower

friction was achieved as the confining pressure decreased. Azman et al. (1994) classified the

failure patterns of cement-treated clays into “friction-dominated” at very high confining stress,

“cementation-plus-friction” at medium confining stress and “cementation-dominated” at low

confining stresses.

Consoli et al. (2000) found that the curing stress increases the friction angle but does not

affect the cohesive intercept. Chin (2006) found that the peak strengths of cement-treated

marine clay appeared to be a curve, with higher peak stress ratio (q/p’) at low p’ value (Fig.

2.15), which is similar to the findings of Tatsuoka and Kobayashi (1983). This indicates that

the contribution of soil structure during occurrence of peak has a greater relative influence at

lower confining pressures.

2.4.3 Stress-strain behavior of cement-treated soil under triaxial condition

The stress-strain behaviour of the cement-stabilized clay under triaxial condition have

been extensively investigated (e.g. Endo, 1976; Tatsuoka and Kobayashi, 1983; Shibuya et al.,

1992; Uddin et al., 1997; Yu et al., 1997; Yin and Lai, 1998; Miura et al., 2001; Kamruzzaman,

2002; Horpibulsuk et al., 2004; Chin, 2006, Chiu et al., 2008). Figure 2.16 shows the typical

stress-strain behaviour under isotropically consolidated drained (CID) and undrained (CIU)

23
CHAPTER 2

triaxial compression test on laboratory prepared 5% cement treated marine clay (Endo, 1976).

The result showed that higher deviator stress could be achieved by increasing the confining

pressure, in both drained and undrained conditions. However, for the CID samples, the

maximum deviator stress was reached at a relatively high strain level of about 20% or more.

In contrast, the CIU samples reached peak deviator stress at a much lower strain level of about

3%. The CIU samples also failed at stress levels lower than the CID samples, presumably

because of the generation of excess positive pore water pressure. Uddin et al. (1997) reported

that the deviator stress-shear strain relation for cement-treated Bangkok Clay is linear up to 60

to 70% of maximum deviator stress and that at high cement content and low confining pressure,

the treated clay showed distinct strain softening behavior.

Tatsuoka and Kobayashi (1983) suggested that the effective stress principle can be

applied to analyze the behaviour of cement treated clay under drained and undrained

conditions. They noted that undrained samples which were tested under low effective

confining pressure normally failed near to or on the tension cut-off line. Similar observations

have also been reported by Chin (2006). On the other hand, for drained triaxial, as the

confining stress increases, the stress-strain relationship change from strain softening to strain

hardening, while the volumetric strain changes from dilation to contraction (Figure 2.17). Such

behavior in turn caused the drained samples (with high confining stress) to fail at higher

effective stress levels, far away from the tension cut-off line.

2.4.4 Studies on cement-treated Singapore marine clay

Lee (1999) studied the unconfined compressive strength of cement-treated Singapore

marine clay with soil/cement/water ratios which are relevant to jet grouting. It was found that

the strength was found to be dependent not only on the water/cement ratio but also the

soil/cement ratio. Lee (1999) also proposed an empirical relationship for the unconfined

compressive strength of jet grout piles and investigated the cause of ground movement during

jet grouting. Two models for predicting the constituent contents of a jet grout pile and one

method for estimating the pressure gradient of two-phase flow of jet grouting effluent were

24
CHAPTER 2

proposed in Lee’s (1999) study. Lee’s study concluded that ground movement during jet

grouting is due to the pressure built up in the jet grout cavity when the opening in the ground

for discharging the effluent slurry is clogged.

Tan et al (2002) investigated the properties of Singapore marine clay through unconfined

strength test using internal strain measurement techniques. It was shown that a convenient

normalization can produce a consistent pattern for evaluation of improved strength of clays

from different parts of Singapore. It was found that the cement-treated marine clay behave in a

non-linear fashion even at very small strain.

Kamruzzaman (2002) focused on the physico-chemical and microstructural aspects of

cement treated marine clay. The flocculated structure of the treated clay was highlighted in his

study. It was noted that the apparent pre-consolidation pressure, effective shear strength

parameters, reduction in compression index, and stiffness increases with the cement content

and curing time. The treated marine clay has a more brittle nature with a well-defined yield

point, which is associated with the breaking of the cementation bond. At higher stress level,

beyond the yield stress, progressive destructuration (breaking of cementation bond) of the

treated clay particles occurred. However, complete destruction (crushing of clay-cement

cluster) only takes place on the shear plane (in undrained shearing test).

Kamruzzaman (2002) reported that increase in the confining stress changes the behaviour

of the specimens from an “over-consolidated” type, characterized by dilatant and negative

excess pore pressure, to a “normally consolidated”, characterized by compression and positive

excess pore pressure (Figure 2.18). Undrained stress paths were often observed to show a peak

stress either near to the tension cut-off line or on the Hvorslev envelope. However, in

Kamruzzaman’s (2002) study, no drained triaxial test was conducted. Furthermore, no

framework or explanation was provided to generalize both undrained and drained stress paths

under different loading conditions, nor for specimens cured under different curing stresses.

Chin (2006) focused on the microstructural changes and macroscopic behaviour of

cement-treated marine clay under triaxial loadings at low and high stresses. It was found that

samples sheared at stresses much lower than their isotropic yield stress behave elastically up to

25
CHAPTER 2

peak stress, followed by strain softening and volumetric dilation. Though the slip plane is

subjected to large strains, large aggregates still can be seen. It was also found that samples

which were sheared undrained under higher effective stress reaches peak strength shortly after

yielding while samples which were sheared drained at higher effective stresses (i.e. at stresses

close to and higher than isotropic yield stress) displayed loss of structure in two stages. The

first stage is a stable strain-hardening stage, wherein densification was able to offset the loss in

strength due to destructuration. The dominant microstructural changes appear to be aggregate

squashing and void closure. The second stage is characterized by strain-softening and slip

plane formation, with the soil within the slip plane undergoing massive aggregate and particle

break-up.

It was also found that the ultimate angle of friction within the slip plane is not a constant,

but decreases with increase in effective stress. This reduction can be attributed to the finer

particulate texture on the slip plane, which suggests increasing severity of aggregate and

particle break-up with increasing effective stress level. Samples cured under loading show

increases in unconfined compressive strength and isotropic yield points which were well-

correlated to their void ratio. As the curing stress increases, the liquid limit and compressibility

decrease. However, there are several limitations in Chin’s (2006) study. Firstly, it was limited

to only one mix proportion, the variation of the isotropic yield stress and peak strength with

mix proportion and curing loading had not been fully investigated. Secondly, the changes in

the yield surface during the hardening phase of the shearing and the hardening parameters

governing these changes are not clearly manifested. Thirdly, the conditions governing the

onset of strain softening and the existence or otherwise of a critical state line was not fully

clarified.

2.4.5 Primary yielding and post-yield behavior of cemented soil

The phenomenon of yielding in clay soils was well known and the detailed discussion

on concepts involved in yielding of soft clays can be found in Roscoe & Schofield’s (1963)

paper. It has been observed that when natural clays are subjected to changes in effective stress,

26
CHAPTER 2

they show a rather stiff behaviour, while the stress vector remains within a domain in the stress

space, the boundary of which is called the “yield locus” (e.g. Mitchell, 1970; Wong & Mitchell,

1975; Crooks & Graham, 1976). There are different definitions about soil yielding. The

detailed discussion on the definition of yield for bonded materials can be found in Malandraki

and Toll’s (1996) paper. Callisto & Calabresi (1998) proposed that the yield locus can be

regarded as a boundary between stress states that cause a relatively stiff, pseudo-elastic

behaviour, and stress states inducing large deformability. Thus, in laboratory testing, yielding

is more often identified by a substantial decrease in stiffness rather than the onset of

irreversible (plastic) strains. This is actually similar to the conventional definition of soil

yielding. They further pointed out that plasticity and elasticity theories are used as tools in

order to interpret the results and to suggest possible way to model the clay behaviour. Airey

(1993) also suggested a similar concept.

Smith et al. (1992) provided detailed information about yield processes of Bothkennar

clay and suggested a provisional framework for identifying and characterizing yield. As shown

in Fig. 2.19 the stress space within the initial bounding surface may be divided into three zones

separated by three types of yield surfaces, namely Y1, Y2, and Y3 respectively. Within Zone 1,

the soil is linear elastic, while within Zone 2, the soil becomes non-linear (no significant

plastic strains are generated). Large-scale changes in particle packing (yield Y3) are delayed

until the stress path reaches the initial bounding surface, producing what is commonly referred

to as soil yield (conventional definition of soil yield). Y3 envelope represents an initial

bounding surface in that it cannot be crossed by the effective stress paths arising from

undrained events, including cycling loading. However, drained stress paths may be able to

progress beyond the initial bounding surface, through a fourth region Zone 4 towards an outer

state boundary surface. Some natural soils may behave in a less stable way, with the

normalized current boundary surface contracting inwards as straining continues.

Coop & Atkinson (1993) reported that at low confining stresses, shearing of cemented

carbonate sands may result in yield stress higher than the frictional failure envelope of the

uncemented soil and continued loading leads to strain softening while at high confining

27
CHAPTER 2

stresses yield occurs during compression and the soil behavior is strain hardening and its

strength is frictional.

Cuccovillo and Coop. (1997) reported the pre-failure behaviour of two structured sands in

triaxial compression performed over a wide range of pressures. It was found that the bond

degradation resulted in a progressive transformation of the structured soil into a frictional

material, giving rise to changes in the yield stress and shear stiffness. It was suggested that as

bonding degraded, the variation of the shear stiffness with state was seen to depend on which

structural feature was predominant.

However, all of the works mentioned above are studied under certain condition, without

considering various factors such as mix proportion and curing load. Rotta et al. (2003) found

that for artificially cemented soil, the variation in yield stress with void ratio and cement

content is dependent on the curing stress and independent of the stress history (see Fig. 2.20).

Consoli et al. (2006) stated that a unique relationship between incremental yield stress,

unconfined compressive strength and bulk modulus was obtained, regardless of the cement

content (see Fig.2.21). On the other hand, the relationship between unconfined strength, initial

bulk modulus, and the primary yield stress was found to depend on the cement content.

However, their works, as are those of Coop & Atkinson (1993) and Cuccovillo Coop (1997),

are only focused on weakly cemented soil, which were intended to simulated naturally

cemented soil.

As mentioned above, some studies have been conducted on cement-treated soil, which are

often more heavily cemented than natural clay (e.g. Uddin et al., 1997; Tan et al., 2003; Chew

et al., 2004; Horpibulsuk et al., 2004; Lee et al., 2005; and Chin, 2006). Tan et al.’s (2003) and

Lee et al.’s (2005) studies were concerned mainly with unconfined compressive test. Uddin et

al.’s (1997), Horpibulsuk et al.’s (2004) and Chew et al.’s (2004) studies involved some

consolidated undrained tests, but these were not supplemented by consolidated drained tests,

scanning electron micrography and other tests, which, as Chin’s (2006) study showed, are

critical for understanding the microstructural changes that occurred during shearing.

28
CHAPTER 2

Chin’s (2006) study only covered one mix proportion, wherein the mass ratios of

soil:cement:water are 5:1:6. Chin (2006) did not investigate the effects of different mix

proportions and curing stress on the primary yield locus. Similarly, the effect of curing time on

the primary yield locus has also not been investigated.

As the primary yield locus has not been defined until now and the evolution of post yield

locus under different mix proportion has not been studied, the difference between the primary

yield locus and post- yield behaviour remains largely unknown. Thus the development of yield

locus during pre-peak shearing is needed to investigate to describe the shearing behavior at

pre-peak stage. Although the softening behavior of cement treated soil is observed very often,

the governing factor of onset of the behavior is still unclear. This is one part of the

experimental work in this study.

2.4.6 Softening and post-peak shearing behavior of cemented soil

There are significant amounts of research on softening behavior of cemented soil.

Georgiannou and Burland (2001) stated that the brittle shearing behaviour of natural stiff clays

in triaxial compression at low to intermediate confining pressures results from the localisation

of strain at around peak strength and thereafter approximately rigid block sliding along the

resulting slip surface. It was noted that the shear strength drops rapidly from peak strength to a

reasonably constant value after a few millimetres, which is termed as post-rupture strength by

Burland (1990). They suggested that the rapid post-peak loss of strength results largely from

changes in microstructure due mainly to the breaking of inter-particle bonds. Georgiannou and

Burland (2001) suggested that this softening process is likely to be rather more dramatic than

that due to the well-known phenomenon of the formation of shear bands in over-consolidated

reconstituted soils, which is essentially due to dilation, loss of interlocking and local drainage

without the breaking of bonds (e.g. Atkinson & Richardson, 1987). Moreover, they pointed out

that when strain localisation takes place, as it frequently does, measurements of volume change

are no longer representative.

29
CHAPTER 2

However, as shown in Fig. 2.22, Viggiani et al. (1993) suggested that localization is

initiated at the maximum stress ratio q/p’, occurring before and developing up to the peak

strength. Using local measurements of axial displacements (triaxial tests) and of horizontal

displacements (biaxial tests), they also found that the shear band completely forms and

emerges from the boundaries of the specimen well after the onset of localization.

Some studies have been done on the effect of dimension and slenderness ratio (i.e. the

ratio of height to diameter) on the behaviour of cemented soil. Abdulla et al (1997) identified

potential effects of specimen size and slenderness ratio on the stress-strain-strength

characteristic of cemented sands. They stated that smaller specimens with the same slenderness

ratio exhibit stiffer and stronger behaviour than the larger specimens when tested under similar

stress paths (see Fig. 2.23). Shearing of shorter specimens results in a significantly smaller

volume of degraded material thus providing a more representative material response (see Fig.

2.24). They also suggested that large specimens of slenderness ratio equal to one may be

preferable if appropriate end lubrication measures are taken.

Callisto & Calabresi (1998) carried out a conventional triaxial test on a specimen with 1:1

height/diameter ratio and lubricated ends and found that the development of the usual slip

surface was not kinematically possible. Instead, the deformation pattern at failure appeared to

be more uniform (see Fig. 2.25), this being similar to that observed in true triaxial test. It was

also noted that although the initial part of deviator stress-strain curve is very similar to that of

conventional specimen with ratio 2:1 and rough ends, the strengths measured in two tests are

quite different.

Petchgate et al. (2001) studied the effects of specimen dimensions and the ratios of height

to diameter of the specimens on strength characteristics of soft Bangkok clay stabilized with

cement. It was noted that sample with L/D ratio of 1.0, 1.5 and 1.8 showed unconfined

compressive strength (UC strength) of 6%, 3% and 1%, respectively, higher than sample with

L/D ratio of 2.0 while samples with diameters of 27.75, 55.5 and 83.3 mm have UC strengths

of 64%, 30% and 14%, respectively, higher than the sample with diameter 107.3mm. However,

they only studied the effect of specimen dimension and slenderness on UC strength.

30
CHAPTER 2

Some studies have also been done on the ultimate state of cemented soil. Airey (1993)

obtained an ultimate critical state line for naturally cemented carbonate soil. However, this was

done by discounting those samples that developed pronounced rupture planes. Leroueil (1998)

noted that the critical state of natural and reconstituted clay can be different, indicating that

even at larger strains, soil behavior is influenced by the initial structure. He cited Saint-Jean-

Vianney clay as an example where the intrinsic and natural critical state angles of shearing

resistance are 30.5 and 44.4 respectively.

Wood (1990) stated that even for reconstituted clays it is often not easy to discern the

critical state at high over-consolidation ratios. It was noted that strength reduction after peak,

accompanied by dilatancy, tends to be a continuous process leading to stress ratios below the

normally consolidated critical-state value.

Rouainia & Wood (2000) pointed out that unless a mechanical test systematically

destroys the structure of the entire sample, the condition of the natural soil may not approach a

structureless asymptote. The occurrence of localization will prevent any further homogeneous

destructuring: only within the zone of concentrated shearing will loss of structure be complete.

Cotecchia & Chandler (1997) also noted that it will not be possible to deduce the asymptotic

behaviour externally. Burland et al. (1996) suggested that their results may indicate that the

material in the rupture plane has a fabric similar to that of the reconstituted material but this

may not always be the case.

Kasama et al. (2000) found that the failure state line of lightly cemented clay is parallel to

that of an uncemetned clay in the p '− q space whereas its slope is steeper than that of an

uncemented clay in the e − ln( p ') space. In contrast, Chin (2006) noted that ultimate failure

envelope of artificially heavily cemented clay is a curve rather than straight line (see Fig.2.26).

However, the stress state of the material within the slip plane cannot be fully defined. To this

extent, the critical state line cannot be clearly defined. Chin (2006) also suggested that an

appropriate governing criterion for the onset of softening is the stress ratio ( η = q / p ' ).

However, the indications from his data are not definitive and the conclusion is not definitive

31
CHAPTER 2

due to shear band formation and further data are needed to confirm this. Therefore, further

study needs to investigate the relevant issues such as governing criterion of softening and for

the post-peak, softening behavior of cement-treated soil.

2.5 Theoretical studies

2.5.1 Review of theoretical studies for natural or artificial cemented soil

Over the last 30 years, there have been significant developments in the development of

constitutive models incorporating the influence of soil structure (e.g. Wong & Mitchell, 1975;

Oka and Adachi, 1985; Hirai et al., 1989; Gens and Nova, 1993; Chazallon and Hicher, 1995,

1998; Wheeler 1997, Kasama et al., 2000; Rouania and Muir Wood, 2000; Kavvadas and

Amorosi, 2000; Vatsala et al., 2001; Liu and Carter, 2002; Lee et al., 2004; Baudet and

Stallebrass, 2004; Horpibulsuk et al., 2009).

Wong & Mitchell (1975) constructed a quasi-elastic work-hardening plastic model for

sensitive cemented clay. In the quasi-elastic range the soil may be considered isotropic but the

relevant parameters are stress dependent. Initial yielding is associated with the breaking down

of cementation bonds and a unique yield envelope, independent of stress path, is defined. The

post-yield locus is expanding with volumetric hardening, with the shape assumed to be the

same. However, the yield locus is developed based on experimentally defined non-associated

flow rule.

Two examples of constitutive models which make explicit introduction of structure and

damage to structure within a single yield locus elastic-plastic framework are provided by Oka

et al. (1985) and Gens & Nova (1993) separately. Gens and Nova (1993) presented a

conceptual framework for elasto-plastic modelling of bonded soils. They assumed that the

yield surface has the same shape and form as that of the uncemented soil but it was enlarged to

account for the additional strength provided by the bonds (see Fig. 2.27). Hardening parameter

was considered to be made up of two components, namely the hardening of the unbonded soil

and the softening due to degradation of cementation bonds with plastic strain. The flow rule is

32
CHAPTER 2

applied with respect to the combined stress state (the bond stress and the stress carried by the

soil skeleton). Similarly, Lagioia &Nova (1993, 1995) proposed an elasto-plastic strain

hardening model allowing for material degradation due to loss of bonding.

Oka and Adachi (1985) and Adachi and Oka (1995) developed a model to describe the

strain softening response of soft clay or soft rocks by introducing a stress history tensor. The

material strength is made up of two components: frictional strength and cementation strength

or cohesion. The frictional component is considered to be some fraction of the overall stress.

The degradation or softening law and the flow rule are applied with respect to the combined

stresses, a feature which is similar to the one used in the model by Gens and Nova (1993).

Burland (1990), Leroueil & Vaughan (1990), Chandler (2000) and many others have

shown that the properties of a reconstituted soil (termed intrinsic properties) form a useful

frame of reference for interpreting the behaviour of the associated natural undisturbed soil. In

particular, such a frame of reference can be used to assess the influence of microstructure (i.e.

fabric and bonding) in the natural material, and its progressive breakdown during loading and

shearing.

Rouainia & Wood (2000) developed a rate-independent constitutive model for natural

clays, which assumes that the size of the structure surface depends on the value of a pre-

selected damage parameter. As shown in Fig.2.28, this model introduces three elliptical loci in

the (p,q) plane. These three loci include a kinematically hardening bubble separating regions of

elastic and plastic response and moving with the current stress, a structure surface acting as a

bounding surface and containing information about the current magnitude and anisotropy of

structure, and a reference surface representing the behaviour of the reconstituted or completely

remoulded soil. As plastic strain occurs, the structure surface tends to collapse towards the

state boundary surface. Drop in stiffness occurs when the soil reaches the plastic regime and

the current stiffness depends on some measure of the separation of the kinematic surface and

the structure surface.

In the models of Rouainia & Wood (2000), Oka et al. (1985) and Gens & Nova (1993),

the elastic properties have been assumed to remain constant and the damage to the structure is

33
CHAPTER 2

contained in the plastic formulation. Chazallon & Hicher (1995, 1998), by contrast, consider

the alternative strategy of allowing for damage through a change to the elastic properties

according to a specified thermodynamically acceptable damage energy release rate. The

overall response is reflected in bonds and non-bonded material and modeled respectively.

Similarly, Vatsala et al (2001) presented a model for cemented soils within the framework

of hardening plasticity. The strength of a cemented soil comes from the usual strength of the

soil skeleton and the strength of cementation bonds while the deformation of the soil is

associated with the soil skeleton and cement bond at any given strain level. The overall

response of the soil under loading can be visualized as two stiffness acting in parallel for the

given unit strain. Separate stress-strain relations are defined for the two components and they

are then combined to give the overall response (see Fig.2.29). Modified cam-clay model is

used for the soil skeleton component while a simple elasto-plastic model is proposed for the

‘cemented’ component, which incorporates strain softening.

Liu & Carter (2002) also introduced the influence of soil structure into the Modified Cam

Clay model. They assumed that the void ratio for a structured soil during virgin compression is

composed of two parts: an elastic part which is dependent on the current mean effective stress

and a plastic part which is dependent on the size of the current yield surface. The plastic part is

subdivided into two components: one associated with the intrinsic properties of the soil and the

other associated with the soil structure. During virgin yielding the yield surface includes the

current stress state and expands isotropically causing destructuration of the material. However,

this model is only suitable for naturally structured clays with negligible or light cementation.

Kasama et al. (2000) proposed a constitutive model for lightly cement treated clay by

assuming the failure state of cement treated clay is the same as that of uncemented clay. Base

on the similar assumption, Lee et al. (2004) proposed a constitutive model for cement treated

clay by introducing bonding stress ratio m and critical state parameter λ’(see Fig. 2.30). These

parameters are used to simulate the increase in the initial stiffness and shear strength of cement

treated soil and subsequent progressive reduction in stiffness and strength due to breaking of

the bonding as a result of shearing. Similarly, based on Liu & Carter’s (2002) model, Liu et al.

34
CHAPTER 2

(2006) and Horpibulsuk et al.(2009) proposed a constitutive model for cement treated clay by

introducing a modified mean effective stress. In this model, the new parameter C, which is

related to the shear strength contributed by cementation, is assumed constant before peak point.

As such, the structure surface will expands according to virgin compression before peak point.

After peak strength, the structure surface will move toward right and collapse.

Callisto & Sebastiano (2004) stated that by defining a normalizing pressure that accounts

for the progressive loss of structure occurring during a test, the state of natural clay during

loading can be completely described in a normalized stress space.

Cuccovillo &Coop (1999) investigated the behavior of two natural sands (calcarenite and

silica sandstone) by triaxial testing over a wide range of pressures. They suggested that to

develop a general framework which is valid for both frictional and cohesive behavior, the

strain-softening and strain-hardening modes of shear behavior should be distinguished on the

basis of the location of the state of the soil relative to its isotropic boundary rather than to its

critical-state or intrinsic normal compression line.

McDowell (2000) proposed work equation which accounts for energy dissipated in

particle fracture and frictional rearrangement, where the relative proportion of plastic work

dissipated in fracture and friction is a simple function of stress ratio, and the normality

principle is applied to generate a new family of yield loci (see Fig. 2.31). However, normality

should not be used in sands at high stress ratios where extensive particle rearrangement is

occurring. In this case, a non-associated flow rule could be used together with the family of

yield loci. Furthermore, McDowell & Hau (2004) suggested a generalized Modified Cam Clay

model for clay and sand incorporating kinematic hardening and bounding surface plasticity.

So far, much of the researches have been focused on natural or artificially lightly

cemented soil. There is much less work on constitutive modeling of heavily cemented soils

such as cement-treated soft clay of the type encountered in deep mixing or jet grouting. At the

present, jet-grouted and deep-mixed soil are often treated as a linear elastic-perfectly plastic

material in finite element modeling. A constitutive framework for cement-treated marine clay

is currently not available.

35
CHAPTER 2

2.5.2 Constitutive frameworks for cement-treated soil – a summary

This section summarizes the principles and frameworks adopted by the constitutive

models presented above. Some researchers developed constitutive models by introducing

damage mechanics to consider the structure and structure loss in soil under loading (Shen,

1993a, 1993b; Chazallon and Hicher, 1995, 1998; Yu et al., 1998; Vatsala et al. 2001; Zhao et

al. 2002; Liu & Shen, 2005). Theses models considered the response of structure soil to be the

sum of the soil skeleton (i.e., friction between grains) and cementation bond and modeled them

separately. However, only Yu et al.’s (1998) and Chazallon & Hicher’s (1998) model were

applied to cement-treated clay. Some researchers developed kinematic hardening constitutive

model for natural clays by adding some initial structure, which can then be progressively

destroyed, to an extension of modified Cam-clay within the framework of kinematic hardening

and bounding surface plasticity (Muir Wood, 1995; Rouainia and Muir Wood, 2000). However,

the shape of structure surface was assumed to be the same as that of the remoulded soil. Some

researchers introduced tensile strength to consider bond strength or cohesion and proposed

elasto-plastic model for cemented soils within a single yield locus modeling framework (Gens

& Nova, 1993, Lagioia and Nova, 1993, 1995; Oka et al., 1989; Adachi and Oka, 1993, 1995).

In Gens & Nova’s model (1993), the yield surface has the same shape and form as that of the

uncemented soil but was enlarged to account for the additional strength provided by the bonds.

The current size and location of the yield surface depend on the amount of damage, through a

combination of distortional and volumetric effects. Lagioia and Nova (1995) modified Gens &

Nova’s model (1993) but only compared the modeling results with Gravina Calcarenite. In

Oka et al.’s (1989), and Adachi & Oka’s model (1993, 1995), the material strength was

thought to be made up of friction strength and cementation strength or cohesion. The friction

strength is considered to be some fraction of the overall stress whereas the degradation or

softening is applied with respect to the combined stresses. However, both models were

proposed mainly for soft clay or rocks and have not been validated for cement-treated soils.

36
CHAPTER 2

There are only a few constitutive models developed for cement-treated sand or sandy soil

(Hirai et al., 1989; Abdulla & Kiousis, 1997; Namikawa &Mihira, 2007). In Hirai et al.’s

model (1989), a generalized form of Modified Cam-clay model was applied to plastic potential.

Two yield functions and two hardening functions were used to describe the characteristic of

the plastic behavior of improved sandy soil. A failure criterion was also proposed to define the

ultimate strength of material. In Namikawa and Mihria’s model (2007), two failure criteria are

employed to express tensile and shear failure characteristic. However, as Hirai’s model (1989),

the yield surface does not have compression cap. In Abdulla & Kiousis’s model (1997), a

micro-mechanical approach was used to model clean sand, cementation bond and pore

pressure separately. By assembling all models, the elasto-plastic behavior of cemented sand

was predicted. However, whether this model is applicable to cement-treated clay is still not

clear.

Constitutive models for cement-treated clay soils remain very scarce (Kasama et al., 2000;

Lee et al., 2004; Horpibulsuk et al., 2009). The present models were applied only for light or

low cement content (usually less than 10%). And the structure surfaces are usually assumed to

be an enlarged surface of the unstructured or reconstituted soils, which may not be true for

cement-treated clay soils.

To date, no validated constitutive model exists for cement-treated Singapore marine clay.

This is one part of the theoretical work in this study.

2.6 Outstanding issues

2.6.1 Outstanding issues

Based on the reviews of the previous sections, this section presents the some outstanding

issues in relation to constitutive behaviour studies on cement-treated soft clays.

Firstly, knowledge of the primary yielding and evolution of yield locus under loading, for

example, triaxial loading, is still unclear and triaxial experiment data for cement-treated soft

clay remains limited. Secondly, the effects of soil/cement/water ratio, curing stress and other

37
CHAPTER 2

treatment conditions such as curing time on the yield surface are also unclear. Thirdly, the

conditions governing the onset of strain softening are still largely unknown. Fourthly, the

existence and form of the intrinsic and critical states in improved soil are also unclear. Indeed,

the applicability of the critical state concept to improved soil is still unknown. Hence,

constitutive frameworks for cement-treated soil remain relatively scarce and that for cement-

treated Singapore marine clay is still not available so far.

2.6.2 Scope of work in the current study

This study will focus largely on the pre-peak constitutive behavior of cement-treated

Singapore marine clay. Cement-treated Singapore marine clay specimens with different mix

proportions and curing conditions will be studied firstly to examine the behavior under quite

different mix proportion and curing condition. The effect of mix proportion and curing stress

and curing time on the primary yield locus will be investigated. The post-yield behavior and

post-yield changes in the yield locus will also be investigated for specimens with different mix

proportions. Short sample used in triaxial testing to study post-peak softening behavior while

preventing shear band formation will also be attempted.

The present study will not address the following aspects of cement-treated soil behavior:

1. Pre-yield behavior of the cement-treated soil including small-strain nonlinear behavior.

2. Anisotropic effects.

3. Post-peak softening behavior including the critical state (for standard specimen). However,

the constitutive behavior of remoulded cement-treated marine clay will be investigated in

detail and its use as an intrinsic state will be explored.

Based on the observations of the experiment of this study, a new energy equation will be

proposed for cement-treated marine clay, from which a new yield locus will be derived. This

new yield locus incorporates a cohesion parameter. The initial value of the cohesion can be

easily deduced from unconfined compression tests, thereby allowing the primary yield locus of

cement-treated marine clay to be readily determined from conventional tests. Furthermore, the

decrease in the cohesion parameter as a function of strain can be determined from the

38
CHAPTER 2

experimental data. By decreasing the cohesion parameter after yielding, changes in the yield

locus after primary yielding can be reproduced reasonably well.

Following this Chapter, the next four Chapters will address the main work of this study,

including experimental study and theoretical study.

39
CHAPTER 2

Figure 2.1: Schematic illustrations of improved soil (after Saitoh et al., 1985)

Figure 2.2: SEM micrographs of lime improved soil (after Locat et al., 1990)

40
CHAPTER 2

1 4

2 5

3 6

Figure 2.3: Microstructures on slip planes of several cement-treated specimens at


the end of triaxial shearing tests: ①CID50; ②CIU50; ③CIU500; ④CIU1500;
⑤CID500; ⑥CID1500 (after Chin, 2006)

41
CHAPTER 2

Figure 2.4: Effect of cement type on compressive strength of soil-cement


for: (a) Kanagawa; and (b) Saga soils (after Kawasaki et al., 1981)

Figure 2.5: Effect of different stabilizers on compressive strength of


different soils in Sweden (after Ahnberg et al., 1995)

42
CHAPTER 2

Figure 2.6: Effect of grain size distribution on cement stabilization


(after Niina et al., 1977)

Figure 2.7: Effect of organic content on the unconfined compressive


strength of peat soils (after Huat, 2006)

43
CHAPTER 2

Figure 2.8: Effect of mixing time on cement stabilization


(after Nakamura et al., 1982)

Figure 2.9: Effect of curing time on strength


(after Kawasaki et al., 1981)

44
CHAPTER 2

Figure 2.10: Effect of curing temperature on compressive strength of silt


(after Enami et al., 1985)

1100

Drained isotropic curing stress, p'cure


1000 without curing stress
p'cure=50kPa
900 p'cure=100kPa
p'cure=250kPa
800 p'cure=500kPa
Axial stress, σa (kN/m2)

700

600 Unload elastically Localisation of strain


(sweating disappear) at rupture plane

flaws start to form and propagate


500
Axial stress, σa (kN/m2 )

Rupture plane
400
Intact part

300
Specimen breaks apart

whole intact specimen


200 been compressed

Axial strain, εa (%)


100

0 1 2 3 4 5 6 7 8

Axial strain, εa (%)

Figure 2.11: UCT results for specimens treated under drained isotropic curing stress
(after Chin, 2006)

45
CHAPTER 2

70

CE
60 ME

freshly mixed
untreated clay
50 CV
0CON0
Plasticity index, %

40
MV 250CON0
CH 50CON0

30 500CON0 100CON0

CI MH
20

10 CL MI

ML
0

0 20 40 60 80 100 120 140


Liquid limit, %

Figure 2.12: Plasticity chart for cement treated clay under various
stress states during curing period (after Chin, 2006)

Fig. 2.13: Effect of initial water content and cement content on e-log σ v'
relationship (after Kamruzzaman, 2002)

46
CHAPTER 2

Figure 2.14: Isotropic compression behavior for cement treated clay under various
stress states during curing period (after Chin, 2006)

Figure 2.15: Peak strength envelope for treated specimens with OCR=1 in p’-q stress
plane (after Chin, 2006)

47
CHAPTER 2

Figure 2.16: Triaxial behaviour of treated clay under (a) drained; and (b)
undrained conditions (after Endo, 1976)

Figure 2.17: Consolidated drained triaxial behaviour with different confining


pressures (after Tatsuoka and Kobayashi, 1983)

48
CHAPTER 2

Figure 2.18: Undrained stress path of cement-treated clay for


(a) 10%; (b) 30%; and (c) 50% of cement contents (after Kamruzzman, 2002)

49
CHAPTER 2

Figure 2.19: Definition of yield surfaces (after Smith, 1992

Figure 2.20: Variation of incremental yield stress with curing


void ratio and cement content(after Rotta et al, 2003)

50
CHAPTER 2

Figure 2.21: Relationship between initial curing void ratio, unconfined compressive
strength, incremental yield stress and initial bulk density (after Consoli et al, 2006)

51
CHAPTER 2

Figure 2.22: Undrained triaxial compression of Todi clay: stress-strain response,


pore water pressure difference and strain homogeneity indexes
(after Viggiani et al,1993)

(a) under low confining stress (b) under high confining stress

Figure 2.23: Size effect on triaxial behavior of specimen with height-diameter


ratio 1(after Abdulla & Panos, 1997)

52
CHAPTER 2

Figure 2.24: Effect of slenderness ratio on drained triaxial behavior of specimen


with cement content 4% and confining stress 200kPa(after Abdulla & Panos,
1997)

Figure 2.25: Effect of the slenderness of the specimen on the results obtained
from constant p’ triaxial test (after Callisto & Calabresi, 1998)

53
CHAPTER 2

Figure 2.26: Post-ruptured strength envelope of samples with mix proportion


s:c:w=5:1:6
(after Chin, 2006)

Figure 2.27: Successive yield surfaces for increasing degrees of bonding. Surface A
corresponds to unbonded material (after Gens and Nova, 1993)

54
CHAPTER 2

Figure 2.28: Reference surface, structure surface and bubble (yield surface)
for destructuration model (after Rouaina & Wood, 2000)

Figure 2.29: Representation of composite action of frictional and bond forces in


cemented soil (after Vatsala et al., 2001)

Figure 2.30: yield surface for cement treated clay (after Lee et al., 2004)

55
CHAPTER 2

Figure 2.31: A new family of yield locus (after McDowell, 2000)

56
CHAPTER 3
EXPERIMENTAL METHODOLOGY AND SETUP

This Chapter presents the test programme and experimental aspects of this study. The

materials used in this study will be discussed first, followed by the mix proportions of the

cement-treated marine clay specimens. The sample preparation procedure is then discussed.

Finally the testing procedure and apparatus are described and the engineering properties of the

treated soil specimens are summarized.

3.1 Materials

3.1.1 Untreated marine clay

The untreated marine clay used in this study was collected from 4 to 5m depth below

seabed at a dredge site offshore of Pulau Tekong and belongs to the Singapore Upper marine

clay of Kallang formation. According to Pitts (1992) and Tan et al. (1983), the sediment was

deposited around 10,000 years ago. The undrained shear strength and compression index Cc of

Singapore upper marine clay ranges from 5~20kPa and 0.7~1.3, respectively (Tan et al. 1983).

The basic properties of the marine clay in this study are summarized in Table 3.1. This is

essentially the same as the clay used by Chin (2006).

Table 3.1: Basic properties of Singapore upper marine clay


Properties Values Properties Values

Liquid Limit, LL (%) 88 Grain Size Distribution:


Plastic Limit, LL (%) 38 Sand (%) 8
Plasticity Index, PI (%) 50 Silt (%) 72
In-situ Moisture Content, m 72% Clay (%) 20
(%)
Liquidity Index, IL 0.68 Total Unit weight, γb (kN/m3) 15
Initial void ratio, eo 1.93 Dry Unit weight, γd (kN/m3) 8.72
Specific gravity, Gs 2.70 K0 0.62
Note: In-situ Moisture Content, initial void ratio, Unit weight and K0 were referred to Chin
(2006), which were determined upon collection from the site.

3.1.2 Ordinary Portland cement

57
CHAPTER 3

The additive used to treat Singapore upper marine clay in this study is Ordinary Portland

Cement (OPC). The definition of cement content ( Aw ) used in this study is synonymous with

the mass ratio of cement: soil (c/s), which is the ratio of dry weight of cement to the dry weight

of clay particles and is expressed in percentage. The soil: cement (s/c) mass ratio is therefore

the reciprocal of the cement content. The total water content ( Cw ) is defined as the ratio of the

mass of water in the resulting mix to the mass of dry soil solid and cement and is expressed in

percentage. These definitions are the same as those used by Chin (2006). By using Lee et al.’s

(2005) definition, the water-cement ratio is defined as the ratio of the mass of water in the

resulting mix to the mass of cement solids while the soil-cement ratio is defined as the ratio of

the mass of soil solids in the mix to the mass of cement solids. The chemical composition and

physical properties of OPC are given in Table 3.2.

Table 3.2: Chemical compositions and physical properties of Ordinary Portland Cement (after
Sun, 2008)
Chemical compositions Values Physical Properties Values

Lime Saturation, Factor (L.S.F) 0.93 Consistency (%) 29.0


Magnesia, MgO (%, m/m) 3.25 Penetration (mm) 7
Sulphuric Anhydride as SO3 (%, m/m) 2.06 Initial Setting Time (min) 180
Loss on Ignition (%, m/m) 2.53 Final Setting Time (min) 210
Silica, SiO2 (%, m/m) 20.26 Soundness (mm) <1
Calcium Oxide, CaO (%, m/m) 63.19 Fineness (m2/kg) 363
Iron Oxide, Fe2O3 (%, m/m) 3.61 Specific Gravity, Gs 3.15
Aluminium Oxide, Al2O3 (%, m/m) 4.31 2-days Strength (N/mm2) 22.7
Sodium Oxide, Na2O (%, m/m) 0.25 28-days Strength (N/mm2) 55.5
Potassium Oxide, K2O (%, m/m) 0.3

3.2 Variables investigated

The main objective of the experiment in this study is to investigate the constitutive

behavior of cement-treated soil under triaxial loading conditions, which provides the basis for

the constitutive framework of this study. For this reason, various cement contents, water

contents, curing stresses and curing time were used during the test program. For load-cured

specimens, the curing stress was applied isotropically. During load-curing, water was allowed

to drain into and out of the specimens via the ends of the specimens, which were covered by

58
CHAPTER 3

filter paper. The isotropic curing stresses p'cure used were 0 (atmospheric curing) ~350kPa.

The curing period is up to 210days.

Chin (2006) used a mix proportion with mass ratio of soil:cement:water (S:C:W) of 5:1:6.

This is equivalent to a cement content of 20% and total water content of 100%. In order to

investigate the general stress-strain behavior under triaxial loading for cement-treated

Singapore marine clay, 10% to 100% cement contents and 72% to 183% total water contents

were used, which resulted in soil:cement ratio ranging from 1 to 10 and water-cement ratios

ranging from 2 to 11. The soil-cement-water ratios of the different composition sets used are

summarized in Table 3.3. Chin (2006) reported that his cement-treated soil composition is

representative of deep mixing operation in Singapore Marine Clay. The mix proportions

selected for this study are representative approximately of the full range of water-cement and

soil-cement ratios used in a number of deep mixing and jet-grouting studies and projects

involving soft, fine-grained soils, as reported by Lee et al. (2005), Figure 3.1. As Figure 3.2

shows, all the mix proportions are in the workability range, which was provided by Lee et al.

(2005). According to Lee et al. (2005), the “bleeding limit” is defined as the water content at

which the settlement reaches 1% of the original height of the slurry. Slurry clay was prepared

by mixing the required amount of water with Singapore marine clay in its natural state. Dried-

pulverized clay was prepared by ovendrying the clay at 105°C and then crushing it into a fine

powder, before reconstituting it with water. They also reported that there is a significant

reduction in the Atterberg limits of the clay upon drying, which indicates that the activity of the

clay had been lowered in the process of drying and subsequent crushing. As such, the slurry clay

behaves different from dried-pulverized clay upon mixing with cement. Consequently, the liquid

and bleeding limits of dried-pulverized clay-cement mixes are significantly lower than those of

slurry clay-cement mixes, and its “workability range,” that is the region between the liquid

limit and the bleeding limit, is much narrower than that of slurry clay cement mixes (see

Figure 3.2). It should be noted that the definition of cement content used herein is different

from that used by Lee et al. (2005) and in Figure 3.2. In Lee et al. (2005) and Figure 3.2,

59
CHAPTER 3

cement content is defined as the mass of cement divided by the mass of soil and cement.

Strength and basic properties assessments were made after specified curing period.

Table 3.3: Mixture proportions and experiments in this study


Mix Soil- Water- Cement Total curing curing Test in this study
proportion cement cement content Water load time
(s:c:w) ratio ratio Aw(%) content (kPa) (days)
(s:c) (w:c) Cw(%)
10:1:11 10:1 11:1 10 100 0~350 7~28 UCT,ICT,CIU,CID,k0,
TST
20:3:23 20:3 23:3 15 100 0~250 7~210 UCT,ICT,CIU,CID,k0, η
, TST
5:1:6 5:1 6:1 20 100 0~250 7~180 UCT,ICT,CIU,CID,k0, η
, SEM, TST
10:3:13 10:3 13:3 30 100 0 7~28 UCT,ICT,CIU,CID,k0, η
, TST
10:3:17.3 10:3 17.3:3 30 133 0~100 7~28 UCT,ICT
10:3:19.5 10:3 19.5:3 30 150 0~100 7~28 UCT,ICT
2:1:3 2:1 3:1 50 100 0 7~28 UCT,ICT,CIU,CID,k0, η
, TST
2:1:4 2:1 4:1 50 133 0~250 7~180 UCT,ICT,CIU,CID,k0, η
,SEM, TST
2:1:4.5 2:1 4.5:1 50 150 0~100 7~100 UCT,ICT
2:1:5 2:1 5:1 50 167 0~200 7~28 UCT,ICT,CIU,CID
2:1:5.5 2:1 5.5:1 50 183 0 7~28 UCT,ICT,CIU,CID, TST
1.3:1:3.06 1.3:1 3.06:1 77 133 0~100 7~28 UCT,ICT
1.3:1:3.45 1.3:1 3.45:1 77 150 0~100 7~28 UCT,ICT
1.3:1:3.5 1.3:1 3.5:1 77 152 0~100 7~28 UCT,ICT,CIU,CID
1:1:2 1:1 2:1 100 100 0~100 7~28 UCT,ICT,CIU,CID
1:1:2.66 1:1 2.66:1 100 133 0~100 7~28 UCT,ICT
1:1:3 1:1 3:1 100 150 0~100 7~90 UCT,ICT
10:1:7.9 10:1 7.9:1 10 72 0~100 7~28 UCT,ICT
6:1:5 6:1 5:1 17 71 0~100 7~28 UCT,ICT
4:1:3.6 4:1 3.6:1 25 72 0~100 7~28 UCT,ICT
remarks 1. basic property tests including after-curing water content, after-curing unit weight,
after-curing specific gravity were conducted on all samples;
2. CIU/CID: isotropic consolidated undraiend/drained compression test;
3. ICT: isotropic compression test; UCT: unconfined compressive strength test;
4. η : constant stress ratio test; k0: k0 consolidation test;
5. SEM: scanning electronic microscopy test;
6. TST: tensile strength test

3.3 Sample preparation procedure

Untreated marine clay was prepared from its natural wet state without pre-drying. Lee et

al. (2005) noted that pre-drying causes largely irreversible changes to the Atterberg’s Limits of

60
CHAPTER 3

the clay, which are indicated of chemical changes. Distilled and de-aired water was first added

to the natural marine clay to bring its water content up to 100%. The resulting clay slurry and

the prescribed amount of cement slurry were then mixed in a Hobart Mixer using a rotational

speed of 125 rpm for about 10 minutes.

Apart from the mix composition, the procedure for sample preparation largely follows

that used by Chin (2006). For samples with zero curing loads, the specimen was placed into a

cylindrical PVC split-mould and then submerged in distilled water for curing. For samples

with curing load, the specimen was placed into a cylindrical PVC split-mould, which has been

pre-lined with a thin latex sleeve (membrane), and then put into triaxial chamber for drained

load-curing. The latex sleeve and cylindrical mould each has a diameter of 50mm and a height

of 100mm. No mechanical compaction of any form was applied. Before application of cell

pressure, the split mould was removed. O-rings were used to seal the membrane to the top cap

and base pedestal, so as to prevent slurry mix from flowing out when confining stress was

applied. Porous stones were provided at both ends of the specimen to allow double drainage.

Microscopic air pockets that might have been trapped within the specimen during preparation

were dissolved into solution by the application of 350kPa of back-pressure. The use of de-aired

water also encourages dissolution of air pockets and enhances saturation of the specimens.

Drained condition was ensured by opening the drainage valves during consolidation.

The elapsed time taken from mixing clay with cement to the application of curing stress is

slightly less than 30 minutes. The short time was used so as to minimize the formation of

structure within the treated clay.

It should be pointed out that during mixing and placement of cement-clay mixture into the

PVC mould, some of the air might have been trapped inside. Tan et al. (2002) showed that the

presence of air voids has an adverse effect on the strength of cement-treated clay. Generally

the UCT strength reduced about 5% for every 1% increase in air voids (Tan et al. 2002).

Therefore, it is important to control the amount of air voids within the specimen. Fig. 3.3 using

the same method as that of Chin (2006) shows a statistical analysis on the air content of

cement-clay mixtures just after mixing. The statistical analysis was carried out assuming the

61
CHAPTER 3

air content follows a normal distribution curve. As can be seen, the mean of the air content in

the mixtures is about 1.45% with the standard deviation (fluctuation about the mean) of 0.43%.

The probability of the air content within the mixtures exceeds 2.15% is about 0.08. As such,

the presence of air voids within the specimens was properly controlled in the present study.

On the other hand, excessive water in the soil-cement mixture may lead to “bleeding” (e.g.

Tan et al. 1997), or segregation of cement and soil solids from the water. Such bleeding occurs

when the mixture exceed its bleeding limit and may induce inhomogeneity into the specimens,

which should be avoided. Figure 3.2 shows the liquid and bleeding limits for cement-clay

mixtures with different admixture proportions as reported by Lee et al. (2005). Also plotted in

the same figure are the soil-cement-water ratios used in the present study. As can be seen, the

compositions used in the current study all lie within the workable range bounded by liquid and

bleeding limits.

3.4 Testing procedure and apparatus

After the specified curing time, the cement-treated marine clay specimens were taken

out and tested. Most of the specimens were trimmed to standard dimension of 38mm diameter

by 76 mm height and tested under triaxial loading system. The triaxial test included

unconfined compressive strength test (UCT), isotropic compression test (ICT), isotropic

consolidated undrained (CIU) and drained (CID) triaxial tests, K0 consolidation test and

constant stress ratio ( η ) test. Some of the specimens were tested for tensile strength using the

Brazilian test or radial compression test. In addition, some of them were used for scanning

electron Microscopy (SEM) test before and after triaxial loading. Some of the specimens were

also remoulded and tested under triaxial system. It should be noted that, following the

guidelines in BS1377 (1990), all specimens were also back-pressured to saturation prior to

triaxial loading. In accordance with (Black and Lee, 1973), the back-pressure was increased

until a B-value of at least 0.9 was obtained. The B value in this study is usually higher than

0.95. Tests conducted in this study are summarized in Table 3.3.

62
CHAPTER 3

3.4.1 Basic properties

The basic properties, which include moisture content, bulk density, initial void ratio and

specific gravity, were determined for all specimens after 7 days of curing. The basic properties

of most of samples with 28 days or longer curing time were also examined. These tests were

carried out according to the procedures and apparatuses described in BS1377 (1990). The

results are summarized in table 3.4.

3.4.2 Isotropic compression

Isotropic compression tests were carried out on the treated specimens as to assess

compressibility properties and isotropic compression yielding of the cement- treated clay, by

using a computer controlled triaxial stress path apparatus supplied by GDS as shown in Figure

3.4 or high pressure test system as shown in Figure 3.5a. Unless otherwise stated, the

specimens all have a diameter of 38mm and a height of 76mm. Prior to isotropic compression,

the specimen was saturated under back-pressure until the B value approaching or higher than

0.9 (Black and Lee, 1973); this often requires a back-pressure of about 400kPa. Such a high

back pressure was needed due to the high stiffness of the treated specimen. A constant rate of

confining stress which equals to 0.5~1 kPa/min was then applied to the specimen. Side

drainage was permitted to facilitate dissipation of pore water pressure through an outlet line

connected to the top cap. During the test, volumetric as well as pressure changes were

monitored by volume change and pressure transducers, respectively. This enables the effective

stress and void ratio to be determined accordingly. At the end of compression test, isotropic

unloading was carried out with a stress rate similar to that applied for the loading rate. The test

results are summarized in Table 3.4.

3.4.3 Triaxial compression and strength

63
CHAPTER 3

The stress-strain behavior and strength were studied using triaxial tests. The treated

specimens were tested using isotropic consolidated undrained (CIU) and drained (CID) triaxial

tests, K0 consolidation test and constant stress ratio ( η ) test. For CIU /CID tests, K0

consolidation test and constant stress ratio (η ) test, strip side drains were used along with filter

paper at the top and bottom of the specimen. Specimens sheared under high effective confining

stress were tested using the GDS equipment or normal triaxial test system with GDS pressure

controller shown in Fig. 3.5a. Tests involving effective confining stress less than 250kPa were

conducted using conventional triaxial equipment shown in Fig.3.5b. Strains rates of

0.01mm/min and 0.005mm/min were selected for undrained and drained triaxial shearing,

respectively. According to BS1377: Part 8: 1990, in order to calculate the rate of axial

displacement (shearing rate), triaxial consolidation tests were conducted to estimate the

significant testing time. The test results showed that the time tf is less than 2h. Thus 2h is used

as significant testing time. However, due to the brittle behavior of cement treated soil, the

significant strain interval for the test specimen is significantly smaller than that of reconstituted

clay, which is around 2.0%. By calculation, the shearing rate for CIU is around 0.01mm/min.

On the other hand, CID test needs to run slowly enough to ensure that pore pressure changes

due to shearing are negligible. Thus the sharing rate should be lower than that of CIU test.

Preliminary CID tests showed that a shearing rate of 0.005mm/min is suitable for this study.

Ko consolidation test is based on volumetric control by GDS test system, where the axial

displacement of the sample will be slowly adjusted thus ensuring the diameter of the

specimen remains constant, where the specimen diameter change is calculated from the

back pressure volume change. The loading procedure followed GDS test menu. In

constant stress ratio test, a stress path p-q was defined according to expected constant stress

ratio (q/p'), which will be completed within specified time. The GDS test system will attempt

to perform linear stress paths between the current values and the target values. For test

conducted under GDS system, the axial and volumetric strains were monitored by an internal

LVDT within the ram motor and volume change gauge in back-pressure controller,

64
CHAPTER 3

respectively. For tests conducted under normal triaxial system, the axial and volumetric strains

were measured by external LVDT and automatic volume change gauge and monitored by a

data logger. The test results are summarized in Table 3.4.

Unconfined Compressive Tests (UCT) were also conducted on some specimens.

Unconfined compressive strength is a commonly used strength parameter for cement-treated

soil. For UCT test, the specimen was sheared at a strain rate of 1.0mm/min, according to the

procedures and apparatuses described in BS1377 (1990) and Head (1986).

3.4.4 Microstructural properties

Scanning electron Microscopy (SEM) investigations were used to study the

microstructure of cement-treated marine clay specimens and remoulded cement-treated marine

clay specimens. SEM tests were conducted by using Hitachi Tabletop microscope TM-1000 as

shown in Figure 3.6. Two types of specimens were prepared for SEM tests. Type One was

taken from intact specimens before triaxial loading or after triaxial loading. Type Two was

taken from remoulded samples after triaxial loading. For Type One specimen, Following Chin

(2006), the specimens were first air dried (Mitchell, 1993) and the observation surface was

prepared by breaking specimen length-wise using finger pressure, while the other sides of the

specimen were then trimmed to a dimension of about 5 mm square by 3 mm thick. The

observation surface was then coated with gold for conductivity. Care was taken to minimize

disturbance on the observation surface during handling of these specimens. For Type Two

tests, post-triaxial loading specimens were broken into pieces and remoulded with a small

amount of water so that the remoulded specimens were scanned in a dispersed state. The

remoulded soil was then allowed to precipitate onto a small plate in a container. The

precipitated soil was then air-dried for about a week and then sputter-coated with gold for

SEM. The final thickness of the precipitate was usually about 1mm. This process causes the

soil grains and aggregates to be more dispersed and more readily discerned under SEM.

During scanning, magnification levels of 1.0k ~5.0k were used.

65
CHAPTER 3

3.4.5 Tensile splitting strength test

As an indirect method, Brazilian test is widely used to test the tensile strength of quasi-

brittle material (e.g. concrete, rock). Bazant & Planas (1998) noted that the Brazilian test (also

termed as radial compression test or tensile splitting strength test), “… has been used as a

measure of tensile strength (even though the stress state is not uniaxial) and has been included

in most standards as a routine method of estimating the tensile strength (ASTM C496 ,

BS1881-117,ISO 4108)”.

In this study, 50mm by 50mm cylindrical specimens were used for tensile strength test. A

triaxial test system was modified to conduct tensile strength test (Figure 3.7a). Displacement

controlled loading is used in this test. As can be seen, a stainless steel ram, with a load cell

mounted at its end, was modified to transfer the load to the specimen. The jig comprises two

loading platens, which were parallel to each other and put on the top and the bottom of the

specimen separately (Figure 3.7b). The top platen was designed to incorporate a steel strip

with rectangle section of 3.5mm by 4mm whereas the bottom one was designed to incorporate

steel strip with cylinder section with thickness 2mm and string length 8mm. This ensures that

the load is applied along a center line on the surface of the specimen and that the latter does

not move out of line during loading.

The specimen was loaded under water so as to simulate fully drain condition during test,

see Figure 3.7c. Very fast loading rate of 0.5mm/s and very slow loading rate of 0.005 mm/s

was used in this study to investigate the loading rate effect. Specimen was also tested under

dry condition to investigate the effect of drainage during test. The sample preparation

procedure is the same as section 3.3. The samples was placed into PVC model with dimension

of 50mm by 50mm and cured under water. Some of the specimens were cured under pressure

and then trimmed to dimension of 38mm by 38mm before test. The dimensions of each

specimen were measured before the test.

66
CHAPTER 3

3.4.6 Remoulded cement-treated marine clay

Cement-treated marine clay specimens were also remoulded and traixial tested to assess

the applicability of the “intrinsic state” concept. These cement-treated marine clay specimens,

with cement content 10%-50% and total water content 100%-167.7%, were prepared using the

procedure presented in section 3.3 and cured under different confining pressure for 7days or 28

days. After 7 days or 28 days of curing, the cement-treated marine clay specimens were taken

out and cut into small pieces before put into oven for 24 hours. The dried specimen pieces

were then pounded manually by using pestle and ground by using electronic grinder separately

(Figure 3.8). The ground specimens were sieved using a 75μm sieve before mixing with water.

After mixing, the remoulded specimens were put into a perforated steel tube and consolidated

under dead loading 50N or 80N for about 2 days. During consolidation, the water was allowed

to pass through both ends as well as tube side. The tube was specially perforated with small

holes, each approximately 2mm diameter, on the side. Before the mixed specimen was put into

tube, a thin coating of grease was put inside the tube to reduce wall friction during

consolidation and a filter tissue(towel) was wrapped around the tube on its outside to allow

water passing through the side holes. Another thin adhesive type was fully wrapped around the

tissue to ensure soil particles not flow out during consolidation while the water flows along the

tissue. The setup for preloading is shown in Figure 3.8.

67
CHAPTER 3

Table 3.4a Basic and engineering properties in this study

Notation Aw Cw pcur ' p 'py Gs w γ qu


(%) (%) (kPa) (%) (kN / m3 ) (kPa)
(kPa)

10-1-11 10 100 0 135 2.69 92.81 14.43 220


50 220 2.69 86.42 14.83 286
100 300 2.69 79.18 15.21 369
250 590 2.69 70.33 15.50 647
350 800 2.69 65.03 15.80 840
20-3-23 15 100 0 210 2.68 92.74 14.58 320
50 275 2.68 83.55 14.85 398
100 400 2.68 79.15 15.12 535
250 700 2.68 70.53 15.47 851
5-1-6 20 100 0 260 2.67 91.24 14.65 412
50 350 2.67 84.63 14.87 509
100 450 2.67 77.73 15.13 618
250 720 2.67 71.43 15.49 900
10-3-13 30 100 0 350 2.66 89.09 14.77 559
10-3-17.3 30 133 0 130 2.66 118.99 13.76 195
50 460 2.66 92.79 14.46 /
100 / 2.66 87.05 14.69 677
10-3-19.5 30 150 0 110 2.66 133.14 13.30 158
50 / 2.66 100.86 14.20 509
100 / 2.66 94.08 14.47 647
2-1-3 50 100 0 580 2.65 84.92 14.84 920
2-1-4 50 133 0 275 2.65 114.6 13.84 385
50 600 2.65 83.73 14.85 939
100 800 2.65 78.84 15.05 1206
250 1350 2.65 67.02 15.59 1691
2-1-4.5 50 150 0 160 2.65 127.58 13.48 264
50 / 2.65 101.66 14.31 782
100 / 2.65 84.48 14.83 1143
2-1-5 50 167 0 100 2.65 132.58 13.36 149
50 475 2.65 94.35 14.42 783
100 700 2.65 86.60 14.65 1059
200 1050 2.65 80.15 15.11 1427
2-1-5.5 50 183 0 80 2.65 168.75 13.14 114
Remarks 1. Gs , w(%) , γ is after-curing specific gravity, water content and unit
weight of specimen tested respectively.
2. pcur ' , p 'py , qu is curing stress, primary isotropic yield stress and
unconfined compressive strength respectively.

68
CHAPTER 3

Table 3.4b Basic and engineering properties in this study

Notation Aw Cw pcur ' p 'py Gs w γ qu


(%) (%) (kPa) (%) 3
(kN / m ) (kPa)
(kPa)

1.3-1-3.06 77 133 0 400 2.64 110.35 13.95 650


50 / 2.64 76.56 15.15 /
100 / 2.64 68.62 15.39 2400
1.3-1-3.45 77 150 0 250 2.64 125.05 13.49 438
50 / 2.64 81.94 14.72 1964
100 / 2.64 72.89 15.34 2650
1.3-1-3.5 77 152 0 350 2.64 114.13 13.84 602
25 650 2.64 92.61 14.59 1132
50 800 2.64 85.24 14.67 1392
100 1400 2.64 74.12 14.90 2349
1-1-2 100 100 0 950 2.63 78.54 14.91 2251
50 / 2.63 60.56 15.11 4153
100 / 2.63 58.60 15.50 /
1-1-2.66 100 130 0 500 2.63 108.40 14.04 823
50 / 2.63 66.72 15.38 3298
100 / 2.63 64.4 15.52 /
1-1-3 100 150 0 325 2.63 122.68 13.62 541
50 1100 2.63 72.64 15.29 1967
100 / 2.63 58.68 15.62 /
10-1-7.9 10 72 0 270 2.69 66.47 15.45 430
30 400 2.69 64.97 15.83 640
100 525 2.69 62.11 16.09 737
6-1-5 16.7 71.4 0 500 2.68 64.73 15.66 844
30 600 2.68 63.80 15.90 975
100 700 2.68 62.65 16.11 1005
4-1-3.6 25 72 0 600 2.67 63.38 15.85 980
50 800 2.67 62.00 16.02 1332
100 950 2.67 59.96 16.19 1424
Remarks 1. Gs , w(%) , γ is after-curing specific gravity, water content and unit
weight of specimen tested respectively.
2. pcur ' , p 'py , qu is curing stress, primary isotropic yield stress and
unconfined compressive strength respectively.

69
CHPATER 3

Figure 3.1: Soil-cement and water-cement ratios for some previous studies
on deep mixing and jet grouting (after Lee et al., 2005) and for this study

Figure 3.2: Working ranges of dried-pulverized and slurry clay cement mixes
(after Lee et al., 2005, mix proportion for this study were added into the same
plot)

70
CHPATER 3

Figure 3.3: Distribution curve of air content (%)

Figure 3.4: A fully computer controlled triaxial stress path

71
CHPATER 3

(a)

(b)

Figure 3.5: Normal triaxial test apparatus (a) high pressure with GDS
pressure controller (b) lower pressure

72
CHPATER 3

Figure 3.6: Hitachi Tabletop microscope TM-1000

73
CHPATER 3

(a) (b)

(c)

Figure 3.7: Tensile strength test setup (a) setup without sample (b)
loading platform with sample (c) loading platform with sample under water

74
CHPATER 3

(a) (b)

(c) (d)

(e) (f)

Figure 3.8: Preparing remoulded cement-treated marine clay specimen (a)


cement-treated marine clay specimens, (b) cutting into pieces, (c) dried pieces,
(d) powder of remoulded specimen, (e) setup for preloading of remoulded
specimen, (f) preloading of remoulded specimen

75
CHAPTER 4
PREPEAK BEHAVIOR OF CEMENT-TREATED
SINGAPORE MARINE CLAY

This and the next chapter discuss the experiment results for cement-treated Singapore

marine clay. These results allow the primary yielding, yield locus and post-yield behavior

under different mix proportions, curing stress and curing period to be highlighted. The

experiment was conducted mainly under triaxial loading condition. In addition, results

obtained from K0 and isotropic compression tests as well as unconfined compression test,

tensile splitting strength test and SEM results are also presented. The organization of this

Chapter is as follows:

Section 4.1 introduces the objectives of the various tests in the light of work which has

been done previously.

Section 4.2 discusses the definitions of artificial structure and primary yielding of

cement treated soil.

Sections 4.3 to 4.6 present results relating to isotropic compression behavior and

triaxial shearing, with emphasis on primary yielding and post-yield behaviors and factors

affecting them.

Sections 4.7 introduce and discuss the tensile strength of cement-treated soil.

4.1. Purpose of experiment

Since this study follows on directly from Chin’s (2006) study, its objectives have to be

77
CHAPTER 4

discussed against the backdrop of Chin’s (2006) findings, which can be summarized as

follows:

1. Cement-treated soil possesses a yield stress in the same manner as natural soils do. The

yield stress can be characterized by a decrease in stiffness under loading, which was

similar to the recommendations of Cotecchia & Chandler (2000). Microstructurally, the

yield stress appears to be associated mainly with breakage of inter-aggregate, but not

with large-scale break-up of intra-aggregate, bonds.

2. Triaxial shearing of cement-treated soil produces different behavioral traits under

different conditions. When samples are sheared undrained under low effective confining

stress (typically less than one-third of the current isotropic yield stress), the stress paths

tend to be constrained by the tensile cut-off envelope (that is the line given by q = 3p')

and the failure is accompanied by vertical cracking of the specimens which is consistent

with the attainment of zero lateral effective stress. Peak strength is reached on or near

the tensile cut-off envelope and is characterized by shear banding. Microstructurally, the

soil within the slip band still contains relatively large and unbroken aggregates.

3. Samples which were sheared drained at confining stresses much lower than their

isotropic yield stress (say 1/10 of the current isotropic yield stress) show stiff response

up to peak strength, followed by strain softening and volumetric dilation. Large

aggregates remained visible within the rupture band.

4. Samples which were sheared undrained under higher effective confining stress reaches

peak strength shortly after yielding. The post-peak behavior of the sample is marked by

shear band development and decrease in mean effective stress. Microstructurally, the

78
CHAPTER 4

soil within the shear band shows signs of massive aggregate breakage and crushing into

smaller fragments.

5. Samples which were sheared drained under higher effective confining stress undergo

strain hardening, which is marked large volumetric compression, prior to reaching peak

strength. Peak strength is reached at much larger value of shear strain than that in

undrained tests. The post-peak behavior of drained samples is marked by shear band

development and decrease in mean effective stress. Microstructurally, the soil within the

shear band shows signs of very severe aggregate breakage and crushing into much

smaller, often sub-micron, fragments. For the same confining stress, drained samples

show more severe aggregate breakage than undrained specimens. For a given type of test

(i.e. drained or undrained), the severity of aggregate breakage increases with confining

stress.

6. Samples which were cured under loading show increases in unconfined compressive

strength and gross yield which appear to be well-correlated to their post-curing void

ratio.

Chin’s (2006) study has several limitations, viz.

a) It is limited to only one mix proportion. There is a need to study if similar

behavioral trends apply to other mix proportions.

b) Chin’s (2006) parametric studies were not sufficiently extensive to permit their use

alone in the development of constitutive model. For instance, the variation of the

isotropic yield stress and unconfined compressive strength with mix proportion and

curing loading had not been investigated. Furthermore, the shape of the yield

79
CHAPTER 4

surfaces may not remain the same for different mix proportions, this needs further

quantitative investigations. As such, a constitutive model was not developed in

Chin’s study (2006).

c) Chin (2006) noted that shear banding prevented definitive measurements of the

stress state of the material within the slip plane. This is a common bifurcation

problem encountered by investigators in strain softening soils. Given this, it is

uncertain whether the specimens actually reached critical state, post-critical state

(e.g. some state between critical and residual) or something different altogether.

The main objective of the current study is to investigate if Chin’s (2006) framework is

applicable to other mix proportions and curing loads, with a view to producing sufficiently

detailed quantitative information for the development of a model for primary yielding and

early post-yield behaviour. In pursuance of this objective, the scope of experiment work

covered in this Chapter and next Chapter is expected to cover the following aspects:

a) Definition of initial or primary yield locus and its dependence on various factors.

Determination of primary yield locus and its possible relationship with other

engineering properties of the cement-treated soil, such as the unconfined

compressive strength.

b) Evolution of the yield locus during the hardening phase (if this exists) of the

shearing, primary how the shape and size of the yield locus change after initial

yielding.

c) The conditions governing the onset of strain softening. Chin’s (2006) data appear to

indicate that an appropriate governing criterion is the stress ratio (η = q/p’).

80
CHAPTER 4

However, the indications from his data are not definitive and further data are needed

to confirm this.

d) The existence or otherwise of a critical state line and, if affirmative, its shape. Chin

(2006) noted that ultimate failure envelope is a curve rather than straight line.

However, the stress state of the material within the slip plane cannot be fully defined.

To this extent, the critical state line cannot be clearly defined. Thus work is still

needed to ascertain the complete stress state of the sample during strain softening.

4.2. Definition of primary yielding and yield locus for

cement-treated soil

In this section, the definition of artificial structure will be discussed firstly. The

definition of primary yielding will then be discussed. The determination of primary yield

locus will be discussed furthermore.

4.2.1 Definition of artificial structure

Previous research works have shown that many natural soils possess some form of

structure (e.g. Burland, 1990; Leroueil & Vaughan, 1990). According to Burland (1990), the

differences in properties of natural and reconstituted soil could be attributed to soil structure.

Similarly, Leroueil & Vaughan (1990) noted that a soil possesses structure if the soil

behaviour cannot be described by void ratio and stress history alone. Burland (1990)

postulated an intrinsic compression line (ICL) and a sedimentation compression line (SCL)

to define the differences in behavior for the reconstituted and natural soils respectively. The

81
CHAPTER 4

ICL defines the compression behavior of reconstituted (i.e. unstructured) soil whereas the

SCL defines the compression behavior of natural deposits which possess certain

sedimentation structure.

Soil improvement processes such as deep mixing and jet grouting which make use of

cement, lime or other chemical agents are also likely to generate structure in soils. In most

cases, this kind of soil structure is characterized by a considerable amount of bonding, which

is likely to be created over and on top of a remoulded soil. For this reason, the kind of soil

structure resulting from chemical soil improvement processes may differ substantially from

those resulting from natural processes. In the discussion which follows, the termed “artificial

soil structure” will be used to define the soil structure resulting from the mixing of cement

(typically in the form of slurry) into a soil matrix.

4.2.2 Definition of yielding and yield locus for cemented soils

Yielding in clay soils is well-known and Roscoe & Schofield (1963) discussed in detail

the concepts involved in yielding of soft clays. Theoretically, the yield locus marks the limit

of elastic behaviour. However, this is often not easy to apply to experimental observation.

Many researchers have reported the difficulties in determining the yield point of soils, for

example, residual soils (Barksdale & Blight 1997), soft rocks (Cecconi et al. 1998),

Bothkennar clay (Smith et al. 1992). As Rotta et al. (2003) stated for artificially cemented

soil, the gradual onset of the breakage of the cement bonds, the grading of the soil and the

micro-features of the cememtitious agent can significantly add difficulties to the

identification of the yield point for cemented soils.

82
CHAPTER 4

The “yielding” in soils was characterized as “single yield” or “multiple yield” by

different researchers. The “single yield” concept has been used by many researchers (e.g.

Mitchell, 1970; Wong & Mitchell, 1975; Atkinson, 1990; Anagnostopoulos et al., 1991;

Coop and Atkinson, 1993; Kavvadas et al., 1993, 1994; Cuccovillo and Coop, 1997; Callisto

& Calabresi, 1998; Cotechia and Chandler, 2000; Rotta et al., 2003). Most of these

researchers define the yielding as a state where a substantial decrease in stiffness occurs. For

instance, Anagnostopoulos et al.(1991) defined “initial yield” as an abrupt decrease in

stiffness observed in the stress-strain curves of Corinth marl under triaxial compression.

Kavvadas et al. (1993, 1994) defined an “initial yield” for a soft rock (lignite) from the

stress-strain curve at the transition between a linear domain to a more plastic region.

However, some of these researches define “yielding” as the limit of elastic range, which is

the same as the classical meaning of “yielding” in mechanics. Atkinson (1990) suggested

that the term “yielding” should be restricted to describing the end of the elastic range. Based

on this concept, Coop and Atkinson (1993) defined the yield of artificially cemented

carbonate sand from stress-strain curve. Furthermore, Cuccovillo and Coop (1997) defined

the yielding of two structured sand (i.e., silica sandstone, calcarenite) from the change in the

tangent shear modulus in G- log ε s or G- log q curve.

Multiple yield points have also been proposed by some researchers (e.g. Vaughan, 1988;

Maccarini, 1987; Bressani, 1990; Leddra et al., 1990; Jardine et al., 1991; Jardine, 1992;

Smith et al, 1992; Malandraki and Toll, 1996). Vaughan (1985, 1990b) defined yielding as a

state at which a material shows a discontinuity in stress-strain behavior. As Malandraki and

Toll (1996) pointed out, this allows for the possibility for more than one yield point.

83
CHAPTER 4

Vaughan (1988) proposed two yield points for residual soil. The first yield point represents

the stress at which some of the bonds start to fail whereas the second yield causes a dramatic

loss in stiffness. Maccarine(1987) and Bressani(1990) defined two yields for an artificially

bonded soil while Leddra et al. (1990) defined two yield points for soft rocks. Based on the

study on the small strain behavior of reconstituted soil, Jardine et al.(1991) and Jardine

(1992) proposed three yields, which were designated “Y1”, “Y2”, “Y3”. “Y1” is the limit

of linear elastic behavior. “Y2” is the limit of recoverable behavior. “Y3” is the onset of

large-strain yielding. Similarly, Smith et al.(1992) suggested a provisional framework for

identifying and characterizing yielding in Bothkennar clay. Malandraki and Toll (1996)

suggested three yields for bonded materials, which are “first yield”, “second yield” and

“final yield”. The “first yield” is the same as “Y1”, which is defined at the point the first

change in stiffness occurs. The “second yield” is defined at the point where a major drop in

stiffness is initiated. The “final yield” is compatible with “Y3”, which is defined at the point

where the material has already lost almost all of its stiffness due to bonding. Malandraki and

Toll (1996) further pointed out that for normally consolidated clays and loose sands, the

“second yield” does not exist whereas for overconsolidated clays, dense sands and soft rocks,

“second yield” can be easily identified from log tangential stiffness-log axial strain curve.

By using local instrumentation, Tatsuoka and Shibuya (1991) suggested that the elastic

strain limit is about 0.0001% for normally consolidated clays but increases with

over-consolidation and that hard rocks may exhibit elastic strain limit of 0.01%. Jardine

(1995) found that the elastic strain limit increases as a result of bounding. Clayton et al.

(1994) also reported that the elastic strain limit for chalk might be as high as 0.03%.

84
CHAPTER 4

Cuccovillo and Coop (1997) found that at a confining stress of 1800kPa, the elastic strain

limit for structured sand (i.e., silica sandstone, calcarenite) is about 0.015%. However, by

using Maccarini’s (1987) and Bressani’s (1990) methods, Malandraki and Toll (1996)

reported that the “second yield” occurred at about 2% of axial strain. Hence, Malandraki and

Toll (1996) suggested that the “second yield” in some researcher’s method (e.g. Maccarini,

1987; Bressani, 1990; Vaughan, 1988; Vaughan et al., 1988) is most similar to the “final

yield”. According to Malandraki and Toll’s (1996) definition, the “second yield” from log

tangential stiffness-log axial strain curve coincides with the definitions of “initial yield” by

Anagnostopoulos et al.(1991) and Kavvadas et al.(1993, 1994), and “yield” by Coop and

Atkinson(1993).

Malandraki and Toll (1996) also postulated that “Y2” defined by some researchers (e.g.,

Jardine et al., 1991 and Jardine, 1992) may coincide with “first yield” defined by other

researchers (e.g., Vaughan, 1988 and Bressani, 1990) for strongly bonded materials, which

indicates that the material is elastic before a major drop in stiffness is initiated. Therefore,

the yield point from Coop and Atkinson(1993) and Cuccovillo and Coop(1997), “Y2” from

Jardine (1992) and Smith et al.(1992), and “second yield” from Malandraki and Toll (1996)

seems to be able to define the “yielding” of strongly bonded material such as cement treated

soil. This is also applicable to Rotta et al’s (2003) definition of “primary yielding”. In

Rotta’s (2003) method, the “primary yielding” is defined as the point in which the

stress-strain (i.e. p'-v) starts to deviate from the initial linear behavior.

The “yielding” of cemented soils can be identified and determined by different methods

according to “single yield” or “multiple yield” concept, namely “stress-strain method”,

85
CHAPTER 4

“stiffness method” and other method. In the “stress-strain method” method, the deviator

stress or mean effective stress is plotted with deviator strain or volumetric strain in natural or

log scale (e.g. Mitchell, 1970; Wong & Mitchell, 1975; Maccarini, 1987; Bressani, 1990;

Airey, 1993; Coop and Atkinson, 1993; Cotechia and Chandler, 2000; Rotta et al., 2003). For

instance, Mitchell (1970) plotted a one-dimensional consolidation curve on a linear stress

scale and defined a yield point by using Standard Casagrande method (Figure 4.1a). Mitchell

noted that “…at stresses less than the yield stress, the specimen compression is small, mostly

recoverable, and nearly a linear function of the applied stress. At stresses exceeding the

yield point stress, the specimen compression is relatively large and mainly irrecoverable”.

However, on the same plot, Mitchell also pointed out an actual yield point, which is the

same as that defined by Rotta’s method discussed later.

Wong & Mitchell (1975) used Taylor & Quinney’s (1931) method to define the yield

point from a given experimental stress-strain curve for sensitive cemented clay (Figure 4.1b).

They stated that “this method of defining the yield envelope is considered appropriate for

cemented clays since there is always a fairly sharp break in the stress-strain curves”.

Furthermore, they pointed out “the existence of a unique yield surface is of great importance

for sensitive cemented soils. After initial breakdown of the cementation bonds (initial

yielding) a small increment of applied stress produces a relatively large amount of strain

that is analogous to plastics flow. There appears to be a progressive breakdown of

cementation bonds following initial yield and it would be an oversimplification to identify

the post-yield sensitive clay with an insensitive normally consolidated clay. Exceptionally

large volumetric strain (up to 50%) is required to reach the normally consolidated or

86
CHAPTER 4

unstructured state.”

Rotta et al.’s (2003) proposes the point where the stress-strain curve deviates from its

initial linear trend as the “primary yield” point, at which breakage of bonds first commenced

(Figure 4.1c). Coop and Atkinson (1993) used the same method to identify the “yield” for

cemented carbonate sands on the plot of deviator stress-strain and mean effective

stress-volumetric strain. Airey (1993) also applied the same method to define the “yield” for

naturally cemented carbonate sand on the plot of deviator stress-axial strain.

Cotechia and Chandler’s (2000) also proposed ‘gross yield point’ as the point of

tangency between the compression curve and a line drawn parallel to the intrinsic

compression line (Burland 1990), at which the ratio of the stress on the compressive curve to

that on the intrinsic compression line is maximum (Figure 4.1d). Chin (2006) applied

Cotechia and Chandler’s (2000) method to determine the gross yield locus for

cement-treated Singapore marine clay.

Maccarini (1987) identified “first yield” at the end of the linear part of the curve of

deviator stress versus axial strain and “second yield” as the point of maximum curvature on

the same graph plotted at natural scale. The “first yield” is similar to Rotta’s “primary yield”

while the “second yield” is similar to Wong & Mitchell’s yield point. Bressani (1990) used

the same graph plotted on a log-log scale in an attempt to identify a clearer change in

behavior (see Figure 4.1e).

In the “stiffness method” method, the tangential modulus, usually E or G, is plotted

against strain in log-log scale or log-natural scale (e.g. Malandraki and Toll, 1996;

Cuccovillo and Coop, 1997). Malandraki and Toll (1996) identified yield of an artificially

87
CHAPTER 4

bonded soil from changes in the tangential stiffness (Etan) versus axial strain plotted to a

log-log scale (see Figure 4.1f). Cuccovillo and Coop (1997) defined the yield of structured

sand from changes in tangential shear modulus G versus log deviator strain or stress (see

Figure 4.1g).

There are also other methods for defining the “yielding” of soils. For example, Bergado

et al. (2006) define “yielding” of cement-treated soft Bangkok clay on the plot of shear

strain versus volumetric strain. They found that the shear strain versus volumetric strain is

bi-linear relationship and the transition point is defined as the yield point (see Figure 4.1h).

Airey (1993) used stress path of CIU specimen to define yield point for naturally cemented

carbonate soil. On the stress space, the transition point of effective stress path is defined as

yield point for most cases (see Figure 4.1i).

In this study, Coop and Atkinson’s method was used to determine the primary yield

points. As Figure 4.2 shows, by combining the primary yielding points obtained from

isotropic compression test, K0-consolidation test, constant stress ratio test and CID test, a

reasonably self-consistent primary yield locus of cement-treated marine clay can be obtained.

It should be pointed out that as shown in Figure 4.2, for CID specimen with lower confining

stress (e.g. 50kPa), the defined primary yield point is much lower than that defined by other

CID specimens with higher confining stress. This indicates that primary yield point for CID

specimen with lower confining stress is not well defined. Although consolidated undrained

(CIU) tests at effective confining pressures less than the isotropic primary yield stress shows

different stress path upon the compression pressure, Figure 4.2 shows that all peak strength

points are lower than or close to the primary yield locus formed by the primary yield points

88
CHAPTER 4

from isotropic compression test, K0-test, constant stress ratio test and CID test. Thus the

primary yield locus may be considered as the initial boundary surface in that it can not be

crossed by the effective stress path developed due to undrained events.

4.2.3 Summary

The definition of “yielding” of soils was discussed in detail. In this study, Coop and

Atkinson’s method was introduced to determine the yielding points for the primary yield

locus. The results showed that a reasonably consistent primary yield locus of cement-treated

marine clay can be obtained.

4.3. Isotropic compression behavior of cement-treated soil

In this section, isotropic compression behavior of cement-treated marine clay,

especially primary yielding, and the factors affecting it, is discussed.

4.3.1 General behavior under isotropic compression

Figures 4.3, 4.5, 4.7 - 4.11 and 4.13 - 4.17 show the isotropic compression curves for

specimens with different cement contents, total water contents, curing stresses and curing

periods, respectively. The first point plotted on each curve represents the specific volume

and the isotropic stress at the beginning of isotropic compression. Examination of all the

compression curves indicates that the initial void ratio decreases with the increase in cement

content, curing stress and curing time and the decrease in total water content. The specimens

show stiff behavior initially and become gradually softer as the isotropic stress increases,

89
CHAPTER 4

which is similar to behaviours reported in the literature for naturally/artificially structured or

cemented soils (e.g. Leroueil & Vaughan, 1990).

4.3.2 Cement content effect

Figure 4.3 shows the isotropic compression curve for specimens with different cement

content and 100% total water content cured under atmospheric pressure for 7 days. From

Figure 4.3, it can be seen that all compression curves seem to be parallel. The compression

index of specimens does not vary significantly with different cement content. This is

consistent with Huang & Airey’s (1998) results. Huang & Airey’s (1998) stated that the

slopes of the compression lines are practically independent of the amount of cement added.

As the cement content increases, the normal consolidation line shifts to the right in a

v : ln p ' plot, which was attributed to the change in grading of the soil. Figures 4.3 to 4.4

show that the primary yielding stress Ppy' under isotropic compression increases with the

increase in cement content of cement-treated soil. Thus cement-treated soil becomes stiffer

as the cement content increases. This is consistent with Kamruzzaman’s (2002) finding that

the apparent pre-consolidation pressure in one dimensional consolidation increases with

cement content.

4.3.3 Total water content effect

Figure 4.5 shows the isotropic compression curve for specimens with 50% cement

content and different total water content and cured under atmospheric pressure for 7 days.

From Figures 4.5 and 4.6, it can be seen that primary yielding stress under isotropic

90
CHAPTER 4

compression increases while the post-curing void ratio decreases with the decrease in total

water content. However, samples with different total water content appear to converge to the

same normal compression line, which indicates that the compression index seems to be

constant when only the total water content is varied. This is consistent with Lorenzo and

Bergado’s (2004) results. Lorenzo and Bergado (2004) stated that only the amount of cement

and soil solids or minerals controls the amount of new cementing products being formed

which eventually decide the initial structure of the soil. As a result, for a given cement

content, the higher total water content, the higher after-curing void ratio, the lesser the

bonded contacted area, the lesser the inherent cohesion, which results in a lower primary

yield stress.

4.3.4 Curing stress effect

Figures 4.7 to 4.11 show the isotropic compression curve for specimens with different

mix proportions cured under different curing stresses for 7 days. As Figure 4.12 shows, for

the same mix proportion, primary yield stress Ppy' under isotropic compression increases

while the post-curing void ratio decreases with increase in curing load (Figures 4.7 to 4.11).

For a given mix proportion, all samples appear to converge to the same normal compression

line regardless of curing stress. Rotta et al. (2003) and Chin (2006) also noted the same trend

for lower cement contents. Rotta et al. (2003) stated that an increase in density resulted in an

increase in the number of contact points between the soil particles where the cement can

form a bond. Consequently, the primary yield stress increases with the curing stress which

induced the decrease in the post-curing void ratio. However, as Lorenzo and Bergado (2004)

91
CHAPTER 4

noted, the initial structure is only controlled by the amount of cement and soil solids or

minerals which controls the amount of new cementing products being formed. As such,

when the compression pressure is much higher than the primary yield stress, the destructured

soil tends to converge with a unique compression line regardless of after–curing void ratio or

curing stress.

4.3.5 Curing period effect

Figures 4.13 to 4.17 show the isotropic compression curve for specimens with different

mix proportions and curing period under atmospheric pressure. From these figures, it can be

seen that for the same mix proportion, primary yield stress increases while post-curing void

ratio decreases with the curing time (Figure 4.18). However, the compression index seems to

be independent on curing period.

4.3.6 Summary

The results of isotropic compression tests in this section showed that the post-yield

compression index seems to be independent on cement content and curing period. The

post-yield compression index is also independent on total water content and curing stress.

However, the primary yielding stress Ppy' under isotropic compression increased with the

increase in cement content, curing load and curing time as well as decrease in total water

content.

92
CHAPTER 4

4.4. Triaxial compression behavior of samples consolidated at

pressures below the primary isotropic yield stress Ppy'

4.4.1 General behavior under triaxial compression

The naming convention used in this section (Figures 4.19 to 4.50) is

Aw-Cw-Pcur-CIU/CID-Pc'-Tcur, i.e., cement content (in percentage)-total water content (in

percentage)-curing stress-CIU/CID reconsolidation pressure-curing period. For example, test

10%-100%-0-ciu50-7ds (Figure 4.19) involves a sample with cement content 10% and total

water content 100% (mix proportion 10:1:11), which was cured under atmospheric pressure

for 7 days before test and then subjected to an isotropic consolidation pressure of 50 kPa and

sheared under undrained condition.

Figures 4.19 to 4.40 illustrate stress-strain behavior of cement-treated marine clay

specimens with different mix proportions, curing stress and curing time, which were

re-consolidated to a mean effective stress that is lower than the primary isotropic yield stress

prior to shearing. As can be seen, all samples sheared under undrained condition show stiff

response up to peak strength. Samples which were sheared undrained at high yield ratio

(ratio of consolidation stress Pc' to primary isotropic yield stress Ppy' ) >2 or ~2 show

approximately vertical stress path at the initial stages and reaches their peak strength along

the tensile failure envelope, which is represented by a straight line with a gradient of 3,

passing through the Origin. Stress states on this tensile failure envelope are associated with

zero effective radial stress on the triaxial specimen. This will be discussed further below.

Specimens sheared at lower yield ratio ~1 failed at peak point close to the tensile failure

envelope, with stress path trending toward the tensile failure envelope. All these suggest an

93
CHAPTER 4

stiff behavior up to the peak strength. The peak point in CIU test appears between 1% and

2% of shear strain or axial strain.

It should be noted that, owing to the limitations of the triaxial test, the undrained stress

path cannot cross the tensile failure envelope. For undrained triaxial test under low confining

stress, the build-up of excess pore pressure can easily lead to a state of zero effective radial

stress in the specimen. When this occurs, the pore pressure within the specimen becomes

equal to or larger than the total radial confining stress. This causes the rubber membrane to

detach from the specimen, thereby allowing pore water to drain out from the sides of the

specimen. Consequently, the state of the specimen is akin to that under unconfined

compression, with water being allowed to drain from the sides. Many researchers have

already reported the phenomenon of water being expelled from the sides of the specimens

under unconfined compression (e.g. Chin, 2006). This explains why all CIU specimens

reconsolidated under pressure lower than isotropic compression primary yield stress cannot

cross the tension failure envelope illustrated in Figures 4.19 to 4.40. This may not reflect the

true behavior of the cement-treated soil because unlike unstructured soils, one would expect

all cement-treated soil to contain some amount of inherent cohesion, which may also be

reflected as a tensile strength. If there is indeed true cohesion, then the yield envelope should

cross the q-axis at a positive intersect. This will be discussed later.

Samples which were tested drained at high yield ratio (>>2) showed nearly elastic

behavior before peak strength. Samples which were tested drained at low yield ratio showed

large volumetric compression up to the peak strength, which is often reached at larger shear

strain than CIU tests. Under higher effective confining stress, the point of peak strength is

94
CHAPTER 4

reached well after primary yield point, which indicates significant pre-peak strain hardening.

The above behavior was also reported by Chin (2006) for his cement treated marine clay

specimens.

4.4.2 Cement content effect

Figures 4.19 - 4.28 and 4.41 - 4.42 show triaxial stress-strain behavior for specimens

prepared by using different cement contents Aw ranging from 10% to 50%, total water

'
content Cw of 100% , zero curing stress Pcur and curing period 7 days. The test parameters of

the specimens are reflected in the specimen identifier. The results show that the cement

content has a significant effect on peak strength in CIU test and compression index in CID

test for specimen with the same consolidation stress. As shown in Figure 4.41, in CIU tests,

specimens with higher cement content show higher peak strength, which is reached at shear

strain of about 1% for all specimens. On the other hand, Figure 4.42 shows that for CID tests

the peak strength increases while the strain to peak (for same effective confining stress)

decreases as cement content increases. Moreover, specimens with higher cement content

show lower initial void ratio and lower compression index. These indicate that specimens

with higher cement content are also stiffer.

4.4.3 Total water content effect

Figures 4.27 to 4.34 and Figures 4.43 to 4.44 show triaxial stress-strain behavior for

specimens prepared by using different total water contents ranging from 100% to 183.3%,

with a constant cement content of 50% ,zero curing stress and 7 days of curing period. The

95
CHAPTER 4

results in Figures 4.43 and 4.44 show that in both CIU and CID tests, peak strength increases

as total water content decreases, which is not surprising. In addition, Figure 4.44 shows that

CID specimens with lower total water cement content also show smaller strain to peak (for

same effective confining stress). On the compression space, specimens with lower cement

content show lower initial void ratio and lower compression index.

4.4.4 Curing stress effect

Figures 4.35 to 4.40 and Figures 4.45 to 4.46 shows the stress-strain curves for

specimens with cement content 50% and total water content 133.3% (mix proportion 2:1:4)

cured under different curing stresses ranging from 0 to 250kPa for 7 days. The results in

Figures 4.45 and 4.46 show that in both CIU and CID tests, peak strength increases with

curing stress, which is not surprising. In addition, Figure 4.44 shows that CID specimens

with higher curing stress also show smaller strain to peak (for same effective confining

stress). Thus, increasing the curing stress has a similar effect as decreasing the total water

content.

4.4.5 Curing period effect

Figures 4.47- 4.50 show the stress-strain behavior for specimens with mix proportion

2:1:4 (cement content 50% and total water content 133.3%) and 5:1:6 (cement content 20%

and total water content 100%) cured under atmospheric pressure for different curing periods

ranging from 7days to 180days. As Figures 4.47 to 4.50 show, both CIU and CID specimens

show increase in strength with curing period. In addition, in CID specimens, the shear strain

96
CHAPTER 4

to peak strength decreases as curing period increases. On the compression space, specimens

with longer curing period show lower initial void ratio and lower compression index.

4.4.6 Summary

The results of triaxial compression test with consolidation stress lower than isotropic

compression primary yield stress in this section showed that for specimens with different

mix proportion under different curing condition, in CIU test all specimens show stiff

behavior with peak strength appearing around 1%~2% of axial strain whereas in CID test the

stress-strain behavior is depend on the effective confining stress or yield ratio. The results

also showed that increasing the cement content, curing stress and curing period or

decreasing the total water content all has a similar effect in increasing the peak strength of

the specimens. In all these cases, the increase in strength seems to be correlated at least

partially to a decrease in post-curing void ratio.

4.5. Yielding behavior of cement-treated Singapore marine clay

In this section, the yielding behavior of cement-treated Singapore marine clay is

discussed in detail. In this discussion, yielding is defined by the primary yield locus as

proposed by Coop and Atkinson (1993). A normalized primary yield locus will be fitted to

the experimental data. Following this, a correlation will be presented between the

unconfined compressive strength and the primary yield stress under isotropic compression.

This relation allows primary yield stress under isotropic compression to be deduced

indirectly. The effect of curing period on the primary yield locus of cement-treated

97
CHAPTER 4

Singapore marine clay will also be discussed.

4.5.1 Primary yielding and yield locus

Figures 4.19 to 4.50 demonstrate the primary yield points of specimens with different

mix proportion, total water content, curing stress and curing time. The primary yield point

was obtained by fitting data obtained largely from CID tests, isotropic compression test and

to a lesser extent K0-consolidation tests and constant stress ratio tests by using the method

described in Section 4.2. It was observed that the stress paths for specimens in CIU tests

generally tend to trend along or near to the tension failure envelope. Some CIU specimens

also reach peak strength around the primary yield locus. In such cases, the peak value of

deviator stress of CIU specimen was taken to be a primary yield point.

Figures 4.41 to 4.50 illustrate the effect of cement content, total water content, curing

stress and curing time on the primary yielding locus of cement-treated Singapore marine clay.

As these figures show, the primary yield locus expands with an increase in cement content,

curing stress and curing period as well as the decrease in total water content. The expansion

of the primary yield locus with cement content can be explained by the increased

cementation. The expansion of the primary yield locus with curing period can be explained

by the formation of secondary cementitious products through pozzolanic reaction. The

expansion of the primary yield locus with curing stress or the decrease in total water content

can be explained by the decrease in the post-curing void ratio.

Figures 4.51 to 4.52 show the yield points from a large number tests, with the mean

effective and deviator stresses normalized by their respective primary yield stress under

98
CHAPTER 4

isotropic compression. As can be seen, although the size of the yield locus changes with the

four factors as discussed above, they have virtually the same shape. As Figures 4.51 to 4.52

show, the normalized primary yield locus can be fitted approximately by either of these

relationships:

2
q ⎛ p' ⎞ ⎛ p' ⎞
= 3.65 ⎜⎜ ' ⎟⎟ − ⎜⎜ ' ⎟⎟ (4-1)
p 'py ⎝ p py ⎠ ⎝ p py ⎠

q p'
= 3.65 ' [−1.5ln( p ' / p 'py )]1/1.5 (4-2)
p 'py p py

However, it was noted that these relationships are both empirically fitted relations and

are only valid within the range represented. As will be shown in Chapter 6, the yield locus

may also be represented by a relationship which is derived from a theoretically motivated

energy equation. That yield locus will have more theoretical underpinning.

4.5.2 Relation of primary yield locus to other parameters

In the last section, it was shown that, once the isotropic primary yield stress Ppy' is

known, the corresponding primary yield locus can be determined. In this section, the yield

stress under isotropic compression will be correlated to two other engineering parameters,

which if known, can potentially allow the isotropic yield stress to be estimated. The first

parameter is the unconfined compressive strength qu and the second is the post-curing

void ratio. A relationship between the isotropic yield stress Ppy' , unconfined compressive

strength qu and cement content and total water content will also be proposed. This allows

the size of the primary yield locus to be estimated from the cement and total water contents.

99
CHAPTER 4

4.5.2.1 Correlation of isotropic yield stress to unconfined compressive strength

Figures 4.53 to 4.54 present the results of unconfined compressive strength test

respectively for specimens with different mix proportion and cured under different curing

stress for 7 days. By combining the results from these two figures with the isotropic yield

stress from Figures 4.3 to 4.11, a correlation can be obtained. As shown in Figure 4.57a, a

remarkably linear relationship is obtained between the unconfined compressive strength and

the difference between the isotropic yield stress and curing stress. Following Rotta et al.

(2003), the latter quantity will be termed hereafter as the incremental isotropic yield stress.

This allows the isotropic yield stress to be correlated to the unconfined compressive strength

via the relationship:

p 'py − pcur
'
= Aqu + B (4-3)

Where A=0.5476, B=19.679 in this study.

Figures 4.55 to 4.56 show the unconfined compressive strength test for specimens with

different mix proportions and curing period. By combining the results in these figures with

isotropic primary yield stress results from Figures 4.13 – 4.17, a similar correlation can be

obtained for different mix proportions and different curing period. As Figure 4.57b shows,

Equation 4-3 also fits remarkably well for specimens with different curing time. Similar

relationships have been proposed by previous researchers. For instance, Rotta et al. (2003)

showed that for artificially slightly cemented soil, incremental isotropic yield stress

p 'py − pcur
'
can be related to post-curing void ratio ecur and cement content Aw. Lorenzo &

Bergado(2004) also proposed that for cement-admixed clay at high water content,

unconfined compressive strength can be correlated to post-curing void ratio ecur obtained at

100
CHAPTER 4

the end of the curing stage and cement content. Thus incremental isotropic yield stress can

be related to unconfined compressive strength as described above. Therefore, the isotropic

primary yield stress may be estimated if the unconfined compressive strength is known.

4.5.2.2 Correlation of isotropic yield stress to post-curing void ratio

As shown in Figure 4.58, the incremental yield stress p 'py − pcur


'
can also be related to

the post-curing void ratio via a logarithmic function of the form

ecur = A1 ln( p 'py − pcur


'
) + B1 (4-4)

As Figure 4.58 shows, specimens with the same cement content can still have different

post-curing void ratio and incremental yield stress. Thus, the cement content is not the only

parameter affecting the incremental yield stress. Miura et al. (2001) has also reported that,

for cement-stabilized clay at high water content, the ratio of the initial water content of the

clay to the cement content is the prime parameter for analysis of strength and deformation

behavior.

Figure 4.59 shows the variation of the coefficients A1 and B1 in Eq. 4-4 with the ratio of

the pre-curing total water content to the cement content, which is Cw/Aw. The relationship

can be fitted by the functions:

A1 = −0.566 − 0.461(Cw / Aw) −0.8919 , R 2 = 0.987 (4-5)

B1 = 9.1372(Cw / Aw)0.2198 , R 2 = 0.981 (4-6)

Substituting Equations (4-5) and (4-6) into Eq. 4-4 leads to

ecur − 9.1372(Cw / Aw) −0.2198


p 'py − pcur
'
= exp( ) (4-7)
−0.566 − 0.461(Cw / Aw) −0.8919

101
CHAPTER 4

4.5.2.3 Direct correlation of isotropic yield stress to cement and total water

content

Since the post-curing void ratio may also be related to the cement content and total

water content, it is possible that the isotropic yield stress may be related directly to the

cement and total water contents. Figure 4.60 plots the relationship between isotropic

compression primary yield stress p 'py and cement content Aw for specimens with different

total water content Cw without curing stress. As it can be seen from Figure 4.60a, the

isotropic primary yield stress p 'py can be expressed as below:

p 'py
= A2 Aw + B2 (4-8a)
Pap

Where Pap is atmospheric pressure.

Pap = 101.325kPa (4-8b)

Figure 4.60b - c show that parameter A2 and B2 can be related to total water content Cw

in the formula below respectively:

A2 = cCw d (4-9)

B2 = eCw + f (4-10)

Where c=9.8652, d=-2.5042, e=-0.8201, f=1.189 in this study.

Thus the isotropic primary yield stress p 'py can be written as below:

p 'py
= cCw d Aw + eCw + f (4-11)
Pap

Figure 4.60d illustrates the predicted p 'py by using Equation 4-11. It can be seen that

the difference between the predicted value and experimental value of p 'py is very small

(generally less than about 10%).

102
CHAPTER 4

4.5.2.4 Correlation of unconfined compressive strength to mix proportions

By combining Eqs. (4-11) and (4-3), the unconfined compressive strength of

cement-treated marine clay can be written as below:

Pap (cCwd Aw + eCw + f ) − Pcur


'
−B
qu = (4-12a)
A

Or

qu = Pap (c ' Cwd Aw + e ' Cw + f ') − Pcur


'
/A (4-12b)

In which c ' = c / A , e ' = e / A , f ' = ( f − B / Pap ) / A

For specimens cured under atmospheric pressure, formula (4-12a) or (4-12b) becomes

qu = Pap (c ' Cwd Aw + e ' Cw + f ') (4-12c)

Where c'=18.015, d=-2.504, e'=-1.498, f'=1.817 in this study

Figure 4.61a shows good match between the experimental and predicted qu by Eq.

(4-12c). As can be seen, the difference between the experimental and predicted qu is lower

than 15%. Lee et al. (2005) has also proposed a relationship between unconfined

compressive strength and soil/cement ratio and water/cement ratio for cement treated

Singapore marine clay as below:

em( s / c )
qu = q0 (4-13)
( w / c)n

Where q0 =4000, m=0.62, n=3.0 for 7 days of curing period.

Figure 4.61b shows the unconfined compressive strength qu of cement-treated

Singapore marine clay specimens with different mix proportions and 7 days of curing period.

Also plotted in this figure are Lee’s (1999) data. It can be seen from Figure 4.61, the value

of qu from this study is higher than that of Lee’s(1999). Figure 4.62 shows the simulated

103
CHAPTER 4

results of Lee’s (1999) data in Figure 4.61 by using formula (4-13) and (4-12c) respectively.

The simulation curve by Lee et al.’s method in Figure 4.62 is the same as that of Figure 4.61.

As it can be seen, the simulated qu by formula (4-12c) and by formula (4-13) is generally

consistent for various of s/c and w/c except that the value of qu simulated by formula (4-12c)

is generally higher than that by formula (4-13) when water content is lower than 150 %( i.e.

w/c is 2.2 to 7.5 corresponding s/c 0.47 to 4.0). This is not surprising because the marine

clay used by Lee (1999) was taken from a different location in Singapore, and the property

of marine clay may have a significant effect on the experiment results. In addition, Lee

(1999) investigated a range of mix ratios which is not exactly the same as the range

investigated in this study. The maximum cement content and water content used in this study

is 100% and 150% respectively, which is lower than 213% and 280% respectively used in

Lee’s study (1999). As Figure 4.63a shows, by modifying the formula (4-12c) proposed by

this study, the simulated results are very close to those obtained from Lee et al.’s (2005)

formula. The modified formula is as below:

qu = Pap (c ' Cwd Aw + e ' Cwg ) (4-14)

Where Pap = 101.325kPa , c’=6.994, d=-3, e’=1.835, g=-3.

Alternatively, as Figure 4.63b shows, an improved matching with the experiment

results of this study can also be obtained if q0 in Lee et al.’s (2005) relationship is increased

from 4000kPa to 7000kPa. As Lee et al. (2005) noted, the value of q0 is dependent upon the

properties of the untreated soil, especially the plasticity index, with a higher plasticity index

giving a higher value of q0. Marine clay from different locations in Singapore has somewhat

different plasticity index and this could have accounted for the differences in q0. This fitted

104
CHAPTER 4

value of q0 also lies within Gallavresi’s (1992) range of 5000kPa to 10000kPa.

4.5.2.5 Summary

The above discussion shows that

1. The primary yield locus can be approximately normalized by the isotropic primary yield

stress.

2. The isotropic primary yield stress can be determined by

a) unconfined compressive strength qu (considering curing stress), or

b) cement content and total water content (only for specimens without curing stress), or

c) post-curing void ratio together with the ratio of total water content to the cement

content (considering curing stress).

4.6. Triaxial compression behavior of samples consolidated at

effective pressure higher than the isotropic primary yield

stress Ppy'

In this section, the triaxial compression behavior of cement-treated marine clay

specimens reconsolidated to effective confining stress p0' higher than isotropic primary

yield stress will be presented. The effective confining stress p0' is termed as

pre-compression pressure hereafter. The evolution of yield locus with the pre-compression

pressure p0' will be examined here. The effects of confining stress, cement content, total

water content on post-yield stress-strain behavior of cement-treated marine clay will be also

discussed.

105
CHAPTER 4

4.6.1 General behavior under triaxial compression

The naming convention used in this section (Figures 4.64 to 4.68) is

Aw-Cw-CIU/CID p0' - pc' , i.e., cement content-total water content-CIU/CID isotropic

pre-compression pressure-effective confining pressure during shearing. All specimens

discussed in this section were cured under atmospheric pressure for 7 days before testing.

The naming convention used in section 4.4 is not applicable to this section as all of the

specimens here were reconsolidated under a stress higher than isotropic yield stress and then

most of them were swelled to a lower stress. For example, test 10%-100%-ciu300-50

(Figure 4.61) involves a sample with cement content 10% and total water content 100%

(mix proportion 10:1:11), which was reconsolidated to 350kPa and then swelled to 50kPa

and sheared under undrained condition.

Figures 4.64 to 4.68 illustrate stress-strain behavior of cement-treated marine clay

specimens with different mix proportion, which were isotropically reconsolidated to an

effective confining stress which is higher than the isotropic primary yield stress. These

samples were either sheared at their isotropic pre-compression (IPC) pressure or else

allowed to swell to a lower pressure before shearing. The latter will be characterized by an

isotropic overconsolidation ratio (IOCR), which is the ratio of the IPC pressure to the mean

effective stress after swelling. The isotropic overconsolidation ratios of the specimens in

these figures are ranged from 1 to 50. As Figures 4.64 to 4.68 show, these samples exhibit

similar trend of stress-strain behavior to those discussed in section 4.4, which have not been

consolidated past the isotropic primary yield stress Ppy' . In all samples, by using the method

106
CHAPTER 4

described in sections 4.2 and 4.4, yield loci can be drawn through the IPC pressures and the

yield points deduced from CID and stress path tests.

However, there are several differences. Firstly, samples which have been isotropically

consolidated past the isotropic primary yield points show much more abrupt drop in stiffness

and change in the compression path upon yielding, especially for the drained samples. The

drained samples which were sheared at their IPC pressure all show small elastic zones in

their stress-strain curves. This may be attributed to the relatively short curing period of the

samples, which gives rise to further formation of structure during and after consolidation.

Secondly, both CIU and CID samples which were sheared under high IOCRs, such as

CIU1500-50 and CID1500-50, show stiff response up to the peak strength, followed by

dilation and strain softening. In such cases, the yield points are taken to be coincident with

the peak strength (see Figures 4.64-4.68). CIU specimens which were sheared at lower

IOCR of ~2 also show stiff response up to peak strength. Their stress paths are also roughly

vertical, which suggest largely elastic behavior up to the peak strength. The above behavior

also is same as that reported by Chin (2006) for specimens with mix proportion 5:1:6 and

maximum consolidation pressure 1500kPa, which is lower than that used in this study, i.e.

2500kPa.

For CIU specimen with high IPC pressure (e.g. 1500kPa) and low IOCR of ~2 (e.g.

swelling from 1500kPa to 750kPa), the peak strength may be slightly higher than the yield

locus deduced from CID and stress path tests. However, as can be seen from Figures 4.64 to

4.68, it was found that the transition point of stress path for this case is close to the yield

locus deduced from CID and stress path test. It was also found that this point occurs at about

107
CHAPTER 4

0.55% of axial strain and seems to coincide with the transition point on the deviator

stress-strain curve, in which the deviator stress-strain relationship starts to deviate from

linear behavior. The definition of yield stress from stress path here is similar to Airey’s

(1993) method for CIU test.

4.6.2 Evolution of yield locus

Figures 4.64 to 4.68 illustrate the evolution of yield locus for specimens with different

mix proportions. As expected, the yield locus expands with IPC pressure, owing to

densification effects. However, as figures 4.69 to 4.70 show, the normalized yield locus

tends to contract, changing from a steep arch to an ellipse. This is due to kinematic softening

induced by the breakage of inter-aggregate bonds or destructuration. The destructuration

mainly comes from isotropically reconsolidation and appears to continue after the yield

locus has been reached during shearing. Thus, the post-yield behavior of the cement-treated

soil appears to be influenced by densification effects as well as breakage of inter-aggregate

bonds. This is similar to that proposed by Gens and Nova (1993) and highlighted by Huang

and Airey(1998). However, Figures 4.70a to 4.70b also show that even under higher IPC

pressure (2500kPa), the friction coefficient M of the yield locus is still much higher than that

of reconstituted marine clay (with M value of 0.9) This indicates that even under severe

shearing, the cemented-treated soil is not equivalent to that of reconstituted marine clay. This

is not surprising since cement-treatment does not merely superimpose bonding on the marine

clay particles, and the cement actually reacts with the clay mineral and changes its

mineralogy. Thus, it is highly possible that reconstituted marine clay does not represent an

108
CHAPTER 4

intrinsic state of the cement-treated clay.

Some SEM tests were conducted to investigate the destructuration of cement treated

specimens under triaxial loading condition. The specimen was isotropically reconsolidated

under 2500kPa and then swelled back to 1250kPa before undrained shearing or drained

shearing. Two types of samples were obtained from cement- treated specimens post-test.

One is taken from the outside of the slip band whereas another one from the inside of the

slip band. Figures 4.71 to 4.72 show that dominant microstructure changes outside of slip

band appear to be aggregate squashing and void closure with less aggregate or particle

break-up. On the other hand, massive aggregate and particle break-up appeared to have

occurred within the slip band, especially for CID specimens. This indicates that

destructuration may only complete within the slip band. Results also show that

microstructure change in CIU specimens is less serious than that in CID specimens due to

undrained condition during shearing stage. Kamruzzaman (2002) and Chin (2006) also

reported similar results.

4.7. Tensile splitting strength test and modification of primary yield

locus

This section discusses the results of tensile strength measurement by the Brazilian test,

the procedure of which has been described in Chapter 3.

4.7.1 Stresses along radial loading section

For a solid cylinder (disk) with thickness t and diameter D under two equal and

109
CHAPTER 4

opposite line loading P along central vertical section, the stress along horizontal direction(x)

σ x at any point within vertical central section can be expressed as (Timoshenko and Goodier,

1970):

2P
σx = − (4-15a)
π Dt
The stress along vertical direction (y) σ y at any point within vertical central section can be

considered as superposition of two kinds of stresses (Timoshenko and Goodier, 1970):

2P
a. A uniform tension in the plane of the disk : −
π Dt
b. Two simple radial stress distributions: 2 P /(π r1t ) and 2 P /(π r2t ) . r1 , r2 is the distance

of the point from the top loading and bottom loading separately. For any point with y

D D
distance from the center of the disk, we have r1 = − y and r2 = + y . Therefore,
2 2
σ y can be expressed as below:

2P 4D2
σy = ( 2 − 1) (4-15b)
π Dt D − 4 y 2

For quasi-brittle material (e.g. concrete, rock), the tensile splitting strength is suggested

as (e.g. BS1881-117, 1983):

2P
σt = (4-16)
π tD
The theoretical solution to the stress distribution along the central vertical section is

shown in Figure 4.73.

4.7.2 Tensile strength of cement-treated marine clay

As shown in Figure 4.74, cement-treated marine clay specimens show similar brittle

behavior as rock or concrete, with cracking along the central vertical plane under radial

compression. Table 4.1 summarizes the tensile strength results.

110
CHAPTER 4

As can be seen from Figures 4.75a to 4.75c, the tensile strength σ t of cement-treated

marine clay specimen increases with the increase in cement content Aw, the decrease in total

water content Cw, as well as the increase in curing time and curing stress. Additionally, to

investigate the effect of radial compression rate on the tensile strength, two compression

rates, viz. 0.005mm/min and 0.5mm/min, were used for each specimen configuration.

Besides this, the Brazilian test was also conducted under dry condition to study the drainage

condition effect on the tensile strength. Figure 4.76 shows the strengths obtained under the

two compression rates, plotted against each other. As can be seen, all points plot within a

±15% sector, with the tensile strength obtained under 0.5mm/min usually being equal to or

slightly higher than that obtained under 0.005mm/min. This shows that the hundred-fold

increase in compression rate does not significantly increase the measured tensile strength.

Figure 4.77 shows the comparisons between compression test under dry condition and water

condition. It was found that the tensile strength obtained under dry testing is usually equal to

or slightly higher than that obtained under water, the relative difference between different

loading environments being no more than 15% range for most cases.

The evidence presented above points to the existence of a true tensile strength which is

largely unaffected over a range of compression rates and drainage conditions. This would

suggest the presence of some inherent or “true” cohesion, as opposed to that arising from

negative pore pressure or suction; the latter would be expected to be more dependent upon

drainage conditions and compression rates. The presence of such inherent or “true” cohesion

can be demonstrated by a simple prolonged submergence test, in which a sample is simply

submerged in water for a prolonged period of time. As shown in Figure 4.78, one remoulded

111
CHAPTER 4

and preloaded marine clay sample and a cement-treated marine clay sample was placed into

different containers and left submerged in water. After several weeks, the remoulded marine

clay sample has largely disintegrated. On the other hand, the cement-treated sample remains

largely intact.

4.7.3 Correlation of tensile strength to unconfined compressive strength

Figure 4.79 shows the relationship between tensile strength and unconfined

compressive strength. As can be seen, a good linear relationship can be obtained. As such,

the tensile strength can also be correlated to the unconfined compressive strength via the

relation

qu = 7.87σ t (4.17a)

Or

σ t = 0.127 qu (4.17b)

The ratio of tensile to compressive strength lies within the ranges reported in other studies.

For instance, Huang and Airey (1998) reported values between 0.1 and 0.3 for artificially

cemented carbonate sand. Mitchell (1976) reported values between 0.2 and 0.3 for cement

treated soils. Clough et al.(1981) gave a value of about 0.1 for cemented sand.

4.7.4 Modification of primary yield locus of cement-treated marine clay under

triaxial loading condition

As mentioned before, for CIU cement-treated specimens with low compression

pressure or effective confine stress, the stress path tends to trend along or close to the tension

112
CHAPTER 4

cut-off line passing through the origin of stress space p’-q when the specimens reach peak

value of deviator stress. This is due to the occurrence of outward drainage through the sides

of the specimens when the rubber membrane detaches from the specimen under condition of

zero lateral effective stress. This may not reflect the true behaviour of cement-treated soil. In

this section, the tensile strength measured by Brazilian test will be used to redefine the

tension failure envelope which could not be determined by triaxial tests. Assuming that the

specimens are isotropic and a single tension strength is applicable to all stress situations, the

condition for the actual tension cut-off under triaxial condition can be expressed as

σ 3' = −σ t (4-18)

Then

q = σ 1' − σ 3' = σ 1' + σ t (4-19)

And

1 1
p ' = (σ 1' + 2σ 3' ) = (σ 1' − 2σ t ) (4-20)
3 3

Eliminating σ1'from Eq. 4-15 and 4-16 leads to

q = 3( p '+ σ t ) (4-21)

Figures 4.80 to 4.83 illustrate the redefined tension failure envelope based on the

tensile strength form Brazilian test. Base on the relationship between the primary yield locus

determined earlier and the tension cut-off line passing through origin of the stress plane, the

primary yield locus may be extrapolated to match the redefined tension failure envelope.

Figures 4.80 to 4.83 illustrate the modified primary yield locus for cement-treated marine

clay specimens by extrapolating if true undrained conditions can be enforced within the CIU

specimens without allowing water to egress from the sides.

113
CHAPTER 4

Table 4.1 Summary of tensile splitting strength test results

Mix Curing Curing Tensile


Tensile strength σ t (kPa) Tensile strength σ t (kPa)
proportion period stress strength
Slow Compression rate(0.005mm/min) fast Compression rate(0.5mm/min)
S-C-W (days) (kPa) σ t (kPa)
sample under water Sample Under water Without water
(average)
7 0 30.97 31.00 29.34 30.33 28.94 33.61 26.14 28.78 29.43 30.80 30.26 NA 29.97
10-1-11
28 0 NA 43.30 52.66 47.78 44.78 46.49 48.97 47.33

7 0 54.64 48.14 50.78 45.70 49.36 55.04 51.95 50.45 52.82 NA 50.98
5-1-6
28 0 NA 66.62 84.54 72.20 83.22 73.4 80.25 76.72

7 0 72.78 70.26 65.02 68.27 62.78 68.47 62.70 74.46 62.22 NA 67.44
10-3-13
28 0 NA 111.63 101.91 114.02 125.79 109.25 112.18 111.87

7 0 117.23 116.02 118.32 119.08 121.69 111.14 116.87 126.31 114.55 117.72 128.88 128.61 NA 119.71
2-1-3
28 0 158.30 174.76 182.07 174.13 NA 172.31

7 0 55.34 51.04 52.27 52.17 55.11 56.55 58.68 59.08 62.20 62.63 57.29 61.71 53.12 56.71
2-1-4
28 0 NA 83.15 84.74 80.15 76.62 81.31 74.84 80.14

7 50 117.95 130.06 137.11 120.55 116.25 125.14 121.41 126.67 125.60 NA 124.53

7 100 152.62 165.68 167.53 145.03 168.56 156.59 145.81 NA 157.40

7 250 280.06 255.11 310.35 253.33 258.15 281.35 311.21 257.30 NA 275.86

7 0 22.90 16.89 20.37 22.85 20.19 19.30 21.76 20.94 NA 20.65


2-1-5
28 0 41.35 34.26 38.61 36.49 40.16 36.81 36.65 37.85 37.77

2-1-5.5 7 0 18.23 22.56 16.38 18.19 17.37 NA 18.55

114
CHAPTER 4

(a) After Mitchell (1970) (b) After Wong and Mitchell (1975)

parallel to ICL
3.6 ICL
p'gy
Specific volume

3.2

2.8

2.4

1 10 100 1000 10000


Consolidation pressure, p' (log scale)

(c) After Rotta et al. (2003) (d) After Cotechia and Chandler (2000)

Figure 4.1: Determination of initial yield stress (a-d)

115
CHAPTER 4

(e) After Bressani (1990) (f) After Malandraki and Toll (1996)

Yield point

Yield point

Yield point

(g) After Cuccovillo and Coop (1997) (h) After Bergado et al. (2006)

(i) After Airey (1993)

Figure 4.1: Determination of initial yield stress (e-i)

116
CHAPTER 4

750 750 10-3-13 CIU TEST


10-3-13 CIU TEST
CIU50 CIU50
CIU100 CIU100
CIU150
CIU150
CIU300
CIU350 CIU300
600 600
tension cut-off line CIU350
primary yield point
primary yield locus
ko test
constant stress ratioη=1.5
deviator stress q(kPa)

deviator stress q(kPa)


450 450

300 300

150 150

0 0
0 200 400 600 0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 0.18 0.2 0.22
mean normal effective stressP'(kPa)
axial strain εs(%)

0 100 200 300 400 500 600 700


3.4

10-3-13 CID TEST


CID50
CID150
CID200
CID250
3.3 primary yield point
specific volume v

3.2

3.1

Figure 4.2: Determination of primary yield locus for cement-treated marine clay
specimen (mix proportion 10:3:13 cured under atmospheric pressure for 7days)

117
CHAPTER 4

10 100 p'(kPa) 1000 10000


Ppy'
3.6
Aw=10%,ppy'=130kPa
Aw=15%,ppy'=210kPa
Aw=20%,ppy'=250kPa
Aw=30%,ppy'=355kPa
Aw=50%,ppy'=580kPa
3.2
specific volume v

2.8

2.4

Figure 4.3: Isotropic compression curves for specimens with different cement content
and 100% total water content cured under atmospheric pressure for 7 days

600

500
primary yielding stress Ppy'(kPa)

400

300

200

100

10 20 30 40 50
cement content Aw(%)

Figure 4.4: Effect of cement content on primary yielding stress under isotropic
compression for specimens with 100% total water content cured under atmospheric
pressure for 7 days

118
CHAPTER 4

10 100 p'(kPa) 1000 10000


5
Ppy'

4.75 Cw183%,ppy'=80kPa

Cw167%,ppy'=105kPa

Cw150%,ppy'=180kPa
4.5 Cw133%,ppy'=275kPa
specific volume v

Cw100%,ppy'=580kPa

4.25

3.75

3.5

3.25

Figure 4.5: Isotropic compression curves for specimens with different total water
content and 50% cement content cured under atmospheric pressure for 7 days
600

500
primary yielding stress Ppy'(kPa)

400

300

200

100

100 120 140 160 180 200


total water content Cw(%)

Figure 4.6: Effect of total water content on primary yielding stress under isotropic
compression for specimens with 50% cement content cured under atmospheric
pressure for 7 days

119
CHAPTER 4

10 100 1000 10000


p'(kPa)
3.6
Ppy'
mix proportion 10-1-11
pcur'=0kPa, ppy'=130kPa

pcur'=50kPa, ppy'=220kPa

pcur'=100kPa, ppy'=300kPa

pcur'=250kPa, ppy'=590kPa

pcur'=350kPa, ppy'=1000kPa
3.2
specific volume v

2.8

2.4

Figure 4.7: Isotropic compression curves for specimens with mix proportion 10-1-11
cured under different curing stress for 7 days
10 100 1000 10000
p'(kPa)
3.6
mix proportion 20-3-23
Ppy' pcur'=0kPa, ppy'=210kPa

pcur'=50kPa, ppy'=275kPa

pcur'=100kPa, ppy'=400kPa

pcur'=250kPa, ppy'=700kPa

3.2
specific volume v

2.8

2.4

Figure 4.8: Isotropic compression curves for specimens with mix proportion 20-3-23
cured under different curing stress for 7 days

120
CHAPTER 4

10 100 1000 10000


p'(kPa)
3.6

Ppy' mix proportion 5-1-6


pcur'=0kPa, ppy'=260kPa

pcur'=50kPa, ppy'=350kPa
3.4
pcur'=100kPa, ppy'=450kPa

pcur'=250kPa, ppy'=720kPa

3.2
specific volume v

2.8

2.6

2.4

Figure 4.9: Isotropic compression curves for specimens with mix proportion 5-1-6
cured under different curing stress for 7 days

10 100 1000 10000


p'(kPa)
4

mix proportion 2-1-4


Ppy'
pcur'=0kPa, ppy'=275kPa

pcur'=50kPa, ppy'=600kPa

pcur'=100kPa, ppy'=800kPa

pcur'=250kPa, ppy'=1250kPa
3.6
specific volume v

3.2

2.8

Figure 4.10: Isotropic compression curves for specimens with mix proportion 2-1-4
cured under different curing stress for 7 days

121
CHAPTER 4

10 100 1000 10000


p'(kPa)
4.4
Ppy'
mix proportion 2-1-5
pcur'=0kPa, ppy'=100kPa

pcur'=50kPa, ppy'=475kPa

pcur'=100kPa, ppy'=700kPa

4 pcur'=200kPa, ppy'=1050kPa
specific volume v

3.6

3.2

2.8

Figure 4.11: Isotropic compression curves for specimens with mix proportion 2-1-5
cured under different curing stress for 7 days

1600
curing stress effect
mix proportion 10-1-11
mix proportion 20-3-23
mix proportion 5-1-6
mix proportion 2-1-4
mix proportion 2-1-5
1200
primary yielding stress Ppy'(kPa)

800

400

0 50 100 150 200 250


curing stress Pcur'(kPa)

Figure 4.12: Effect of curing stress on primary yielding stress under isotropic
compression for specimens with different mix proportion cured for 7 days

122
CHAPTER 4

10 100 1000 10000


p'(kPa)
3.6
Ppy'
mix proportion 20-3-23
tcur'=7days, ppy'=210kPa
tcur=28days, ppy'=285kPa
tcur=90days, ppy'=385kPa
3.4
tcur'=210days, ppy'=500kPa

3.2
specific volume v

2.8

2.6

Figure 4.13: Isotropic compression curves for specimens with mix proportion 20-3-23
and different curing time under atmospheric pressure
10 100 1000 10000
p'(kPa)
3.6

mix proportion 5-1-6


Ppy' tcur'=7days, ppy'=260kPa
tcur=28days, ppy'=370kPa
tcur=90days, ppy'=480kPa
3.4 tcur'=180days, ppy'=535kPa
specific volume v

3.2

2.8

Figure 4.14: Isotropic compression curves for specimens with mix proportion 5-1-6
and different curing time under atmospheric pressure

123
CHAPTER 4

10 100 1000 10000


p'(kPa)
4

Ppy'

3.8

mix proportion 2-1-4


3.6
tcur'=7days, ppy'=275kPa
tcur=28days, ppy'=385kPa
specific volume v

tcur=90days, ppy'=500kPa
tcur'=180days, ppy'=580kPa
3.4

3.2

2.8

Figure 4.15: Isotropic compression curves for specimens with mix proportion 2-1-4
and different curing time under atmospheric pressure (test for curing period 90 days
did not complete)

10 100 1000 10000


p'(kPa)

Ppy' mix proportion 2-1-4.5


tcur'=7days, ppy'=210kPa

4.4 tcur=35days, ppy'=300kPa


tcur=90days, ppy'=350kPa
specific volume v

3.6

3.2
Figure 4.16: Isotropic compression curves for specimens with mix proportion 2-1-4.5
and different curing time under atmospheric pressure

124
CHAPTER 4

10 100 1000 10000


p'(kPa)
4.4
mix proportion 1-1-3
tcur'=7days, ppy'=300kPa
tcur=28days, ppy'=525kPa
Ppy' tcur=90days, ppy'=600kPa

4.2
specific volume v

3.8

3.6

Figure 4.17: Isotropic compression curves for specimens with mix proportion 1-1-3
and different curing time under atmospheric pressure

600

500
primary yielding stress Ppy'(kPa)

400

curing time effect


mix proportion 20-3-23
mix proportion 5-1-6
mix proportion 2-1-4
mix proportion 2-1-4.5
300
mix proportion 1-1-3

200

0 50 100 150 200 250


curing time tcur(day)

Figure 4.18: Effect of curing time on primary yielding stress under isotropic
compression for specimens with different mix proportion cured under atmospheric
pressure

125
CHAPTER 4

10-1-11 CIU TEST


270 10%-100%-0-CIU50-7ds
10%-100%-0-CIU75-7ds
10%-100%-0-CIU100-7ds
10%-100%-0-CIU120-7ds
tension cut-off line
primary yield point
primary yield locus

180
deviator stress q(kPa)

90

0
0 40 80 120 160
mean normal effective stressP'(kPa)

(a)
270 10-1-11 CIU TEST
10%-100%-0-CIU50-7ds
10%-100%-0-CIU75-7ds
10%-100%-0-CIU100-7ds
10%-100%-0-CIU120-7ds

180
deviator stress q(kPa)

90

0
0 0.01 0.02 0.03 0.04 0.05
axial strain εs

(b)
Fig. 4.19: Undrained triaxial shearing behavior of specimens with mix proportion
10-1-11 cured under atmospheric pressure for 7 days (a) stress path (b) deviator
stress-strain curve

126
CHAPTER 4

600 600
10-1-11 CID TEST
10%-100%-0-CID25-7ds
10%-100%-0-CID50-7ds
10%-100%-0-CID75-7ds
500
10%-100%-0-CID100-7ds

400 400

deviator stress q(kPa)


deviator stress q(kPa)

300

200 10-1-11 CID TEST 200


10%-100%-0-CID25-7ds
10%-100%-0-CID50-7ds
10%-100%-0-CID75-7ds
10%-100%-0-CID100-7ds 100
tension cut-off line
primary yield point
primary yield locus
0 0
0 100 200 300 0 0.05 0.1 0.15 0.2 0.25 0.3
mean normal effective stressP'(kPa) deviator strain εs
0 100 200 300

10-1-11 CID TEST


3.6 10%-100%-0-CID25-7ds
10%-100%-0-CID50-7ds
10%-100%-0-CID75-7ds
10%-100%-0-CID100-7ds
primary yield point
3.5
specific volume v

3.4

3.3

3.2

Figure 4.20: Drained triaxial shearing behavior of specimens with mix proportion 10-1-11 cured
under atmospheric pressure for 7 days

127
CHAPTER 4

450 20-3-23 CIU TEST


15%-100%-0-CIU50-7ds
15%-100%-0-CIU100-7ds
15%-100%-0-CIU150-7ds
15%-100%-0-CIU200-7ds
360 tension cut-off line
primary yield point
primary yield locus

deviator stress q(kPa)


270

180

90

0
0 50 100 150 200 250
mean normal effective stressP'(kPa)

(a)
450
20-3-23 CIU TEST
15%-100%-0-CIU50-7ds
15%-100%-0-CIU100-7ds
15%-100%-0-CIU150-7ds
360 15%-100%-0-CIU200-7ds
deviator stress q(kPa)

270

180

90

0
0 0.05 0.1 0.15 0.2 0.25
axial strain εs

(b)

Figure 4.21: Undrained triaxial shearing behavior of specimens with mix proportion
20-3-23 cured under atmospheric pressure for 7 days (a) stress path (b) deviator
stress-strain curve

128
CHAPTER 4

1000 1000
20-3-23 CID TEST
15%-100%-0-CID50-7ds
900 15%-100%-0-CID100-7ds
15%-100%-0-CID150-7ds
800 800

700

deviator stress q(kPa)


deviator stress q(kPa)

600 600

500

400 400

20-3-23 CID TEST 300


15%-100%-0-CID50-7ds
15%-100%-0-CID100-7ds
200 200
15%-100%-0-CID150-7ds
tension cut-off line
primary yield point 100
primary yield locus

0 0
0 100 200 300 400 500 0 0.05 0.1 0.15 0.2 0.25 0.3
mean normal effective stressP'(kPa) deviator strain εs
0 100 200 300 400 500
3.6

20-3-23 CID TEST


15%-100%-0-CID50-7ds
15%-100%-0-CID100-7ds
3.5
15%-100%-0-CID150-7ds
primary yield point
specific volume v

3.4

3.3

3.2

3.1

Figure 4.22: Drained triaxial shearing behavior of specimens with mix proportion
20-3-23 cured under atmospheric pressure for 7 days

129
CHAPTER 4

630
5-1-6 CIU TEST
20%-100%-0-CIU50-7ds
20%-100%-0-CIU100-7ds
540
20%-100%-0-CIU200-7ds
tension cut-off line
primary yield point
450 primary yield locus

deviator stress q(kPa)


360

270

180

90

0
0 50 100 150 200 250
mean normal effective stressP'(kPa)

(a)

450
5-1-6 CIU TEST
20%-100%-0-CIU50-7ds
20%-100%-0-CIU100-7ds
20%-100%-0-CIU200-7ds
360
deviator stress q(kPa)

270

180

90

0
0 0.05 0.1 0.15 0.2 0.25
axial strain εs

(b)

Figure 4.23: Undrained triaxial shearing behavior of specimens with mix proportion
5-1-6 cured under atmospheric pressure for 7 days (a) stress path (b) deviator
stress-strain curve

130
CHAPTER 4

1100 1100
5-1-6 CID TEST
20%-100%-0-CID50-7ds
1000 1000
20%-100%-0-CID100-7ds
20%-100%-0-CID200-7ds
900 900

800 800

deviator stress q(kPa)


700 700
deviator stress q(kPa)

600 600

500 500

400 400

300 5-1-6 CID TEST 300


20%-100%-0-CID50-7ds
20%-100%-0-CID100-7ds
200 200
20%-100%-0-CID200-7ds
tension cut-off line
100 100
primary yield point
primary yield locus
0 0
0 200 400 600 0 0.05 0.1 0.15 0.2 0.25 0.3
mean normal effective stressP'(kPa) deviator strain εs

0 100 200 300 400 500 600


3.5

5-1-6 CID TEST


20%-100%-0-CID50-7ds
3.4 20%-100%-0-CID100-7ds
20%-100%-0-CID200-7ds
primary yield point

3.3
specific volume v

3.2

3.1

2.9

Figure 4.24: Drained triaxial shearing behavior of specimens with mix proportion 5-1-6
cured under atmospheric pressure for 7 days

131
CHAPTER 4

750 10-3-13 CIU TEST


30%-100%-0-CIU50-7ds
30%-100%-0-CIU100-7ds
30%-100%-0-CIU150-7ds
30%-100%-0-CIU300-7ds
30%-100%-0-CIU350-7ds
600 tension cut-off line
primary yield point
primary yield locus
deviator stress q(kPa)

450

300

150

0
0 100 200 300 400
mean normal effective stressP'(kPa)

(a)

750
10-3-13 CIU TEST
30%-100%-0-CIU50-7ds
30%-100%-0-CIU100-7ds
30%-100%-0-CIU150-7ds
600 30%-100%-0-CIU300-7ds
30%-100%-0-CIU350-7ds
deviator stress q(kPa)

450

300

150

0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 0.18 0.2 0.22
axial strain εs

(b)

Fig. 4.25: Undrained triaxial shearing behavior of specimens with mix proportion
10-3-13 cured under atmospheric pressure for 7 days (a) stress path (b) deviator
stress-strain curve

132
CHAPTER 4

1600 1600
10-3-13 CID TEST
30%-100%-0-CID50-7ds
1400 1400 30%-100%-0-CID150-7ds
30%-100%-0-CID200-7ds
30%-100%-0-CID250-7ds
1200 1200

deviator stress q(kPa)


deviator stress q(kPa)

1000 1000

800 800

600 600
10-3-13 CID TEST
30%-100%-0-CID50-7ds
400 30%-100%-0-CID150-7ds 400
30%-100%-0-CID200-7ds
30%-100%-0-CID250-7ds
200 tension cut-off line 200
primary yield point
primary yield locus
0 0
0 200 400 600 800 0 0.05 0.1 0.15 0.2 0.25 0.3
mean normal effective stressP'(kPa) deviator strain εs

0 100 200 300 400 500 600 700 800


3.4

10-3-13 CID TEST


30%-100%-0-CID50-7ds
30%-100%-0-CID150-7ds
30%-100%-0-CID200-7ds
30%-100%-0-CID250-7ds
3.3 primary yield point
specific volume v

3.2

3.1

Fig. 4.26: Drained triaxial shearing behavior of specimens with mix proportion 10-3-13
cured under atmospheric pressure for 7 days

133
CHAPTER 4

1100 2-1-3 CIU TEST


50%-100%-0-CIU50-7ds
50%-100%-0-CIU250-7ds
1000
50%-100%-0-CIU300-7ds
50%-100%-0-CIU450-7ds
900 tension cut-off line
primary yield point
800 primary yield locus

700
deviator stress q(kPa)

600

500

400

300

200

100

0
0 200 400 600
mean normal effective stressP'(kPa)

(a)

1100
2-1-3 CIU TEST
50%-100%-0-CIU50-7ds
1000 50%-100%-0-CIU250-7ds
50%-100%-0-CIU300-7ds
900 50%-100%-0-CIU450-7ds

800

700
deviator stress q(kPa)

600

500

400

300

200

100

0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 0.18 0.2 0.22 0.24
axial strain εs

(b)

Fig. 4.27: Undrained triaxial shearing behavior of specimens with mix proportion
2-1-3 cured under atmospheric pressure for 7 days (a) stress path (b) deviator
stress-strain curve

134
CHAPTER 4

2600 2600 2-1-3 CID TEST


50%-100%-0-CID50-7ds
2400 2400
50%-100%-0-CID150-7ds
2200 2200
50%-100%-0-CID250-7ds
50%-100%-0-CID450-7ds
2000 2000

1800 1800

deviator stress q(kPa)


deviator stress q(kPa)

1600 1600

1400 1400

1200 1200

1000 1000

800 2-1-3 CID TEST 800


50%-100%-0-CID50-7ds
600 50%-100%-0-CID150-7ds 600
50%-100%-0-CID250-7ds
400 50%-100%-0-CID450-7ds 400
tension cut-off line
200 primary yield point 200
primary yield locus
0 0
0 100 200 300 400 500 600 700 800 900 1000 1100 1200 1300 0 0.05 0.1 0.15 0.2 0.25 0.3
mean normal effective stressP'(kPa) deviator strain εs
0 100 200 300 400 500 600 700 800 900 1000 1100 1200 1300
3.4
2-1-3 CID TEST
50%-100%-0-CID50-7ds
50%-100%-0-CID150-7ds
3.3 50%-100%-0-CID250-7ds
50%-100%-0-CID450-7ds
primary yield point

3.2
specific volume v

3.1

2.9

2.8

Figure 4.28: Drained triaxial shearing behavior of specimens with mix proportion 2-1-3
cured under atmospheric pressure for 7 days

135
CHAPTER 4

2-1-4 CIU TEST


600
50%-133%-0-CIU50-7ds
50%-133%-0-CIU100-7ds
50%-133%-0-CIU150-7ds
50%-133%-0-CIU200-7ds
500 50%-133%-0-CIU250-7ds
tension cut-off line
primary yield point
primary yield locus

deviator stress q(kPa) 400

300

200

100

0
0 50 100 150 200 250 300
mean normal effective stressP'(kPa)

(a)
600
2-1-4 CIU TEST
50%-133%-0-CIU50-7ds
50%-133%-0-CIU100-7ds
50%-133%-0-CIU150-7ds
50%-133%-0-CIU200-7ds
50%-133%-0-CIU250-7ds
450
deviator stress q(kPa)

300

150

0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 0.18 0.2 0.22
axial strain εs

(b)

Figure 4.29: Undrained triaxial shearing behavior of specimens with mix proportion
2-1-4 cured under atmospheric pressure for 7 days (a) stress path (b) deviator
stress-strain curve

136
CHAPTER 4

2-1-4 CID TEST


1600 1600
50%-133%-0-CID50-7ds
50%-133%-0-CID100-7ds
50%-133%-0-CID200-7ds
1400 1400 50%-133%-0-CID250-7ds

1200 1200

deviator stress q(kPa)


deviator stress q(kPa)

1000 1000

800 800

600 600

2-1-4 CID TEST


400 50%-133%-0-CID50-7ds 400
50%-133%-0-CID100-7ds
50%-133%-0-CID200-7ds
50%-133%-0-CID250-7ds
200 tension cut-off line 200
primary yield point
primary yield locus

0 0
0 200 400 600 800 0 0.05 0.1 0.15 0.2 0.25 0.3
mean normal effective stressP'(kPa) deviator strain εs

0 100 200 300 400 500 600 700 800


3.9

2-1-4 CID TEST


50%-133%-0-CID50-7ds
3.8
50%-133%-0-CID100-7ds
50%-133%-0-CID200-7ds
50%-133%-0-CID250-7ds
3.7 primary yield point

3.6
specific volume v

3.5

3.4

3.3

3.2

3.1

Figure 4.30: Drained triaxial behaviors of specimens with mix proportion 2-1-4 cured
under atmospheric pressure for 7 days

137
CHAPTER 4

300
2-1-5 CIU TEST
50%-167%-0-CIU25-7ds
50%-167%-0-CIU50-7ds
50%-167%-0-CIU75-7ds
tension cut-off line
primary yield point
primary yield locus

200
deviator stress q(kPa)

100

0
0 50 100 150
mean normal effective stressP'(kPa)

(a)
300
2-1-5 CIU TEST
50%-167%-0-CIU25-7ds
50%-167%-0-CIU50-7ds
50%-167%-0-CIU75-7ds
deviator stress q(kPa)

150

0
0 0.05 0.1 0.15 0.2 0.25 0.3
axial strain εs

(b)

Figure 4.31: Undrained triaxial shearing behavior of specimens with mix proportion
2-1-5 cured under atmospheric pressure for 7 days (a) stress path (b) deviator
stress-strain curve

138
CHAPTER 4

600 600
2-1-5 CID TEST
2-1-5 CID TEST 50%-167%-0-CID25-7ds
50%-167%-0-CID25-7ds 50%-167%-0-CID50-7ds
50%-167%-0-CID50-7ds 50%-167%-0-CID75-7ds
50%-167%-0-CID75-7ds
tension cut-off line
primary yield point
primary yield locus
400 400

deviator stress q(kPa)


deviator stress q(kPa)

200 200

0 0
0 50 100 150 200 250 0 0.05 0.1 0.15 0.2 0.25 0.3
mean normal effective stressP'(kPa) deviator strain εs
0 50 100 150 200 250
4.6

2-1-5 CID TEST


50%-167%-0-CID25-7ds
50%-167%-0-CID50-7ds
50%-167%-0-CID75-7ds
primary yield point
4.5
specific volume v

4.4

4.3

4.2

Figure 4.32: Drained triaxial shearing behavior of specimens with mix proportion
2-1-5 cured under atmospheric pressure for 7 days

139
CHAPTER 4

180
2-1-5.5 CIU TEST
50%-183%-0-CIU25-7ds
50%-183%-0-CIU50-7ds
50%-183%-0-CIU75-7ds
150
tension cut-off line
primary yield point
primary yield locus

120
deviator stress q(kPa)

90

60

30

0
0 50 100
mean normal effective stressP'(kPa)

(a)

180

150

120
deviator stress q(kPa)

90

2-1-5.5 CIU TEST


50%-183%-0-CIU25-7ds
60 50%-183%-0-CIU50-7ds
50%-183%-0-CIU75-7ds

30

0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 0.18 0.2 0.22 0.24 0.26 0.28 0.3
axial strain εs

(b)

Figure 4.33: Undrained triaxial shearing behavior of specimens with mix proportion
2-1-5.5 cured under atmospheric pressure for 7 days (a) stress path (b) deviator
stress-strain curve

140
CHAPTER 4

400 400
2-1-5.5 CID TEST
50%-183%-0-CID25-7ds
50%-183%-0-CID50-7ds

300

deviator stress q(kPa)


deviator stress q(kPa)

200 200

100 2-1-5.5 CID TEST


50%-183%-0-CID25-7ds
50%-183%-0-CID50-7ds
tension cut-off line
primary yield point
primary yield locus
0 0
0 40 80 120 160 0 0.05 0.1 0.15 0.2 0.25
mean normal effective stressP'(kPa) deviator strain εs
0 40 80 120 160
4.9

2-1-5.5 CID TEST


50%-183%-0-CID25-7ds
50%-183%-0-CID50-7ds
primary yield point

4.8
specific volume v

4.7

4.6

4.5

Fig 4.34: Drained triaxial shearing behavior of specimens with mix proportion
2-1-5.5 cured under atmospheric pressure for 7 days

141
CHAPTER 4

1200 2-1-4-50 CIU TEST


50%-133%-50-CIU50-7ds
1100 50%-133%-50-CIU100-7ds
50%-133%-50-CIU250-7ds
1000 50%-133%-50-CIU400-7ds
tension cut-off line
primary yield point
900
primary yield locus

800
deviator stress q(kPa)
700

600

500

400

300

200

100

0
0 50 100 150 200 250 300 350 400 450 500 550 600
mean normal effective stressP'(kPa)

(a)

1200 2-1-4-50 CIU TEST


50%-133%-50-CIU50-7ds
50%-133%-50-CIU100-7ds
1050 50%-133%-50-CIU250-7ds
50%-133%-50-CIU400-7ds

900
deviator stress q(kPa)

750

600

450

300

150

0
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35
axial strain εs

(b)

Figure 4.35: Undrained triaxial shearing behavior of specimens with mix proportion
2-1-4 cured under 50kPa effective confining pressure for 7 days (a) stress path (b)
deviator stress-strain curve

142
CHAPTER 4

2400 2200 2-1-4-50 CID TEST


2-1-4-50 CID TEST 50%-133%-50-CID50-7ds
50%-133%-50-CID50-7ds
50%-133%-50-CID250-7ds
2200 50%-133%-50-CID250-7ds 2000
50%-133%-50-CID450-7ds
50%-133%-50-CID450-7ds
2000 tension cut-off line 1800
primary yield point
1800 primary yield locus
1600

1600

deviator stress q(kPa)


1400
deviator stress q(kPa)

1400
1200
1200
1000
1000
800
800

600
600

400
400

200 200

0 0
0 100 200 300 400 500 600 700 800 900 1000 1100 1200 0 0.05 0.1 0.15 0.2 0.25 0.3 0.35
mean normal effective stressP'(kPa) deviator strain εs

0 100 200 300 400 500 600 700 800 900 1000 1100 1200
3.4

2-1-4-50 CID TEST


3.3
50%-133%-50-CID50-7ds
50%-133%-50-CID250-7ds
50%-133%-50-CID450-7ds
primary yield point
3.2
specific volume v

3.1

2.9

2.8

2.7

Fig 4.36: Drained triaxial shearing behavior of specimens with mix proportion 2-1-4 cured
under 50KP effective confining stress for 7 days

143
CHAPTER 4

2-1-4-100 CIU TEST


50%-133%-100-CIU100-7ds
1500
50%-133%-100-CIU250-7ds
1400 50%-133%-100-CIU400-7ds
tension cut-off line
1300
primary yield point
1200 primary yield locus

1100

1000
deviator stress q(kPa)

900

800

700

600

500

400

300

200

100

0
0 50 100 150 200 250 300 350 400 450 500 550 600 650 700 750 800
mean normal effective stressP'(kPa)

(a)
1500
2-1-4-100 CIU TEST
50%-133%-100-CIU100-7ds
1350 50%-133%-100-CIU250-7ds
50%-133%-100-CIU400-7ds
1200

1050
deviator stress q(kPa)

900

750

600

450

300

150

0
0 0.05 0.1 0.15 0.2 0.25
axial strain εs
(b)
Figure 4.37: Undrained triaxial shearing behavior of specimens with mix proportion
2-1-4 cured under 100kPa effective confining pressure for 7 days (a) stress path (b)
deviator stress-strain curve

144
CHAPTER 4

2600 2600
2-1-4-100 CID TEST
2400 2400 50%-133%-100-CID100-7ds
50%-133%-100-CID200-7ds
2200 2200 50%-133%-100-CID250-7ds
50%-133%-100-CID400-7ds
2000 2000

1800 1800

deviator stress q(kPa)


deviator stress q(kPa)

1600 1600

1400 1400

1200 1200

1000 1000

800 2-1-4-100 CID TEST 800


50%-133%-100-CID100-7ds
600 50%-133%-100-CID200-7ds 600
50%-133%-100-CID250-7ds
400 50%-133%-100-CID400-7ds 400
tension cut-off line
primary yield point
200 200
primary yield locus

0 0
0 100 200 300 400 500 600 700 800 900 1000 1100 1200 1300 0 0.05 0.1 0.15 0.2 0.25 0.3
mean normal effective stressP'(kPa) deviator strain εs

100 200 300 400 500 600 700 800 900 1000 1100 1200 1300

3.2 2-1-4-100 CID TEST


50%-133%-100-CID100-7ds
50%-133%-100-CID200-7ds
50%-133%-100-CID250-7ds
3.1
50%-133%-100-CID400-7ds
primary yield point
specific volume v

2.9

2.8

2.7

Figure 4.38: Drained triaxial shearing behavior of specimens with mix proportion
2-1-4 cured under 100KP effective confining stress for 7 days

145
CHAPTER 4

2600
2-1-4-250 CIU TEST
2400 50%-133%-250-CIU100
50%-133%-250-CIU250
2200 tension cut-off line
primary yield point
2000 primary yield locus

1800
deviator stress q(kPa)
1600

1400

1200

1000

800

600

400

200

0
0 100 200 300 400 500 600 700 800 900 1000 1100 1200 1300 1400
mean normal effective stressP'(kPa)

(a)

2400
2-1-4-250 CIU TEST
2200 50%-133%-250-CIU100
50%-133%-250-CIU250
2000

1800

1600
deviator stress q(kPa)

1400

1200

1000

800

600

400

200

0
0 0.05 0.1 0.15 0.2 0.25
axial strain εs

(b)
Figure 4.39: Undrained triaxial shearing behaviors of specimens with mix proportion
2-1-4 cured under 250kPa effective confining pressure for 7 days (a) stress path (b)
deviator stress-strain curve

146
CHAPTER 4

3800 3800
3600 3600 2-1-4-250 CID TEST
50%-133%-250-CID100-7ds
3400 3400
50%-133%-250-CID250-7ds
3200 3200 50%-133%-250-CID500-7ds
3000 3000 50%-133%-250-CID650-7ds
2800 2800 50%-133%-250-CID900-7ds
2600 2600

deviator stress q(kPa)


deviator stress q(kPa)

2400 2400
2200 2200
2000 2000
1800 1800
1600 1600
1400 2-1-4-250 CID TEST 1400
1200 50%-133%-250-CID100-7ds 1200
50%-133%-250-CID250-7ds
1000 1000
50%-133%-250-CID500-7ds
800 50%-133%-250-CID650-7ds 800
600 50%-133%-250-CID900-7ds 600
tension cut-off line
400 400
primary yield point
200 primary yield locus 200
0 0
0 200 400 600 800 1000 1200 1400 1600 1800 2000 2200 0 0.05 0.1 0.15 0.2 0.25
mean normal effective stressP'(kPa) deviator strain εs
0 200 400 600 800 1000 1200 1400 1600 1800 2000 2200

2-1-4-250 CID TEST


2.9 50%-133%-250-CID100-7ds
50%-133%-250-CID250-7ds
50%-133%-250-CID500-7ds
50%-133%-250-CID650-7ds
2.8 50%-133%-250-CID900-7ds
primary yield point
specific volume v

2.7

2.6

2.5

2.4

Figure 4.40: Drained triaxial shearing behavior of specimens with mix proportion
2-1-4 cured under 250KP effective pressure for 7 days

147
CHAPTER 4

cement content effect-CIU TEST


Aw=10%-ciu50
Aw=10%-ciu75
Aw=10%-ciu100
Aw=10%-ciu120
Aw=15%-ciu50
Aw=15%-ciu100
Aw=15%-ciu150
1200 Aw=15%-ciu200
Aw=20%-ciu50
Aw=20%-ciu100
Aw=20%-ciu200
Aw=20%-ciu50
Aw=30%-ciu100
1050 Aw=30%-ciu150
Aw=30%-ciu300
Aw=30%-ciu350
Aw=50%-CIU50
Aw=50%-CIU250
Aw=50%-CIU300
Aw=50%-CIU450
900 tension cut-off line
primary yield locus
primary yield point-10%
primary yield point-15%
primary yield point-20%
primary yield point-30%
deviator stress q(kPa)

750 primary yield point-50%

600

450

300

150

0
0 200 400 600
mean normal effective stressP'(kPa)
(a)
cement content effect-CIU TEST
Aw=10%-ciu50
Aw=10%-ciu75
Aw=10%-ciu100

1050 Aw=10%-ciu120
Aw=15%-ciu50
Aw=15%-ciu100
Aw=15%-ciu150
Aw=15%-ciu200
Aw=20%-ciu50
Aw=20%-ciu100
Aw=20%-ciu200
900 Aw=30%-ciu50
Aw=30%-ciu100
Aw=30%-ciu150
Aw=30%-ciu300
Aw=30%-ciu350
Aw=50%-ciu50
Aw=30%-ciu250
750 Aw=50%-ciu300
Aw=50%-ciu450
deviator stress q(kPa)

600

450

300

150

0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 0.18 0.2 0.22 0.24
axial strain εs
(b)
Figure 4.41: Cement content effect on undrained stress-strain behavior of specimens
with 100% water content cured under atmospheric pressure for 7 days of curing period
(a) stress path (b) deviator stress-strain curve

148
CHAPTER 4

2800 2600
cement content effect-CID TEST
Aw=10%-CID25
2600 2400 Aw=10%-CID50
Aw=10%-CID75
2400 Aw=10%-CID100
2200 Aw=15%-CID50
Aw=15%-CID100
2200 Aw=15%-CID150
2000 Aw=20%-CID50
Aw=20%-CID100
2000 Aw=20%-CID200
1800 Aw=30%-CID50
Aw=30%-CID150

deviator stress q(kPa)


1800
deviator stress q(kPa)

Aw=30%-CID200
1600 Aw=30%-CID250
cement content effect -CID TEST
1600 Aw=10%-CID25 Aw=50%-CID50
Aw=10%-CID50 Aw=50%-CID150
Aw=10%-CID75 1400 Aw=50%-CID250
Aw=10%-CID100
1400 Aw=15%-CID50
Aw=50%-CID450
Aw=15%-CID100 1200
Aw=15%-CID150
1200 Aw=20%-CID50
Aw=20%-CID100
Aw=20%-CID200 1000
1000 Aw=30%-CID50
Aw=30%-CID150
Aw=30%-CID200 800
800 Aw=30%-CID250
Aw=50%-CID50
Aw=50%-CID150
600
600 Aw=50%-CID250
Aw=50%-CID450
tension cut-off line
400 primary yield locus 400
primary yield point-10%
primary yield point-15%
200 primary yield point-20% 200
primary yield point-30%
primary yield point-50%

0 0
0 100 200 300 400 500 600 700 800 900 1000 1100 1200 1300 0 0.05 0.1 0.15 0.2 0.25 0.3
mean normal effective stressP'(kPa) deviator strain εs

0 100 200 300 400 500 600 700 800 900 1000 1100 1200 1300
3.6
cement content effect-CID TEST
Aw=10%-CID25
Aw=10%-CID50
Aw=10%-CID75
3.5 Aw=10%-CID100
Aw=15%-CID50
Aw=15%-CID100
Aw=15%-CID150
Aw=20%-CID50
3.4 Aw=20%-CID100
Aw=20%-CID200
Aw=30%-CID50
Aw=30%-CID150
Aw=30%-CID200
3.3 Aw=30%-CID250
specific volume v

Aw=50%-CID50
Aw=50%-CID150
Aw=50%-CID250
Aw=50%-CID450
3.2 primary yield point-10%
primary yield point-15%
primary yield point-20%
primary yield point-30%
primary yield point-50%

3.1

2.9

2.8

Figure 4.42: Cement content effect on drained stress-strain behavior of specimens


cured under atmospheric pressure for 7 days of curing period

149
CHAPTER 4

water content effect-CIU TEST


100%-CIU50
100%-CIU250
100%-CIU300
100%-CIU450
133.3%-CIU50
1200 133.3%-CIU100
133.3%-CIU150
133.3%-CIU200
133.3%-CIU250
166.7%-CIU25
166.7%-CIU50
1050 166.7%-CIU75
183.3%-CIU25
183.3%-CIU50
183.3%-CIU75
tension cut-off line
primary yield locus
900 primary yield point-100%
primary yield point-133.3%
primary yield point-166.7%
primary yield point-183.3%
deviator stress q(kPa)

750

600

450

300

150

0
0 200 400 600
mean normal effective stressP'(kPa)
(a)

water content effect-CIU TEST


1050
100%-CIU50
100%-CIU250
100%-CIU300
100%-CIU450
900 133.3%-CIU50
133.3%-CIU100
133.3%-CIU150
133.3%-CIU200
133.3%-CIU250
750 166.7%-CIU25
166.7%-CIU50
166.7%-CIU75
deviator stress q(kPa)

183.3%-CIU25
183.3%-CIU50
600 183.3%-CIU75

450

300

150

0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 0.18 0.2 0.22 0.24 0.26 0.28 0.3
axial strain εs

(b)
Figure 4.43: Total water content effect on undrained stress-strain behavior of
specimens with 50% cement content cured under atmospheric pressure for 7 days of
curing period (a) stress path (b) deviator stress-strain curve

150
CHAPTER 4

2800 2600 water content effect-CID TEST


100%-CID50-7ds
2600 100%-CID150-7ds
2400
100%-CID250-7ds
100%-CID450-7ds
2400 2200 133.3%-CID50-7ds
133.3%-CID100-7ds
2200 133.3%-CID200-7ds
2000
133.3%-CID250-7ds
2000 166.7%-CID25-7ds
1800 166.7%-CID50-7ds
166.7%-CID75-7ds

deviator stress q(kPa)


1800
deviator stress q(kPa)

1600 183.3%-CID25-7ds
183.3%-CID50-7ds
1600 water content effect-CID TEST
100%-CID50 1400
100%-CID150
1400
100%-CID250
100%-CID450 1200
1200 133.3%-CID50
133.3%-CID100
133.3%-CID200
1000
1000 133.3%-CID250
166.7%-CID25
800
800 166.7%-CID50
166.7%-CID75
183.3%-CID25 600
600 183.3%-CID50
tension cut-off line
400 primary yield locus 400
primary yield point-100%
primary yield point-133.3%
200 primary yield point-166.7% 200
primary yield point-183.3%

0 0
0 100 200 300 400 500 600 700 800 900 1000 1100 1200 1300 0 0.05 0.1 0.15 0.2 0.25 0.3
mean normal effective stressP'(kPa) deviator strain εs

0 100 200 300 400 500 600 700 800 900 1000 1100 1200 1300
4.9
4.8 water content effect-CID TEST
100%-CID50-7ds
4.7
100%-CID150-7ds
4.6 100%-CID250-7ds
4.5 100%-CID450-7ds
133.3%-CID50-7ds
4.4
133.3%-CID100-7ds
4.3 133.3%-CID200-7ds
4.2 133.3%-CID250-7ds
166.7%-CID25-7ds
4.1
specific volume v

166.7%-CID50-7ds
4 166.7%-CID75-7ds
3.9 183.3%-CID25-7ds
183.3%-CID50-7ds
3.8
primary yield point-100%
3.7 primary yield point-133.3%
3.6 primary yield point-166.7%
primary yield point-183.3%
3.5
3.4
3.3
3.2
3.1
3
2.9
2.8

Fig 4.44: Total water content effect on drained stress-strain behavior of specimens with
50% cement content cured under atmospheric pressure for 7 days of curing period

151
CHAPTER 4

curing stress effect-CIU TEST


0-CIU50
0-CIU100
0-CIU150
0-CIU200
0-CIU250
50-CIU50
50-CIU100
2600 50-CIU250
50-CIU400
100-CIU100
100-CIU250
2400
100-CIU400
250-CIU100
250-CIU250
2200 tension cut-off line
primary yield locus
primary yield point-0kPa
2000 primary yield point-50kPa
primary yield point-100kPa
primary yield point-250kPa
1800
deviator stress q(kPa)

1600

1400

1200

1000

800

600

400

200

0
0 100 200 300 400 500 600 700 800 900 1000 1100 1200 1300 1400
mean normal effective stressP'(kPa)
(a)
2400 curing stress effect-CIU TEST
0-CIU50
0-CIU100
2200 0-CIU150
0-CIU200
2000 0-CIU250
50-CIU50
50-CIU100
1800 50-CIU250
50-CIU400
100-CIU100
1600
100-CIU250
deviator stress q(kPa)

100-CIU400
1400 250-CIU100
250-CIU250

1200

1000

800

600

400

200

0
0 0.05 0.1 0.15 0.2 0.25 0.3
axial strain εs
(b)
Figure 4.45: Curing stress effect on stress-strain behavior of specimens with mix
proportion 2-1-4 cured for 7 days (a) stress path (b) deviator stress-strain curve

152
CHAPTER 4

3800 3800 curing stress effect-CID TEST


0-CID50
3600 3600 0-CID100
0-CID150
3400 3400 0-CID200
0-CID250
3200 3200 50-CID50
50-CID250
3000 3000 50-CID450
100-CID50
2800 2800 100-CID200
100-CID250
2600 2600 100-CID400
250-CID100

deviator stress q(kPa)


deviator stress q(kPa)

2400 curing stress effect-CID TEST 2400 250-CID250


0-CID50 250-CID500
2200 0-CID100 2200 250-CID650
0-CID150 250-CID900
2000 0-CID200 2000
0-CID250
1800 50-CID50 1800
50-CID250
1600 50-CID450 1600
100-CID100
1400 100-CID200 1400
100-CID250
1200 100-CID400 1200
250-CID100
1000 250-CID250 1000
250-CID500
800 250-CID650 800
250-CID900
600 tension cut-off line 600
primary yield locus
400 primary yield point-0kPa 400
primary yield point-50kPa
200 primary yield point-100kPa 200
primary yield point-250kPa
0 0
0 200 400 600 800 1000 1200 1400 1600 1800 2000 2200 0 0.05 0.1 0.15 0.2 0.25 0.3
mean normal effective stressP'(kPa) deviator strain εs

0 200 400 600 800 1000 1200 1400 1600 1800 2000 2200
3.9
curing stres effect-CID TEST
3.8 0-CID50
0-CID100
3.7 0-CID150
0-CID200
0-CID250
3.6 50-CID50
50-CID250
3.5 50-CID450
100-CID100
100-CID200
3.4 100-CID250
100-CID400
specific volume v

3.3 250-CID100
250-CID250
3.2 250-CID500
250-CID650
250-CID900
3.1 primary yield point-0kPa
primary yield point-50kPa
3 primary yield point-100kPa
primary yield point-250kPa

2.9

2.8

2.7

2.6

2.5

2.4

Figure 4.46: Curing stress effect on drained stress-strain behavior of specimens with
mix proportion 2-1-4(cement content 50%, water content 133.3%) cured for 7 days

153
CHAPTER 4

2-1-4 CIU TEST


CIU50-7ds
CIU100-7ds
CIU200-7ds
CIU50-28ds
CIU100-28ds
CIU200-28ds
1100 CIU50-90ds
CIU250-90ds
CIU450-90ds
1000 CIU50-180ds
CIU250-180ds
CIU450-180ds
900 tension cut-off line
primary yield point-7ds
primary yield point-28ds
primary yield point-90ds
800
primary yield point-180ds
primary yield locus

700
deviator stress q(kPa)

600

500

400

300

200

100

0
0 50 100 150 200 250 300 350 400 450 500 550
mean normal effective stressP'(kPa)

(a)
2-1-4 CIU TEST
CIU50-7ds
1000 CIU100-7ds
CIU200-7ds
CIU50-28ds
900
CIU100-28ds
CIU200-28ds
800 CIU50-90ds
CIU250-90ds
CIU450-90ds
700 CIU50-180ds
CIU250-180ds
deviator stress q(kPa)

CIU450-180ds
600

500

400

300

200

100

0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 0.18 0.2 0.22
axial strain εs

(b)
Figure 4.47: Curing time effect on undrained triaxial behavior of specimens with mix
proportion 2-1-4 cured under atmospheric pressure (a) stress path (b) deviator
stress-strain curve

154
CHAPTER 4

2200 2200
2-1-4 CID TEST
CID50-7ds
2000 2000 CID100-7ds
CID200-7ds
1800 1800 CID250-7ds
CID50-28ds
CID100-28ds
1600 1600 CID200-28ds
2-1-4 CID TEST CID50-90ds

deviator stress q(kPa)


CID50-7ds CID100-90ds
1400 1400
deviator stress q(kPa)

CID100-7ds CID200-90ds
CID200-7ds CID250-90ds
1200 CID250-7ds 1200 CID450-90ds
CID50-28ds
CID100-28ds
CID50-180ds
CID200-28ds CID250-180ds
1000 1000
CID50-90ds CID450-180ds
CID100-90ds
800 CID200-90ds 800
CID250-90ds
CID450-90ds
600 CID50-180ds 600
CID250-180ds
CID450-180ds
400 tension cut-off line 400
primary yield points-7ds
primary yield points-28ds
200 primary yield points-90ds 200
primary yield points-180ds
primary yield locus
0 0
0 100 200 300 400 500 600 700 800 900 1000 1100 1200 0 0.05 0.1 0.15 0.2 0.25 0.3 0.35
mean normal effective stressP'(kPa) deviator strain εs

0 100 200 300 400 500 600 700 800 900 1000 1100 1200
3.9
2-1-4 CID TEST
CID50-7ds
3.8 CID100-7ds
CID200-7ds
CID250-7ds
3.7 CID50-28ds
CID100-28ds
CID200-28ds
CID50-90ds
3.6
CID100-90ds
specific volume v

CID200-90ds
CID250-90ds
3.5 CID450-90ds
CID50-180ds
CID250-180ds
3.4 CID450-180ds
primary yield point-7ds
primary yield point-28ds
primary yield point-90ds
3.3
primary yield point-180ds

3.2

3.1

Figure 4.48: Curing time effect on drained triaxial behavior of specimens with mix
proportion 2-1-4(50% cement content, 133.3% water content) cured under
atmospheric pressure

155
CHAPTER 4

5-1-6 CIU TEST


CIU50-7ds
CIU100-7ds
CIU200-7ds
CIU50-28ds
CIU100-28ds
1100 CIU200-28ds
CIU250-28ds
CIU50-90ds
1000 CIU250-90ds
CIU50-180ds
CIU200-180ds
900 CIU400-180ds
tension cutoff line
primary yield points-7ds
primary yield points-28ds
800 primary yield points-90ds
primary yield points-180ds
primary yield locus
700
deviator stress q(kPa)

600

500

400

300

200

100

0
0 50 100 150 200 250 300 350 400 450 500 550
mean normal effective stressP'(kPa)

(a)
5-1-6 CIU TEST
1000
CIU50-7ds
CIU100-7ds
900 CIU200-7ds
CIU50-28ds
CIU100-28ds
800 CIU200-28ds
CIU250-28ds
CIU50-90ds
700 CIU250-90ds
CIU50-180ds
deviator stress q(kPa)

CIU200-180ds
600
CIU400-180ds

500

400

300

200

100

0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 0.18 0.2 0.22
axial strain εs

(b)
Figure 4.49: Curing time effect on undrained triaxial behavior of specimens with mix
proportion 51-6 cured under atmospheric pressure (a) stress path (b) deviator
stress-strain curve

156
CHAPTER 4

2000 2000 5-1-6 CID TEST


CID50-7ds
CID100-7ds
1800 1800 CID200-7ds
CID50-28ds
CID100-28ds
1600 1600
CID200-28ds
CID250-28ds
1400 1400 CID50-90ds
CID250-90ds

deviator stress q(kPa)


5-1-6 CID TEST
CID400-90ds
deviator stress q(kPa)

CID50-7ds
1200 1200 CID50-180ds
CID100-7ds
CID200-7ds
CID200-180ds
CID50-28ds CID400-180ds
1000 1000
CID100-28ds
CID200-28ds
CID250-28ds
800 800
CID50-90ds
CID250-90ds
CID400-90ds
600 600
CID50-180ds
CID200-180ds
CID400-180ds
400 400
tension cut-off line
primary yield points-7ds
primary yield points-28ds
200 primary yield points-90ds 200
primary yield points-180ds
primary yield locus
0 0
0 100 200 300 400 500 600 700 800 900 1000 1100 0 0.05 0.1 0.15 0.2 0.25 0.3
mean normal effective stressP'(kPa) deviator strain εs

0 100 200 300 400 500 600 700 800 900 1000 1100
3.5
5-1-6 CID TEST
CID50-7ds
CID100-7ds
3.4 CID200-7ds
CID50-28ds
CID100-28ds
CID200-28ds
3.3 CID250-28ds
CID50-90ds
CID250-90ds
specific volume v

CID400-90ds
3.2
CID50-180ds
CID200-180ds
CID400-180ds
primary yield points-7ds
3.1
primary yield points-28ds
primary yield points-90ds
primary yield points-180ds

2.9

2.8

Figure 4.50: Curing time effect on drained triaxial behavior of specimens with mix
proportion 5-1-6(20% cement content, 100% water content) cured under
atmospheric pressure

157
CHAPTER 4

mix proportion-curing stress-curing time


2
10-1-11-0-7ds
q ⎛ p' ⎞ ⎛ p' ⎞
20-3-23-0-7ds
= 3.65 ⎜⎜ ' ⎟⎟ − ⎜⎜ ' ⎟⎟
5-1-6-0-7ds p 'py ⎝ p py ⎠ ⎝ p py ⎠
1.6 10-3-13-0-7ds
2-1-3-0-7ds
2-1-4-0-7ds
2-1-5-0-7ds
2-1-5.5-0-7ds
1.2 2-1-4-50-7ds
2-1-4-100-7ds
q/ppy'

2-1-4-250-7ds
normalized primary yield locus
tension cutoff line
0.8

0.4

0 0.2 0.4 0.6 0.8 1


p'/ppy'

(a)
2
mix proportion-curing stress-curing time
10-1-11-0-7ds
20-3-23-0-7ds q p'
5-1-6-0-7ds
'
= 3.65 ' [−1.5ln( p ' / p 'py )]1/1.5
1.6 10-3-13-0-7ds p py p py
2-1-3-0-7ds
2-1-4-0-7ds
2-1-5-0-7ds
2-1-5.5-0-7ds
2-1-4-50-7ds
1.2
2-1-4-100-7ds
2-1-4-250-7ds
q/ppy'

normalized primary yield locus


tension cutoff line

0.8

0.4

0 0.2 0.4 0.6 0.8 1


p'/ppy'

(b)

Figure 4.51: Primary yield loci normalized by isotropic primary yield stress for
specimens with different mix proportion and curing stress and cured for 7 days(a)
fitted by function (4-1) (b) fitted by function (4-2)

158
CHAPTER 4

2
q ⎛ p' ⎞ ⎛ p' ⎞
= 3.65 ⎜⎜ ' ⎟⎟ − ⎜⎜ ' ⎟⎟
1.6 p 'py ⎝ p py ⎠ ⎝ p py ⎠
mix proportion-curing stress-curing time
10-1-11-0-7ds
20-3-23-0-7ds
5-1-6-0-7ds
1.2 10-3-13-0-7ds
2-1-3-0-7ds
2-1-4-0-7ds
q/ppy'

2-1-5-0-7ds
2-1-5.5-0-7ds
2-1-4-50-7ds
0.8 2-1-4-100-7ds
2-1-4-250-7ds
2-1-4-0-28ds
2-1-4-0-90ds
2-1-4-0-180ds
5-1-6-0-28ds
0.4
5-1-6-0-90ds
5-1-6-0-180ds
normalized primary yield locus
tension cutoff line

0 0.2 0.4 0.6 0.8 1


p'/ppy'

(a)

q p'
1.6 '
= 3.65 ' [−1.5ln( p ' / p 'py )]1/1.5
p py p py
mix proportion-curing stress-curing time
10-1-11-0-7ds
20-3-23-0-7ds
5-1-6-0-7ds
1.2
10-3-13-0-7ds
2-1-3-0-7ds
q/ppy'

2-1-4-0-7ds
2-1-5-0-7ds
2-1-5.5-0-7ds
0.8 2-1-4-50-7ds
2-1-4-100-7ds
2-1-4-250-7ds
2-1-4-0-28ds
2-1-4-0-90ds
2-1-4-0-180ds
0.4 5-1-6-0-28ds
5-1-6-0-90ds
5-1-6-0-180ds
normalized primary yield locus
tensin cutoff line

0 0.2 0.4 0.6 0.8 1


p'/ppy'

(b)
Figure 4.52: Primary yield loci normalized by isotropic primary yield stress with
different curing period (a) fitted by function (4-1) (b) fitted by function (4-2)

159
CHAPTER 4

900
10-1-11 UCS test 900
10-1-11-7ds-0kPa
20-3-23 UCS test
800 10-1-11-7ds-50kPa
20-3-23-7ds-0kPa
10-1-11-7ds-100kPa 800 20-3-23-7ds-50kPa
10-1-11-7ds-250kPa 20-3-23-7ds-100kPa
700 10-1-11-7ds-350kPa 20-3-23-7ds-250kPa
700

600
deviator stress q(kPa)

600

deviator stress q(kPa)


500
500

400
400

300
300

200
200

100
100

0
0
0 0.01 0.02 0.03 0.04 0.05 0.06
0 0.01 0.02 0.03 0.04 0.05 0.06
axial strain εs axial strain εs

(a) mix proportion 10-1-11 (b) mix proportion 20-3-23

1000 2200
5-1-6 UCS test 2-1-4 UCS test
5-1-6-7ds-0kPa 2-1-4-7ds-0kPa
900 5-1-6-7ds-50kPa 2000 2-1-4-7ds-50kPa
5-1-6-7ds-100kPa 2-1-4-7ds-100kPa
5-1-6-7ds-250kPa 1800 2-1-4-7ds-250kPa
800

1600
700

1400
deviator stress q(kPa)

deviator stress q(kPa)

600
1200
500
1000
400
800

300
600

200
400

100 200

0 0
0 0.01 0.02 0.03 0.04 0.05 0.06 0 0.01 0.02 0.03 0.04 0.05 0.06
axial strain εs axial strain εs

(c) mix proportion 5-1-6 (d) mix proportion 2-1-4

Figure 4.53: Unconfined compressive strength tests for specimens with different curing
stress and 7 days of curing time period (a-d)

160
CHAPTER 4

1200 1600
2-1-4.5 UCS test 2-1-5 UCS test
2-1-4.5-7ds-0kPa 2-1-5-7ds-0kPa
2-1-5-7ds-50kPa
2-1-4.5-7ds-50kPa
1400 2-1-5-7ds-100kPa
2-1-4.5-7ds-100kPa 2-1-5-7ds-250kPa
1000

1200

800

deviator stress q(kPa)


deviator stress q(kPa)

1000

600 800

600

400

400

200
200

0
0
0 0.01 0.02 0.03 0.04 0.05
0 0.01 0.02 0.03 0.04
axial strain εs
axial strain εs

(e) mix proportion 2-1-4.5 (f) mix proportion 2-1-5

2800 2400
1.3-1-3.45 UCS test 1.3-1-3.5 UCS test
2600 1.3-1-3.45-7ds-0kPa 1.3-1-3.5-7ds-0kPa
2200
1.3-1-3.45-7ds-50kPa 1.3-1-3.5-7ds-50kPa
2400 1.3-1-3.45-7ds-100kPa 1.3-1-3.5-7ds-100kPa
2000
2200
1800
2000
1600
1800
deviator stress q(kPa)

deviator stress q(kPa)

1600 1400

1400 1200

1200 1000

1000
800
800
600
600
400
400

200 200

0 0
0 0.01 0.02 0.03 0.04 0.05 0 0.01 0.02 0.03 0.04 0.05
axial strain εs axial strain εs

(g) mix proportion 1.3-1-3.45 (h) mix proportion 1.3-1-3.5

Figure 4.54: Unconfined compressive strength tests for specimens with different curing
stress and 7 days of curing time period (e-h)

161
CHAPTER 4

400 800
10-1-11 UCS test 20-3-23 UCS test
10-1-11-7ds 20-3-23-7ds
10-1-11-28ds 20-3-23-28ds
700 20-3-23-90ds
20-3-23-210ds

300 600
deviator stress q(kPa)

deviator stress q(kPa)


500

200 400

300

100 200

100

0 0
0 0.01 0.02 0.03 0.04 0 0.01 0.02 0.03 0.04
axial strain εs axial strain εs

(a) mix proportion 10-1-11 (b) mix proportion 20-3-23

900 1400
5-1-6 UCS test 2-1-4 UCS test
5-1-6-7ds 1300 2-1-4-7ds
800 5-1-6-28ds 2-1-4-28ds
5-1-6-90ds 1200 2-1-4-90ds
5-1-6-180ds
2-1-4-180ds
700 1100 2-1-4-730ds

1000
600
900
deviator stress q(kPa)
deviator stress q(kPa)

800
500

700

400
600

500
300
400

200 300

200
100
100

0 0
0 0.01 0.02 0.03 0.04 0.05 0 0.01 0.02 0.03 0.04 0.05
axial strain εs axial strain εs

(c) mix proportion 5-1-6 (d)mix proportion 2-1-4

Figure 4.55: Unconfined compressive strength tests for specimens with different curing
time and zero curing stress (a-d)

162
CHAPTER 4

600
2-1-4.5 UCS test 700
2-1-4.5-7ds 1.3-1-3.45 UCS test
2-1-4.5-28ds 1.3-1-3.45-7ds
2-1-4.5-100ds 1.3-1-3.45-28ds
500
600

400 500
deviator stress q(kPa)

deviator stress q(kPa)


400
300

300
200

200

100

100

0
0 0.01 0.02 0.03 0.04 0
0 0.01 0.02 0.03 0.04
axial strain εs
axial strain εs

(e) mix proportion 2-1-4.5 (f) mix proportion 1.3-1-3.45

1000
1-1-3 UCS test
1-1-3-7ds
900 1-1-3-28ds
1-1-3-90ds

800

700
deviator stress q(kPa)

600

500

400

300

200

100

0
0 0.01 0.02 0.03 0.04
axial strain εs

(g) mix proportion 1-1-3

Fig 4.56: Unconfined compressive strength tests for specimens with different curing
time and zero curing stress (e-g)

163
CHAPTER 4

1350

1200

1050

900
ppy'-pcur'(kPa)

750 10-1-11
20-3-23
5-1-6
10-3-13
600 2-1-3
2-1-4
2-1-4.5
2-1-5
450 2-1-5.5
10-3-17.3
10-3-19.5
300 1.3-1-3.06
1.3-1-3.45
1.3-1-3.5
1-1-3
150 1-1-2.66
fitted line p0'-pcur'=0.5476qu+19.679

0
0 500 1000 1500 2000 2500
unconfined compressive strength qu(kPa)

(a)
1350

1200

1050
10-1-11
20-3-23
900 5-1-6
10-3-13
2-1-3
2-1-4
ppy'-pcur'(kPa)

750 2-1-4.5
2-1-5
2-1-5.5
10-3-17.3
600 10-3-19.5
1.3-1-3.06
1.3-1-3.45
450 1.3-1-3.5
1-1-3
1-1-2.66
fitted line p0'-pcur'=0.5476qu+19.679
300
2-1-4(7ds,28ds,90ds,180ds)
5-1-6(7ds,28ds,90ds,180ds)
10-1-11(7ds,28ds)
150 20-3-23(7ds,28ds,90ds,210ds)
1-1-3(7ds,28ds,90ds)
1.3-1-3.45(7ds,28ds)

0
0 500 1000 1500 2000 2500
unconfined compressive strength qu(kPa)

(b)
Fig 4.57: Relationship between Isotropic primary yield stress and unconfined
compressive strength for specimens (a) with 7 days of curing and (b) different curing
time period

164
CHAPTER 4

4 incremental yield stress-post-curing voild ratio


10-1-11
20-3-23
5-1-6
2-1-4
2-1-5
3.5 1.3-1-3.5
Fit 1.3-1-3.5:Y = -0.8014* X + 7.9223,R2=0.96
Fit 10-1-11: Y = -0.6260 * X + 5.6232,R2=0.966
Fit 20-3-23: Y = -0.6472* X + 5.9538,R2=0.985
Fit 5-1-6: Y = -0.6790* X + 6.2352,R2=0.999
2
3 Fit 2-1-4: Y = -0.7790* X + 7.3357,R =0.994

Fit 2-1-5: Y = -0.7186* X + 7.1273,R2=0.959


ecur

2.5

1.5

4 5 6 7 8
Ln(ppy'-pcur')

Fig 4.58 Relationship between incremental Isotropic primary yield stress and
post-curing void ratio for cement-treated marine clay specimens cured for 7 days

-0.6 8

A1-Cw/Aw B1-Cw/Aw
A1=-0.566-0.461(Cw/Aw)-0.8919, R=0.987 Fit : Y=9.1355X-0.2196,R2=0.9811

-0.64
7.5

-0.68

7
B1
A1

-0.72

6.5

-0.76

6
-0.8

-0.84 5.5

2 4 6 8 10 2 4 6 8 10
Cw/Aw Cw/Aw

(a) (b)

Figure 4.59: Relationship between parameters and ratio of pre-curing total water
content to the cement content Cw/Aw (a) A1-Cw/Aw, (b) B1-Cw/Aw

165
CHAPTER 4

12 p py'-Aw relationship
Cw=72%
Cw=100%
Cw=133.3%
Cw=150%
10 fit 4: Cw=150%-- p0'=3.4455Aw-0.0543,R2=0.985
fit 2: Cw=100%-- p0'=10.52Aw+0.3737,R2=0.999
fit 3: Cw=133.3%-- p0'=4.8754Aw+0.1129,R2=0.990
fit 1: Cw=72%-- p0'=21.642Aw+0.5923,R2=0.992

8
p'py/Pap

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 1.1 1.2
cement content Aw

(a)
25
A2-Cw
Fit 1: A2=9.8652Cw(-2.5042),R2=0.997 0.6
B2-Cw
fit 1: B2=-0.8201Cw+1.189,R2=0.998

20
0.5

0.4
15
A2(kPa)

0.3
B2(kPa)

10
0.2

5 0.1

0
0.7 0.8 0.9 1 1.1 1.2 1.3 1.4 1.5
0
0.7 0.8 0.9 1 1.1 1.2 1.3 1.4 1.5 total water content Cw
-0.1
total water content Cw

(b) (c)

166
CHAPTER 4

1300 p py'-Aw relationship


Cw=72%
Cw=100%
1200
Cw=133.3%
Cw=150%
1100 1:1 line
+10% line
1000 -10% line

900

experimental p' py (kPa) 800

700

600

500

400

300

200

100

0
100 200 300 400 500 600 700 800 900 1000 1100
predicted Ppy'(kPa)

(d)

Fig 4.60 Relationship between Isotropic primary yield stress and cement content and
total water content for cement treated specimens cured under atmospheric pressure for
7days (a) p 'py versus cement content (b) parameter A2 versus total water content (c)
parameter B2 versus total water content (d) comparison between experimental data and
simulation results by using formula (4-11)
2000 qu-Aw relationship
Cw=72%
Cw=100%
1800 Cw=133.3%
Cw=150%
1:1 line
1600 +15% line
-15% line

1400
experimental qu (kPa)

1200

1000

800

600

400

200

0
0 200 400 600 800 1000 1200 1400 1600 1800 2000 2200
predicted qu(kPa)
(a)

167
CHAPTER 4

(b)

Figure 4.61: (a) Comparison between experimental and predicted qu by formula


(4-12c); (b) Comparison between unconfined compressive strength obtained from this
study and Lee’s study (1999)

10000 10000
simulation of qu---s/c=0.47
simulation of qu ---s/c=1.0
simulation by using Lee et al.'s(2005) Formula
simulation by using Lee et al.'s(2005) Formula
simulation by using Formula of this study
simulation by using Formula of this study

1000 1000
qu (kPa)
qu (kPa)

100 100

10 10

1 1.2 1.4 1.6 1.8 2 2.2 2.4 2.6 1.4 1.6 1.8 2 2.2 2.4 2.6 2.8 3 3.2 3.4 3.6 3.8

w/c w/c

(a) S/C=0.47 (b)S/C=1.00

168
CHAPTER 4

1000 simulation of qu ---s/c=3.14


10000
simulation of q u---s/c=1.54 simulation by using Lee et al.'s(2005) Formula
simulation by using Lee et al.'s(2005) Formula simulation by using Formula of this study
simulation by using Formula of this study

1000

q u (kPa)
100
q u (kPa)

100

10
10
3.5 4 4.5 5 5.5 6 6.5
2 2.4 2.8 3.2 3.6 4 4.4 4.8
w/c
w/c

(c) S/C=1.54 (d) S/C=3.14

1000
simulation of q u---s/c=4.0
simulation by using Lee et al.'s(2005) Formula
simulation by using Formula of this study
q u (kPa)

100

10
4.5 5 5.5 6 6.5 7 7.5
w/c

(e) S/C=4

Figure 4.62: Simulated unconfined compressive strength by using Lee et al.’s (2005)’s
formula and this study’s formula

169
CHAPTER 4

Figure 4.63a: Simulated unconfined compressive strength by using modified formula


of this study and experimental unconfined compressive strength from Lee’s (1999)
study 10000
simulation of qu
test s/c=3.33
test s/c=2.00
test s/c=1.30
test s/c=1.00
simulation by Lee's(2005) formula
with different q0

1000
qu(kPa)

100

10
1.6 2 2.4 2.8 3.2 3.6 4 4.4 4.8 5.2 5.6 6 6.4 6.8 7.2

w/c

Figure 4.63b: Simulated unconfined compressive strength by using Lee’s (2005)


formula with different q0 and experimental unconfined compressive strength from this
study

170
CHAPTER 4

2200

2000

1800

1600

1400
deviator stress q(kPa)

10-1-11 post yielding behaivor


1200 10%-100%-CIU300-50
10%-100%-CIU300-150
10%-100%-CIU500-50
10%-100%-CIU500-250
1000 10%-100%-CIU1000-50
10%-100%-CIU1000-500
10%-100%-CIU1500-50
10%-100%-CIU1500-750
800 10%-100%-CIU2000-50
10%-100%-CIU2000-500
10%-100%-CIU2000-1000
10%-100%-CIU2500-50
600 10%-100%-CIU2500-1250
tension cut-off line
yield point-P'0=300kPa
yield point-P'0=500kPa
400 yield point-P'0=1000kPa
yield point-P'0=1500kPa
yield point-P'0=2000kPa
200 yield point-P'0=2500kPa
yield locus
primary yield locus
primary yield point from CIU test

0
0 250 500 750 1000 1250 1500 1750 2000 2250 2500
mean normal effective stressP'(kPa)

(a)

2200 10-1-11 post yielding behavior


10%-100%-CIU300-50
10%-100%-CIU300-150
10%-100%-CIU500-50
2000 10%-100%-CIU500-250
10%-100%-CIU1000-50
10%-100%-CIU1000-500
1800 10%-100%-CIU1500-50
10%-100%-CIU1500-750
10%-100%-CIU2000-50
10%-100%-CIU2000-500
1600
10%-100%-CIU2000-1000
10%-100%-CIU2500-50
10%-100%-CIU2500-1250
deviator stress q(kPa)

1400 primary yield point from CIU test

1200

1000

800

600

400

200

0
0 0.05 0.1 0.15 0.2
deviator strain εs

(b)

Figure 4.64a: Post-yield stress-strain behavior of CIU specimen with mix proportion
10:1:11 cured under atmospheric pressure for 7days (a) stress path, (b) deviator
stress-strain curve

171
CHAPTER 4

10-1-11 post yielding behavior


10-1-11 post yielding behaivor
10%-100%-CID300-100 10%-100%-CID300-100
5500 10%-100%-CID300-150 5500 10%-100%-CID300-150
10%-100%-CID500-50
10%-100%-CID500-50
10%-100%-CID500-100
10%-100%-CID500-150 10%-100%-CID500-100
5000 10%-100%-CID500-250
5000 10%-100%-CID500-150
10%-100%-CID1000-50 10%-100%-CID500-250
10%-100%-CID1000-200
10%-100%-CID1000-250 10%-100%-CID500
10%-100%-CID1000-500 10%-100%-CID1000-50
4500 10%-100%-CID1500-50 4500 10%-100%-CID1000-200
10%-100%-CID1500-375
10%-100%-CID1500-500
10%-100%-CID1000-250
10%-100%-CID1500-750 10%-100%-CID1000-500
4000 10%-100%-CID2000-500 4000 10%-100%-CID1000
10%-100%-CID2000-1000
10%-100%-CID2000-1250
10%-100%-CID1500-50
10%-100%-CID2500-50 10%-100%-CID1500-375

deviator stress q(kPa)


10%-100%-CID2500-500 10%-100%-CID1500-500
3500 3500
deviator stress q(kPa)

10%-100%-CID2500-1250
10%-100%-CID2500-1750
10%-100%-CID1500-750
tension cut-off line 10%-100%-CID2000-500
yield point-P'0=300kPa 10%-100%-CID2000-1000
3000 yield point-P'0=500kPa 3000 10%-100%-CID2000-1250
yield point-P'0=1000kPa
yield point-P'0=1500kPa 10%-100%-CID2500-50
yield point-P'0=2000kPa 10%-100%-CID2500-500
yield point-P'0=2500kPa
2500 yield locus
2500 10%-100%-CID2500-1250
primary yield locus 10%-100%-CID2500-1750

2000 2000

1500 1500

1000 1000

500 500

0 0
0 250 500 750 1000 1250 1500 1750 2000 2250 2500 2750 3000 3250 3500 0 0.05 0.1 0.15 0.2
mean normal effective stressP'(kPa) deviator strain εs

0 250 500 750 1000 1250 1500 1750 2000 2250 2500 2750 3000 3250 3500
3.4
10-1-11 post yielding behaivor
10%-100%-CID300-100
10%-100%-CID300-150
10%-100%-CID500-50
3.2 10%-100%-CID500-100
10%-100%-CID500-150
10%-100%-CID500-250
10%-100%-CID500
10%-100%-CID1000-50
10%-100%-CID1000-200
3 10%-100%-CID1000-250
10%-100%-CID1000-500
10%-100%-CID1000
10%-100%-CID1500-50
10%-100%-CID1500-375
specific volume v

2.8 10%-100%-CID1500-500
10%-100%-CID1500-750
10%-100%-CID2000-500
10%-100%-CID2000-1000
10%-100%-CID2000-1250
10%-100%-CID2500-50
2.6 10%-100%-CID2500-500
10%-100%-CID2500-1250
10%-100%-CID2500-1750
yield point-P'0=300kPa
yield point-P'0=500kPa
2.4 yield point-P'0=1000kPa
yield point-P'0=1500kPa
yield point-P'0=2000kPa
yield point-P'0=2500kPa

2.2

Figure 4.64b: Post-yield stress-strain behavior of CID specimen with mix proportion
10:1:11 cured under atmospheric pressure for 7days

172
CHAPTER 4

2500

2250

2000

1750
deviator stress q(kPa)

5-1-6 post yielding behaivor


20%-100%-CIU380-50
1500 20%-100%-CIU380-190
20%-100%-CIU500-50
20%-100%-CIU500-250
20%-100%-CIU1000-50
1250 20%-100%-CIU1000-250
20%-100%-CIU1000-500
20%-100%-CIU1500-50
20%-100%-CIU1500-500
1000 20%-100%-CIU1500-750
20%-100%-CIU2000-50
20%-100%-CIU2000-1000
20%-100%-CIU2500-50
20%-100%-CIU2500-500
750 20%-100%-CIU2500-1250
tension cut-off line
yield point-P'0=380kPa
yield point-P'0=500kPa
500
yield point-P'0=1000kPa
yield point-P'0=1500kPa
yield point-P'0=2000kPa
250 yield point-P'0=2500kPa
yield locus
primary yield locus
primary yield point from CIU test

0
0 250 500 750 1000 1250 1500 1750 2000 2250 2500
mean normal effective stressP'(kPa)

(a)

2500 5-1-6 post yielding behavior


20%-100%-CIU380-50
20%-100%-CIU380-190
20%-100%-CIU500-50
2250 20%-100%-CIU500-250
20%-100%-CIU1000-50
20%-100%-CIU1000-250
2000 20%-100%-CIU1000-500
20%-100%-CIU1500-50
20%-100%-CIU1500-500
20%-100%-CIU1500-750
1750 20%-100%-CIU2000-50
20%-100%-CIU2000-500
deviator stress q(kPa)

20%-100%-CIU2000-1000
1500 20%-100%-CIU2500-50
20%-100%-CIU2500-500
20%-100%-CIU2500-1250
primary yield point from CIU test
1250

1000

750

500

250

0
0 0.05 0.1 0.15 0.2 0.25
deviator strain εs
(b)

Fig 4.65a: Post-yield stress-strain behavior of CIU specimen with mix proportion 5:1:6
cured under atmospheric pressure for 7days (a) stress path, (b) deviator stress-strain
curve

173
CHAPTER 4

5-1-6 post yielding behavior


5-1-6 post yielding behaivor
20%-100%-CID380-50 20%-100%-CID380-50
20%-100%-CID380-190 20%-100%-CID380-190
6500 20%-100%-CID500-50 6500
20%-100%-CID500-100
20%-100%-CID500-50
6250 20%-100%-CID500-250
6250 20%-100%-CID500-100
6000 20%-100%-CID500 6000 20%-100%-CID500-250
20%-100%-CID1000-50
20%-100%-CID500
5750 20%-100%-CID1000-250 5750
20%-100%-CID1000-500 20%-100%-CID1000-50
5500 20%-100%-CID1000 5500 20%-100%-CID1000-250
20%-100%-CID1500-50
5250 20%-100%-CID1500-500 5250 20%-100%-CID1000-500
20%-100%-CID1500-750 20%-100%-CID1000
5000 20%-100%-CID2000-50
5000 20%-100%-CID1500-50
4750 20%-100%-CID2000-500 4750 20%-100%-CID1500-500
20%-100%-CID2000-1000
4500 20%-100%-CID2000-1500 4500 20%-100%-CID1500-750
20%-100%-CID2500-50 20%-100%-CID2000-50
4250 4250

deviator stress q(kPa)


20%-100%-CID2500-500
20%-100%-CID2000-500
deviator stress q(kPa)

20%-100%-CID2500-850
4000 20%-100%-CID2500-1250 4000 20%-100%-CID2000-1000
3750 20%-100%-CID2500-1750 20%-100%-CID2000-1500
20%-100%-CID2500-2000
3750
20%-100%-CID2500-50
3500 tension cut-off line
yield point-P'0=380kPa
3500 20%-100%-CID2500-500
3250 yield point-P'0=500kPa 3250 20%-100%-CID2500-850
yield point-P'0=1000kPa
20%-100%-CID2500-1250
3000 yield point-P'0=1500kPa 3000
yield point-P'0=2000kPa 20%-100%-CID2500-1750
2750 yield point-P'0=2500kPa 2750 20%-100%-CID2500-2000
yield locus
2500 primary yield locus 2500
2250 2250
2000 2000
1750 1750
1500 1500
1250 1250
1000 1000
750 750
500 500
250 250
0 0
0 250 500 750 1000125015001750200022502500275030003250350037504000 0 0.05 0.1 0.15 0.2 0.25 0.3
mean normal effective stressP'(kPa) deviator strain εs

0 250 500 750 1000125015001750200022502500275030003250350037504000


3.4
5-1-6 post yielding behavior
3.3 20%-100%-CID380-50
20%-100%-CID380-190
20%-100%-CID500-50
3.2 20%-100%-CID500-100
20%-100%-CID500-250
20%-100%-CID500
3.1 20%-100%-CID1000-50
20%-100%-CID1000-250
20%-100%-CID1000-500
3 20%-100%-CID1000
20%-100%-CID1500-50
20%-100%-CID1500-500
2.9 20%-100%-CID1500-750
20%-100%-CID2000-50
specific volume v

20%-100%-CID2000-500
2.8
20%-100%-CID2000-1000
20%-100%-CID2000-1500
2.7 20%-100%-CID2500-50
20%-100%-CID2500-500
20%-100%-CID2500-850
2.6 20%-100%-CID2500-1250
20%-100%-CID2500-1750
20%-100%-CID2500-2000
2.5 yield point-P'0=380kPa
yield point-P'0=500kPa
yield point-P'0=1000kPa
2.4
yield point-P'0=1500kPa
yield point-P'0=2000kPa
2.3 yield point-P'0=2500kPa

2.2

2.1

1.9

Fig 4.65b: Post-yield stress-strain behavior of CID specimen with mix proportion
5:1:6 cured under atmospheric pressure for 7days

174
CHAPTER 4

2750

2500

2250

2000

1750
deviator stress q(kPa)

10-3-13 post yielding behaivor


1500 30%-100%-CIU500-50
30%-100%-CIU500-150
30%-100%-CIU500-250
30%-100%-CIU1000-50
1250 30%-100%-CIU1000-500
30%-100%-CIU1500-50
30%-100%-CIU1500-750
1000 30%-100%-CIU2000-50
30%-100%-CIU2000-500
30%-100%-CIU2000-1000
30%-100%-CIU2500-50
750 30%-100%-CIU2500-500
tension cut-off line
yield point-P'0=500kPa
500 yield point-P'0=1000kPa
yield point-P'0=1500kPa
yield point-P'0=2000kPa
yield point-P'0=2500kPa
250
yield locus
primary yield locus
primary yield point from CIU test
0
0 250 500 750 1000 1250 1500 1750 2000 2250 2500
mean normal effective stressP'(kPa)

(a)
2500 10-3-13 post yielding behavior
30%-100%-CIU500-50
30%-100%-CIU500-150
30%-100%-CIU500-250
2250 30%-100%-CIU1000-50
30%-100%-CIU1000-500
30%-100%-CIU1500-50
30%-100%-CIU1500-750
2000 30%-100%-CIU2000-50
30%-100%-CIU2000-500
30%-100%-CIU2000-1000
1750 30%-100%-CIU2500-50
30%-100%-CIU2500-500
deviator stress q(kPa)

primary yield point from CIU test

1500

1250

1000

750

500

250

0
0 0.05 0.1 0.15 0.2
deviator strain εs

(b)

Fig 4.66a: Post-yield stress-strain behavior of specimen with mix proportion 10:3:13
cured under atmospheric pressure for 7days (a) stress path, (b) deviator stress-strain
curve

175
CHAPTER 4

10-3-13 post yielding behavior


30%-100%-CID500-50
6500 10-3-13 post yielding behaivor
30%-100%-CID500-50
6500 30%-100%-CID500-150
30%-100%-CID500-150 30%-100%-CID500-250
30%-100%-CID500-250 30%-100%-CID1000-50
6000 30%-100%-CID1000-50 6000
30%-100%-CID1000-250
30%-100%-CID1000-250
30%-100%-CID1000-500 30%-100%-CID1000-500
5500 30%-100%-CID1500-50 5500 30%-100%-CID1500-50
30%-100%-CID1500-375 30%-100%-CID1500-375
30%-100%-CID1500-500 30%-100%-CID1500-500
30%-100%-CID1500-750
5000 30%-100%-CID2000-500 5000 30%-100%-CID1500-750
30%-100%-CID2000-1000 30%-100%-CID2000-500
30%-100%-CID2000-1250 30%-100%-CID2000-1000
4500 30%-100%-CID2500-50 4500 30%-100%-CID2000-1250
30%-100%-CID2500-500 30%-100%-CID2500-50

deviator stress q(kPa)


30%-100%-CID2500-1250
deviator stress q(kPa)

30%-100%-CID2500-500
30%-100%-CID2500-1500
4000 tension cut-off line
4000 30%-100%-CID2500-1250
yield point-P'0=500kPa 30%-100%-CID2500-1500
yield point-P'0=1000kPa
3500 yield point-P'0=1500kPa 3500
yield point-P'0=2000kPa
yield point-P'0=2500kPa
yield locus
3000 primary yield locus
3000

2500 2500

2000 2000

1500 1500

1000 1000

500 500

0 0
0 250 500 750 1000 1250 1500 1750 2000 2250 2500 2750 3000 3250 3500 3750 0 0.05 0.1 0.15 0.2
mean normal effective stressP'(kPa) deviator strain εs

0 250 500 750 1000 1250 1500 1750 2000 2250 2500 2750 3000 3250 3500 3750
3.4
10-3-13 post-yielding behavior
30%-100%-CID500-50
3.3 30%-100%-CID500-150
30%-100%-CID500-250
30%-100%-CID1000-50
3.2
30%-100%-CID1000-250
30%-100%-CID1000-500
3.1 30%-100%-CID1500-50
30%-100%-CID1500-375
30%-100%-CID1500-500
3 30%-100%-CID1500-750
30%-100%-CID2000-500
30%-100%-CID2000-1000
2.9 30%-100%-CID2000-1250
specific volume v

30%-100%-CID2500-50
30%-100%-CID2500-500
2.8 30%-100%-CID2500-1250
30%-100%-CID2500-1500
yield point-P'0=500kPa
2.7 yield point-P'0=1000kPa
yield point-P'0=1500kPa
yield point-P'0=2000kPa
2.6 yield point-P'0=2500kPa

2.5

2.4

2.3

2.2

2.1

Fig 4.66b: Post-yield stress-strain behavior of specimen with mix proportion 10:3:13
cured under atmospheric pressure for 7days

176
CHAPTER 4

2750

2500

2250

2000

1750
deviator stress q(kPa)

2-1-3 post yielding behaivor


50%-100%-CIU1000-50
1500
50%-100%-CIU1000-500
50%-100%-CIU1500-50
50%-100%-CIU1500-500
1250
50%-100%-CIU1500-750
50%-100%-CIU2000-50
50%-100%-CIU2000-750
1000
50%-100%-CIU2000-1000
50%-100%-CIU2500-50
50%-100%-CIU2500-750
750 tension cut-off line
yield point-P'0=1000kPa
yield point-P'0=1500kPa
500
yield point-P'0=2000kPa
yield point-P'0=2500kPa
250 yield locus
primary yield locus
primary yield point from CIU test

0
0 250 500 750 1000 1250 1500 1750 2000 2250 2500
mean normal effective stressP'(kPa)

(a)

2750 2-1-3 post yielding behavior


50%-100%-CIU1000-50
50%-100%-CIU1000-500
2500 50%-100%-CIU1500-50
50%-100%-CIU1500-500
50%-100%-CIU1500-750
50%-100%-CIU2000-50
2250
50%-100%-CIU2000-750
50%-100%-CIU2000-1000
50%-100%-CIU2500-50
2000 50%-100%-CIU2500-750
primary yield point from CIU test
deviator stress q(kPa)

1750

1500

1250

1000

750

500

250

0
0 0.05 0.1 0.15 0.2
deviator strain εs

(b)

Fig 4.67a: Post-yield stress-strain behavior of specimen with mix proportion 2:1:3
cured under atmospheric pressure for 7days (a) stress path, (b) deviator stress-strain
curve

177
CHAPTER 4

2-1-3 post yielding behavior


2-1-3 post yielding behaivor
50%-100%-CID1000-50 50%-100%-CID1000-50
5750 50%-100%-CID1000-250
5750 50%-100%-CID1000-250
50%-100%-CID1000-500
5500 50%-100%-CID1000-500 5500
50%-100%-CID1000 50%-100%-CID1000
5250 50%-100%-CID1500-250 5250 50%-100%-CID1500-250
50%-100%-CID1500-375 50%-100%-CID1500-375
5000 50%-100%-CID1500-500
5000 50%-100%-CID1500-500
4750 50%-100%-CID1500-750 4750 50%-100%-CID1500-750
50%-100%-CID1500-1000 50%-100%-CID1500-1000
4500 50%-100%-CID2000-500 4500 50%-100%-CID2000-500
50%-100%-CID2000-1000 50%-100%-CID2000-1000
4250 50%-100%-CID2000-1250
4250
50%-100%-CID2000-1250
4000 50%-100%-CID2500-500 4000 50%-100%-CID2500-500
50%-100%-CID2500-750 50%-100%-CID2500-750
3750 3750

deviator stress q(kPa)


50%-100%-CID2500-1250 50%-100%-CID2500-1250
deviator stress q(kPa)

tension cut-off line


3500 yield point-P'0=1000kPa
3500
3250 yield point-P'0=1500kPa 3250
yield point-P'0=2000kPa
3000 yield point-P'0=2500kPa 3000
yield locus
2750 primary yield locus
2750
2500 2500
2250 2250
2000 2000
1750 1750
1500 1500
1250 1250
1000 1000
750 750
500 500
250 250
0 0
0 250 500 750 1000 1250 1500 1750 2000 2250 2500 2750 3000 3250 0 0.05 0.1 0.15 0.2 0.25
mean normal effective stressP'(kPa) deviator strain εs

0 250 500 750 1000 1250 1500 1750 2000 2250 2500 2750 3000 3250
3.2
2-1-3 post-yielding behavior
50%-100%-CID1000-50
50%-100%-CID1000-250
3.1 50%-100%-CID1000-500
50%-100%-CID1000
50%-100%-CID1500-250
50%-100%-CID1500-375
3 50%-100%-CID1500-500
50%-100%-CID1500-750
50%-100%-CID1500-1000
50%-100%-CID2000-500
50%-100%-CID2000-1000
2.9
50%-100%-CID2000-1250
50%-100%-CID2500-500
50%-100%-CID2500-750
specific volume v

50%-100%-CID2500-1250
2.8 yield point-P'0=1000kPa
yield point-P'0=1500kPa
yield point-P'0=2000kPa
yield point-P'0=2500kPa
2.7

2.6

2.5

2.4

2.3

2.2

Fig 4.67b: Post-yield stress-strain behavior of specimen with mix proportion 2:1:3
cured under atmospheric pressure for 7days

178
CHAPTER 4

2-1-4 post yielding behaivor


50%-133%-CIU500-50
50%-133%-CIU500-100
50%-133%-CIU500-250
50%-133%-CIU1000-50
2500 50%-133%-CIU1000-500
50%-133%-CIU1500-50
50%-133%-CIU1500-375
50%-133%-CIU1500-750
50%-133%-CIU2000-50
50%-133%-CIU2000-400
50%-133%-CIU2000-1000
50%-133%-CIU2500-50
50%-133%-CIU2500-500
2000 50%-133%-CIU2500-1250
tension cut-off line
yield point-P'0=500kPa
yield point-P'0=1000kPa
yield point-P'0=1500kPa
yield point-P'0=2000kPa
yield point-P'0=2500kPa
deviator stress q(kPa)

yield locus
primary yield locus
1500 primary yield point from CIU test

1000

500

0
0 500 1000 1500 2000 2500
mean normal effective stressP'(kPa)

(a)

2500 2-1-4 post yielding behavior


50%-133%-CIU500-50
50%-133%-CIU500-100
50%-133%-CIU500-250
50%-133%-CIU1000-50
50%-133%-CIU1000-500
50%-133%-CIU1500-50
2000 50%-133%-CIU1500-375
50%-133%-CIU1500-750
50%-133%-CIU2000-50
50%-133%-CIU2000-400
50%-133%-CIU2000-1000
50%-133%-CIU2500-50
deviator stress q(kPa)

50%-133%-CIU2500-500
50%-133%-CIU2500-1250
1500
primary yield point from CIU test

1000

500

0
0 0.05 0.1 0.15 0.2 0.25
deviator strain εs

(b)

Fig 4.68a: Post-yield stress-strain behavior of CIU specimen with mix proportion 2:1:4
cured under atmospheric pressure for 7days (a) stress path, (b) deviator stress-strain
curve

179
CHAPTER 4

2-1-4 post yielding behavior


2-1-4 post yielding behaivor 50%-133%-CID500-50
8000 50%-133%-CID500-50 8000 50%-133%-CID500-100
50%-133%-CID500-100
50%-133%-CID500-250 50%-133%-CID500-250
7500 50%-133%-CID500-350 7500 50%-133%-CID500-350
50%-133%-CID1000-50
50%-133%-CID1000-50
50%-133%-CID1000-250
7000 50%-133%-CID1000-350 7000 50%-133%-CID1000-250
50%-133%-CID1000-500 50%-133%-CID1000-350
50%-133%-CID1000-650
50%-133%-CID1000-500
6500 50%-133%-CID1500-50 6500
50%-133%-CID1500-375 50%-133%-CID1000-650
50%-133%-CID1500-500 50%-133%-CID1500-50
6000 50%-133%-CID1500-750 6000 50%-133%-CID1500-375
50%-133%-CID2000-50
50%-133%-CID2000-400 50%-133%-CID1500-500
5500 50%-133%-CID2000-500 5500 50%-133%-CID1500-750
50%-133%-CID2000-1000 50%-133%-CID2000-50

deviator stress q(kPa)


deviator stress q(kPa)

50%-133%-CID2000-1500
5000 50%-133%-CID2500-50 5000 50%-133%-CID2000-400
50%-133%-CID2500-500 50%-133%-CID2000-500
50%-133%-CID2500-1250 50%-133%-CID2000-1000
4500 50%-133%-CID2500-1750 4500
50%-133%-CID2500-2000
50%-133%-CID2000-1500
tension cut-off line 50%-133%-CID2500-50
4000 yield point-P'0=500kPa 4000 50%-133%-CID2500-500
yield point-P'0=1000kPa
yield point-P'0=1500kPa
50%-133%-CID2500-1250
3500 yield point-P'0=2000kPa 3500 50%-133%-CID2500-1750
yield point-P'0=2500kPa 50%-133%-CID2500-2000
yield locus
3000 primary yield locus 3000

2500 2500

2000 2000

1500 1500

1000 1000

500 500

0 0
0 500 1000 1500 2000 2500 3000 3500 4000 4500 0 0.05 0.1 0.15 0.2 0.25 0.3 0.35
mean normal effective stressP'(kPa) deviator strain εs

0 500 1000 1500 2000 2500 3000 3500 4000 4500


3.8
2-1-4 post yielding behavior
3.7 50%-133%-CID500-50
50%-133%-CID500-100
3.6 50%-133%-CID500-250
50%-133%-CID500-350
50%-133%-CID1000-50
3.5
50%-133%-CID1000-250
50%-133%-CID1000-350
3.4 50%-133%-CID1000-500
50%-133%-CID1000-650
3.3 50%-133%-CID1500-50
50%-133%-CID1500-375
3.2 50%-133%-CID1500-500
50%-133%-CID1500-750
3.1 50%-133%-CID2000-50
specific volume v

50%-133%-CID2000-400
3 50%-133%-CID2000-500
50%-133%-CID2000-1000
50%-133%-CID2000-1500
2.9 50%-133%-CID2500-50
50%-133%-CID2500-500
2.8 50%-133%-CID2500-1250
50%-133%-CID2500-1750
2.7 50%-133%-CID2500-2000
yield point-P'0=500kPa
2.6 yield point-P'0=1000kPa
yield point-P'0=1500kPa
2.5 yield point-P'0=2000kPa
yield point-P'0=2500kPa
2.4
2.3
2.2
2.1
2
1.9

Fig 4.68b: Post-yield stress-strain behavior of CID specimen with mix proportion
2:1:4 cured under atmospheric pressure for 7days

180
CHAPTER 4

2
10-1-11 evolution of yield locus
primary yield point
yield point-p'0=300kPa
yield point-p'0=500kPa
yield point-p'0=1000kPa
1.6
yield point-p'0=1500kPa
yield point-p'0=2000kPa
yield point-p'0=2500kPa
primary yield locus
yield locus
1.2
tension cut-off line
q/p 0 '

0.8

0.4

0 0.2 0.4 0.6 0.8 1


p'/p0'

(a)

2
5-1-6 evolution of yield locus
primary yield point
yield point-p'0=380kPa
yield point-p'0=500kPa
yield point-p'0=1000kPa
1.6
yield point-p'0=1500kPa
yield point-p'0=2000kPa
yield point-p'0=2500kPa
primary yield locus
yield locus
1.2 tension cut-off line
q/p 0 '

0.8

0.4

0 0.2 0.4 0.6 0.8 1


p'/p0'

(b)

181
CHAPTER 4

2
10-3-13 evolution of yield locus
primary yield point
yield point-p'0=500kPa
yield point-p'0=1000kPa
yield point-p'0=1500kPa
1.6
yield point-p'0=2000kPa
yield point-p'0=2500kPa
primary yield locus
yield locus
tension cut-off line
1.2
q/p 0 '

0.8

0.4

0 0.2 0.4 0.6 0.8 1


p'/p0'

(c)

2
2-1-3 evolution of yield locus
primary yield point
yield point-p'0=1000kPa
yield point-p'0=1500kPa
yield point-p'0=2000kPa
1.6
yield point-p'0=2500kPa
primary yield locus
yield locus
tension cut-off line

1.2
q/p 0 '

0.8

0.4

0 0.2 0.4 0.6 0.8 1


p'/p0'

(d)

182
CHAPTER 4

2
2-1-4 evolution of yield locus
primary yield point
yield point-p'0=500kPa
yield point-p'0=1000kPa
yield point-p'0=1500kPa
1.6
yield point-p'0=2000kPa
yield point-p'0=2500kPa
primary yield locus
yield locus
tension cut-off line
1.2
q/p 0 '

0.8

0.4

0 0.2 0.4 0.6 0.8 1


p'/p0'

(e)

Fig 4.69 Evolution of yield locus for cement treated marine clay specimens on
normalized stress space (a) mix proportion 10:1:11 (b) 5:1:6 (c) 10:3:13 (d) 2:1:3 (e)
2:1:4

183
CHAPTER 4

2
cement content effect on evolution of yield locus
tension cutoff line
yield locus-Aw=10%
yield locus-Aw=20%
yield locus-Aw=50%
1.6 remoulded marine clay with M=0.9

1.2
q/p 0 '

0.8

0.4

0 0.2 0.4 0.6 0.8 1


p'/p0 '
(a)
2
total water content effect on evolution of yield locus
tension cutoff line
yield locus-2-1-3
yield locus-2-1-4
remoulded marine clay with M=0.9
1.6

1.2
q/p 0 '

0.8

0.4

0 0.2 0.4 0.6 0.8 1


p'/p0 '

(b)
Fig 4.70 Comparison between evolutions of yield locus for cement treated marine clay
specimens on normalized stress space (a) different cement content (b) different total
water content

184
CHAPTER 4

(a)

(b)

Figure 4.71 Microstructure of cement treated marine clay specimens with mix
proportion 5:1:6 under IPC pressure 2500kPa and drained shearing (CID2500-1250) (a)
outside of slip band (b) inside of slip band

185
CHAPTER 4

(a)

(b)
Figure 4.72: Microstructure of cement treated marine clay specimens with mix
proportion 5:1:6 under IPC pressure 2500kPa and undrained shearing (2500-1250CU)
(a) outside of slip band (b) inside of slip band

186
CHAPTER 4

y/D/2
1

0.5

0
-2 -1 0 1 2 3 4 5 6 7 8 9 10
2P/πDt
-0.5

-1

Figure 4.73: Theoretical solutions to the stress distribution of cylinder specimen with
thickness t and diameter D under two radial and opposite line loading [after
Timoshenko and Goodier, 1970]

(a) (b) (c)

Figure 4.74: The failure pattern of cement-treated marine clay cylinder specimen
(a)split specimen (b) top loading face after Brazilian test (c) specimen with central
crack

187
CHAPTER 4

180
curing time 7 days
curing time 28 days

160

140

120

100
σt(kPa)

80

60

40

20

0
0 0.1 0.2 0.3 0.4 0.5

cement content Aw

(a)

180 280
curing time 7 days
260
160
curing time 7 days
240
curing time 28 days
140 220

200
120
180

160
100
σt(kPa)

σt(kPa)

140
80
120

100
60
80

40 60

40
20
20

0 0
50 100 150 200 0 50 100 150 200 250

total water content Cw curing strss pcur'

(b) (c)

Figure 4.75: The tensile splitting strength of cement-treated marine clay cylinder
specimen (a) tensile strength versus cement content (b) tensile strength versus total
water content (c) tensile strength versus curing stress

188
CHAPTER 4

curing stress-s-c-w-curing time


200 0-10-1-11-7days
0-5-1-6-7days
0-10-3-13-7days
180 0-2-1-3-7days
0-2-1-5-7days
0-2-1-5.5-7days
160 50-2-1-4-7days
100-2-1-4-7days

tensil strength under compresson rate 0.005mm/Min


250-2-1-4-7days
140 0-10-3-13-28days
0-2-1-3-28days
0-2-1-5-28days
1:1 line
120
15% line
-15% line

100

80

60

40

20

0
0 20 40 60 80 100 120 140 160 180 200
tensil strength under compression rate 0.5mm/Min

Figure 4.76: Comparisons between very fast and vey slow compression rate

curing stress-s-c-w-curing time


140 0-10-1-11-28days
0-5-1-6-28days
0-10-3-13-28days
0-2-1-5-28days
120 0-2-1-4-7days
0-2-1-4-28days
1:1 line
15% line
-15% line
100
compression under water

80

60

40

20

0
0 20 40 60 80 100 120 140
compression under dry condition

Figure 4.77: Comparisons between compression under dry condition and water

189
CHAPTER 4

(a) (b)

Figure 4.78: Prolonged submerged test (a) remoulded marine clay sample (b)
cement-treated marine clay sample

2500
qu-σt
Fit 2: qu=7.87σt, R2=0.996

2000

1500
qu(kPa)

1000

500

0
0 100 200 300
tensile strength σt(kPa)

Figure 4.79: Relationship between tensile strength and unconfined compressive


strength

190
CHAPTER 4

1200 modified primary yield locus


s-c-w
10-1-11
1100
5-1-6
10-3-13
1000
2-1-3
tension cut-off line
900 10-1-11 tension failure envelop
5-1-6 tension failure envelop
800 10-3-13 tension failure envelop

deviator stress q(kPa)


2-1-3 tension failure envelop
700

600

500

400

300

200

100

0
-200 -100 0 100 200 300 400 500 600
mean normal effective stressP'(kPa)

Figure 4.80: Modified primary yield locus for cement-treated marine clay
specimens with different cement content and under triaxial loading

1200 modified primary yield locus


s-c-w
2-1-3
1100
2-1-4
2-1-5.5
1000
tension cut-off line
2-1-3 tension failure envelop
900
2-1-4 tension failure envelop
2-1-5.5 tension failure envelop
800
deviator stress q(kPa)

700

600

500

400

300

200

100

0
-200 -100 0 100 200 300 400 500 600
mean normal effective stressP'(kPa)

Figure 4.81: Modified primary yield locus for cement-treated marine clay
specimens with different total water content and under triaxial loading

191
CHAPTER 4

2750 modified primary yield locus


cruing stress-s-c-w
0-2-1-4
2500 50-2-1-4
100-2-1-4
2250 250-2-1-4
tension cut-off line
2000 0-2-1-4 tension failure envelop
50-2-1-4 tension failure envelop
100-2-1-4 tension failure envelop
1750

deviator stress q(kPa)


250-2-1-4 tension failure envelop

1500

1250

1000

750

500

250

0
-500 -250 0 250 500 750 1000 1250 1500
mean normal effective stressP'(kPa)

Figure 4.82: Modified primary yield locus for cement-treated marine clay
specimens with different curing stress and under triaxial loading

900 modified primary yield locus


5-1-6-7days
2-1-4-7days
800 5-1-6-28days
2-1-4-28days
tension cut-off line
700
5-1-6-7days tension failure envelop
2-1-4-7days tension failure envelop
5-1-6-28days tension failure envelop
600
2-1-4-28days tension failure envelop
deviator stress q(kPa)

500

400

300

200

100

0
-100 0 100 200 300 400
mean normal effective stressP'(kPa)

Figure 4.83: Modified primary yield locus for cement-treated marine clay
specimens with different curing time period and under triaxial loading

192
CHAPTER 5
ULTIMATE STATE OF CEMENT-TREATED MARINE
CLAY

In this chapter, the post peak behavior and issues related to the ultimate state of

cement-treated marine clay are discussed. First, the previous works on ultimate state of

cemented soil are reviewed briefly. Second, the onset of strain softening and post-peak

behavior of cement-treated marine clay is discussed. Some preliminary results under triaxial

test on special specimen with lubrication and radial excess are presented. Third, a remoulding

process was also used to remould the soil in an attempt to achieve an ultimate state. The

triaxial behavior of remoulded cement-treated marine clay was investigated and the relevant

results are presented in detail. Finally some issues on the ultimate state of cement-treated

marine clay are discussed, based on the experiment observations.

5.1. Introduction

As discussed earlier, natural and artificial cemented soils possess structure, which makes

their behavior very different from that of reconstituted soil. In order to consider the effect of

structure, an intrinsic state is usually introduced by many researchers (e.g. Burland, 1990;

Leroueil & Vaughan, 1990). Many constitutive models for structured soils make use of an

intrinsic state as an “ultimate state” towards which the state of the soil evolves with continual

shearing. However, this ultimate state is often not easy to discern in practice due to the

difficulties in achieving the fully destuctured state of cemented or structured soil. Rouainia &

193
CHAPTER 5

Wood (2000) pointed out that unless a mechanical test systematically destroys the structure of

the entire sample, the condition of the natural soil may not approach a structureless asymptote.

Cotecchia & Chandler (1997) also noted that it will not be possible to deduce the asymptotic

behaviour externally.

Moreover, Leroueil (1998) noted that the critical state of natural and reconstituted clay

can be different, indicating that even at larger strains, soil behavior may be influenced by the

initial structure. He cited Saint-Jean-Vianney clay as an example where the intrinsic and

natural critical state angles of shearing resistance are 30.5 and 44.4 respectively. Wood (1990)

stated that even for reconstituted clays it is often not easy to discern the critical state at high

over-consolidation ratios. He noted that strength reduction after peak, accompanied by

dilatancy, tends to be a continuous process leading to stress ratios below the normally

consolidated critical-state value.

Some studies have also been done on the ultimate state of cemented soil. Airey (1993)

obtained an ultimate critical state line for naturally cemented carbonate soil. However, Airey

(1993) noted that the evidence for the existence of a critical state line in the compression line is

less clear cut. Kasama et al. (2000) found that the failure state line of lightly cemented clay is

parallel to that of the uncemented clay in the p’-q space whereas its slope is steeper than that of

an uncemented clay in the e-ln(p’) space. In contrast, Tatsuoka (1983) and Chin (2006) noted

that ultimate failure envelope of artificially cemented clay is a curve rather than straight line.

However, Chin (2006) also noted that the stress state of the material within the slip plane

cannot be fully defined. As such, the ultimate state of cement-treated marine clay is still not

clear. Based on experiment observations on different specimens, the ultimate state of

194
CHAPTER 5

cement-treated marine clay will be discussed here.

5.2. Post-peak behavior

So far all analyses are focused on pre-peak behavior of cement-treated marine clay. In this

section, the post-peak behavior of cement-treated marine clay will be briefly discussed.

However, it should be noted that the main focus of this study is on pre-peak, not post-peak

behavior.

5.2.1 Onset of strain softening

Strain localization or softening of stiff clay has been studied by many researchers (e.g.

Viggiani et al., 1993a, 1993b; Finno et al., 1994; Saada et al., 1994, 1995; Georgiannou &

Burland, 2001). Saada et al. (1994) also studied strain localization and shear banding in

saturated clayey soils. However, only global measurements of deformations and visual

examination of the photographs were used to interpret the results. Viggiani et al. (1993a, 1993b)

used local measurements of both pore pressures and deformations to study strain localization

in natural stiff overconsolidated clays. Viggiani et al. (1993a) showed that the pore pressure

measured at the base is different from that at the mid-height of the specimen; see Figure 5.1,

with maximum difference being observed at the point of maximum stress ratio q / p ' . By

introducing two indices on the strain homogeneity based on the local measurement of axial

deformation (Figure 5.1), Viggiani et al. (1993a) showed that localization is initiated at the

maximum stress ratio q / p ' and localization is a progressive phenomenon, leading eventually

to the formation of a shear band.

195
CHAPTER 5

Figure 5.2 shows the results of CIU specimens with 50% cement content, 133% total

water content under low consolidation stress (50~250kPa). In these figures, all quantities are

normalized by their maximum values. As can be seen, the normalized stress ratio q / p ' usually

coincides with normalized pore pressure until maximum stress ratio q / p ' , which occurred

before or at peak value of deviator stress q. After maximum stress ratio q / p ' , the normalized

stress ratio curve and pore pressure curve starts to deviate from each other. This phenomenon

can also be seen in other cases with maximum consolidation pressure of 500kPa, 2000kPa

(Figures 5.3 and 5.4). Figure 5.5 shows similar results for CIU specimens with 20% cement

content, 100% total water content under low to high consolidation stress. This phenomenon,

which shows the stationary points for the excess pore pressure and stress ratio coinciding with

each other, is similar to that observed by Viggiani et al. (1993a) and suggests that excess pore

pressure and stress ratio may be better indicators for shear band initiation than the deviator

stress itself.

5.2.2 Special specimen with lubrication and radial excess

All the results presented up to now were for specimens with height of 76mm and diameter

of 38mm. All these specimens showed post-peak softening behavior and shear band formation

in CIU and CID test, similar to what has been reported by Chin (2006). As noted by Chin

(2006), once the shear band is formed, the soil specimen is not uniform anymore, and

macroscopic changes in shear strain, specific volume and pore pressure do not reflect those

which occur within the shear band.

Vardoularkis (1978) suggested that shear band formation may be inhibited by using short

196
CHAPTER 5

specimens. The motivation for using short specimens stems from the belief that shear bands

are necessarily inclined at an angle to the horizontal and therefore need a certain minimum

length of specimen to develop. By using short specimens, it was felt that shear band formation

can be inhibited by the close proximity of the end caps. Much work has been conducted on

short specimens. For instance, Abdulla and Kiousis (1997) reported that “…the experimental

program implemented here is based on the following principles: (1) the specimen must have

lubricated ends; (2) the specimen must be as large (in diameter) as possible; and (3) the

specimen must be as short as possible...”. In general, the results of this work indicate that short

specimens alone do not completely inhibit shear banding; they need to be used in conjunction

with lubricated platens in order to enhance the uniformity of strain in the specimen. Where

both conditions can be achieved, what is often observed is the formation of multiple shear

bands rather than completely uniform behaviour (e.g. Samieh and Wong 1997, Wong 1999,

Desrues and Chambon 2002, Desrues and Viggiani 2004). Notwithstanding this, it is felt that,

compared to a single shear band, multiple shear bands allow macroscopic volumetric and shear

strain measurements as well as pore pressure measurements to be more reflective of that within

the shear band, and is therefore preferable. For example, Samieh and Wong (1997) noted that

“…For tests with high H/D ratios and conventional ends, a single shear band is developed

(Figure 5a). The formation of the single shear band not only restricts the dilation developed

within this localized zone, resulting in the least induced shear dilation, but also causes the

dilation rate to drop abruptly close to zero at the residual stress. The use of free ends in long

specimens does increase the total dilation. However, the dilation is still limited at the zones

adjacent to the top and bottom free ends. Series of shear bands in conical configurations are

197
CHAPTER 5

developed near the free ends as illustrated in Figure 5b. The use of a short specimen and free

ends changes the stress–strain response and mode failure drastically. Multiple shear bands are

developed uniformly within the specimen. There is no preferred orientation in these shear

bands. Formation of multiple shear bands results in a remarkable increase in the total dilation.

The occurrence of the residual stress and dilation rate is also delayed. However, the residual

dilation rate is constant with increasing axial strain, showing that the specimen has not

reached the critical state….”.

The implementation of free ends (or frictionless end caps) has been reported by various

researchers (e.g. Rowe & Barden 1964, Barden & McDermott 1965). Lee (1978) discussed the

end restraint effect in CIU test. Abdulla & Panos (1997) discussed the end friction, specimen

size and slenderness for cemented sand in triaxial test. Petchgate et al. (2000) discussed the

effect of height-to-diameter ratio on strength of cement-treated Bangkok Clay. In the

discussion below, the results of some preliminary tests using short specimens to investigate

post-peak behavior will be presented and discussed.

5.2.2.1 Effect of slenderness ratio

Figure 5.6 shows the results of two specimens, with slenderness ratios (ratio of height to

diameter or H/D) of 2 and 1 but with the porous stone ends in CIU tests with 1500kPa

effective confining stress. As can be seen, the shorter specimen (with H/D ratio of 1) shows a

significantly slower rate of strain softening than the taller specimen. However, there is much

less difference in terms of excess pore pressure development and stress path. In both cases, a

single shear band is clearly visible after the test, with post-peak stress path far away from

198
CHAPTER 5

tension cut-off line (see Figure 5.7). This indicates that the post-peak failure may be mainly

shear failure in these cases.

5.2.2.2 Effect of lubrication

Figure 5.8 shows the results of two specimens with different end platens but with the

same H/D ratio of 2 in CIU tests under 1000kPa confining stress. The two types of end platens

are the normal porous stone platen (high friction) and Teflon platens (low friction). As can be

seen, the specimen with Teflon platens also shows a slightly slower rate of strain softening

than that with the porous stone platens. Stress paths and excess pore pressure development are

nearly identical for both specimens. Each specimen shows a single shear band (see Figure5.9),

with post-peak stress path far away from tension cut-off line. This indicates that only low

friction platen does not change the post-peak failure pattern of the specimen.

5.2.2.3 Effect of combination of slenderness and lubrication

Figure 5.10 shows the results of two specimens with different end caps and H/D ratios

(specimen 1 and specimen 2). As can be seen, the short specimen with Teflon end caps shows

significantly slower rate of strain softening than that with porous stones and H/D ratio of 2.

The stress path and maximum excess pore pressure are almost identical, but the shorter

specimen (also with Teflon end caps) shows almost no post-peak pore pressure drop. As Figure

5.11 shows, the shorter specimen now develops two shear bands instead of one. In both cases,

the post-peak stress path are still far away from the tension cut-off line, indicating that short

specimen with low friction shows similar post-peak failure pattern to long specimen.

199
CHAPTER 5

5.2.2.4 Effect of enlarged low-friction end caps

Specimen 3 in Figure 5.10 is a short specimen tested with enlarged Teflon platens. The

use of enlarged low-friction platens was suggested by Rowe and Barden (1964) and arises out

of the observation that, as the specimen is compressed, it expands laterally, allowing the platen

to punch into it and creating a stress-concentration zone, from which a shear band usually

starts. Current tests also show that shear bands tend to initiate from the edge of the end caps.

As Figure 5.10 shows, the short specimen (i.e. Specimen 3) with enlarged low-friction platens

shows slower rate of strain softening than the other two specimens. In this Figure, the

parameter Ls is the difference between the diameter of the platen and the initial diameter of the

specimen, hereafter termed “radial excess”. Furthermore, the excess pore pressure also appears

to continue to increase, albeit slightly, after the peak strength is reached instead of decreasing,

as the other specimens show. Finally, the post-peak stress paths of Specimen 2 and 4 appear to

trend to tension cut-off line at large strain. This may indicate that for short specimen with

enlarged lower friction platens, the post-peak failure tends to change to tensile failure at large

strain, with the fissures over the specimens (see Figure 5.11c). As Figure 5.12 shows, a similar

trend is obtained with a lower effective confining stress of 500kPa (specimen 3 and 4). As

Figure 5.11c shows, the shorter specimen now shows multiple shear bands.

5.2.2.5 Stress state at post-peak stage

In order to check whether the tendency of increase in deviator stress and excess pore

pressure at post-peak stage will continue with axial strain, specimens with slenderness ratio of

1, lubrication and radial excess of 6mm were sheared up to very large axial strain of roughly

200
CHAPTER 5

55% under confining stress of 1000kPa in CIU tests. The tests results are shown in Figure 5.13.

As can be seen, there is a very gradual decrease in deviator stress and excess pore pressure

with axial strain. However, beyond about 20% strain, the deviator stress and excess pore

pressure have reached almost constant values. Similar trend was also observed with 35-mm

diameter short specimen with enlarged Teflon platens under 1000kPa effective confining stress

(see Figure 5.14, specimen 2). All these results suggest that the short specimens probably

reached a critical state at a shear strain of about 20%.

Similarly, Figures 5.13 to 5.14 show that the post-peak stress path of short specimens with

enlarged low friction platens trend to tension cut-off line, indicating tensile failure occurred at

large strain. As shown in Figure 5.15, instead of a single shear band, the short specimens with

enlarged Teflon platens after shearing to very large axial strain of about 55% show multiple

shear bands or fissures distributed all over the specimen surface. This is in accordance with the

observations of previous researchers and indicates a more uniform post-peak behavior than

that exhibited by long specimens showing a single shear band.

The differences between the conventional long specimens and the short specimens with

enlarged Teflon platens can be attributed to the following reasons:

a) In conventional specimens, shear strains are concentrated into the shear band.

Thus, for the same axial compression, the shear strain inside the shear band is much

higher. Microstructure study may account for this, which showed that destructuration

in the slip plane is much more serious than that of outside of the slip plane (section 4.6

and also Chin (2006)). Thus, the post-peak stress-strain curve of a conventional

specimen, plotted using the macroscopic shear strain, is reflective of that with much

201
CHAPTER 5

higher level of shear strain inside the shear band. This explains the apparently faster

drop in stress at post-peak stage in the conventional specimens.

b) Chin(2006) reported that conventional samples which were sheared at lower

IOCR(e.g. specimen 1, Figure 5.13c) tend to reach peak strength below the tensile

failure envelope, this usually occurring at shear strain levels of about 2.5% or less.

Such samples develop slip planes at peak strength. Strain softening is accompanied by

a decrease in mean normal effective stress. As the soil outside the slip plane is

unloading, the excess pore pressure is likely to be generated by the soil within the slip

plane. The typical duration of the post-peak shearing stage is about 90minutes, which

is less than the typical time required for equilibration of excess pore pressure during

isotropic swelling. Therefore, excess pore pressure within slip plane may not have

completely equilibrated to the rest of the specimen, and the measured excess pore

pressure may be an underestimate of that within the slip plane. The rest of the

specimens are actually swelling as the load reduces and this swelling, together with the

relatively small positive pore pressure reservoir which was confined to a single shear

band, leads to the rapid drop in post-peak excess pore pressure.

5.2.2.6 Results of CID short specimens

All the results presented above were obtained from CIU tests. Stress-strain behavior of

short cement-treated specimen in CID test is still unknown. To investigate the stress-strain

behavior of short sample in CID test, low consolidation stress (250kPa) and high consolidation

stress (500kPa) were applied before shearing. Consolidation stress higher than 500kPa was not

202
CHAPTER 5

applied due to the limitation of the current setup. Figures 5.16 to 5.17 show the results for

conventional specimen and short specimen in CID test. As can be seen, compared with

conventional specimens, the short CID samples show hardening behavior over a much larger

range of shear strain, with much larger volumetric strain at peak strength. The shear strains at

peak strength in short CID samples are about 55%, which are about three times of

conventional samples. Post-peak decrease in deviator stress persists up to about 65% shear

strain, at which point the test was terminated. Multiple shear bands or fissures were also

clearly distributed all over the specimen surface. Shear strain larger than 65% was not

achievable due to the limitation of the current setup; this is already much larger than that

observed by other researchers (e.g., Wong & Mitchell, 1975). Short CID specimens did not

show any stabilization of the post-peak strength reduction, even at very large shear strain under

the current setup. Hence the critical state could not be observed in short CID samples with the

current setup if it does exist. This finding is similar to those reported by other researchers (e.g.,

Samieh and Wong, 1997).

5.2.3. Critical state and residual state

The results on the short specimens raise the question on what is the critical state for

cement-treated marine clay. Chin (2006) regarded the critical state to be that reached by long

specimens after about 10% to 20% strain. He based this conclusion on the observation that the

deviator stress and pore pressure has more or less stabilized and scanning electron micrographs

which showed highly broken-up particles. As discussed above and shown in Figures 5.12 and

5.13, the stress ratio q/p’ reached in such circumstances is much lower than that achieved by

203
CHAPTER 5

short specimens at ~20% strain. As highlighted earlier, for the short specimens, deviator stress

and pore pressure have also stabilized after ~20% strain, but with the stress ratio at a much

higher level than that reached by the longer specimens. As shown in Figure 5.18, the “critical

state” line for the CIU short specimens has a gradient of about 2.60 (critical state of CID short

specimens could not be reached in this study as explained earlier) and is evidently much

steeper than that for conventional long specimens reported by Chin (2006) (mix proportion

5-1-6) as well as in this study (mix proportion 2-1-4), which has a gradient of about 1.12. The

Figure 5.18 also shows that the critical states reported by Chin (2006) and this study are very

close within the same stress range. It also must be pointed out that the maximum confining

stress for CIU specimen in Figure 5.18 is 1500kPa while that for CID specimen is 500kPa. As

Chin’s (2006) study and this study noted, the friction coefficient at critical state decreases with

the increase in effective confining stress, especially for CID specimens. As such for those CID

long specimens with maximum confining stress higher than 500kPa, the critical state line is

lower than that in Figure 5.18. However, for the stress range covered in Figure 5.18, the short

specimen and long specimens have gradients which are higher than that for marine clay critical

state line with M value of about 0.9(Chua, 1990). The question is therefore: which is closer to

the definition and concept of critical state?

It is clear that, in the case of Chin’s (2006) specimens, the shear strain within the shear

band would have been much higher than 20%, given the small thickness of the shear band.

Indeed, a magnitude of several hundred percent strains would not be unlikely. This would be

highly unusual for a critical state, which is often reached within about 20% strain. On the other

hand, as mentioned above, the gradient of the critical line for the short specimens is much

204
CHAPTER 5

higher than that normally encountered for natural soils. Moreover, scanning electron

micrograph of short specimens after testing (see Figure 5.19) do not show the high degree of

particle break-up seen in Chin’s (2006) specimens and Figure 4.70.

Given this anomaly, one possible way is to regard the short specimens’ critical state as an

engineering critical state, which is useful for engineering testing and design. At a

microstructural level, one may surmise that, at this critical state, particle break-up is not

necessarily completely and clusters may still remain. However, the destructuration has reached

a point where plastic flow can occur on a macroscopic scale. As this plastic flow occurs,

further destructuration may still continue to take place. On the other hand, Chin’s (2006)

“critical state” may be more representative of an “intrinsic” or truly unstructured state, rather

than critical state. This would then necessarily imply that the critical state is not necessarily an

“intrinsic” or truly unstructured state.

5.3. Constitutive behavior of remoulded cement-treated marine clay

5.3.1. Effects of remoulding on cement-treated marine clay – previous works

As discussed above, the reconstituted or remoulded state of a soil is often regarded as its

destructured state. As mentioned above, the “ultimate state” reached by short and long

specimens appear to be very different. In this section, the behaviour of remoulded

cement-treated marine clay are presented and discussed to assess its relation to the “critical

state” observed from short and long CIU specimens. Some of the properties of remoulded

cement-treated marine clay have been reported by Chin (2006), who noted that, in contrast to

many natural structural soils, it is not easy to define an unstructured or “intrinsic” state for

205
CHAPTER 5

cement-treated marine clay, even with prolonged hand remoulding. This is because the particle

size distribution of the remoulded cement-treated clay becomes finer with continuous

remoulding, and this was observed to persist even after a few hours of remoulding. Chin (2006)

attributed this to the cluster structure of cement-treated clay. In natural cemented soils, the

bonding often develops long after the soil has been sedimented. In other words, the bonding is

overlaid on top of a pre-existing sedimentation structure.

In the case of cement-treated soils, the situation is rather different. Chew et al. (2004)

noted that early bonding in cemented-treated clay is facilitated by the hydration reaction,

which also releases lime into the pore water, thereby changing its pH. This causes flocculation

of the clay particles, resulting in the formation of largely hollow particle clusters. Subsequent

cementation is overlaid on top of this cluster structure.

Based on scanning electron micrograph observations, Chin (2006) concluded that yielding,

and indeed shearing up to the peak strength, appears to be associated with the break-up of

inter-cluster bonds, but not intra-cluster bonds. In other words, the bonds between clusters

were broken up but the clusters themselves were not. The clusters behave as much larger

particles than the marine clay particles, but are nonetheless small enough to allow plastic flow

to take place.

Chin (2006) further reported that break-up of the individual clusters only occurred within

the shear band after a significant amount of strain softening. The degree of cluster break-up

also depends upon the effective confining pressure. At high confining pressure, clusters

break-up into very fine particles, whereas at lower confining pressure, larger fragments remain.

Given this situation, it is not surprising that a large amount of remoulding effort is required to

206
CHAPTER 5

complete break-up the cluster.

Chin (2006) also noted that, even under high degree of cluster break-up, the friction

coefficient of the cement-treated soil remains much higher than that of the untreated marine

clay. This is not surprising. The action of the cement is not merely that of a binder; it actually

changes the mineralogical make-up of the clay. Chew et al. (2004) noted that the lime

produced by the cement consumes virtually all of the kaolinite, but relatively little of the illite.

In return, the cement covers the soil particles with layers of calcium silicate hydrates (CSH)

and calcium aluminate silicate hydrates (CASH), which are the cementitious products; this is

likely to lead to a decrease in the level of surface activity of the particles.

The discussion highlights several difficulties with identifying or using the intrinsic state.

These are

1. A fully destructured state, wherein all clusters are broken up, may not be reached

under normal condition in a triaxial test.

2. Depending upon the effective confining pressure, one may reach different ultimate

states characterized by different degrees of cluster break-up.

Notwithstanding the above, however, many constitutive models for structured soils do

make use of an intrinsic state as an “ultimate state” towards which the state of the soil evolves

with continual shearing. However, given the philosophical difficulties associated with defining

an ultimate state discussed earlier, the approach adopted in herein is to define a reference

“remoulded” state as a state towards which the microstructure of the cement-treated clay will

evolve with continual shearing, but not necessarily the real ultimate state. Indeed, it is

uncertain if there is a unique ultimate state towards which all specimens will transit towards. In

207
CHAPTER 5

this section, the constitutive behavior of the cement-treated clay which has been subjected to a

fixed remoulding effort will be studied under isotropic compression and undrained triaxial. As

described in Chapter 3, the remoulding procedure used in this study involves drying the treated

soil and then pounding and grinding the soil into a powder form. This is different from the wet

remoulding approach used by Chin (2006) and is probably more severe.

5.3.2. Isotropic compression behavior of remoulded cement-treated marine clay

In order to investigate isotropic compression behavior of remoulded cement-treated

marine clay, cement-treated marine clay specimens with 10% to 50% cement content, 100% to

167% total water content and cured under 0 to 250kPa confining pressure for 7 to 28 days were

remoulded by following the procedure described in Chapter 3. It should be noted that the

remoulded treated marine clay was consolidated in a tube under preloading 80N (i.e.

pre-consolidation pressure 70kPa) for about 2 days. Figure 5.20 shows the isotropic

compression curves for remoulded cement-treated marine clay specimens with different

cement contents and 100% total water content cured under atmospheric pressure for 7 days

before remoulding. As can be seen, the isotropic compression index increases with the increase

in cement content, albeit not by a very significant margin. The initial specific volume of the

remoulded specimen also increases with cement content. This can be explained by higher

liquid limit as cement content increases. Chin (2006) and Chew et al. (2004) among others

found that the liquid limit of cemented soil increase with cement content. As such, for the same

amount of remoulded soil solid, the mass of water needed for mixing increases with cement

content. In other words, the remoulded specimens with higher cement content require more

208
CHAPTER 5

water for mixing to get the expected uniform mixture.

Figure 5.21 compares the isotropic compression behaviors of remoulded marine clay,

remoulded cement-treated marine clay and intact cement-treated marine clay. As this Figure

shows, when the cement content is low (e.g. 10%~20%), the compression index λ of the

remoulded treated marine clay is lower than that of untreated marine clay. On the other hand,

when the cement content is high (e.g., 30%~50%), it’s slightly higher than that of untreated

marine clay. For all the four cement contents tested, the isotropic compression index of

remoulded treated marine clay is much lower than that of intact (i.e. not remoulded) treated

marine clay; the trend being more obvious when the cement content is low. The position of the

isotropic compression curve of remoulded treated marine clay lies between those of the

remoulded untreated marine clay and the corresponding intact treated marine clay.

Figure 5.22 shows the isotropic compression curves of remoulded treated marine clay,

which were obtained from cement-treated marine clay with 50% cement content and different

total water contents. As can be seen, all isotropic compression curves appear to converge

together with the increase in the compression pressure. This is consistent with the observation

on the isotropic compression behavior of cement-treated marine clay in Chapter 4, wherein it

was noted that the isotropic compression curves of cement-treated marine clay with 50%

cement content and 100% to 183% total water content tend to converge together when the

isotropic pressure increases due to destructuration. This would suggest that even the remoulded

cement-treated clay still has some residual structure and can undergo further destructuration

with increasing compression pressure.

Figure 5.23 shows the isotropic compression curve of remoulded treated marine clay,

209
CHAPTER 5

which was obtained from cement-treated marine clay with 50% cement content and 133% total

water content cured under 0~250kPa confining stress for 7 days. As can be seen, the isotropic

compression index for all cases seems to be almost the same. This is also consistent with the

observation on the isotropic compression behavior of cement-treated marine clay in Chapter 4.

In Chapter 4, it was pointed out that the isotropic compression curve of cement-treated marine

clay with the same cement content and total water content and different curing stress seems to

converge on the same curve with the increase in the isotropic pressure due to the fact that the

curing stress only induced densification during curing. And the densification may only

contribute to the post-curing void ratio and isotropic compression primary yield stress. As such,

when the isotropic pressure is higher than the isotropic compression primary yield stress, the

isotropic compression follows the same behavior, which indicates that after destructuration, the

isotropic compression behavior should be the same for specimens with different curing stress.

Figure 5.24 demonstrates the isotropic compression behavior of remoulded treated marine

clay, which was obtained from treated marine clay cured for 7 days and 28 days respectively.

As can be seen from Figure 5.24, the isotropic compression index of remoulded treated marine

clay after 7 days curing is generally slightly higher than that with 28 days curing but the

difference is small. This parallels the finding in Chapter 4 that the curing time does not have a

significant effect on the isotropic compression index of the intact treated clay.

Therefore, in general, the isotropic compression curve of the remoulded treated marine

clay depends mainly on the cement content and lies between the compression curves of the

untreated marine clay and the corresponding intact treated marine clay.

210
CHAPTER 5

5.3.3 Triaxial compression behavior

5.3.3.1 Remoulded marine clay

The stress-strain behavior of remoulded untreated marine clay has been studied by Chua

(1990). He noted that remoulded, isotropically reconsolidated Singapore marine clay behaves

in a manner that is reasonably well-described by the modified Cam-clay model with M~0.9. To

check whether the remoulded marine clay used in this study has the same characteristics as that

in Chua’s study, some CIU tests were conducted on remoulded marine clay with

preconsolidation pressure 70kPa.

Figure 5.25 shows the CIU test results of remoulded marine clay specimens. The naming

convention used here is CIUPc'. Pc' is effective confining stress during shearing. For example,

CIU100 involves a remoulded cement-treated marine clay specimen, which was reconsolidated

to 100kPa and then sheared under undrained condition. As Figure 5-25b and c show, the

critical state is reached after about 13% strain. Furthermore, as Figure 5-25a shows, the friction

coefficient M for remoulded marine clay is about 0.88, which agrees well with Chua’s

recommended value of 0.9. Chua (1990) reported an isotropic compression index λ of 0.30,

which also agrees well with the value of 0.231 shown in Figure 5.21. Thus, the properties of

the remoulded marine clay used in this study are very similar to that of Chua (1990). Figure

5-25a also compares the experimental stress paths with the theoretical undrained stress paths of

the modified Cam Clay model. These stress paths were calculated based on the measured

friction coefficient and compression index, with an assumed swelling index κ of 0.05

recommended by Chua (1990). As can be seen, above a re-consolidation pressure of 70kPa, the

fit between the measured stress paths and the modified Cam Clay predictions is remarkably

211
CHAPTER 5

good. This supports Chua’s conclusion that remoulded marine clay behavior is well-described

by the modified Cam Clay.

5.3.3.2 Remoulded cement-treated marine clay with 100% pre-remoulding total

water content

Figures 5.26 to 5.43 illustrate the results of remoulded treated marine clay with 100%

total water content and 10% to 50% cement content. It was found that the trend is basically the

same regardless of cement content. It should be noted that the remoulded treated marine clay

was reconsolidated under 100kPa to 1250kPa before shearing. Preconsolidation pressure effect

and pre-remoulding curing time effect is highlighted here. As mentioned in chapter 3, the

preconsolidation pressure is 70kPa (i.e. preload 80N) and 44kPa (i.e. preload 50N) respectively.

And the pre-remoulding curing period is 7 days and 28 days before the cement-treated soil was

remoulded.

Figure 5.26 shows undrained triaxial behavior of remoulded cement-treated marine clay

specimen with 70 kPa preconsolidation pressure, which was obtained from cement-treated

marine clay with 10% cement content and 100% total water content cured under atmospheric

pressure for 7 days. As can be seen, the undrained stress path is very similar to that of

remoulded untreated marine clay before reaching the critical state line; it is also well-described

by the modified Cam Clay’s undrained stress path with appropriate M, λ and κ values. The

deviator stress-strain curve is much more ductile than that of intact cement-treated marine clay

and quite similar to that of remoulded untreated marine clay. It was also found that the

undrained traxial shearing behavior is dependent upon the confining stress. When the

212
CHAPTER 5

confining stress is low (e.g. lower than 250kPa), the stress-strain curve shows a slight peak at

around 6%~8% axial strain, with only a small amount of post-peak softening. When the

confining stress is high (e.g., higher than 500kPa), the peak value of deviator stress occurs at

around 10%-12% axial strain, after which obvious shear band may be seen, especially at

confining stress 1000kPa. As can be seen later, other cases also indicated that the shear band

usually can be seen at high confining pressure like 1000kPa.

Figure 5.27 shows undrained triaxial behavior of remoulded cement-treated marine clay

specimens with 10% cement content and different preconsolidation pressure (i.e., 70kPa and

44kPa respectively). As can be seen, the peak value of specimens with 70kPa preconsolidation

pressure is usually slightly higher than that with preconsolidation pressure of 44kPa. However,

the difference in the undrained stress path is very small and the friction coefficient M is nearly

the same at about 1.5. This indicates that the preconsolidation pressure effect on undrained

shearing behavior of remoulded treated marine clay is very small when the cement content is

low, say at about 10%.

Figure 5.28 shows undrained triaxial behavior of remoulded cement-treated marine clay

specimens with 10% cement content and 70kPa preconsolidation pressure, with different

pre-remoulding curing period. As can be seen, the peak value of specimens remoulded after 28

days of curing is slightly higher than that remoulded after 7 days of curing. However, the

difference in the undrained stress path is small and the friction coefficient M is nearly the same,

at about 1.5. This indicates that the pre-remoulding curing time has little effect on the

undrained behavior of remoulded treated marine clay when the cement content is low (e.g.,

10%), and for pre-remoulding curing periods between 7days and 28days.

213
CHAPTER 5

Figure 5.29 shows the undrained triaxial behavior of remoulded cement-treated marine

clay specimens with 10% cement content and 44kPa preconsolidation pressure, with different

curing periods. The trend is similar to that shown in Fig. 5.28.

Figures 5.30 to 5.33 and figures 5.34 to 5.37 show the undrained triaxial behavior of

remoulded cement-treated marine clay specimens with 20% and 30% cement content

respectively. As these figures show, the trend is similar to that shown in figures 5.26 to 5.29.

Similarly, Figures 5.30a and 5.34a show that the undrained stress path is also well-described

by the modified Cam Clay’s undrained stress path with appropriate M, λ and κ values. The

results also show that the preconsolidation pressure and the pre-remoulding curing period have

little effect on the undrained behavior of remoulded treated marine clay when the cement

content is between 20% and 30%. However, Figure 5.31 and Figure 5.35 show that the friction

coefficient increases with preconsolidation pressure, with a slightly higher increase for 30%

cement content than 20% cement content. This indicates that the preconsolidation effect on

undrained shearing behavior of remoulded treated marine clay is dependent on cement content.

Figures 5.30 to 5.33 and figures 5.34 to 5.37 show the friction coefficient is 1.61 to 1.65

and 1.67 to 1.73 for 20% and 30% cement content respectively. This indicates that the

introducing cement to remoulded marine clay changed the shearing property of remoulded

marine clay even after destructuration. As such, unlike untreated marine clay, the critical state

line for remoulded cement-treated marine clay is dependent on the cement content introduced

to the marine clay.

Figures 5.38 to 5.39 show undrained triaxial behavior of remoulded cement-treated

marine clay specimen with 70 kPa preconsolidation pressure and 50% cement content. As can

214
CHAPTER 5

be seen, the behaviour is similar to that with 10%-30% cement content. However, as figure

5.39 shows, the critical state line of remoulded cement-treated marine clay with 50% cement

content is higher than that with 10% cement content. The peak strength of specimens with 50%

cement content is higher than that with 10% cement content. However, the pore pressure

stabilized at about the same value for the same confining pressure.

Figure 5.38 also shows the simulated stress path by modified Cam clay model. As can be

seen, the simulated stress path is lower than that of experiment stress path due to the very high

cement content. This is also evident even for specimens with 28 days of pre-remoulding curing

period as shown in Figure 5.40a. However, as Figures 5.40b and 5.40c show, for the specimens

with 44kPa preconsolidation pressure, the simulated stress path is closer to the experimental

results. This indicates that preconsolidation pressure does have effect on the stress path

behavior of remoulded cement-treated marine clay when the cement content is very high (e.g.

50%).

Figure 5.41 shows the preconsolidation pressure effect on stress-strain behavior of

remoulded cement-treated marine clay specimens with 50% cement content. As can be seen,

the peak value of specimens with 70kPa preconsolidation pressure is usually a bit higher than

that with 44kPa preconsolidation pressure, which is the same as that with 10%-30% cement

content. However, the preconsolidation pressure effect on the stress-strain curves is more

evident than that with 10%-30% cement content. The increase in friction coefficient with

preconsolidation pressure is also slightly higher than that with 10%-30% cement content.

Figure 5.42 shows undrained triaxial behavior of remoulded cement-treated marine clay

specimens with 50% cement content and 70kPa preconsolidation pressure, with different

215
CHAPTER 5

pre-remoulding curing periods. As can be seen, the peak strength of specimens remoulded after

28 days of curing is generally a bit higher than that from samples after 7 days of curing.

However, the difference in the undrained stress path is quite small and the friction coefficient

is practically the same, at 1.77 and 1.74 respectively. This indicates that the pre-remoulding

curing time effect on undrained shearing behavior of remoulded treated marine clay is also

quite small even under high cement content of, say 50%.

Similarly, Figure 5.43 shows undrained triaxial behavior of remoulded cement-treated

marine clay specimens with 50% cement content and 44kPa preconsolidation pressure, and

different pre-remoulding curing periods. As can be seen, the trend is the same as that with

70kPa, with a friction coefficient of about 1.70. This indicates that under lower

preconsolidation pressure, the curing period has little effect on the undrained behavior of

remoulded treated marine clay even at high cement content, of say 50%.

5.3.3.3 Remoulded treated marine clay with 50% cement content –effect of

pre-remoulding total water content

Figures 5.44 to 5.46 show the undrained behavior of remoulded treated marine clay with

50% cement content and different pre-remoulding total water contents. As can be seen, the

trend is the same for all cases and similar to those discussed above. The critical state line is

virtually the same regardless of total water content, with a friction coefficient of 1.70. This

indicates that the pre-remoulding total water content does not affect the undrained shearing

behavior and critical state of remoulded cement-treated marine clay, which is consistent with

the isotropic compression results in the previous section. The isotropic compression results

216
CHAPTER 5

also show that the compression line tends to converge with isotropic pressure regardless of

pre-remoulding total water content.

Figures 5.44 show that the undrained stress paths are also well-described by the modified

Cam Clay’s undrained stress paths with appropriate M, λ and κ values, especially for confining

stress between 250kPa and 500kPa.

5.3.3.4 Remoulded treated marine clay with 50% cement content and 133%

pre-remoulding total water content –effects of curing stress

Figures 5.47 to 5.50 show the undrained shearing behavior of remoulded treated marine

clay with 50% cement content and 133% total water content and different pre-remoulding

curing stress. As can be seen, the trend is the same for all cases and similar to those discussed

above. The critical state line is nearly the same regardless of curing stress, with a friction

coefficient of 1.70. This indicates that the pre-remoulding curing stress does not significantly

affect undrained shearing behavior and critical state of remoulded cement-treated marine clay,

which is consistent with the isotropic compression results. The isotropic compression results

also show that the compression lines of remoulded treated marine clay specimens with

different curing stress tend to converge with isotropic pressure. As stated before, the curing

stress only contributes to densification. Therefore, after destructuration, the specimens with

different curing stress show the same behavior of undraiend triaxial behavior.

Consequently, it may be concluded that the undrained triaxial behavior as well as critical

state of remoulded cement-treated marine clay is affected mainly by the cement content.

217
CHAPTER 5

5.3.3.5 Summary of findings

As such, the undrained triaxial behavior of remoulded treated marine clay with different

cement content can be summarized as below.

a) The undrained stress path of remoulded cement-treated marine clay has a similar shape

as those of remoulded marine clay, which is well-described by the modified Cam Clay. The

deviator stress-strain is much ductile than that of intact treated marine clay and similar to that

of remoulded marine clay. The excess pore pressure changes very small after peak point. The

peak value usually occurs at about 6%~8% of axial strain when the confining stress is lower

than 250kPa, after which very small fluctuation can be seen. The peak value usually occurs at

about 10%~12% of axial strain when the confining stress is higher than 500kPa, after which

shear band may be seen.

b) The preconsolidation pressure effect on the undrained triaxial behavior increases with

cement content. When cement content is low (e.g., 10%), the preconsolidation pressure effect

is very small and the friction coefficient is almost the same. For higher cement content, the

difference appears to be more significant.

c) Pre-remoulding curing stress and pre-remoulding total water content does not

significantly affect undrained shearing behavior and critical state of remoulded cement-treated

marine clay.

d) Pre-remoulding curing period has little effect on the undrained triaxial behavior

regardless of cement content and preconsolidation pressure.

e) The M value is between 1.50 and 1.77 for preconsolidation pressure 70kPa while it is

between 1.50 and 1.7 for preconsolidation pressure 44kPa when cement content is between

218
CHAPTER 5

10% and 50%.

5.4. Discussion

As mentioned earlier, the real ultimate state of cement-treated soil is very difficult to

discern with conventional triaxial test because the ultimate state obtained from triaxial test is

dependent on mean effective stress. This may result in a friction coefficient lower than that of

untreated soil as observed in Chin’s (2006) and this study.

In this chapter, three different ultimate states were discussed. Notwithstanding this,

however, one can surmise that a trend actual exists and there appears to be a correlation

between the degree of break-up of the particle clusters and the friction coefficient M. Both

Chin’s (2006) study and this study showed that an ultimate state may be achieved within the

rupture plane. However, the ultimate state line in the p'-q plane is a curve rather than straight

line. This curve may be interpreted as a manifestation of the fact that ultimate secant friction

coefficient M actually decreases with mean effective stress, as illustrated in Figure 5.51.

Moreover, Chin’s (2006) study and this study showed that for CID specimens, particle

break-up within the rupture plane may be very severe under high effective confining pressure,

as seen in specimen 6 of Figure 5.51 and Figure 4.71. On the other hand, Chin (2006) also

showed that at low effective confining stress, the break-up is less severe and aggregates are

still discernible from the SEM, as seen specimens 1 and 2 of Figure 5.51.

SEMs of remoulded cement-treated marine clay specimens prepared in this study show

that, while there is some particle break-up, it does not appear to be as severe as that within the

rupture plane for specimens under high effective confining stress (see Figure 5.52). Finally,

219
CHAPTER 5

SEMs of short CIU specimens after failure show that large aggregates are still discernible.

Table 5-1 shows the values of M and microstructural state after different processes. As can be

seen, the values of M appear to decrease as particle break-up increases in severity.

As Table 5.1 shows, short CIU specimens seem to inflict the lightest damage to the

aggregates. Long CIU specimens under low effective confining stress (inside the rupture plane)

and remoulded, pulverized cement-treated marine clay seem to inflict similar degree of

damage to the particles and the resulting friction coefficient is also similar. Long CID

specimens under high effective confining stress (inside the rupture plane) inflict the heaviest

damage to the aggregates, with massive break-up and low resulting friction coefficient. This

may be due to the large shear strain within the rupture plane coupled within the large effective

normal stress arising from the high effective confining pressure.

Table 5.1 Summary of Friction Coefficient and Microstructure after Different Processes
Specimen Description Microstructure Description M
Short CIU specimens at Large aggregates still 2.6
“critical state” discernible
Long CIU specimens under Aggregates still visible but 1.6-1.88
low effective confining stress there seems to be some
(inside rupture plane) amount of break-up
Remoulded, pulverized Aggregates still visible but 1.5-1.70
cement-treated marine clay there seems to be some
amount of break-up
Long CID specimen under Aggregates appears to be 0.75-0.96
high effective confining stress severely broken up into fine
(inside rupture plane) particles

In the discussion that follows, the fully destructured state will be taken to be that of the

remoulded, pulverized cement-treated marine clay. The “critical state” of short CIU specimens

was not chosen because the friction coefficient is high, indeed almost equal to the peak friction

coefficient of the long specimens under low effective confining stress. The friction coefficient

of the remoulded cement-treated marine clay, on the other hand, is much lower. The state of

220
CHAPTER 5

the material within the rupture of long CID specimens under high confining state is even more

destructured, but this state of the material cannot be achieved through any means of

remoulding or sample preparation, it can only be reached with a rupture plane under high

effective stresses. This makes it unsuitable for use as a reference state. Furthermore, the

ultimate state of remoulded cement treated specimen seems to be only dependent on the

cement content, thereby making it a convenient reference state.

221
CHPATER 5

Fig. 5.1: Undrained triaxial compression of Todi clay: stress-strain response, pore
water pressure difference and strain homogeneity indexes(after Viggiani et al,1993a)

222
CHPATER 5

2-1-4-50CIU
0.8 (q/p')/(q/p')max~εa
q/qmax~εa
1
u/umax~εa

(q/p')/(q/p')max,q/q max,u/u max


(q/p')/(q/p')max,q/q max,u/u max 0.9
2-1-4-50CIU
0.4 (q/p')/(q/p')max~εa
q/qmax~εa 0.8
u/umax~εa

0.7

0 0.6

0 0.01 0.02 0.03 0.04 0.05


axial strain

-0.4

0 0.04 0.08 0.12 0.16 0.2


axial strain

(a)
1 2-1-4-100CIU
2-1-4-100CIU
(q/p')/(q/p')max~εa
(q/p')/(q/p')max~εa
q/qmax~εa
q/qmax~εa
u/umax~εa 1 u/umax~εa
0.8

(q/p')/(q/p')max,q/q max,u/u max


0.9
(q/p')/(q/p')max,q/qmax,u/umax

0.6 0.8

0.7

0.4

0.6

0 0.01 0.02 0.03 0.04


axial strain
0.2

0 0.04 0.08 0.12 0.16 0.2


axial strain

(b)
1

0.8
1
(q/p')/(q/p')max,q/q max,u/u max
(q/p')/(q/p')max,q/qmax,u/umax

0.9

0.6

2-1-4-250CIU 0.8 2-1-4-250CIU


(q/p')/(q/p')max~εa (q/p')/(q/p')max~εa
q/qmax~εa q/qmax~εa
0.4 0.7
u/umax~εa u/umax~εa

0.6

0 0.01 0.02 0.03 0.04 0.05


0.2 axial strain

0 0.05 0.1 0.15 0.2


axial strain
(c)
Figure 5.2: Localization of cement-treated marine clay specimens with mix proportion
2-1-4 in CIU test (a) 50CIU, (b) 100CIU, (c) 250CIU

223
CHPATER 5

2-1-4-500-50CIU
1
(q/p')/(q/p')max~εa
0 q/qmax~εa

(q/p')/(q/p')max,q/q max,u/u max


2-1-4-500-50CIU
0.9 u/umax~εa
(q/p')/(q/p')max~εa
(q/p')/(q/p')max,q/qmax,u/umax

q/qmax~εa
u/umax~εa 0.8

-1

0.7

0.6

-2 0 0.01 0.02 0.03


axial strain

-3

0 0.04 0.08 0.12 0.16 0.2


axial strain

(a)

1
2-1-4-500-250CIU
(q/p')/(q/p')max~εa
q/qmax~εa
u/umax~εa
1
0.8
2-1-4-500-250CIU
(q/p')/(q/p')max,q/qmax,u/u max

0.96 (q/p')/(q/p')max~εa
q/qmax~εa
(q/p')/(q/p')max,q/qmax,u/umax

0.92 u/umax~εa
0.6

0.88

0.84
0.4

0.8

0 0.01 0.02 0.03 0.04 0.05


axial strain
0.2

0 0.04 0.08 0.12 0.16 0.2


axial strain

(b)

Figure 5.3: Localization of cement-treated marine clay specimens (mix proportion


2-1-4) in CIU test with maximum consolidation stress 500kPa (a) 500-50CIU, (b)
500-250CIU

224
CHPATER 5

2-1-4-2000-50CIU
2-1-4-2000-50CIU 1 (q/p')/(q/p')max~εa
(q/p')/(q/p')max,q/qmax,u/umax

(q/p')/(q/p')max~εa q/qmax~εa

(q/p')/(q/p')max,q/qmax,u/u max
-2 q/qmax~εa 0.9
u/umax~εa
u/umax~εa

0.8

-4 0.7

0.6

-6 0.5

0 0.01 0.02 0.03 0.04


axial strain

-8

0 0.02 0.04 0.06 0.08 0.1


axial strain

(a)

1
2-1-4-2000-400CIU
(q/p')/(q/p')max~εa
q/qmax~εa
u/umax~εa
0.8
1 2-1-4-2000-400CIU
(q/p')/(q/p')max~εa
(q/p')/(q/p')max,q/q max,u/umax

q/qmax~εa
(q/p')/(q/p')max,q/qmax,u/umax

0.9 u/umax~εa
0.6

0.8

0.4 0.7

0.6

0.2 0 0.01 0.02 0.03 0.04


axial strain

0 0.02 0.04 0.06


axial strain

(b)

Figure 5.4: Localization of cement-treated marine clay specimens (mix proportion


2-1-4) in CIU test with maximum consolidation stress 2000kPa (a) 2000-50CIU, (b)
2000-400CIU

225
CHPATER 5

1 5-1-6-50CIU
(q/p')/(q/p')max~εa
q/qmax~εa
5-1-6-50CIU
u/umax~εa
(q/p')/(q/p')max~εa
1 q/qmax~εa
0.8
u/umax~εa

(q/p')/(q/p')max,q/qmax,u/u max
0.9
(q/p')/(q/p')max,q/q max,u/umax

0.6
0.8

0.7
0.4

0.6

0 0.01 0.02 0.03


axial strain
0.2

0 0.01 0.02 0.03 0.04 0.05


axial strain

(a)

5-1-6-1000-250CIU
(q/p')/(q/p')max~εa
q/qmax~εa
0.8
u/umax~εa 5-1-6-1000-250CIU
(q/p')/(q/p')max~εa
1
q/qmax~εa
u/umax~εa

(q/p')/(q/p')max,q/qmax,u/u max
(q/p')/(q/p')max,q/qmax,u/umax

0.9

0.4
0.8

0.7

0
0.6

0 0.01 0.02 0.03


axial strain

-0.4

0 0.04 0.08 0.12 0.16


axial strain

(b)

0
5-1-6-2000-50CIU
5-1-6-2000-50CIU 1 (q/p')/(q/p')max~εa
(q/p')/(q/p')max~εa
q/qmax~εa
q/qmax~εa
(q/p')/(q/p')max,q/q max,u/u max

u/umax~εa
(q/p')/(q/p')max,q/q max,u/u max

u/umax~εa 0.9

-2
0.8

0.7

-4 0.6

0 0.01 0.02 0.03


axial strain

-6

0 0.05 0.1 0.15 0.2 0.25


axial strain

(c)
Figure 5.5: Localization of cement-treated marine clay specimens (mix proportion
5-1-6) in CIU test with maximum consolidation stress 50kPa, 1000kPa and 2000kPa
respectively (a) 50CIU, (b) 1000-250CIU, (c) 2000-50CIU

226
CHPATER 5

2000 2000
specimen 1-H/D=2, porous stone,Ls=0
specimen 2-H/D=1, porous stone,Ls=0

1600 1600

1200 1200

q(kPa)
q(kPa)

800 800

400 400
specimen 1-H/D=2, porous stone,Ls=0
specimen 2-H/D=1,porous stone,Ls=0
tension cut-off line

0 0

0 0.05 0.1 εs 0.15 0.2 0.25


0 400 800 1200 1600
p'(kPa)

(a) Stress path (b) Deviator stress-strain curve

1400

1200

1000

800 specimen 1-H/D=2, porous stone,Ls=0


u(kPa)

specimen 2-H/D=1, porous stone,Ls=0

600

400

200

0 0.05 0.1 εs 0.15 0.2 0.25

(c) Excess pore pressure versus axial strain

Figure 5.6: Effect of slenderness ratio (ratio of height to diameter H/D) on specimens
in CIU test (mix proportion 2-1-4, with confining stress 1500kPa, lateral space Ls=0,
sample diameter is the same as that of caps)

(a) Specimen 1-H/D=2, porous stone, Ls=0, (b) Specimen2-H/D=1, porous stone, Ls=0

Figure 5.7: Samples after shearing in CIU test (mix proportion 2-1-4)

227
CHPATER 5

1200 1200

specimen 1-H/D=2, porous stone,Ls=0


specimen 2-H/D=2, teflon,Ls=0

800 800

q(kPa)
q(kPa)

400 400

specimen 1-H/D=2, porous stone,Ls=0


specimen 2-H/D=2, teflon,Ls=0
tension cut-off line

0 0

0 0.05 0.1 εs 0.15 0.2 0.25


0 200 p'(kPa) 400 600 800 1000

(a) Stress path (b) Deviator stress-strain curve


1000

800

600
u(kPa)

specimen 1-H/D=2, porous stone,Ls=0


specimen 2-H/D=2, teflon,Ls=0
400

200

0 0.05 0.1 εs 0.15 0.2 0.25

(c) Excess pore pressure versus axial strain

Figure 5.8: Effect of end restraint (lubrication) on specimens in CIU test (mix
proportion 2-1-4, with confining stress 1000kPa, Ls=0, sample diameter is the same as
that of caps), slenderness ratio = 2.

(a) Specimen 1-H/D=2, porous stone, Ls=0, (b) Specimen 2-H/D=2, Teflon, Ls=0

Figure 5.9: Samples after shearing in CIU test (mix proportion 2-1-4)

228
CHPATER 5

1200 1200

800
q(kPa) 800

q(kPa)
400 400

specimen 1-H/D=2, porous stone,Ls=0 specimen 1-H/D=2, porous stone,Ls=0


specimen 2-H/D=1, teflon,Ls=0 specimen 2-H/D=1, teflon,Ls=0
specimen 3-H/D=1, teflon,Ls=6mm specimen 3-H/D=1, teflon,Ls=6mm
tension cut-off line

0 0

0 200 400 p'(kPa) 600 800 1000 0 0.04 0.08 εs 0.12 0.16 0.2

(a) Stress path (b) Deviator stress-strain curve

1000

800

600
u(kPa)

400
specimen 1-H/D=2, porous stone,Ls=0
specimen 2-H/D=1, teflon,Ls=0
specimen 3-H/D=1, teflon,Ls=6mm
200

0
εs
0 0.04 0.08 0.12 0.16 0.2

(c) Excess pore pressure versus axial strain

Figure 5.10: Effect of slenderness and end restraint on specimens in CIU test (mix
proportion 2-1-4, with confining stress 1000kPa)

(a) (b) (c)

Figure 5.11: Samples after shearing in CIU test (a) Specimen H/D=1, porous stone,
Ls=0, (b) Specimen H/D=1, Teflon, Ls=0 (c) Specimen -H/D=1, Teflon, Ls=6mm

229
CHPATER 5

specimen 1-H/D=2, porous stone,Ls=0


specimen 2-H/D=1, teflon,Ls=6mm
specimen 3-H/D=2, porous stone,Ls=0
specimen 4-H/D=1, teflon,Ls=6mm
tension cut-off line specimen 1-H/D=2, porous stone,Ls=0
1200 1200 specimen 2-H/D=1, teflon,Ls=6mm
specimen 3-H/D=2, porous stone,Ls=0
specimen 4-H/D=1, teflon,Ls=6mm

800 800
q(kPa)

q(kPa)
400 400

0 0
εs
0 200 400 p'(kPa) 600 800 1000 0 0.04 0.08 0.12 0.16 0.2

(a) Stress path (b) deviator stress-strain curve

1000

800

600
u(kPa)

400

specimen 1-H/D=2, porous stone,Ls=0


200
specimen 2-H/D=1, teflon,Ls=6mm
specimen 3-H/D=2, porous stone,Ls=0
specimen 4-H/D=1, teflon,Ls=6mm
0
εs
0 0.04 0.08 0.12 0.16 0.2

(c) Excess pore pressure versus axial strain

Figure 5.12: Effect of slenderness, end restraint and lateral space on specimens in CIU
test (mix proportion 2-1-4, specimen 1 and 2 with confining stress 1000kPa, specimen
3 and 4 with confining stress 500kPa)

230
CHPATER 5

specimen 1-H/D=2, porous stone,Ls=0


specimen 2-H/D=1, teflon,Ls=6mm
specimen 3-H/D=1, teflon,Ls=6mm, very large strain
tension cut-off line
1200

1000

800

q(kPa)

600

400

200

0 200 400 p'(kPa) 600 800 1000 1200

(a) Stress path

1200 specimen 1-H/D=2, porous stone,Ls=0


specimen 2-H/D=1, teflon,Ls=6mm
specimen 3-H/D=1, teflon,Ls=6mm,very large strain

800
q(kPa)

400

0 0.2 εs 0.4 0.6

(b) Deviator stress-strain curve

1000

800

600
u(kPa)

specimen 1-H/D=2, porous stone,Ls=0


400 specimen 2-H/D=1, teflon,Ls=6mm
specimen 3-H/D=1, teflon,Ls=6mm,very large strain

200

0
εs
0 0.2 0.4 0.6

(c) Excess pore pressure versus axial strain


Figure 5.13: Post-peak stress statuses at different range of axial strain in CIU test (mix
proportion 2-1-4, with confining stress 1000kPa)

231
CHPATER 5

specimen 1-H/D=1, teflon,Ls=6mm, specimen diameter 38mm


specimen 2-H/D=1, teflon,Ls=9mm, specimen diameter 35mm
tension cut-off line
1200

1000

800

q(kPa)
600

400

200

0 200 400 600 p'(kPa)800 1000 1200

(a) Stress path

1200

800

specimen 1-H/D=1, teflon,Ls=6mm, specimen diameter 38mm


q(kPa)

specimen 2-H/D=1, teflon,Ls=9mm, specimen diameter 35mm

400

0 0.2 εs 0.4 0.6

(b) Deviator stress-strain curve

1000

800

600 specimen 1-H/D=1, teflon,Ls=6mm, specimen diameter 38mm


specimen 2-H/D=1, teflon,Ls=9mm, specimen diameter 35mm
u(kPa)

400

200

0
εs
0 0.2 0.4 0.6

(c) Excess pore pressure versus axial strain


Figure 5.14: Post-peak stress statuses at very larger axial strain in CIU test (mix
proportion 2-1-4, with confining stress 1000kPa)

232
CHPATER 5

(a) (b)
Figure 5.15: Short samples with enlarged low friction platens after shearing to axial
strain of about 55% in CIU test (a) confining stress 1000kPa (b) confining stress
500kPa
3000 mix proportion 2-1-4
CID250
specimen 1-H/D=1, teflon,Ls=10mm
specimen 2-H/D=2, poroston,Ls=0mm

2000
q(kPa)

1000

0 0.2 0.4 εs 0.6 0.8

(a) Deviator stress-strain curve


p'(kPa)
200 300 400 500 600 700 800 900 1000 1100

mix propotion 2-1-4


CID 250
specimen 1-H/D=1, teflon,Ls=10mm
3.6 specimen 2-H/D=2, poroston,Ls=0mm

3.2
v

2.8

2.4
(b) Compression space
Figure 5.16: Comparison between conventional specimen and short specimen in CID
test (mix proportion 2-1-4, effective confining stress 250kPa)

233
CHPATER 5

mix proportion 2-1-4


CID500
4000 specimen 1-H/D=1, teflon,Ls=10mm
specimen 2-H/D=2, poroston,Ls=0mm

3000
q(kPa)

2000

1000

0 0.2 0.4 εs 0.6 0.8

(a) Deviator stress-strain curve

p'(kPa)
500 600 700 800 900 1000 1100 1200 1300 1400 1500 1600 1700 1800

mix propotion 2-1-4


CID 500
3.6 specimen 1-H/D=1, teflon,Ls=10mm
specimen 2-H/D=2, poroston,Ls=0mm

3.2
v

2.8

2.4

(b) Compression space

Figure 5.17: Comparison between conventional specimen and short specimen in CID
test (mix proportion 2-1-4, effective confining stress 500kPa)

234
CHPATER 5

short CIU specimen with mix proprotion 2-1-4


conventional specimen with mix proportion 5-1-6
2000 conventional specimen with mix proportion 2-1-4
critical state line for short specimen
critical state line for conventional specimen

1600

1200
q(kPa)

800

400

0 200 400 600 800 1000


p'(kPa)

Figure 5.18: Critical state of conventional specimen and short specimen

(a) Microscopy within multiple shear bands

235
CHPATER 5

(b) Microscopy outside of multiple shear bands (without remoulding)

(c) Microscopy outside of multiple shear bands (after remoulding)

Figure 5.19: Microstructure of cement treated marine clay short specimens with mix
proportion 2:1:4 under effective confining pressure 500kPa and drained shearing
(500CID)

236
CHPATER 5
10 100 p'(kPa)
1000 10000

Aw=50%
Aw=30%
3.2 Aw=20%
Aw=10%
Aw=50%-v=4.74-0.285Ln(p')
Aw=30%-v=4.19-0.25Ln(p')
Aw=20%-v=3.885-0.23Ln(p')
Aw=10%-v=3.52-0.20Ln(p')
2.8
specific volume V

2.4

Figure 5.20: Isotropic compression curves for remoulded cement-treated marine clay
specimens with different cement content and 100% total water content cured under
atmospheric pressure for 7 days before remoulding
10 100 p'(kPa)
1000 10000
3.6

3.2

2.8
specific volume V

2.4

1.6

Aw=10%-unremoulded-v=6.44-0.54Ln(p')
Aw=20%-unremoulded-v=6.52-0.52Ln(p')
Aw=30%-unremoulded-v=6.80-0.53Ln(p')
Aw=50%-unremoulded-v=6.715-0.50Ln(p')
remoulded marine clay- v=3.23-0.231Ln(p')
Aw=10%-remoulded-v=3.52-0.20Ln(p')
Aw=20%-remoulded-v=3.885-0.23Ln(p')
Aw=30%-remoulded-v=4.19-0.25Ln(p')
Aw=50%-remoulded-v=4.74-0.285Ln(p')

Figure 5.21: Isotropic compression curves for remoulded and unremoulded


cement-treated marine clay specimens and remoulded marine clay

237
CHPATER 5

10 100 p'(kPa)
1000 10000
3.4
Cw=100%
Cw=133%
Cw=167%
3.2 Cw=100%-v=4.74-0.285Ln(p')
Cw=133%-v=4.85-0.30Ln(p')
Cw=167%-v=4.86-0.30Ln(p')

3
specific volume V

2.8

2.6

2.4

Figure 5.22: Isotropic compression curves for remoulded cement-treated marine clay
specimens with different total water content and 50% cement content cured under
atmospheric pressure for 7 days before remoulding

10 100 p'(kPa) 1000 10000

p'cur=0kPa
p'cur=50kPa
p'cur=100kPa
3.2
p'cur=250kPa
p'cur=0kPa-v=4.85-0.30Ln(p')
p'cur=50kPa-v=4.78-0.31Ln(p')
p'cur=100kPa-v=4.71-0.30Ln(p')
p'cur=250kPa-v=4.715-0.30Ln(p')
specific volume v

2.8

2.4

Figure 5.23: Isotropic compression curves for remoulded cement-treated marine clay
specimens with different curing stress and 50% cement content and 133% total water
content before remoulding

238
CHPATER 5

10 100 p'(kPa)
1000 10000
3.2
Aw=50%-7ds
Aw=30%-7ds
Aw=20%-7ds
Aw=10%-7ds
Aw=50%-v=4.74-0.285Ln(p')
Aw=30%-v=4.19-0.25Ln(p')
2.8 Aw=20%-v=3.885-0.23Ln(p')
Aw=10%-v=3.52-0.20Ln(p')
Aw=50%-28ds
Aw=30%-28ds
specific volume V

Aw=20%-28ds
Aw=10%-28ds
Aw=50%-v=4.60-0.28Ln(p')
Aw=30%-v=4.12-0.24Ln(p')
2.4 Aw=20%-v=3.855-0.218Ln(p')
Aw=10%-v=3.62-0.22Ln(p')

Figure 5.24: Isotropic compression curves for remoulded cement-treated marine clay
specimens with different cement content and curing time before remoulding

239
CHPATER 5

700 remoulded marine clay


CIU100-experiment
CIU250-experiment
600 CIU500-experiment
CIU1000-experiment
M=0.88
CIU100-simulated
500
CIU250-simulated

deviator stress q(kPa)


CIU500-simulated
CIU1000-simulated
400

300

200

100

0
0 100 200 300 400 500 600 700 800 900 1000
mean normal effective stressP'(kPa)

(a) Simulated stress path (by modified Cam clay model) and experimental stress path
600

500
remoulded marine clay CIU TEST
CIU100
400 CIU250
deviator stress q(kPa)

CIU500
CIU1000
300

200

100

0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14
axial strain εs(%)

(b) Deviator stress-strain curve


600

500
remoulded marine clay CIU TEST
CIU100
CIU250
400
pore pressure u(kPa)

CIU500
CIU1000

300

200

100

0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14
axial strain εs(%)

(c) Pore pressure versus axial strain


Figure 5.25: Undrained triaxial shearing behaviors of remoulded marine clay
specimens with 70 kPa preconsolidation pressure

240
CHPATER 5
Aw=10% CIU TEST
CIU100-experiment
CIU250-experiment
*CIU500-experiment
900 *CIU750-experiment
*CIU1000-experiment

800 M=1.50
CIU100-simulated
CIU250-simulated
700 CIU500-simulated
CIU750-simulated
600 CIU1000-simulated

deviator stress q(kPa)


500

400

300

200

100

0
0 100 200 300 400 500 600 700 800 900 1000

(a) Experiment and simulated stress path by MCC (M=1.5, λ =0.20, κ =0.024)
mean normal effective stressP'(kPa)

Aw=10% CIU TEST


900
CIU100
CIU250
800 *CIU500
*CIU750
700 *CIU1000

600
deviator stress q(kPa)

500

400

300

200

100

0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14
axial strain εs(%)

(b) Deviator stress-strain curve


Aw=10% CIU TEST
800 CIU100
CIU250
*CIU500
700
*CIU750
*CIU1000
600
pore pressure u(kPa)

500

400

300

200

100

0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14
axial strain εs(%)

(c) Pore pressure versus axial strain


Figure 5.26: Undrained triaxial shearing behavior of remoulded cement-treated marine
clay specimens with 10% cement content and 70 kPa preconsolidation pressure
(*indicates shear band was formed after peak value at about 12% axial strain)

241
CHPATER 5
Aw=10% CIU TEST
1000 P=70kPa,CIU250
P=70kPa,*CIU500
900 P=70kPa,*CIU1000
P=44kPa,CIU250
800 P=44kPa,CIU500
P=44kPa,*CIU1000
M=1.50
700
M=1.52

deviator stress q(kPa)


600

500

400

300

200

100

0
0 100 200 300 400 500 600 700 800 900 1000
mean normal effective stressP'(kPa)

(a) Stress path of specimen with preconsolidation pressure 70kPa and 44kPa
respectively 900

Aw=10% CIU TEST


800
P=70kPa,CIU250
P=70kPa,*CIU500
700 P=70kPa,*CIU1000
P=44kPa,CIU250
600 P=44kPa,CIU500
deviator stress q(kPa)

P=44kPa,*CIU1000
500

400

300

200

100

0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14
axial strain εs(%)
(b) Deviator stress-strain curve
800

700

Aw=10% CIU TEST


600 P=70kPa,CIU250
P=70kPa,*CIU500
pore pressure u(kPa)

P=70kPa,*CIU1000
500
P=44kPa,CIU250
P=44kPa,CIU500
400 P=44kPa,*CIU1000

300

200

100

0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14
axial strain εs(%)

(c) Pore pressure versus axial strain


Figure 5.27: Undrained triaxial shearing behavior of remoulded cement-treated marine
clay specimens with 10% cement content and different preconsolidation pressure (*
indicates shear band was formed after peak value at about 12% axial strain)

242
CHPATER 5

Aw=10%, P=70kPa, CIU TEST


CIU250-7ds
900
*CIU500-7ds
*CIU1000-7ds
800 CIU250-28ds
CIU500-28ds
700 *CIU1000-28ds
M=1.50

600

deviator stress q(kPa)


500

400

300

200

100

0
0 100 200 300 400 500 600 700 800 900 1000 1100
mean normal effective stressP'(kPa)
(a) Stress path of specimen with curing period 7 days and 28 days respectively
900

800 Aw=10%,P=70kPa, CIU TEST


CIU250-7ds
*CIU500-7ds
700
*CIU1000-7ds
CIU250-28ds
600
deviator stress q(kPa)

CIU500-28ds
*CIU1000-28ds
500

400

300

200

100

0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14
axial strain εs(%)

(b) Deviator stress-strain curve


800

700

Aw=10%,P=70kPa, CIU TEST


600 CIU250-7ds
*CIU500-7ds
pore pressure u(kPa)

*CIU1000-7ds
500
CIU250-28ds
CIU500-28ds
400 *CIU1000-28ds

300

200

100

0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14
axial strain εs(%)

(c) Pore pressure versus axial strain


Figure 5.28: Undrained triaxial shearing behavior of remoulded cement-treated marine
clay specimens with 10% cement content and 70kPa preconsolidation pressure and
different curing period (* indicates shear band was formed after peak value at about
12% axial strain)

243
CHPATER 5
Aw=10%, P=44kPa, CIU TEST
CIU250-7ds
1000 CIU500-7ds
*CIU1000-7ds
900 CIU250-28ds
CIU500-28ds
800 CIU750-28ds
M=1.52
M=1.50
700

deviator stress q(kPa)


600

500

400

300

200

100

0
0 100 200 300 400 500 600 700 800 900 1000
mean normal effective stressP'(kPa)

(a) Stress path of specimen with curing period 7 days and 28 days respectively
900 Aw=10%, P=44kPa,CIU TEST
CIU250-7ds
CIU500-7ds
800
*CIU1000-7ds
CIU250-28ds
700
CIU500-28ds
CIU750-28ds
600
deviator stress q(kPa)

500

400

300

200

100

0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14
axial strain εs(%)

(b) Deviator stress-strain curve Aw=10%, P=44kPa,CIU TEST


CIU250-7ds
800
CIU500-7ds
*CIU1000-7ds

700 CIU250-28ds
CIU500-28ds
CIU750-28ds
600
pore pressure u(kPa)

500

400

300

200

100

0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14
axial strain εs(%)

(c) Pore pressure versus axial strain


Figure 5.29: Undrained triaxial shearing behavior of remoulded cement-treated marine
clay specimens with 10% cement content and 44kPa preconsolidation pressure and
different curing period (* indicates shear band was formed after peak value at about
12% axial strain)

244
CHPATER 5
Aw=20% CIU TEST
CIU100-experiment
CIU250-experiment
CIU500-experiment
*CIU750-experiment
*CIU1000-experiment
1100 M=1.65
CIU100-simulated
1000 CIU250-simulated
CIU500-simulated
900 CIU750-simulated
CIU1000-simulated
800

deviator stress q(kPa)


700

600

500

400

300

200

100

0
0 100 200 300 400 500 600 700 800 900 1000 1100
mean normal effective stressP'(kPa)

(a) Experiment and simulated stress path by MCC (M=1.65, λ =0.23, κ =0.018)

Aw=20% CIU TEST


CIU100
1100 CIU250
CIU500
1000 *CIU750
*CIU1000
900

800
deviator stress q(kPa)

700

600

500

400

300

200

100

0
0 0.02 0.04 0.06 0.08 0.1 0.12
axial strain εs(%)
(b) Deviator stress-strain curve
800

700 Aw=20% CIU TEST


CIU100
CIU250
600 CIU500
*CIU750
*CIU1000
pore pressure u(kPa)

500

400

300

200

100

0
0 0.02 0.04 0.06 0.08 0.1 0.12
axial strain εs(%)

(c) Pore pressure versus axial strain


Figure 5.30: Undrained triaxial shearing behavior of remoulded cement-treated marine
clay specimens with 20% cement content and 70 kPa preconsolidation pressure
(*indicates shear band was formed after peak value or about 12% axial strain)

245
CHPATER 5

Aw=20% CIU TEST


1100 P=70kPa,CIU250
P=70kPa,CIU500
1000 P=70kPa,*CIU1000
P=44kPa,CIU225
900 P=44kPa,CIU500
P=44kPa,CIU1000
800 M=1.65
M=1.62

deviator stress q(kPa)


700

600

500

400

300

200

100

0
0 100 200 300 400 500 600 700 800 900 1000 1100
mean normal effective stressP'(kPa)

(a) Stress path of specimen with preconsolidation pressure 70kPa and 44kPa
respectively 1100

1000

900 Aw=20% CIU TEST


P=70kPa,CIU250
800 P=70kPa,CIU500
P=70kPa,*CIU1000
deviator stress q(kPa)

700
P=44kPa,CIU225
P=44kPa,CIU500
600
P=44kPa,*CIU1000

500

400

300

200

100

0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14
axial strain εs(%)

(b) Deviator stress-strain curve


800

700
Aw=20% CIU TEST
P=70kPa,CIU250
600 P=70kPa,CIU500
P=70kPa,*CIU1000
pore pressure u(kPa)

500 P=44kPa,CIU225
P=44kPa,CIU500
P=44kPa,*CIU1000
400

300

200

100

0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14
axial strain εs(%)
(c) Pore pressure versus axial strain
Figure 5.31: Undrained triaxial shearing behavior of remoulded cement-treated marine
clay specimens with 20% cement content and different preconsolidation pressure (*
indicates shear band was formed after peak value at about 12% axial strain)

246
CHPATER 5
Aw=20%, P=70kPa, CIU TEST
1000 CIU100-7ds
CIU500-7ds
900 *CIU750-7ds
CIU100-28ds
800 CIU500-28ds
CIU750-28ds
M=1.65
700
M=1.61

deviator stress q(kPa)


600

500

400

300

200

100

0
0 100 200 300 400 500 600 700 800
mean normal effective stressP'(kPa)

(a) Stress path of specimen with pre-remoulding curing period 7 days and 28 days
respectively 900

800

700

600
deviator stress q(kPa)

500
Aw=20%,P=70kPa, CIU TEST
CIU100-7ds
400 CIU500-7ds
*CIU750-7ds
300 CIU100-28ds
CIU500-28ds
200 CIU750-28ds

100

0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14
axial strain εs(%)
(b) Deviator stress-strain curve
600

500

400
pore pressure u(kPa)

300 Aw=20%,P=70kPa, CIU TEST


CIU100-7ds
CIU500-7ds
200 *CIU750-7ds
CIU100-28ds
CIU500-28ds
CIU750-28ds
100

0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14
axial strain εs(%)
(c) Pore pressure versus axial strain
Figure 5.32: Undrained triaxial shearing behavior of remoulded cement-treated marine
clay specimens with 20% cement content and 70kPa preconsolidation pressure and
different curing periods (* indicates shear band was formed after peak value at about
12% axial strain)

247
CHPATER 5
Aw=20%, P=44kPa, CIU TEST
CIU225-7ds
1100 CIU500-7ds
*CIU1000-7ds
1000 CIU250-28ds
CIU500-28ds
900 *CIU1000-28ds
M=1.62
800 M=1.61

deviator stress q(kPa)


700

600

500

400

300

200

100

0
0 100 200 300 400 500 600 700 800 900 1000 1100
mean normal effective stressP'(kPa)

(a) Stress path of specimen with pre-remoulding curing period 7 days and 28 days
respectively 1100

1000

900 Aw=20%, P=44kPa,CIU TEST


CIU225-7ds
800 CIU500-7ds
*CIU1000-7ds
deviator stress q(kPa)

700
CIU250-28ds
CIU500-28ds
600
*CIU1000-28ds

500

400

300

200

100

0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14
axial strain εs(%)

(b) Deviator stress-strain curve


800

700

Aw=20%, P=44kPa,CIU TEST


600 CIU225-7ds
CIU500-7ds
*CIU1000-7ds
deviator stress q(kPa)

500 CIU250-28ds
CIU500-28ds
*CIU1000-28ds
400

300

200

100

0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14
axial strain εs(%)
(c) Pore pressure versus axial strain
Figure 5.33: Undrained triaxial shearing behavior of remoulded cement-treated marine
clay specimens with 20% cement content and 44kPa preconsolidation pressure and
different curing periods (* indicates shear band was formed after peak value at about
12% axial strain)

248
CHPATER 5

Aw=30% CIU TEST


CIU250-experiment
1300 CIU500-experiment
*CIU750-experiment
1200
*CIU1000-experiment

1100 M=1.73
CIU250-simulated
1000 CIU500-simulated
CIU750-simulated
900 CIU1000-simulated

deviator stress q(kPa)


800

700

600

500

400

300

200

100

0
0 100 200 300 400 500 600 700 800 900 1000 1100

(a) Experiment and simulated stress path by MCC (M=1.73, λ =0.25, κ =0.016)
mean normal effective stressP'(kPa)

Aw=30% CIU TEST


CIU250
1200
CIU500
*CIU750
1100
*CIU1000

1000

900

800
deviator stress q(kPa)

700

600

500

400

300

200

100

0
0 0.02 0.04 0.06 0.08 0.1 0.12
axial strain εs(%)

(b) Deviator stress-strain curve


800

700

Aw=30% CIU TEST


600 CIU250
CIU500
*CIU750
*CIU1000
pore pressure u(kPa)

500

400

300

200

100

0
0 0.02 0.04 0.06 0.08 0.1 0.12
axial strain εs(%)

(c) Pore pressure versus axial strain


Figure 5.34: Undrained triaxial shearing behavior of remoulded cement-treated marine
clay specimens with 30% cement content and 70 kPa preconsolidation pressure
(*indicates shear band was formed after peak value at about 12% axial strain)

249
CHPATER 5

1300 Aw=30% CIU TEST


P=70kPa,CIU250
1200 P=70kPa,CIU500
P=70kPa,*CIU1000
1100 P=44kPa,CIU215
P=44kPa,CIU500
1000
P=44kPa,*CIU1000
900 M=1.73

deviator stress q(kPa)


M=1.67
800

700

600

500

400

300

200

100

0
0 100 200 300 400 500 600 700 800 900 1000 1100
mean normal effective stressP'(kPa)
(a) Stress path of specimen with preconsolidation pressure 70kPa and 44kPa
respectively 1200

1100

1000 Aw=30% CIU TEST


P=70kPa,CIU250
900 P=70kPa,CIU500
P=70kPa,*CIU1000
800
deviator stress q(kPa)

P=44kPa,CIU215

700 P=44kPa,CIU500
P=44kPa,*CIU1000
600

500

400

300

200

100

0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14
axial strain εs(%)

(b) Deviator stress-strain curve


800

700
Aw=30% CIU TEST
600 P=70kPa,CIU250
P=70kPa,CIU500
P=70kPa,*CIU1000
pore pressure u(kPa)

500 P=44kPa,CIU215
P=44kPa,CIU500
400 P=44kPa,*CIU1000

300

200

100

0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14
axial strain εs(%)

(c) Pore pressure versus axial strain


Figure 5.35: Undrained triaxial shearing behavior of remoulded cement-treated marine
clay specimens with 30% cement content and different preconsolidation pressure (*
indicates shear band was formed after peak value at about 12% axial strain)

250
CHPATER 5

Aw=30%, P=70kPa, CIU TEST


CIU250-7ds
1300 CIU500-7ds
*CIU750-7ds
1200
*CIU1000-7ds
1100 CIU250-28ds
CIU500-28ds
1000 CIU750-28ds
CIU1000-28ds
900
M=1.73

deviator stress q(kPa)


800 M=1.71

700

600

500

400

300

200

100

0
0 100 200 300 400 500 600 700 800 900 1000 1100
mean normal effective stressP'(kPa)

(a) Stress path of specimen with pre-remoulding curing period 7 days and 28 days
respectively 1200 CIU250-7ds
Aw=30%,P=70kPa, CIU TEST

CIU500-7ds
1100 *CIU750-7ds
*CIU1000-7ds
1000
CIU250-28ds
900 CIU500-28ds
*CIU750-28ds
800
deviator stress q(kPa)

*CIU1000-28ds

700

600

500

400

300

200

100

0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14
axial strain εs(%)
(b) Deviator stress-strain curve
Aw=30%,P=70kPa, CIU TEST
800
CIU250-7ds
CIU500-7ds
700 *CIU750-7ds
*CIU1000-7ds

600
CIU250-28ds
CIU500-28ds
*CIU750-28ds
pore pressure u(kPa)

500 *CIU1000-28ds

400

300

200

100

0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14
axial strain εs(%)

(c) Pore pressure versus axial strain


Figure 5.36: Undrained triaxial shearing behavior of remoulded cement-treated marine
clay specimens with 30% cement content and 70kPa preconsolidation pressure and
different curing periods (* indicates shear band was formed after peak value at about
12% axial strain)

251
CHPATER 5

1200 Aw=30%, P=44kPa, CIU TEST


CIU215-7ds
1100 CIU500-7ds
*CIU1000-7ds
1000 CIU250-28ds
CIU500-28ds
900 *CIU1000-28ds
M=1.67
800

deviator stress q(kPa)


M=1.67
700

600

500

400

300

200

100

0
0 100 200 300 400 500 600 700 800 900 1000 1100
mean normal effective stressP'(kPa)

(a) Stress path of specimen with pre-remoulding curing period 7 days and 28 days
1100
respectively
1000 Aw=30%, P=44kPa,CIU TEST
CIU215-7ds
900 CIU500-7ds
*CIU1000-7ds
800
CIU250-28ds
deviator stress q(kPa)

CIU500-28ds
700
*CIU1000-28ds

600

500

400

300

200

100

0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14
axial strain εs(%)

(b) Deviator stress-strain curve


800

700

Aw=30%, P=44kPa,CIU TEST


600 CIU215-7ds
CIU500-7ds
deviator stress q(kPa)

*CIU1000-7ds
500
CIU250-28ds
CIU500-28ds
400 *CIU1000-28ds

300

200

100

0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14
axial strain εs(%)

(c) Pore pressure versus axial strain


Figure 5.37: Undrained triaxial shearing behavior of remoulded cement-treated marine
clay specimens with 30% cement content and 44kPa preconsolidation pressure and
different curing periods (* indicates shear band was formed after peak value at about
12% axial strain)

252
CHPATER 5

Aw=50% CIU TEST


1400 CIU100-experiment
1300 CIU250-experiment
CIU500-experiment
1200 *CIU750-experiment
*CIU1000-experiment
1100
M=1.77
1000 CIU250-simulated
CIU500-simulated

deviator stress q(kPa)


900
CIU750-simulated
800 CIU1000-simulated

700

600

500

400

300

200

100

0
0 100 200 300 400 500 600 700 800 900 1000 1100
mean normal effective stressP'(kPa)

(a) Experiment and simulated stress path by MCC (M=1.77, λ =0.29, κ =0.015)
1300

1200

1100
Aw=50% CIU TEST
1000 CIU100
CIU150
900 CIU200
CIU500
deviator stress q(kPa)

800 *CIU750
*CIU1000
700

600

500

400

300

200

100

0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14
axial strain εs(%)

(b) Deviator stress-strain curve


800

700

600
Aw=50% CIU TEST
CIU100
CIU150
pore pressure u(kPa)

500
CIU200
CIU500
*CIU750
400
*CIU1000

300

200

100

0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14
axial strain εs(%)

(c) Pore pressure versus axial strain


Figure 5.38: Undrained triaxial shearing behavior of remoulded cement-treated marine
clay specimens with 50% cement content and 70 kPa preconsolidation pressure
(*indicates shear band was formed after peak value at about 12% axial strain)

253
CHPATER 5
Aw=10% CIU TEST
CIU100-10%
CIU250-10%
1400 *CIU500-10%
*CIU750-10%
1300 *CIU1000-10%

1200 M=1.50
CIU100-50%
1100 CIU250-50%
CIU500-50%
1000
CIU750-50%

deviator stress q(kPa)


900 CIU1000-50%
M=1.77
800

700

600

500

400

300

200

100

0
0 100 200 300 400 500 600 700 800 900 1000 1100
mean normal effective stressP'(kPa)

(a) Stress path of specimen with 10% and 50% cement content respectively
1300 Aw=10% CIU TEST
CIU100-10%
1200 CIU250-10%
*CIU500-10%
1100
*CIU750-10%
1000 *CIU1000-10%

900 CIU100-50%
CIU250-50%
deviator stress q(kPa)

800 CIU500-50%
700 *CIU750-50%
*CIU1000-50%
600

500

400

300

200

100

0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14

axial strain εs(%)

(b) Deviator stress-strain curve Aw=10% CIU TEST


800 CIU100-10%
CIU250-10%
*CIU500-10%
700
*CIU750-10%
*CIU1000-10%
600 CIU100-50%
CIU250-50%
pore pressure u(kPa)

CIU500-50%
500
*CIU750-50%
*CIU1000-50%
400

300

200

100

0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14
axial strain εs(%)

(c) Pore pressure versus axial strain


Figure 539:: Undrained triaxial shearing behavior of remoulded cement-treated marine
clay specimens with different cement content and 70 kPa preconsolidation pressure
(*indicates shear band was formed after peak value at about 12% axial strain)

254
CHPATER 5
Aw=50%, P=70kPa, 28days, CIU TEST
CIU250-experiment
CIU500-experiment
CIU750-experiment
1400 *CIU1000-experiment
CIU250-simulated
1300
CIU500-simulated
1200 *CIU750-simulated
*CIU1000-simulated
1100
M=1.74
1000

deviator stress q(kPa)


900

800

700

600

500

400

300

200

100

0
0 100 200 300 400 500 600 700 800 900 1000 1100
mean normal effective stressP'(kPa)

(a) With preconsolidation pressure 70kPa and curing period 28days (M=1.74, λ =0.29,
κ =0.015) Aw=50%, P=44kPa, 7days, CIU TEST
1200
CIU250-experiment
1100 CIU500-experiment
*CIU1000-experiment
1000 CIU250-simulated
CIU500-simulated
900 CIU1000-simulated
M=1.70
800
deviator stress q(kPa)

700

600

500

400

300

200

100

0
0 100 200 300 400 500 600 700 800 900 1000
mean normal effective stressP'(kPa)
(b) With preconsolidation pressure 44kPa and curing period 7days (M=1.70, λ =0.29,
κ =0.015)
Aw=50%, P=44kPa, 28days, CIU TEST
1200 CIU250-experiment
CIU500-experiment
1100
CIU1000-experiment
CIU250-simulated
1000
CIU500-simulated
900 *CIU1000-simulated
M=1.70
800
deviator stress q(kPa)

700

600

500

400

300

200

100

0
0 100 200 300 400 500 600 700 800 900 1000
mean normal effective stressP'(kPa)

(b) With preconsolidation pressure 44kPa and curing period 28days (M=1.70, λ =0.29,
κ =0.015)
Figure 5.40: Simulation of stress path of remoulded cement-treated marine clay
specimens with 50% cement content by modified Cam clay model

255
CHPATER 5

1400 Aw=50% CIU TEST


P=70kPa,CIU250
1300 P=70kPa,CIU500
P=70kPa,*CIU1000
1200
P=44kPa,CIU250
1100 P=44kPa,CIU500
P=44kPa,*CIU1000
1000
M=1.77

deviator stress q(kPa)


900 M=1.70

800

700

600

500

400

300

200

100

0
0 100 200 300 400 500 600 700 800 900 1000 1100
mean normal effective stressP'(kPa)
(a) Stress path of specimen with preconsolidation pressure 70kPa and 44kPa
respectively 1300
Aw=50% CIU TEST
1200
P=70kPa,CIU250
1100 P=70kPa,CIU500
P=70kPa,*CIU1000
1000 P=44kPa,CIU250
P=44kPa,CIU500
900
deviator stress q(kPa)

P=44kPa,*CIU1000
800

700

600

500

400

300

200

100

0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14
axial strain εs(%)
(b) Deviator stress-strain curve
800

700

Aw=50% CIU TEST


600 P=70kPa,CIU250
P=70kPa,CIU500
pore pressure u(kPa)

P=70kPa,*CIU1000
500
P=44kPa,CIU250
P=44kPa,CIU500
400 P=44kPa,*CIU1000

300

200

100

0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14
axial strain εs(%)

(c) Pore pressure versus axial strain


Figure 5.41: Undrained triaxial shearing behavior of remoulded cement-treated marine
clay specimens with 50% cement content and different preconsolidation pressure (*
indicates shear band was formed after peak value at about 12% axial strain)

256
CHPATER 5

Aw=50%, P=70kPa, CIU TEST


CIU100-7ds
CIU250-7ds
CIU500-7ds
*CIU750-7ds
1400
*CIU1000-7ds
1300 CIU100-28ds
CIU250-28ds
1200 CIU500-28ds
1100 CIU750-28ds
*CIU1000-28ds
1000 M=1.77

deviator stress q(kPa)


M=1.74
900

800

700

600

500

400

300

200

100

0
0 100 200 300 400 500 600 700 800 900 1000 1100
mean normal effective stressP'(kPa)
(a) Stress path of specimen with pre-remoulding curing period 7 days and 28 days
respectively 1400
Aw=50%,P=70kPa, CIU TEST
1300 CIU100-7ds
CIU250-7ds
1200
CIU500-7ds
1100 *CIU750-7ds
*CIU1000-7ds
1000 CIU100-28ds
deviator stress q(kPa)

900 CIU250-28ds
CIU500-28ds
800 CIU750-28ds
700 *CIU1000-28ds

600

500

400

300

200

100

0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14
axial strain εs(%)

(b) Deviator stress-strain curve


800
Aw=50%,P=70kPa, CIU TEST
CIU100-7ds
700
CIU250-7ds
CIU500-7ds
600 *CIU750-7ds
*CIU1000-7ds
pore pressure u(kPa)

CIU100-28ds
500
CIU250-28ds
CIU500-28ds
400 CIU750-28ds
*CIU1000-28ds

300

200

100

0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14
axial strain εs(%)

(c) Pore pressure versus axial strain


Figure 5.42: Undrained triaxial shearing behavior of remoulded cement-treated marine
clay specimens with 50% cement content and 70kPa preconsolidation pressure and
different curing period (* indicates shear band was formed after peak value at about
12% axial strain)

257
CHPATER 5
Aw=50%, P=44kPa, CIU TEST
1300 CIU250-7ds
CIU500-7ds
1200
*CIU1000-7ds
1100 CIU250-28ds
CIU500-28ds
1000 CIU1000-28ds
M=1.70
900
M=1.70

deviator stress q(kPa)


800

700

600

500

400

300

200

100

0
0 100 200 300 400 500 600 700 800 900 1000 1100
mean normal effective stressP'(kPa)

(a) Stress path of specimen with pre-remoulding curing period 7 days and 28 days
respectively CIU250-7ds
Aw=50%, P=44kPa,CIU TEST

CIU500-7ds
*CIU1000-7ds

1300
CIU250-28ds
CIU500-28ds
1200 CIU1000-28ds

1100

1000

900
deviator stress q(kPa)

800

700

600

500

400

300

200

100

0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14
axial strain εs(%)

(b) Deviator stress-strain curve


800

700

600 Aw=50%, P=44kPa,CIU TEST


CIU250-7ds
CIU500-7ds
deviator stress q(kPa)

500 *CIU1000-7ds
CIU250-28ds
CIU500-28ds
400
CIU1000-28ds

300

200

100

0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14
axial strain εs(%)

(c) Pore pressure versus axial strain


Figure 5.43: Undrained triaxial shearing behavior of remoulded cement-treated marine
clay specimens with 50% cement content and 44kPa preconsolidation pressure and
different curing period (* indicates shear band was formed after peak value at about
12% axial strain)

258
CHPATER 5

1200 Aw=50%, P=44kPa, 7days, CIU TEST


CIU250-experiment
1100 CIU500-experiment
*CIU1000-experiment
1000 CIU250-simulated
CIU500-simulated
900 CIU1000-simulated
M=1.70
800

deviator stress q(kPa)


700

600

500

400

300

200

100

0
0 100 200 300 400 500 600 700 800 900 1000
mean normal effective stressP'(kPa)

(a) Remoulded from mix proportion 2-1-3(M=1.70, λ =0.29, κ =0.015)


1200
2-1-4-P=44kPa-CIU TEST
CIU250-experiment
1100
CIU500-experiment
CIU1000-experiment
1000 M=1.70
CIU250-simulated
900 CIU500-simulated
CIU1000-simulated
800
deviator stress q(kPa)

700

600

500

400

300

200

100

0
0 100 200 300 400 500 600 700 800 900 1000 1100
mean normal effective stressP'(kPa)
(b) Remoulded from mix proportion 2-1-4 (M=1.7, λ =0.30, κ =0.015)
1200
2-1-5 CIU TEST
CIU250-experiment
1100 CIU500-experiment
CIU1000-experiment
1000 M=1.70
CIU250-simulated
900 CIU500-simulated
CIU100-simulated
800
deviator stress q(kPa)

700

600

500

400

300

200

100

0
0 100 200 300 400 500 600 700 800 900 1000
mean normal effective stressP'(kPa)

(c) Remoulded from mix proportion 2-1-5 (M=1.7, λ =0.30, κ =0.015)


Figure 5.44: Experiment and simulated Stress paths of remoulded cement-treated
marine clay specimens with preloading 44kPa (from cement-treated marine clay with
50% cement content and different total water content, cured under atmospheric
pressure for 7 days)

259
CHPATER 5

1200

1100

1000
2-1-3 CIU TEST
CIU250
900 CIU500
*CIU1000
800

deviator stress q(kPa)


700

600

500

400

300

200

100

0
0 0.02 0.04 0.06 0.08 0.1 0.12
axial strain εs(%)

(a) Remoulded from mix proportion 2-1-3


1100

1000

900 2-1-4 CIU TEST


CIU250
800 CIU500
CIU1000
deviator stress q(kPa)

700

600

500

400

300

200

100

0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16
axial strain εs(%)

(b) Remoulded from mix proportion 2-1-4


1100

1000

900 2-1-5 CIU TEST


CIU100
800 CIU250
CIU500
CIU1000
deviator stress q(kPa)

700

600

500

400

300

200

100

0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14
axial strain εs(%)

(c) Remoulded from mix proportion 2-1-5


Figure 5.45: Stress-strain curve of remoulded cement-treated marine clay specimens
with preloading 44kPa (from cement-treated marine clay with 50% cement content and
different pre-remoulding total water content, cured under atmospheric pressure for 7
days)

260
CHPATER 5

800

700

600
2-1-3 CIU TEST
CIU250
CIU500

pore pressure u(kPa)


500
CIU1000

400

300

200

100

0
0 0.02 0.04 0.06 0.08 0.1 0.12
axial strain εs(%)

(a) Remoulded from mix proportion 2-1-3


800

700

2-1-4 CIU TEST


CIU250
600 CIU500
CIU1000
pore pressure u(kPa)

500

400

300

200

100

0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16
axial strain εs(%)

(b) Remoulded from mix proportion 2-1-4


800

700

2-1-5 CIU TEST


CIU100
600 CIU250
CIU500
CIU1000
pore pressure u(kPa)

500

400

300

200

100

0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14
axial strain εs(%)

(c) Remoulded from mix proportion 2-1-5


Figure 5.46: Excess pore pressure-strain curve of remoulded cement-treated marine
clay specimens with preloading 44kPa (from cement-treated marine clay with 50%
cement content and different pre-remoulding total water content, cured under
atmospheric pressure for 7 days)

261
CHPATER 5

0-2-1-4-P=70kPa CIU TEST


CIU250-experiment
CIU500-experiment
CIU1000-experiment
CIU750-experiment
CIU1250-experiment
1300 M=1.70
CIU250-simulated
1200 CIU500-simulated
CIU750-simulated
1100
CIU1000-simulated
1000 CIU1250-simulated

900

deviator stress q(kPa)


800

700

600

500

400

300

200

100

0
0 100 200 300 400 500 600 700 800 900 1000 1100 1200 1300
mean normal effective stressP'(kPa)

(a) Experiment and simulated stress path by MCC (M=1.7, λ =0.30, κ =0.015)
1100

1000

900 0-2-1-4 CIU TEST


CIU250
800 CIU500
CIU1000
deviator stress q(kPa)

700

600

500

400

300

200

100

0
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1 0.11 0.12 0.13 0.14 0.15
axial strain εs(%)

(b) Deviator stress-strain curve


800

700

0-2-1-4 CIU TEST


CIU250
600
CIU500
CIU1000
pore pressure u(kPa)

500

400

300

200

100

0
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1 0.11 0.12 0.13 0.14 0.15
axial strain εs(%)

(c) Pore pressure versus axial strain


Figure 5.47: Undrained triaxial shearing behavior of remoulded cement-treated marine
clay specimens with preconsolidation pressure 70kPa (remoulded from cement-treated
marine clay with 50% cement content and 133% total water content and cured under
atmospheric pressure for 7 days)

262
CHPATER 5

50-2-1-4 CIU TEST


CIU100-experiment
CIU250-experiment
1200 CIU500-experiment
CIU750-experiment
1100 CIU1000-experiment
M=1.70
1000
CIU250-simulated
CIU500-simulated
900
CIU750-simulated
800 CIU1000-simulated

deviator stress q(kPa)


700

600

500

400

300

200

100

0
0 100 200 300 400 500 600 700 800 900 1000 1100
mean normal effective stressP'(kPa)

(a) Experiment and simulated stress path by MCC (M=1.7, λ =0.30, κ =0.015)
1100

1000

900 50-2-1-4 CIU TEST


CIU100
800 CIU250
CIU500
CIU1000
deviator stress q(kPa)

700

600

500

400

300

200

100

0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 0.18
axial strain εs(%)

(b) Deviator stress-strain curve


800

700

50-2-1-4 CIU TEST


CIU100
600
CIU250
CIU500
CIU1000
pore pressure u(kPa)

500

400

300

200

100

0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 0.18
axial strain εs(%)

(c) Pore pressure versus axial strain


Figure 5.48: Undrained triaxial shearing behavior of remoulded cement-treated marine
clay specimens with preconsolidation pressure 70kPa (remoulded from cement-treated
marine clay with 50% cement content and 133% total water content and 50kPa curing
stress)

263
CHPATER 5
1200
100-2-1-4 CIU TEST
CIU250-experiment
1100 CIU500-experiment
CIU750-experiment
1000 CIU1000-experiment
M=1.70
900 CIU250-simulated
CIU500-simulated
800 CIU750-simulated
CIU1000-simulated

deviator stress q(kPa)


700

600

500

400

300

200

100

0
0 100 200 300 400 500 600 700 800 900 1000 1100
mean normal effective stressP'(kPa)

(a) Experiment and simulated stress path by MCC (M=1.7, λ =0.30, κ =0.015)
1200

1100

1000

900

800
deviator stress q(kPa)

700

600

500

400

300
100-2-1-4 CIU TEST
CIU250
200
CIU500
CIU750
100
CIU1000
0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 0.18
axial strain εs(%)

(b) Deviator stress-strain curve


800

700

600

100-2-1-4 CIU TEST


CIU250
pore pressure u(kPa)

500
CIU500
CIU750
CIU1000
400

300

200

100

0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 0.18
axial strain εs(%)

(c) Pore pressure versus axial strain


Figure 5.49: Undrained triaxial shearing behavior of remoulded cement-treated marine
clay specimens with preconsolidation pressure 70kPa (remoulded from cement-treated
marine clay with 50% cement content and 133% total water content and 100kPa curing
stress)

264
CHPATER 5
1200
250-2-1-4 CIU TEST
CIU250-experiment
1100 CIU500-experiment
CIU750-experiment
1000 CIU1000-experiment
M=1.70
900 CIU250-simulated
CIU500-simulated
800 CIU750-simulated
CIU1000-simulated

deviator stress q(kPa)


700

600

500

400

300

200

100

0
0 100 200 300 400 500 600 700 800 900 1000 1100
mean normal effective stressP'(kPa)

(a) Experiment and simulated stress path by MCC (M=1.7, λ =0.30, κ =0.015)
1100

1000

900

800

700
deviator stress q(kPa)

600

500

400

300
250-2-1-4 CIU TEST
200 CIU250
CIU500
100 CIU750
CIU1000
0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16
axial strain εs(%)

(b) Deviator stress-strain curve


800

700

600

250-2-1-4 CIU TEST


CIU250
pore pressure u(kPa)

500
CIU500
CIU750
400 CIU1000

300

200

100

0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16
axial strain εs(%)

(c) Pore pressure versus axial strain


Figure 5.50: Uundrained triaxial shearing behavior of remoulded cement-treated
marine clay specimens with preconsolidation pressure 70kPa (remoulded from
cement-treated marine clay with 50% cement content and 133% total water content
and 250kPa curing stress)

265
CHPATER 5

5000 4
peak strength line (OCR=1)
OCR=1 (peak strength) p'≅po'/2
varies YSR (peak strength)

varies OCR (peak strength) 3.6 4.1


1 8.2
4000 8.2 1.6
post-rupture strength 2
4.1
10
3
5
1 2
Deviatoric stress, q (kN/m2)

3.2
tension cut-off line 20 1.6
30

Specific volume, ν
3000 peak strength envelope 20 2

(OCR=1) 2
30
2.8 3
4 1 1

5 2 3

2000
SEM points SEM points
3
2.4
2
1 0CID50 1 0CID50

30
30 6 2 0CIU50 2 0CIU50 1
3
post-ruptured
2 2
3 0CIU500 3 0CIU500 isotropic compression
1000 20 strength envelope
20 1.6
2 of 0CON0 specimen
4 0CIU1500 4 0CIU1500 6 1
2 10 PYCL
5
5 0CID500
8.2 4.1 5 0CID500 5 CCL
1.6
4.1 4 post-ruptured strengths
8.2
3 6 0CID1500 6 0CID1500
0 2 1.6
1
20 30 50 200 300 500 2000 3000 5000
0 1000 2000 3000 10 100 1000

Mean normal effective stress, p' (kN/m2) Mean normal effective stress, p' (kN/m2)

1 4

2 5

3 6

Figure 5.51: Microstructure at post-rupture states, as shown in the mechanical behaviour (after
Chin (2006))
266
CHPATER 5

(a) (b)

Figure 5.52 Comparison between SEM of remoulded cement-treated marine clay and
cement-treated marine clay after drained shearing test (a) remoulded cement-treated
marine clay specimen, (b) rupture plane of cement-treated marine clay
(CID2500-1250)

267
CHAPTER 6
A CONSTITUTIVE MODEL FOR CEMENT-TREATED
MARINE CLAY

In this chapter, a constitutive model will be proposed for cement-treated marine clay. First,

general framework for describing the behavior of structured soil such as natural and artificially

cemented soil is discussed. Second, by introducing hardening parameter and structure-relevant

parameter, an empirical framework for cement-treated soil based on experimental study in

Chapter 4 was proposed. Third, by considering inherent cohesion and internal friction, a

theoretical framework based on the observation of constitutive behavior of cement-treated soil

and energy equation was proposed furthermore. In order to determine the structure effect and

the degradation of structure during loading, based on the experiment study of the previous

chapter, the remoulded cement-treated marine clay was used as a reference state of

cement-treated marine clay. Finally, the hardening parameter and structure-relevant parameter

were investigated and appropriate relationship between these parameters and plastic strains and

works were proposed.

6.1. Introduction

Many constitutive models have been proposed for natural or artificial structured or

cemented soils. In the discussion below, the constitutive models are grouped into a few broad

categories, namely

1. those using expanded structure surfaces,

269
CHAPTER6

2. those using a superposition approach, as well as

3. other methods such as ‘history stress tensor’ method and micromechanical method.

6.1.1. Expanded structure surface approach

Many researchers used an enlarged yield surface to model the effect of soil structure (e.g.

Rouainia and Muir Wood, 2000; Kavvadas and Amoros, 2000; Liu and Carter, 2002; Asaoka et

al, 2000; Baudet and Stallebrass et al, 2004).

Many models do not model the tensile strength (e.g. Muir Wood, 1995; Liu and Carter,

2002; Asaoka et al, 2000; Baudet and Stallebrass et al, 2004). In these models, the structure

surface is enlarged yield surface of uncemented or reconstituted soil. Since the structure

surface passes through the origin of the stress space, the effect of the expanded structure

surface is essentially to increase the “pre-consolidation” pressure (often called isotropic yield

stress since there is really no over-consolidation), thereby inducing an over-consolidation

effect on the soil. Thus, philosophically, the effect of structure is considered to be equivalent to

that of over-consolidation.

The position and size of the structure surface will change with loss of structure or

cementation. When the soil becomes totally unstructured, the structure surface will coincide

with the yield surface of the unstructured or reconstituted soil. For example, Muir Wood (1995)

proposed a constitutive model for structured soil based on the kinematic hardening modified

Cam-clay model. The structure yield surface is expanded elliptic surface of reference surface

or intrinsic surface. The evolution of the structure surface is controlled by a structure

parameter and pre-consolidation stress, which changes with plastic strain.

270
CHAPTER6

Similarly, Asaoka et al. (2000) proposed a constitutive model for highly structured soil by

introducing superloading yield surface, which is an enlarged original Cam-clay yield locus.

The state of the structure is defined as the ratio of the size of the original Cam-clay yield

surface and structure yield surface, and this ratio changes with plastic strain. The soil finally

reaches a remoulded and normally consolidated state, with the structure surface coinciding

with the original Cam-clay yield surface.

Liu and Carter (2002) also proposed a Structured-Cam Clay model expanding the

Modified Cam Clay yield surface into a structure surface. In this model, the structured Cam

Clay is idealized as an isotropic hardening and destructuring material with elastic and virgin

yielding behavior.

More recently, Baudet and Stallebrass (2004) proposed a constitutive model for structured

clays by introducing sensitivity to the existing structure. Sensitivity is, in reality, a structure

parameter and is assumed to decrease exponentially with damage strain, which equals to the

magnitude of the vector of plastic strain increment. In the computation of the “damage” strain,

both volumetric and shear strains are accorded equal weightage. The evolution of the structure

surface is controlled by the sensitivity and pre-consolidation stress, which changes with plastic

strain. At ultimate failure, the structure surface coincides with the intrinsic (i.e. unstructured)

surface. All these models do not model any tensile strength or true cohesion of cemented soil

imparted by cementation or bonding. In this discussion, the true cohesion refers to the ability

of the soil to withstand shear or deviator stress even in the absence of effective stress. As

shown in Chapter 4, cement-treated marine clay submerged unconfined in a bowl of water for

a prolonged period of time still maintains its shape; this is suggestive of the presence of true

271
CHAPTER6

cohesion. These models also assume the shape of structure surface is the same as that of

uncemented or reconstituted soil, which as discussed in Chapter 4, is not the case for

cement-treated marine clay. Thus, expanded yield surface alone is unlikely to be able to model

the behavior of cement-treated marine clay.

Some constitutive models do model tensile strength (e.g. Liu et al., 2006; Horpibulsuk et

al., 2009; Lee et al., 2004; Rouainia and Muir Wood, 2000; Kavvadas and Amoros, 2000; Gens

and Nova, 1993; Lagioia and Nova, 1993; 1995). A common method of doing this is to

left-shift the structure surface or yield locus so that part of it lies in the tensile region of the

stress space. As in the models without tensile strength, the position and size of the structure

surface also changes with loss of structure or bonding and finally coincides with the yield

surface of unstructured or reconstituted soil. For example, Gens and Nova (1993) proposed a

conceptual constitutive model for bonded soils and weak rocks by including a single yield

surface and non-associated flow rule. Two parameters were used to define the new enlarged

yield locus. The first parameter, P 'c 0 , controls the yielding of the bonded soil in isotropic

compression while the second parameter, Pt ' , relates to the cohesion and tensile strength. The

evolution of yield locus is controlled by volumetric hardening (as in unstructured soil) and

bond degradation. Under high confining pressure, the yield locus initially moves rightwards (to

model a decrease in tensile strength) and expands. As plastic strain continues to increase, the

yield locus continues moving rightwards but decreases in size. Under low confining pressure,

the initial expansion phase does not occur. Lagioia and Nova (1993, 1995) also proposed a

similar constitutive model for soft rocks.

Kavvadas and Amorosi (2000) proposed a constitutive model for structured soil by

272
CHAPTER6

introducing a bond strength envelope (BSE) which is effectively an outer structure surface.

Compared to Gens and Nova’s (1993) model, this model requires two surfaces. The inner

surface delimits elastic and plastic state. The external surface is controlled by bond strength.

The evolution of yield surface is controlled by two hardening rules. The first rule, kinematic

hardening, controls the position of the surfaces while the second rule, isotropic hardening,

controls the size of the structure surface. The later is composed of two components. The

volumetric component models the intrinsic volumetric hardening and the

volumetric-strain-induced structure degradation, which is similar to Gens and Nova’s (1993),

and Lagioia and Nova’s (1995) model. The deviator component models shear-induced

structure degradation. This enables the structure surface to decrease in size with plastic shear

strain.

Based on Wood’s (1995) model, Rouainia and Wood (2000) proposed a kinematic

hardening constitutive model for natural clays that considered tensile strength. Three yield

surfaces were used in this model. The reference surface represents the behavior of

reconstituted or completely remoulded soil and will move with plastic volumetric strain. A

kinematically hardening bubble delimits the elastic and plastic regions and moves with the

current stress. Once the current stress approaches the bubble, plastic strain occurs. A structure

surface contains information about the magnitude and anisotropy of structure. As plastic strain

occurs, it tends to collapse towards the reference surface. The evolution of structure surface is

similar to that proposed in Gens and Nova (1993)’s model. When the structure is totally lost,

the reference surface coincides with structure surface.

The expanded yield surface approach, with tensile strength, has also been used for

273
CHAPTER6

modeling cement treated soils (e.g. Liu et al., 2006; Horpibulsuk et al., 2009; Lee et al., 2004).

Lee et al. (2004) proposed a constitutive model for cement treated clay by introducing bonding

stress ratio m and critical state parameter λ ' . These parameters are used to simulate the

increase in the initial stiffness and shear strength of cement treated soil and subsequent

progressive reduction in stiffness and strength due to breaking of the bonding as a result of

shearing. The initial yield locus is an elliptic surface, including tensile strength. The evolution

of yield locus is similar to Gens and Nova (1993)’s model. The yield locus will move

rightwards and expands with plastic strain. After peak strength, the yield locus tends to shrink.

Finally, the structure surface becomes a “critical state” surface, which is the same as the yield

surface of the unstructured or reconstituted material.

Based on Liu and Carter’s(2002) model, Liu et al (2006) and Horpibulsuk et al. (2009)

proposed a constitutive model for cement treated clay based on concept similar to Lee et al.’s

(2004) model by introducing a modified mean effective stress. In this model, the new

parameter C, which is related to the shear strength contributed by cementation, is assumed

constant before peak point. As such, the structure surface expands according to virgin

compression curve before peak point. After peak strength, the structure surface will move

rightwards and collapse. Finally the structure surface will be the same as that of the

unstructured or reconstituted material. Obviously, the shortcomings of these models are the

same as those for natural cemented soils.

It may be noted that left-shifting the yield locus to allow for negative p' actually preserves

the shape of the yield locus. This may be difficult to justify on philosophical grounds since

different strength and failure mechanisms come into play in compression and tension. Yielding

274
CHAPTER6

under compression is likely to mobilize friction, interlocking (or densification) and kinematic

hardening. On the other hand, tensile failure is probably more appropriate related to cracking

and fracture. It is difficult to see how friction and interlocking can exist under tensile

conditions. For this reason, although left-shifted yield surface with the same shape may be able

to model the effect of tensile stress on a phenomenological level, they may ultimately lack

physical underpinning

6.1.2. Superposition method

Another approach which have been used for constitutive models of cemented soils is the

superposition approach (e.g. Chazallon and Hicher, 1998; 1995; Vatsala et al., 2001; Yu et al.,

1998; Shen, 1993a, 1993b; Liu and Shen, 2005). In these models, the response of cemented

soil was assumed to be the sum of the soil skeleton (i.e., friction between grains) and

cementation bond, which were modeled separately. In most cases, the response variable used is

either stress or stress-strain matrix.

Vatsala et al. (2001) proposed an elastoplastic model for cemented soils in which the

strength is assumed to be made of two components: the usual strength of the remoulded soil

skeleton, and the strength of cementation bond. The overall response of the cemented soil

under loading are reflected in these two components. The modified Cam-clay model is used for

soil skeleton stress-strain relationship whereas a simple elastic-plastic model is proposed for

cemented component.

Similarly, Chazallon and Hicher (1995, 1998) proposed a constitutive model for bonded

geomaterials. The overall response is composed from that of the bonds and non-bonded

275
CHAPTER6

material, which are modeled by an elasto-plastic model and an elastic model with damage

separately. However, Chazallon and Hicher (1998) stated that their model is only applicable for

cemented soil in which the cement bonds exist only at the particle contacts and not in the

interstitial pore spaces. This is likely to severely affect their applicability to cement-treated

marine clay.

Shen (1933a, 1993b) and Liu and Shen (2005) proposed nonlinear elastic model and

elasto-plastic model for structured clays, which involves superimpose the stress-strain matrices

of an unstructured component and of the cementation. The philosophical basis of their models

is the assumption that, at any stage during loading, a structured soil body is a binary medium

consisting of bonded blocks and weakened bands which are idealized as elastic-brittle

elements and elasto-plastic elements, respectively. Before destructuration, the soil response is

reflected in elasto-brittle material. After the elasto-brittle material is damaged, it is

considered to be transformed into an elasto-plastic material and can be modeled by modified

Cam clay model. The superposition process therefore treats structured and unstructured

portions to be physical entities occupying different points within the soil body, rather than as

components of strengths and stiffness within a same material. A weighting function is used to

define the relative proportion of damaged to undamaged material within the soil body. The

validity of this philosophy has not been established and it is difficult to see how structured and

unstructured parts can be delineated in space within a soil body.

Yu et al. (1998) proposed a constitutive model for cement treated soil (with clay, sand

and gravel) using a similar philosophy but representing both the structured and unstructured

components by Mohr-Coulomb models. The strength of cement treated soil is assumed to be

276
CHAPTER6

composed of structural strength and frictional resistance strength, both of which satisfy

Mohr-Coulomb strength criterion.

In these models, direct superposition often involves the superposition of two independent

constitutive models. This often leads to a large number of material parameters. For instance,

Yu et al.’s (1998) model involves 12 parameters without even considering compression

yielding. The evaluation of these parameters may be difficult and without a proper test

programme, parametric values may not be uniquely defined. In severe cases, this may lead to

highly counter-intuitive and unreasonable values. For example, Yu et al.’s (1998) requires two

cohesion values, one for the structured part and another one for the frictional resistance, that is

unstructured part. Yu et al. stated that “…for the structural part, the value of c is generally

fairly large and that of ϕ is fairly low; otherwise, for the frictional resistance part, c is

generally fairly low and ϕ is fairly large”. This is however, not reflected in their proposed

parameters for 4 different mixes. In three out of the four mixes, the value of c ranges from

200kPa to 500kPa for the frictional resistance part. This is counter-intuitive for an unstructured

part which is not supposed to have any cohesion. Furthermore, in one of the mixes, the friction

angle for the structural part is larger than the value for frictional resistance part; this clearly

contradicts Yu et al.’s statement. Thirdly, comparison of parameters of two different mixes

shows that a mix which has a higher cement content and lower water content has a lower

structural c value than another with lower cement content and higher water content, which is

counter-intuitive. Finally, the friction angles for the structural part shows large fluctuations

ranging from 11.5° to 40.7°. This highlights the danger of models which have a large number

of parameters and are based largely on mathematical devices with little or no physical basis.

277
CHAPTER6

In addition, in cases such as cemented-treated soil, where the truly unstructured state may

not be readily defined, the unstructured constitutive model may also not be evaluated readily.

6.1.3. Other methods

Other methods were also used in the constitutive modeling of cemented soil by some

researches (e.g. Wong and Mitchell, 1975; Oka et al., 1989; Hirai et al., 1989; Adachi T. and

Oka F., 1993, 1995; Abdulla and Kiousis, 1997; Namikawa and Mihria, 2007). For instance,

Oka et al. (1989) proposed a constitutive model for natural soft clay by introducing stress

history tensor and including tensile strength. The material strength is composed of frictional

strength and cementation. The frictional strength is relatively small at the early stages of the

straining process and the strength due to cementation decreases with the advances of

deformation and eventually frictional strength develops. Adachi and Oka (1993, 1995)

proposed a similar constitutive model for soft rock by introducing stress history tensor.

Abdulla and Kiousis (1997) proposed a constitutive model for lightly cement-treated sand by

using micromechanical method. In this model, the clean sand, the cementing bond and the pore

pressure are modeled independently. Some constitutive models were also proposed for cement

treated sand or sandy soil (Hirai et al., 1989; Namikawa and Mihria, 2007). In most of these

models, yield locus does not feature compression caps.

6.1.4. Difficulties in using existing models for cement-treated marine clays

The models discussed above suffer from two shortcomings when applied to

cement-treated marine clay. These are as follows:

278
CHAPTER6

1. The experimental data clearly shows a change in the shape, or at least the aspect ratio,

of the yield surface. This is not modeled by most of the models.

2. Many of the models have numerous parameters which cannot be readily evaluated.

As discussed above, in extreme cases, this can lead to highly counter-intuitive or

unreasonable parametric values.

6.2. An empirical constitutive framework for cement-treated marine

clay

In this section, an empirical constitutive framework for cement-treated marine clay is

introduced, which is basically based on the analysis of experimental results in Chapter 4. Two

different yield functions were introduced in this framework. One is mainly based on original

modified Cam-clay, with elliptical yield locus shape passing through the origin of the stress

plane p’-q and change in stress ratio on the top of the yield locus. This is the approach used by

Lee et al. (2004). In Lee et al’s (2004) model, a bonding stress ratio was introduced to consider

the change in stress ratio on the top of the yield locus. The bonding stress ratio is composed of

two components. The first component is the critical stress ratio M in the modified Cam clay

model. The second component is related to structure effect and destructuration process.

Another one is based on McDowell’s (2000) yield function, with yield locus passing through

the origin of the stress plane and change in stress ratio on the top of the yield locus. In both

yield functions, the yield locus expands with the increase in the plastic volumetric strain and

flattens with destructruation due to plastic strain. The former is actually volumetric hardening

whereas the later is referred to kinematic softening.

279
CHAPTER6

6.2.1. Empirical yield function one

Based on formula (4-1) and experiment results (e.g., Figures 4-63 to Figures 4-69), the

yield function is given below:

q = N p ' p0' − p '2 (6-1)

Where N is the stress ratio at the top of the yield locus

It must be pointed out that N is always higher than critical state friction dissipation

constant M at unstructured status due to the structure effect. Clearly, when N=M, which is

critical state friction coefficient, formula (6-1) is modified Cam-clay model yield function. N

may equal to M only after the cementation bond is destroyed fully, which may be only found

within the shear band in triaxial compression test. Consequently, there are two hardening

parameters. One parameter is po' , which is responsible for volumetric hardening and nothing

new. However, the conventional way is not suitable for determination of po' as a reference

state is needed when po' is higher than isotropic compression primary yield stress p 'py . The

determination of p 'py was discussed in detail in Chapter 4, which will also coved in section

6.4. The determination of po' after p 'py will be introduced in detail in section 6.4.

Another parameter is N, which is responsible for destructuration or kinematic softening.

For simplicity and convenience, N is considered as a function of plastic volumetric strain ε vp .

Figures 6-1 shows the relationship between N and ε vp for different mix proportion. As can be

seen, N can be express as below:

N = a1ε vpb1 (6-2)

Where a1=1.2299, b1=-0.2543 in this study.

280
CHAPTER6

Place formula (6-2) in formula (6-1), the yield function in empirical method one is as

below:

q = a1ε vpb1 p ' p0' − p '2 (6-3)

By using formula (6-3), at the certain po' and ε vp , a corresponding yield locus can be obtained.

Figures 6-2 shows the simulated yield locus and experimental yield points. Therefore, by using

two hardening parameters, the yield locus and its evolution can be obtained approximately.

6.2.2. Empirical yield function two

McDowell (2000) derived a family of yield loci, using a proposed work equation together

with the application of the normality principle. The yield function is as below:

η = M [α ln( p0' / p ')]1/ α (6-4)

McDowell found that when parameter α equals to 1.5, the yield locus is quite close to that of

original Modified Cam-clay. Therefore, another empirical yield function can be as below:

η = N [1.5ln( p0' / p ')]1/1.5 (6-5)

N has similar meaning as that in formula (6-1) and can also be related to volumetric plastic

strain. Also, when N=M, formula (6-5) reduces to McDowell’s yield function. Figures 6-3

show the relationship between N and ε vp for different mix proportion. Finally N can be

expressed as below:

N = a2ε vpb2
(6-6)

Where, a2=1.2748, b2=-0.2479 in this study.

Combining formula (6-5) and (6-6), the yield function in empirical method 2 is given below:

η = a2ε vpb [1.5ln( p0' / p ')]1/1.5


2
(6-7)

281
CHAPTER6

Similarly, for a given po' and ε vp , a yield locus can be obtained. Figures 6-4 show the

simulated yield locus and experimental yield points. Therefore, by using two hardening

parameters, the yield locus and its evolution can also be obtained approximately.

It must be pointed out that in empirical yield functions, the inherent cohesion or tensile

strength were not be able to be considered, which do not reflect the real status of cemented

soils. As such these yield function are only suitable for some loading conditions, for example,

triaxial loading condition. In order to overcome the shortcomings of these empirical yield

functions, a conceptual yield function was proposed based on theoretical study. This will be

addressed in next section.

6.3. A theoretical constitutive framework for cement-treated marine

clay

As discussed in Chapter 4, the observed primary yield locus is different from that of

untreated marine clay. Previous research has shown that the yield locus of remoulded marine

clay can be modeled by original modified Cam-clay. For example, Chua (1990) noted that

remoulded, isotropically reconsolidated Singapore marine clay behaves in a manner that is

reasonably well-described by the modified Cam-clay model with M~0.9. More importantly, as

presented in Chapter 4, upon shearing, the yield locus gradually evolves into an elliptical shape

which is akin to the modified Cam Clay, even though the friction coefficient M remains much

higher than that of remoulded untreated marine clay. As discussed above, using a model which

involves a primary yield locus of the same shape and aspect ratio but different size will not

replicate what the test data show. In the previous section, it has been shown using a series of

282
CHAPTER6

elliptical structure surfaces with changing aspect ratios can give a reasonable well fit to the

evolution of the yield surface during destructuration. However, there are shortcomings in these

two empirical yield functions. Firstly, it is not clear which parameter physically captures the

loss of structure. In the first model, the parameter N is used to change the aspect ratio of the

elliptical yield locus, but N is, in fact, the same parameter as the friction coefficient M in the

modified Cam Clay. Thus, while it may be possible to account for structure by increasing

friction coefficient, this is clearly less than satisfying from a physical viewpoint. The same

goes for the modified McDowell’s function. Secondly, both empirical yield functions do not

prescribe any true cohesion; the deviator stress goes to zero when the mean effective stress

decreases to zero. It is also evident that both yield functions do not prescribe any tensile

strength, which is contrary to what was measured from the Brazilian tests. Attempts to fit the

observed yield loci using a non-zero tensile strength, i.e. by allowing p' to go negative,

following the approach of some researchers (e.g. Liu et al., 2006; Horpibulsuk et al., 2009; Lee

et al., 2004) did not meet with much success. As shown in Figure 6.5, if negative p’ is allowed

in the yield locus, the resulting N (or M) value is exceedingly high, as much as ~5. Using

lower N-values led to very poor fit. In addition, using the same function to describe yielding

under compression and tensile failure (not yielding) may not be theoretically viable. For

instance, the elliptical yield function was originally founded upon the notions that interlocking

and friction are the components giving rise to shear strength in the soil. Left-shifting the yield

locus to allow for negative p' would imply that these components also apply under tension and

it is difficult to see how friction and interlocking can exist under tensile conditions. Similarly,

McDowell’s relation was based on the original Cam Clay energy equation, which is also based

283
CHAPTER6

on interlocking and friction. For this reason, it may be better to acknowledge the existence of

another mechanics, possibly fracture mechanics, under tension and that compressive yielding

and tensile failure may not be governed by the same function.

As mentioned earlier, the use of a yield locus with a constant shape during destructuration

does not match experimental observations herein. A superposition approach with damage

parameter may be more viable. Most of the superposition approaches used to date (e.g. Shen

1993a, 1993b; Chazallon and Hicher, 1998; Yu et al., 1998; Vatsala et al., 2001; Liu and Shen,

2005) consider the cemented soil as the sum or mixture of two components, a remoulded

(unstructured) soil component and a bonding component. The moduli, stress-strain matrices

and, in some cases, the yield loci of the two components are then added up directly or using

theory of mixture (e.g. Yu et al. 1998). Thus direct superposition in effect involves the

superposition of two independent constitutive models. This often leads to a large number of

material parameters. In addition, it is not easy to separate the constitutive model for the

bonding component from that of the unstructured component. Moreover, in cases such as

cemented-treated soil, where the unstructured state may not be readily defined, the

unstructured constitutive model may also not be evaluated readily. For instance, Yu et al.’s

(1998) model for cement-bentonite-treated soil has 12 parameters, all of which appear to

require a considerable amount of fitting to experimental data. Furthermore, some of the

parameters appear to rather counter-intuitive. For example, Yu et al. used the Mohr-Coulomb

envelope for both the frictional (i.e. remoulded) and bonding components. Their parameters for

several of the samples show very high Mohr-Coulomb cohesion for the “frictional”

components, ranging from 200kPa to 500kPa, where one have expected the cohesion to be

284
CHAPTER6

very small given that the bonding has been discounted from this component. Similarly, the

bonding component returns almost as high an angle of friction as the “frictional” component.

Yu et al.’s (1998) study also presented other difficulties. For example, the sample with the

highest cement: soil ratio (~8.8%) has a cohesion value for its bonding component which is

almost half that of another sample with lower cement: soil ratio and higher water content. In

addition, some of the cohesion values of the bonding component appear to be exceedingly high,

e.g. 840kPa, which is very surprising given the relatively low cement content of 9% or less.

This underscores the danger of fitting a large number of parameters to relatively few samples,

especially when the parameters have little or no physical meaning and cannot be independently

evaluated.

Finally, the superposition approach does not consider the interaction between the packing

of the soil and bond strength. As Figure 4.6 shows, the primary yield stress of cement-treated

soil specimens falls very drastically with water content, even if the cement content is kept

constant. The same goes the unconfined compressive strength since as discussed in chapter 4,

the primary yield stress is correlated with unconfined compressive strength. This should not be

so if one applies the ideas of the superposition approach. Since the cement content is kept

constant, the bonding component should be unchanged, the increase in water content would

only lead to an increase in void ratio, which presumably affects the remoulded component, and

one would not expect such a drastic decrease in shear strength.

The above discussion highlights some of the shortcoming of current approaches to model

structure or cementation effects. In this section, a new theoretical approach is attempted. This

approach is based on the observation that the cement-treated soils do have a structure surface

285
CHAPTER6

(or primary yield surface) but that this yield surface can change shape and size during shearing.

The assumption of a structure surface is similar to that for the expanded yield locus approach.

The difference here is that the yield function is derived for cement-treated marine clay using a

modified form of the modified Cam Clay energy equation, which takes into account

interlocking, friction and cohesion.

6.3.1. Basis of theoretical model

The starting point of the theoretical model is the experimental observation that under

triaxial loading, the yield locus of the cement-treated soil evolves into a shape which is

well-fitted by an ellipse. Furthermore, as discussed in the previous chapter, the behavior of the

remoulded cemented-treated marine clay is also well-described by a modified Cam Clay, albeit

with a much higher value of M than the untreated marine clay. This motivates the hypothesis

that loss of structure in the cement-treated marine clay leads eventually to an elliptical yield

locus, which was the hallmark of Roscoe and Burland’s (1968) modified Cam Clay model,

probably the earliest model to propose an elliptical yield locus. The energy equation used by

Roscoe and Burland (1968) was

p' dε pv + qdε sp = M 2 p' 2 dε sp2 + p' 2 dε p2


v (6.8)

In which d ε vp and d ε sp represent the plastic volumetric and shear strain increment,

respectively. In Eq. 6.8, interlocking is embodied within the dilatancy term p ' d ε vp on the

left-hand side while the right-hand side contains the dissipative components, which are the

frictional shear component and plastic volumetric change component. This leads to the

modified Cam Clay yield function

286
CHAPTER6

M 2 p '2 + q 2 = M 2 p0' p ' (6.9)

In which p0' is the pre-compression pressure.

The use of an expanded yield locus for structured soil (e.g. Muir Wood, 1995; Liu and

Carter, 2002; Asaoka et al, 2000; Baudet and Stallebrass et al, 2004) is equivalent to increasing

pre-compression pressure. Seen in this light, using an expanded yield locus is equivalent to

postulating that the effect of structure is to raise the pre-compression pressure and thereby

increasing the yield ratio of the soil. However, if p ' is 0, then q must also be 0. Thus, this

way of representing structure does not prescribe any true cohesion to the structured soil.

Furthermore, the experimental data presented earlier show that, in cement-treated clay, the

shape of the yield locus does indeed change. Hence, merely increasing the pre-compression

pressure cannot fully describe the behavior of cement-treated soil.

The fact that the initial yield locus is always steeper than the subsequent yield loci

suggests that there is another component which has not been considered in Eq. 6.8 and this

component has the effect of increasing the deviator stress at yielding, independently of the

effect on the mean effective stress p’. This component is postulated to be the true cohesion,

denoted herein by the term C. It is also postulated that work done against this component is

irrecoverable, that is, this component is dissipative. This motivates the inclusion of true

cohesion C on the right-hand side of Eq. 6.8 and not the left-hand side. Finally, the fact that

true cohesion C affects only the deviator stress suggests that it is related purely to the plastic

shear strain increment dεsp.

There are many ways of incorporating C into the right-hand side of Eq. 6.8, but one way

that preserves the general form of the equation is

287
CHAPTER6

p ' d ε v p + qd ε s p = (C + Mp ')2 d ε p 2s + p '2 d ε p v2 (6-10)

It will be noted that Eq. 6-10 involves, in effect, a superposition of the dissipative component

related to the plastic shear strain and that this method of superposition is not unlike the

approach in the Mohr-Coulomb strength envelope, in that the cohesive component, C, is added

directly to the frictional component Mp'.

We note that when C = 0 in Eq. 6.10, it reduces to Eq. 6.8. Furthermore when p ' = 0 ,

q = C . Hence C is indeed a measure of the cohesion when p ' = 0 , that is, it is the cohesion

intercept of the prospective yield locus with the q-axis.

6.3.2. Yield locus


6.3.2.1 Deduction of yield locus

Eq. 6-10 can be rewritten as

q 2 d ε p 2s = (C + Mp ') 2 d ε sp 2 − 2 p ' qd ε vp d ε sp (6-11a)

Or

d ε vp
q2 + 2 p ' q = (C + Mp ') 2 (6-11b)
d ε sp

Assuming normality,

d ε vp dq
=− (6-12)
dε s p
dp '

Substituting Eq. 6-12 into 6-11 leads to

dq
q2 − 2 p ' q = (C + Mp ') 2
dp ' (6-13)

q
We can define the stress ratioη = , so that
p'

dq dη
=η + p' (6-14)
dp ' dp '

288
CHAPTER6

Substituting Eq. 6-14 into Eq. 6-13 gives


p '2η 2 + 2η p '3 '
+ (C + Mp ') 2 = 0 (6-15)
dp

Or

dη η (C + Mp ') 2 −1
+ =− η (6-16)
dp ' 2 p ' 2 p '3

We noted that Eq. 6-16 has the form


+ g ( p ')η = h( p ')η −1 (6-17)
dp '

1 (C + Mp ') 2
wherein g ( p ') = , and h( p ') = −
2p' 2 p '3

Which is a Bernoulli Equation.

The general form of the Bernoulli Equation is

dy
+ g ( x ) y = h( x ) y n (6-18)
dx

where g(x) and h(x) are functions of x, and n ≠ 0 or 1.

We now define z = η 2 , so that

dz dη 1 dz dη
= 2η or = (6-19)
dp ' dp ' 2η dp ' dp '

Substituting Eq. 6-19 into Eq. 6-16 leads to

dz z (C + Mp ') 2
+ =− (6-20)
dp ' p ' p '3

This is a first-order linear differential equation.

A first-order linear differential equation has the form

dy
+ f ( x ) y = m( x ) (6-21)
dx

Where f(x) and m(x) are functions of x. The solution of this first-order differential equation has

the form

289
CHAPTER6

− f ( x ) dx ⎛ ∫ f ( x ) dx dx + c ⎞
y= e ∫
⎝∫
⎜ m( x )e 0⎟ (6-22)

In which c0 is the constant of integration.

1 1
Letting f ( p ') = , ∫ f ( p ') dp ' = ∫ dp ' = ln p ' , so that
p' p'

e∫
f ( p ') dp '
= p' (6-23)

And

1
e ∫
− f ( p ') dp '
= (6-24)
p'

We also know that

( C + Mp ')
2

m( p ') = − (6-25)
p '3

In view of Eqs. 6-22 to 6-25, the solution of Eq. 6-20 is given by

1 ⎛ ( C + Mp ')2 ⎞
z=- ⎜∫ dp '− c ⎟
p' ⎜ p '2
0

⎝ ⎠

1 ⎛ C 2 + 2CMp '+ M 2 p '2 ⎞


=- ⎜∫ 2
dp '− c0 ⎟
p' ⎝ p' ⎠

C2 ln p ' c
= '2
− 2 MC −M2 + 0 (6-26)
p p' p'

Or

C2 ln p ' c
η2 = '2
− 2 MC −M2 + 0 (6-27)
p p' p'

To evaluate c0, set p ' = p0' and η = 0 , which gives

C2
c0 = M 2 p0' + 2 MC ln p0' − (6-28)
p0'

Substituting Eq. 6-28 into Eq. 6-27 leads to

290
CHAPTER6

p0'
M 2 ( p0' − p ') + 2MC ln( )
C2 p' C2
η2 = + − (6-29)
p '2 p' p0' p '

Or

C2 '
q 2 = M 2 p '( p0 '− p ') + 2MCp 'ln( p0' / p ') + ( p0 − p ') (6-30)
p0'

which is the plastic potential, and also the yield function if the associated flow rule is assumed.

It should be noted that the associated flow rule is essentially an assumption. Studies on natural

clays (e.g. Wood, 1990) have shown that this assumption of associated flow is reasonably

realistic for clays on the wet side of critical and are undergo strain hardening. In the critical

state framework, this would refer to the Roscoe Surface. For cement-treated soil, the critical

state is only reached at an advanced stage of destructuration. Even then, as discussed in

Chapter 5, there is uncertainty as to which state constitutes the critical state. Anyway, all states

on the primary or early yield surfaces would lie above the fully-destructured yield surface.

Thus, the notion of wet or dry side of critical really does not apply to primary yielding.

Furthermore, as discussed above, cement-treated clay under drained loading conditions

continues to strain hardening considerably after primary yield and the “compression cap”

occupies a significant portion of the stress space. This compression cap is akin to the Roscoe

surface in remoulded soils. In view of this, the assumption of associated flow is unduly

unreasonable when it is applied to the compression cap.

6.3.2.2 Consistency checks


The yield function was subjected to a few consistency checks, as follows:

1. When p ' = p0' , q=0, which gives the isotropic compression condition.

2. When C=0, q = M p '( p0 '− p ') , which is exactly the equation of modified Cam-clay

291
CHAPTER6

model. This means that when the true cohesion C = 0, the yield locus becomes an ellipse..

3. When p’ = 0, then

q 2 = C 2 − 2 MC lim p ' Lnp ' (6-31)


p' → 0

To evaluate the limit, we use L’Hopital’s rule and express

f ( p' )
p ' Lnp ' = (6-32)
g ( p' )

in which f ( p ' ) = Lnp ' and g ( p ' ) = 1/ p '

Using L’Hopital’s rule,

1
f ( p ') f ' ( p ') p'
lim = lim ' = lim = lim(− p ') = 0 (6-33)
p '→ 0 g ( p ') p '→ 0 g ( p ') p '→ 0 − 1 p '→ 0
'2
p

Thus, when p’ = 0, q = C.

6.3.2.3 Comparison with test data


Figure 6.6 demonstrates the proposed yield locus. As can be seen, the cementation effect

is reflected in both the initial cohesion and isotropic primary yield stress. The initial cohesion

is responsible for the increasing in deviator stress upon yielding as shown in Figure 6.6a. The

isotropic primary yield stress is actually the initial value of p0', which is responsible for the

expanding of the primary yield locus with respect to unstructured soil. Figures 6.7 to 6.10

assesses the ability of the proposed yield function to fit the observed primary yield loci for

cement-treated marine clay specimens with different mix proportion and cured under different

confining pressure for different curing period, using different values of M and C. As shown in

Figures 6.7 to 6.10, with appropriately chosen values of M and C, the matching is remarkably

good, considering the theoretical, rather than fitted, nature of the yield loci.

292
CHAPTER6

In these figures, the tension failure envelopes were also included, these being based on

the tensile strengths of cement-treated marine clay from Brazilian tests and a straight line of

gradient 3. The rationale for using this as the tensile failure envelope is based on the idea that,

if the soil has a tensile strength of σt, then under triaxial compression conditions, tensile failure

would not occur until the minor principal effective stress or effective radial stress σ3’ satisfies

the condition

σ 3' = −σ t (6-34)

Then

1 1
p ' = (σ 1' + 2σ 3' ) = (σ 1' − 2σ t ) (6-35)
3 3

And

q = σ 1' − σ 3' = σ 1' + σ t (6-36)

Eq. 6-35 leads to

1
p ' + σ t = (σ 1' + σ t ) (6-37)
3

Dividing Eq. 5-37 by Eq. 5-36 gives

q
=3 (6-38)
p' +σ t

Hence the postulate that cement-treated marine clay specimens have a tensile strength that

is isotropic and can be evaluated via the Brazilian test, would, under triaxial compression

condition, lead to a tensile failure envelope which has the form of straight line with gradient 3

and intercepts the p ' -axis at p ' = −σ t . It should also be noted that, in a triaxial compressive

condition, tensile cracking would intuitively occur longitudinal planes. In Brazilian test, tensile

cracking is also induced along the longitudinal diametral plane. This gives added confidence to

the notion that the tensile strength measured in Brazilian tests would also apply to triaxial

293
CHAPTER6

compression tests.

The use of a tensile failure envelope that has a different shape from the yield locus relates

to the point made earlier on the different mechanisms which govern yield and tensile failure.

The yield locus was derived based on the interaction between friction, interlocking, cohesion

and volumetric dissipation. It is difficult to see how friction, interlocking and volumetric

dissipation can be effective under conditions of tension. Thus, the use of left-shifted yield

locus to represent tensile failure of cemented soil may well be fundamentally problematic in

this respect.

Figures 6.11 to 6.15 also compare the proposed yield function with observed yield loci at

various stages of shearing after initial yielding. In each figure, which is for a given mix, curing

period and curing stress, the value of M is kept constant while C is decreased and p0'

increased. In general, the agreement between the theoretical and observed yield loci is

remarkably good. There was some discrepancy in some of initial yield loci, such as Fig. 6.13,

in which some measured end points on the apparent tensile failure envelope (i.e. the line q =

3p') appear to suggest yield beneath the yield locus. However, these points might have been

affected by the inability of triaxial tests to reach any states with σ3' < 0; this has been discussed

earlier in Chapter 4. Methods for determining C, po' as well as M will be discussed in detail

later.

6.4. Parameters of theoretical yield function

There are three parameters in the proposed yield function, i.e., the friction coefficient M,

the mean effective stress under isotropic compression p0' and cohesion C. In this study, the

initial value of p0' is actually isotropic primary yield stress. After isotropic primary yield

294
CHAPTER6

stress p 'py , p0' is also named preconsolidaiton stress of cement-treated soil, which is similar to

original modified Cam clay model. Based on the concept of reference state of cement-treated

soil, when the cement-treated soil reaches at the reference state, p0' is actually the

preconsolidation stress of remoulded cement-treated soil. As the determination of isotropic

primary yield stress has been discussed in detail in Chapter 4, only preconsolidation stress of

cement-treated soil will be discussed in this section

6.4.1 Friction coefficient M

As shown in Figure 6.2, if the evolution of the yield loci between initial yielding and

peak strength is fitted using an ellipse, the apparent friction coefficient N of the ellipse will

decrease. In the theoretical model, however, a constant friction coefficient M is presumed in

place of a variable friction coefficient N. This is in keeping with a basic premise of the Cam

Clay model, wherein M is taken to be a material constant. In the subsequent discussion, the

friction coefficient M will be assumed to be a constant for a given mix, curing period and

curing stress. The value of M used in each fitting attempt to find C is based on the value of

M of the corresponding remoulded cement-treated soil. This is consistent with the discussion

in Chapter 5 that, while remoulded cement-treated soil may not really be totally unstructured,

it does provide a reasonably good basis for a “critical” state which can be used in practical

engineering.

6.4.2 Preconsolidation stress p0'

Figures 6.16 to 6.20 show the isotropic compression curves of cement-treated marine clay

295
CHAPTER6

and remoulded cement-treated marine clay. As can be seen, these figures are similar to that of

natural structured soil presented by many researches (e.g. Liu and Carter, 1999, 2000). In each

case, the compression of the intact specimen converges towards the corresponding curve of the

remoulded specimen.

Using an approach similar to Liu and Carter’s (2002), we can consider that, under

isotropic compression, the mean effective stress in cement-treated marine clay p0' is composed

'
of two components, i.e., p0u and Δp0s
'
, such that

p0' = p0' u + Δp0' s (6-39)

'
In which p0u is the component which is borne by the remoulded soil skeleton and Δp0s
'
is

the additional stress that can be resisted due to structure, as shown in Figure 6.21a.

It is also assumed that the superposition rule expressed in Eq. 6-39 applies also to all

stress situations, so that, in general

p ' = pu' + Δps' (6-40)

Where pu' is mean effective stress carried by the remoulded or unstructured cement-treated

marine clay and Δps' is that component corresponding to structure effect.

During triaxial shearing or volumetric compression, both p0' and p0' u increases with the

increase in volume change whereas Δp0s


'
deceases with destructuration due to plastic work.

When the soil reaches a fully destructured state, Δp0s


'
equals to zero, so that p0' equals

to p0' u . This is similar to the compression of structured soil described by Liu and Carter (2002)

(Fig 6.21b).

Following Liu and Carter (2002), for any state ( po' , v) of unstructured soil, the specific volume

can be expressed as below:

296
CHAPTER6

v = vIU − λU LN ( p0' u ) (6-41)

Where vIU is initial specific volume of unstructured soil when p0' u =1 and λU is the isotropic

compression index of the unstructured soil.

Hence, the pre-consolidation stress of unstructured or remoulded soil p0' u can be determined

by equation (6-41).

At the isotropic yield point ( p 'py , v py ) , Equations (6-39) and (6-41) can be written as

p 'py = p0' ui + Δp0' si (6-42)

v py = vIU − λU LN ( p0' ui ) (6-43)

The determination of the isotropic compression yield stress p 'py has been addressed in detail

in Chapter 4. Hence, by using the above two equations, the initial value of Δp0s
'
before initial

yielding, i.e. Δp0si


'
can be determined.

Loss of structure is assumed to commence when po' > p 'py ; this is consistent with Liu and

Carter’s (2002) assumption. Following Liu and Carter (2002), the extra pre-consolidation

stress Δp0s
'
due to structure effect may be expressed as a power function of pre-consolidation

stress p0' .

p 'py
Δp0' s = Δp0' si ( )b (6-44)
p0'

The parameter b can be obtained by fitting the isotropic compression curve of cement-treated

marine clay.

Combining Equations (6-39) and (6-44), the pre-consolidation stress of cement-treated soil can

be expressed as below.

p 'py
p0' = pou
'
+ Δp0' si ( )b (6-45)
p0'

The differential form of equation (6-45) is as below:

297
CHAPTER6

p 'py Δp0'
Δp0' = Δpou
'
− bΔp0' si ( )b (6-46a)
p0' p0'

Or

Δp0' u ( p0' )b +1
Δp0' = (6-46b)
bΔp0' si ( p 'py )b + ( p0' )b +1

Following the Cam-clay model (Schofield and Wroth, 1968), for unstructured or remoulded

soil, Δpou
'
can be expressed as below.

Δv p
Δp0' u = − p0' u (6-47)
λU − κU

Where κU and λU are isotropic compression parameters of unstructured soil.

Combining Equations (6-47), (6-46b) and (6-41), the incremental pre-consolidation stress of
cement-treated soil may be written as below.

Δv p vIU − v
− exp( )( p0' )b +1
λU − κU λU
Δp0' = (6-48)
bΔp ( p ) + ( p0' )b +1
'
0 si
'
py
b

We note that

vIU − v
exp( ) = p0' u (6-49)
λU

When cement-treated soil is fully destructured, p0' equals to p0u


'
and is much bigger

than p 'py . Consequently, equation (6-48) reduces to equation (6-47).

6.4.3 Determination of cohesion

Recall yield function (6-30).

C2 '
q 2 = M 2 p '( p0 '− p ') + 2MCp 'ln( p0' / p ') + ( p0 − p ') (6-30)
p0'

So far, the determination of M and p0' has been demonstrated in the previous sections. The

only left main parameter is inherent cohesion C. Since C is only dependent on structure of soil,

the initial value of C, termed hereafter as initial cohesion C0, and degradation of C will be

298
CHAPTER6

discussed separately.

6.4.3.1 Determination of initial cohesion C0


If M and p0' or p 'py is known, any point on the primary yield locus can be used to

determine C0. One way is to use the intersection ( p1' , q1 ) between yield locus and apparent

tensile cut-off line may be used to get C0. On the cut-off line,

q1
p1' = (6-50)
3

Thus the yield function becomes

q1 q q p' C2 q
q12 = M 2 ( p0 '− 1 ) + 2 MC 1 ln( 0 ) + ' ( p0' − 1 )
3 3 3 q1 p0 3
3
Or

M2 q 2 3 p' q
p0' q12 = q1 p0' ( p0 '− 1 ) + Mq1 p0' ln( 0 )C + ( p0' − 1 )C 2 (6-51)
3 3 3 q1 3

Let
q1
g1 ( p0' , q1 ) = p0' − , (6-52a)
3

2 3 p'
g 2 ( p0' , q1 ) = Mq1 p0' Ln( 0 ) , (6-52b)
3 q1

M2 M2
g3 ( p0' , q1 ) = q1 p0'2 − ( + 1) p0' q12 (6-52c)
3 9

Thus the yield function becomes

g1C 2 + g 2C + g3 = 0 (6-52d)

The solution is

g 22 − 4 g1 g 3 − g 2
C= (6-52e)
2 g1

As shown in Figure 6.22, q1 is nearly equal to qu . This suggests that unconfined

compression test specimens generally follow the apparent tensile cut-off line and fails at the

point of intersection of this apparent cut-off line and the primary yield surface and the use of

299
CHAPTER6

the developed yield surface together with a tensile cut-off gives a good prediction of the

unconfined compressive strength. This is not surprising; in the unconfined compression test,

the confining stress is zero. If we further assumed, as was suggested in Chapter 4, that any

positive excess pore pressure will be able to dissipate through the sides of the specimens, then

the effective stress path will have to be along the apparent tensile cut-off line. Therefore, C0

can be estimated by replacing q1 with qu and p0' with p 'py in equations (6-52a) to (6-52d) and

using Equation (6-52e). It should be noted that, in most unconfined compression tests, failure

is accompanied by the formation of a shear band. However, it was also observed that, in most,

if not all specimens, shear bands only started to propagate after the peak unconfined

compressive strength was reached. The qu measured in the unconfined compression tests in

these studies are the peak values and they therefore represent the strength of the specimens

prior to shear band formation, at which point, the specimens are still continuous. Thus, it is not

surprising that a continuum model, like the one developed herein, should give a good

prediction of the unconfined compression strength.

Figure 6.23a shows the deduced C0s by using q1(qu) and C0t obtained by experiment. As

can be seen, C0 s = 0.769Cot , indicating that C0 is generally underestimated if we use Eq. 6-52.

The underestimated C0 may due to test error in qu and some of the unconfined compression

specimens might have failed just before reaching the apparent cut-off line. As Figure 6.23b

shows, if we apply a correction of about 1.3 to the deduced C0, good agreement is obtained

with the experimental C0t.

Alternative, Figure 6.24 shows the relationship between C0t obtained by experiment

and qu . It was found that C0t = 0.95qu . This means that the inherent cohesion is about the same

300
CHAPTER6

as the unconfined compressive strength.

6.4.3.2 Cohesion degradation


This section examines the relationship between the cohesion at any point of time during

testing and plastic work. In this discussion, the plastic work is computed from drained triaxial

tests by the following algorithm or relations:

W = Wvp + Wsp (6-53)

Where Wvp is the component due to plastic volumetric strain ε vp and may be written in the

accumulation form below:

pi' + pi' +1
Wvp = ∑ (ε vp ( i +1) − ε vp ( i ) ) (6-54)
2
and Wsp is the component due to plastic shear strain ε sp and may be written in the

accumulation form below:

qi + qi +1
Wsp = ∑ (ε sp ( i +1) − ε sp ( i ) ) (6-55)
2

For a given drained triaxil test, different C and W values can be obtained corresponding to

each post-yield locus, which was determined by triaxial tests. These (C, W) points are then

plotted on a normalized plane. For each mix proportion, several drained triaxial test were

chosen to obtain the relationship between C and W. Consequently, as can be seen later, there

were several points with the same C/C0 because all these points are on the same yield locus,

which has the same C/C0 value. Undrained test data are much more difficult to incorporate into

the present framework as the specimens usually strain soften fairly rapidly after primary

yielding, with formation of shear band. This violates the continuum assumption which is

needed for application of the model.

Figure 6.25 shows the degradation of C with plastic work for cement-treated marine clay.

301
CHAPTER6

In this figure, C is normalized by the corresponding initial cohesion C0 while the plastic work

on the horizontal axis is also normalized by C0. As can be seen, the degradation curve indicates

that there is nonlinear relationship between normalized C and normalized plastic work W. As

this figure shows, the true cohesion C decreases very rapidly with work done at the beginning,

but the rate of decrease tail off towards the end when the work done becomes very large. This

is intuitively correct since one would expect the rate of destructuration to be maximum at the

start but it should trend towards zero as cohesion approaches zero.

Figures 6.26 to 6.30 show the degradation of C with plastic work for cement-treated

specimens with different mix proportions. In these figures, the normalized plastic work on the

horizontal axis is raised to a best-fit power by fitting the data. As can be seen, the normalized

C and normalized with plastic work W expressed in a power function has a good linear

relationship. This indicates that the degradation of C can be expressed as power function of

plastic work.

C / C0 = a1 (W / C0 )b1 + 1 (6-56)

From these figures, the degradation of C for different mix proportion may have the forms

below.

For mix proportion 10-1-11, degradation of C is

C / C0 = −0.6634(W / C0 )0.199 + 1 (6-57)

Similarly, for mix proportion 5-1-6, degradation of C is

C / C0 = −0.7006(W / C0 )0.186 + 1 (6-58)

For mix proportion 10-3-13, degradation of C is

C / C0 = −0.6522(W / C0 )0.182 + 1 (6-59)

302
CHAPTER6

For mix proportion 2-1-3, degradation of C is

C / C0 = −0.7019(W / C0 )0.042 + 1 (6-60)

For mix proportion 2-1-4, degradation of C is

C / C0 = −0.5579(W / C0 )0.099 + 1 (6-61)

Eq. 6-56 – 6-61 shows that

C / C0 = 1 − a1 (W / C0 ) m (6-62)

Or y = 1 − a1 x m (6-63)

In which x = W / C0 and y = C / C0

Differentiating Eq.6-63 leads to

dy / dx = − a1mx ( m −1) (6-64)

If m ≥ 1.0, then dy / dx will not tend to zero as x tends to ∞ , which would be counter-intuitive.

Hence m should be less than 1, and Eq. 6-57-6.61 show that this is indeed the case.

Figures 6.31 to 6.35 show the cohesion degradation with evolution of yield locus and plastic

strain needed to degrade the cohesion. As can be seen, the cohesion degradation and the

needed plastic strain are dependent on the mix proportion.

303
CHAPTER 6

4 4
10-1-11 5-1-6
Fit : N=1.1609εvp(-0.2616), R2=0.988 Fit : N=1.2284εvp(-0.2654), R2=0.995
3.5 3.5

3 3

N
N

2.5 2.5

2 2

1.5 1.5

1 1
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0 0.05 0.1 0.15 0.2 0.25 0.3 0.35
plastic volumetric strain εpv plastic volumetric strain εpv

(a) (b)
4
10-3-13 4
2-1-3
Fit : N=1.3006εvp(-0.2392), R2=0.977
Fit : N=1.1788εvp(-0.2575), R2=0.980
3.5
3.5

3
3
N

2.5
N

2.5

2
2

1.5
1.5

1
1
0 0.05 0.1 0.15 0.2 0.25 0.3
plastic volumetric strain εpv 0 0.05 0.1 0.15 0.2
plastic volumetric strain εpv

(c) (d)
4 4
2-1-4 cement content 10%-50%,
Fit : N=1.3118εvp(-0.2387), R2=0.974 total water content 100%-133%
3.5 3.5 Fit : N=1.2299εvp(-0.2543), R2=0.97

3 3
N

2.5 2.5

2 2

1.5 1.5

1 1
0 0.05 0.1 0.15 0.2 0.25 0.3 0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4
plastic volumetric strain εp v plastic volumetric strain εpv

(e) (f)
Figure 6.1: Relationship between stress ratio at the top of the yield locus(empirical
method one) and plastic volumetric strain for different mix proportion (a) 10-1-11, (b)
5-1-6, (3) 10-3-13, (d) 2-1-3, (e) 2-1-4, (f) for all mix proportions

304
CHAPTER 6

simuated yield locus


10-1-11
2500 yield point-P'0=130kPa
yield point-P'0=300kPa
yield point-P'0=500kPa
yield point-P'0=1000kPa
yield point-P'0=1500kPa
yield point-P'0=2000kPa
yield point-P'0=2500kPa
simulated yield locus
2000 Extension cut-off line
deviator stress q(kPa)

1500

1000

500

0
0 500 1000 1500 2000 2500
mean normal effective stressP'(kPa)

(a)
simuated yield locus
5-1-6
yield point-P'0=260kPa
yield point-P'0=380kPa
yield point-P'0=500kPa
2500 yield point-P'0=1000kPa
yield point-P'0=1500kPa
yield point-P'0=2000kPa
yield point-P'0=2500kPa
simulated yield locus
Extension cut-off line

2000
deviator stress q(kPa)

1500

1000

500

0
0 500 1000 1500 2000 2500
mean normal effective stressP'(kPa)

(b)

305
CHAPTER 6

simuated yield locus


10-3-13
yield point-P'0=355kPa
yield point-P'0=500kPa
2500
yield point-P'0=1000kPa
yield point-P'0=1500kPa
yield point-P'0=2000kPa
yield point-P'0=2500kPa
simulated yield locus
Extension cut-off line

2000
deviator stress q(kPa)

1500

1000

500

0
0 500 1000 1500 2000 2500
mean normal effective stressP'(kPa)
(c)

simuated yield locus


2-1-3
yield point-P'0=580kPa
yield point-P'0=1000kPa
2500 yield point-P'0=1500kPa
yield point-P'0=2000kPa
yield point-P'0=2500kPa
simulated yield locus
Extension cut-off line

2000
deviator stress q(kPa)

1500

1000

500

0
0 500 1000 1500 2000 2500
mean normal effective stressP'(kPa)

(d)

306
CHAPTER 6

simuated yield locus


2-1-4
yield point-P'0=275kPa
2500 yield point-P'0=500kPa
yield point-P'0=1000kPa
yield point-P'0=1500kPa
yield point-P'0=2000kPa
yield point-P'0=2500kPa
simulated yield locus
Extension cut-off line
deviator stress q(kPa) 2000

1500

1000

500

0
0 500 1000 1500 2000 2500
mean normal effective stressP'(kPa)

(e)

Figure 6.2: Simulated yield loci through empiric method one with experimental data
for different mix proportion (a) 10-1-11, (b) 5-1-6, (3) 10-3-13, (d) 2-1-3, (e) 2-1-4

307
CHAPTER 6

4
10-1-11 4
5-1-6
Fit : N=1.1741εvp(-0.2595), R2=0.989
Fit : N=1.2847εvp(-0.2537), R2=0.991
3.5
3.5

3
3
N

2.5

N
2.5

2
2

1.5
1.5

1
1
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35
plastic volumetric strain εpv
plastic volumetric strain εpv

(a) (b)
4
10-3-13 4
2-1-3
Fit : N=1.2904εvp(-0.2428), R2=0.985
Fit : N=1.1489εvp(-0.2648), R2=0.984
3.5
3.5

3
3
N

2.5
N

2.5

2
2

1.5
1.5

1
0 0.05 0.1 0.15 0.2 0.25 0.3 1
plastic volumetric strain εpv 0 0.05 0.1 0.15 0.2
plastic volumetric strain εpv

(c) (d)
4 4

2-1-4
cement content 10%-50%,
Fit : N=1.3568εvp(-0.2352), R2=0.989 total water content 100%-133%
3.5 3.5 Fit : N=1.2748εvp(-0.2479), R2=0.947

3 3
N

2.5 2.5

2 2

1.5 1.5

1 1
0 0.05 0.1 0.15 0.2 0.25 0.3 0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4
plastic volumetric strain εp v plastic volumetric strain εpv

(e) (f)
Figure 6.3: Relationship between stress ratio at the top of the yield locus(empirical
method two) and plastic volumetric strain for different mix proportion (a) 10-1-11, (b)
5-1-6, (3) 10-3-13, (d) 2-1-3, (e) 2-1-4, (f) for all mix proportions

308
CHAPTER 6

simuated yield locus


2500 10-1-11
yield point-P'0=130kPa
yield point-P'0=300kPa
yield point-P'0=500kPa
yield point-P'0=1000kPa
yield point-P'0=1500kPa
yield point-P'0=2000kPa
yield point-P'0=2500kPa
2000 simulated yield locus
Extension cut-off line

deviator stress q(kPa)

1500

1000

500

0
0 500 1000 1500 2000 2500
mean normal effective stressP'(kPa)

(a)
simuated yield locus
5-1-6
yield point-P'0=260kPa
yield point-P'0=380kPa
yield point-P'0=500kPa
2500 yield point-P'0=1000kPa
yield point-P'0=1500kPa
yield point-P'0=2000kPa
yield point-P'0=2500kPa
simulated yield locus
Extension cut-off line

2000
deviator stress q(kPa)

1500

1000

500

0
0 500 1000 1500 2000 2500
mean normal effective stressP'(kPa)

(b)

309
CHAPTER 6

simuated yield locus


10-3-13
yield point-P'0=355kPa
yield point-P'0=500kPa
yield point-P'0=1000kPa
2500
yield point-P'0=1500kPa
yield point-P'0=2000kPa
yield point-P'0=2500kPa
simulated yield locus
Extension cut-off line

2000
deviator stress q(kPa)

1500

1000

500

0
0 500 1000 1500 2000 2500
mean normal effective stressP'(kPa)

(c)
simuated yield locus
2-1-3
yield point-P'0=580kPa

2500 yield point-P'0=1000kPa


yield point-P'0=1500kPa
yield point-P'0=2000kPa
yield point-P'0=2500kPa
simulated yield locus
Extension cut-off line

2000
deviator stress q(kPa)

1500

1000

500

0
0 500 1000 1500 2000 2500
mean normal effective stressP'(kPa)

(d)

310
CHAPTER 6

simuated yield locus


2-1-4
yield point-P'0=275kPa
2500 yield point-P'0=500kPa
yield point-P'0=1000kPa
yield point-P'0=1500kPa
yield point-P'0=2000kPa
yield point-P'0=2500kPa
simulated yield locus
Extension cut-off line
2000
deviator stress q(kPa)

1500

1000

500

0
0 500 1000 1500 2000 2500
mean normal effective stressP'(kPa)

(e)

Figure 6.4: Simulated yield loci through empiric method two with experimental data
for different mix proportion (a) 10-1-11, (b) 5-1-6, (3) 10-3-13, (d) 2-1-3, (e) 2-1-4

311
CHAPTER 6

450
10-1-11 primary yield locus
theoretical method
emperical method-without tensile strength N=3.65
emperical method-with tensile strength N=5.10
emperical method-with tensile strength N=2.40

deviator stress q(kPa)


primary yield point
tension cut-off line

300

150

0
-40 0 40 80 120 160
mean normal effective stressP'(kPa)
(a)

1050
10-3-13 primary yield locus
theoretical method
emperical method-without tensile strength N=3.65
emperical method-with tensile strength N=4.85
900
emperical method-with tensile strength N=2.42
deviator stress q(kPa)

primary yield point


tension cut-off line

750

600

450

300

150

0
-100 0 100 200 300 400
mean normal effective stressP'(kPa)

(b)

Figure 6.5: Simulated yield loci through empiric method one with tensile strength (a)
mix proportion 10-1-11, (b) mix proportion 10-3-13

312
CHAPTER 6

500 demonstration of yield locus


untreated soil
isotropic expanding
enlarging alonge q axis
Extension cut-off line

deviator stress q(kPa)

Enlarging

Expanding

0
0 100 200 300
mean normal effective stressP'(kPa)

(a)

1.5 demonstration of yield locus


untreated soil
isotropic expanding
enlarging alonge q axis
Extension cut-off line

1
Enlarging
q/po'

0.5

0
0 0.5 1
P'/po'

(b)

Figure 6.6: Demonstration of proposed yield locus (a) p '− q plane (b)

p '/ po' − q / po' plane

313
CHAPTER 6

1300 simulated primary yield locus


s-c-w
1200 10-1-11-M=1.50,C=245.68kPa
5-1-6-M=1.62,C=450.00kPa
1100 10-3-13-M=1.65,C=620.00kPa
2-1-3-M=1.70,C=960.00kPa
1000 Extension cut-off line
10-1-11 tension failure envelop
900 5-1-6 tension failure envelop
10-3-13 tension failure envelop

deviator stress q(kPa)


800 2-1-3 tension failure envelop
primary yield point
700

600

500

400

300

200

100

0
-200 -100 0 100 200 300 400 500 600
mean normal effective stressP'(kPa)

Figure 6.7: Simulated primary yield loci for cement-treated marine clay specimens
with different cement content and cured under atmospheric pressure for 7 days
1300 simulated primary yield locus
s-c-w
1200 2-1-3-M=1.70,C=960.00kPa
2-1-4-M=1.70,C=460.00kPa
1100 2-1-5-M=1.70,C=175.00kPa
2-1-5.5-M=1.70,C=132.00kPa
1000 Extension cut-off line
2-1-3 tension failure envelop
900 2-1-4 tension failure envelop
2-1-5 tension failure envelop
deviator stress q(kPa)

800 2-1-5.5 tension failure envelop


primary yield point
700

600

500

400

300

200

100

0
-200 -100 0 100 200 300 400 500 600
mean normal effective stressP'(kPa)

Figure 6.8: Simulated primary yield loci for cement-treated marine clay specimens
with different total water content and cured under atmospheric pressure for 7 days

314
CHAPTER 6

simulated primary yield locus


2750 cruing stress-s-c-w
0-2-1-4-M=1.70,C=460.00kPa
2500 50-2-1-4-M=1.70,C=1000.00kPa
100-2-1-4-M=1.70,C=1220.00kPa
250-2-1-4-M=1.70,C=2000.00kPa
2250
Extension cut-off line
0-2-1-4 tension failure envelop
2000
50-2-1-4 tension failure envelop
100-2-1-4 tension failure envelop
1750

deviator stress q(kPa)


250-2-1-4 tension failure envelop
primary yield point

1500

1250

1000

750

500

250

0
-500 -250 0 250 500 750 1000 1250 1500
mean normal effective stressP'(kPa)

Figure 6.9: Simulated primary yield loci for cement-treated marine clay specimens
with 50% cement content and 133% total water content and cured under different
confining pressure for 7 days
900 simulated primary yield locus
5-1-6-7days-M=1.62,C=450.00kPa
2-1-4-7days-M=1.70,C=460.00kPa
800 5-1-6-28days-M=1.62,C=680.00kPa
2-1-4-28days-M=1.70,C=680.00kPa
Extension cut-off line
700
5-1-6-7days tension failure envelope
2-1-4-7days tension failure envelope
5-1-6-28days tension failure envelope
600
2-1-4-28days tension failure envelope
deviator stress q(kPa)

primary yield point


500

400

300

200

100

0
-100 0 100 200 300 400
mean normal effective stressP'(kPa)

Figure 6.10: Simulated primary yield loci for cement-treated marine clay specimens
with diffeern mix proportion and cured under atmospheric pressure for different curing
time periods

315
CHAPTER 6

simuated yield locus


10-1-11
yield point-P'0=130kPa

2250 yield point-P'0=300kPa


yield point-P'0=500kPa
yield point-P'0=1000kPa
yield point-P'0=1500kPa
yield point-P'0=2000kPa

2000 yield point-P'0=2500kPa


simulated yield locus
Extension cut-off line

1750

1500
deviator stress q(kPa)

1250

1000

750

500

250

0
0 500 1000 1500 2000 2500
mean normal effective stressP'(kPa)

Figure 6.11: Simulated evolution of yield locus for cement-treated marine clay
specimens with 10% cement content and 100% total water content and cured under
atmospheric pressure for 7days
simuated yield locus
5-1-6
2500 yield point-P'0=260kPa
yield point-P'0=380kPa
yield point-P'0=500kPa
yield point-P'0=1000kPa
yield point-P'0=1500kPa
yield point-P'0=2000kPa
yield point-P'0=2500kPa
simulated yield locus
Extension cut-off line
2000
deviator stress q(kPa)

1500

1000

500

0
0 500 1000 1500 2000 2500
mean normal effective stressP'(kPa)

Figure 6.12: Simulated evolution of yield locus for cement-treated marine clay
specimens with 20% cement content and 100% total water content and cured under
atmospheric pressure for 7days

316
CHAPTER 6

2500 simuated yield locus


10-3-13
yield point-P'0=355kPa
yield point-P'0=500kPa
yield point-P'0=1000kPa
yield point-P'0=1500kPa
yield point-P'0=2000kPa
yield point-P'0=2500kPa
simulated yield locus
Extension cut-off line
2000
deviator stress q(kPa)

1500

1000

500

0
0 500 1000 1500 2000 2500
mean normal effective stressP'(kPa)

Figure 6.13: Simulated evolution of yield locus for cement-treated marine clay
specimens with 30% cement content and 100% total water content and cured under
atmospheric pressure for 7days
simuated yield locus
2-1-3
yield point-P'0=580kPa
2500 yield point-P'0=1000kPa
yield point-P'0=1500kPa
yield point-P'0=2000kPa
yield point-P'0=2500kPa
simulated yield locus
Extension cut-off line

2000
deviator stress q(kPa)

1500

1000

500

0
0 500 1000 1500 2000 2500
mean normal effective stressP'(kPa)

Figure 6.14: Simulated evolution of yield locus for cement-treated marine clay
specimens with 50% cement content and 100% total water content and cured under
atmospheric pressure for 7days

317
CHAPTER 6

simuated yield locus


2-1-4
yield point-P'0=275kPa
2500 yield point-P'0=500kPa
yield point-P'0=1000kPa
yield point-P'0=1500kPa
yield point-P'0=2000kPa
yield point-P'0=2500kPa
simulated yield locus
Extension cut-off line

2000
deviator stress q(kPa)

1500

1000

500

0
0 500 1000 1500 2000 2500
mean normal effective stressP'(kPa)

Figure 6.15: Simulated evolution of yield locus for cement-treated marine clay
specimens with 50% cement content and 133% total water content and cured under
atmospheric pressure for 7days

318
CHAPTER 6

10 100 p'(kPa)
1000 10000

3.6

3.2
specific volume V

2.8

2.4

2
Aw=10%-cement-treated marine clay
Aw=10%-remoulded cement-treated marine clay
Aw=10%-unremoulded-v=6.44-0.54Ln(p')
Aw=10%-remoulded-v=3.52-0.20Ln(p')
1.6

Figure 6.16: Isotropic compression curves of remoulded and unremoulded


cement-treated marine clay specimens with 10% cement content

10 100 p'(kPa)
1000 10000
3.6

3.2
specific volume V

2.8

2.4

2 Aw=20%-cement-treated marine clay


Aw=20%-remoulded cement-treated marine clay
Aw=20%-unremoulded-v=6.52-0.52Ln(p')
Aw=20%-remoulded-v=3.885-0.23Ln(p')

Figure 6.17: Isotropic compression curves of remoulded and unremoulded


cement-treated marine clay specimens with 20% cement content

319
CHAPTER 6

10 100 p'(kPa)
1000 10000
3.6

3.2
specific volume V

2.8

2.4

Aw=30%-cement-treated marine clay


Aw=30%-remoulded cement-treated marine clay
2 Aw=30%-unremoulded-v=6.80-0.53Ln(p')
Aw=30%-remoulded-v=4.19-0.25Ln(p')

Figure 6.18: Isotropic compression curves of remoulded and unremoulded


cement-treated marine clay specimens with 30% cement content

10 100 p'(kPa)
1000 10000
3.6

3.2
specific volume V

2.8

Aw=50%-cement-treated marine clay


Aw=50%-remoulded cement-treated marine clay
Aw=50%-unremoulded-v=6.715-0.50Ln(p')
2.4 Aw=50%-remoulded-v=4.74-0.285Ln(p')

Figure 6.19: Isotropic compression curves of remoulded and unremoulded


cement-treated marine clay specimens with 50% cement content

320
CHAPTER 6

10 100 p'(kPa)
1000 10000
4.5

3.5
specific volume V

2.5

Aw=50%, Cw=133%-cement-treated marine clay


Aw=50%, Cw=133%-remoulded cement-treated marine clay
2 Aw=50%, Cw=133%-unremoulded-v=5.8-0.38Ln(p')
Aw=50%, Cw=133%-remoulded-v=4.50-0.275Ln(p')

Figure 6.20: Isotropic compression curves of remoulded and unremoulded


cement-treated marine clay specimens with 50% cement content and 133% total water
content before remoulding

Δp0s
'

'
p0u

(a) In this study (b) after Liu and Carter, 2002

Figure 6.21: Idealization of the isotropic compression behavior of reconstituted and


structured soils

321
CHAPTER 6

2500
q 1-qu
fitting line q1=1.1qu, R2=0.998

2000

1500
q1(kPa)

1000

500

0
0 500 1000 1500 2000 2500
qu(kPa)

Figure 6.22: Relationships between q1 and qu

1600 2400
simuated Co and tested Co simuated Co after correction and test Co
fitting line C0s=0.769C0t, R2=0.994 fitting line C0s=0.999C0t, R2=0.994

2000

1200

1600
C0s(kPa)

C0s(kPa)

800 1200

800

400

400

0 0
0 400 800 1200 1600 2000 0 400 800 1200 1600 2000
C0t(kPa) C0t(kPa)

(a) Before correction (b) After correction

Figure 6.23: Comparison between Cot and Cos (Cot was obtained from experiment,
Cos was obtained by using qu and yield function)

322
CHAPTER 6

2000
C-qu
fitting line C0=0.95qu,R2=0.99

1600

1200
C0(kPa)

800

400

0
0 500 1000 1500 2000 2500
qu(kPa)

Figure 6.24: Relationship between C0 and qu


1
degradation of C -10-1-11

0.8
C/C 0

0.6

0.4

0.2
0 0.4 0.8 1.2 1.6
W/C0

Figure 6.25: Degradation of C with plastic work on normalized plane for


cement-treated marine clay (with mix proportion 10:1:11)

323
CHAPTER 6

1
degradation of C -10-1-11
Fit line: C/C0=-0.6634(W/C0)0.199+1,R2=0.985

0.8

C/C 0

0.6

0.4

0.2
0 0.4 0.8 1.2
(W/C0)0.199

Figure 6.26: Degradation of inherent cohesion C for cement-treated specimen with


10% cement content and 100% total water content

1
degradation of C -5-1-6
Fit line: C/C0=-0.7006(W/C0)0.186+1,R2=0.985

0.8
C/C 0

0.6

0.4

0.2
0 0.2 0.4 0.6 0.8 1
(W/C0)0.186

Figure 6.27: Degradation of inherent cohesion C for cement-treated specimen with


20% cement content and 100% total water content

324
CHAPTER 6

1
degradation of C -10-3-13
Fit line: C/C0=-0.6522(W/C0)0.182+1,R2=0.988

0.8
C/C 0

0.6

0.4
0 0.2 0.4 0.6 0.8 1
(W/C0)0.182

Figure 6.28: Degradation of inherent cohesion C for cement-treated specimen with


30% cement content and 100% total water content

1
degradation of C -2-1-3
Fit line: C/C0=-0.7019(W/C0)0.042+1,R2=0.994

0.8
C/C 0

0.6

0.4

0.2
0 0.2 0.4 0.6 0.8 1
(W/C0)0.042

Figure 6.29: Degradation of inherent cohesion C for cement-treated specimen with


50% cement content and 100% total water content

325
CHAPTER 6

1
degradation of C -2-1-4
Fit line: C/C0=-0.5579(W/C0)0.099+1,R2=0.949

0.8

C/C 0

0.6

0.4
0 0.2 0.4 0.6 0.8 1
(W/C0)0.099
Figure 6.30: Degradation of inherent cohesion C for cement-treated specimen with
50% cement content and 133% total water content

simuated yield locus


10-1-11
yield point-P'0=130kPa

2250 yield point-P'0=300kPa


yield point-P'0=500kPa
yield point-P'0=1000kPa
yield point-P'0=1500kPa
yield point-P'0=2000kPa

2000 yield point-P'0=2500kPa

0.10 simulated yield locus


Extension cut-off line

1750

0.08
1500 0.32
εsp
deviator stress q(kPa)

1250
0.06 0.29
εvp
1000
0.045 0.03
0.25 75
750
0.02 0.07 C
εsp 0.16 85
0.03
500
εvp 100 135
160
115 245
250
C
0
0 500 1000 1500 2000 2500
mean normal effective stressP'(kPa)

Figure 6.31: Cohesion degradation with evolution of yield locus and plastic strain for
cement-treated marine clay specimens with 10% cement content and 100% total water
content and cured under atmospheric pressure for 7days

326
CHAPTER 6

simuated yield locus


5-1-6
2500 yield point-P'0=260kPa
yield point-P'0=380kPa
yield point-P'0=500kPa
yield point-P'0=1000kPa
yield point-P'0=1500kPa
yield point-P'0=2000kPa
yield point-P'0=2500kPa
simulated yield locus
Extension cut-off line
2000
0.09

0.075
deviator stress q(kPa)

1500
εsp
0.055
0.30
0.265
1000 0.04 0.025 εvp
0.22 0.015
εsp 0.06
0.12 150 C
175 0.02
500 270
εvp 195 300
230 450
C
0
0 500 1000 1500 2000 2500
mean normal effective stressP'(kPa)

Figure 6.32: Cohesion degradation with evolution of yield locus and plastic strain for
cement-treated marine clay specimens with 20% cement content and 100% total water
content and cured under atmospheric pressure for 7days

2500 simuated yield locus


10-3-13
yield point-P'0=355kPa
yield point-P'0=500kPa
yield point-P'0=1000kPa
yield point-P'0=1500kPa
yield point-P'0=2000kPa
yield point-P'0=2500kPa

0.077 simulated yield locus


Extension cut-off line
2000

0.057
deviator stress q(kPa)

1500
0.04 0.035 εsp
0.235

0.035 0.20 0.008


1000 εvp
0.145
εsp 0.00 0.0075
0.045
εvp 0.00 C
270 450
500
300 620
330
350
C
0
0 500 1000 1500 2000 2500
mean normal effective stressP'(kPa)

Figure 6.33: Cohesion degradation with evolution of yield locus and plastic strain for
cement-treated marine clay specimens with 30% cement content and 100% total water
content and cured under atmospheric pressure for 7days

327
CHAPTER 6

simuated yield locus


2-1-3
yield point-P'0=580kPa
2500 yield point-P'0=1000kPa
yield point-P'0=1500kPa
yield point-P'0=2000kPa
yield point-P'0=2500kPa
simulated yield locus

0.05 Extension cut-off line

2000
0.03

0.025
0.16
deviator stress q(kPa)

1500
0.12
0.0065
0.083
1000 εsp
0.0055
0.00 320
350
εvp 370
500 400
960
C
0
0 500 1000 1500 2000 2500
mean normal effective stressP'(kPa)

Figure 6.34: Cohesion degradation with evolution of yield locus and plastic strain for
cement-treated marine clay specimens with 50% cement content and 100% total water
content and cured under atmospheric pressure for 7days

simuated yield locus


2-1-4
yield point-P'0=275kPa
2500 yield point-P'0=500kPa
yield point-P'0=1000kPa
yield point-P'0=1500kPa
yield point-P'0=2000kPa
yield point-P'0=2500kPa
simulated yield locus

0.085 Extension cut-off line

2000

0.075

εsp
deviator stress q(kPa)

1500
0.052
0.28 0.04
0.245
0.04 εvp
1000 0.20
0.02
εsp
0.10 0.05
0.00
200 C
500 εvp 220 0.00
240 275
260
C 460

0
0 500 1000 1500 2000 2500
mean normal effective stressP'(kPa)

Figure 6.35: Cohesion degradation with evolution of yield locus and plastic strain for
cement-treated marine clay specimens with 50% cement content and 133% total water
content and cured under atmospheric pressure for 7days

328
CHAPTER 7
CONCLUSIONS

7.1 Conclusions

This study deals with the post-yield constitutive behavior of cement-treated Singapore

marine clay with emphasis on the pre-peak behavior. Cement-treated Singapore marine clay

specimens with different mix proportions and curing conditions was studied firstly to examine

the behavior under different mix proportion and curing condition. The effect of mix

proportion and curing stress and curing time on the primary yield locus was investigated. The

post-yield behavior and post-yield changes in the yield locus were also investigated for

specimens with different mix proportions. Secondly, short samples were also used in triaxial

testing to study post-peak softening behavior, as part of the measures to prevent shear band

formation. The constitutive behavior of remoulded cement-treated marine clay was

investigated and its use as an intrinsic state was explored. Finally, based on the observations

of the experiment of this study, a new energy equation was proposed for cement-treated

marine clay, from which a new yield locus was derived. Based on these studies, the following

conclusions can be drawn.

7.1.1. Pre-peak behavior of cement-treated Singapore marine clay

1. In this study, Coop and Atkinson’s method was used to define and determine the

primary yield stress and the primary yield locus. The results showed that a reasonably

consistent primary yield locus of cement-treated marine clay can be obtained, using this

method.

2. The results of isotropic compression tests showed that the post-yield compression

index seems to be independent on cement content and curing period. The post-yield

compression index is also independent on total water content and curing stress. However, the

isotropic primary yielding stress increased with the increase in cement content, curing load

and curing time as well as decrease in total water content.

329
CHAPTER 7

3. Triaxial compression test on specimens with consolidation stress lower than isotropic

primary yield stress showed stiff behavior with peak strength appearing around 1%~2% of

axial strain. On the other hand, in CID tests, the stress-strain behavior is dependent upon the

effective confining stress or yield ratio.. Specifically, all CIU samples reach their peak

strength on or close to the tensile failure envelope, with the stress paths trending along or

toward the tensile failure envelope. All CID samples tested at high yield ratio (>>2) showed

highly elastic behavior before peak strength while CID samples tested at low yield ratio

showed large volumetric compression up to the peak strength, which is often reached at larger

shear strain than CIU tests. Under higher effective confining stress, the point of peak strength

is reached well after primary yield point, which indicates significant pre-peak strain hardening.

These behaviors under triaxial compression are consistent with Chin’s (2006) study. The

results also showed that increasing the cement content, curing stress and curing period or

decreasing the total water content all has a similar effect in increasing the peak strength of the

specimens. In all these cases, the increase in strength seems to be correlated, at least partially,

to a decrease in post-curing void ratio.

4. The experimental results showed that although the primary yield locus expands with

an increase in cement content, curing stress and curing period as well as the decrease in total

water content, it can be approximately normalized by the isotropic primary yield stress. The

isotropic primary yield stress can be correlated to the

a) unconfined compressive strength (considering curing stress), or

b) cement content and total water content (without considering curing stress), or

c) post-curing void ratio together with the ratio of total water content to the cement content

(considering curing stress).

5. Samples reconsolidated to effective confining stress higher than isotropic primary

yield stress exhibit similar trend of stress-strain behavior to those which have not been

consolidated past the isotropic primary yield stress. However, samples which have been

isotropically consolidated past the isotropic primary yield points show much more abrupt

changes in stiffness and compression curve upon yielding. Both CIU and CID samples which

330
CHAPTER 7

were sheared under high isotropic over-consolidation ratios show stiff response up to the peak

strength, followed by dilation and strain softening. CIU specimens which were sheared at

lower isotropic over-consolidation ratio of ~2 also show stiff response up to peak strength.

Their stress paths are also roughly vertical, which would be consistent with largely elastic

behavior up to the peak strength. These behaviors under triaxial compression are consistent

with Chin’s (2006) study. In all samples, yield loci can be drawn through the isotropic pre-

compression pressures and the yield points deduced from CID and stress path tests.

6. The experiment results showed that the post-yield locus expands with isotropic pre-

compression pressure, owing to densification effects. However, the normalized yield locus

tends to contract, changing from a steep arch to an ellipse. This can be attributed to kinematic

softening induced by the breakage of inter-aggregate bonds. The destructuration mainly

comes from isotropically reconsolidation and appears to continue after the yield locus has

been reached during shearing. Thus, the post-yield behavior of the cement-treated soil appears

to be influenced by densification effects as well as breakage of inter-aggregate bonds.

However, it was also found that even under higher pre-compression pressure (e.g., 2500kPa),

the friction coefficient M of the yield locus is still much higher than that of reconstituted

marine clay (with M value of 0.9) This indicates that even under sustained shearing, the

cemented-treated soil does not degrade to the same friction coefficient as that for

reconstituted marine clay.

7. The tensile splitting strength test showed that cement-treated marine clay specimens

show similar brittle behavior as rock or concrete, with cracking along the central vertical

plane under radial compression. The tensile strength of cement-treated marine clay specimen

increases with the increase in cement content, the decrease in total water content, as well as

the increase in curing period and curing stress. The experiment results indicated the existence

of a true tensile strength which is almost constant over a range of compression rates and

drainage conditions. The tensile strength can also be correlated to the unconfined compressive

strength, with the ratio of tensile to compressive strength lying within the ranges reported in

other studies. By assuming that the specimens are isotropic and that one single tension

331
CHAPTER 7

strength is applicable to all stress situations, the tensile strength measured by Brazilian test

could be used to redefine the tension failure envelope which could not be determined by

triaxial tests.

7.1.2 Ultimate state of cement-treated marine clay

1. The experimental results suggest that excess pore pressure and stress ratio may be

better indicators for shear band initiation than the deviator stress itself.

2. The short specimens with enlarged low-friction end caps showed significantly slower

rate of strain softening than that of conventional specimens. The short specimens with

enlarged low-friction end caps after shearing to very large axial strain of about 55% also

showed multiple shear bands or fissures distributed all over the specimen surface. This is

consistent with the observations of previous researchers and indicates a more uniform post-

peak behavior than that exhibited by conventional long specimens showing a single shear

band. It was found that the deviator stress and pore pressure stabilize after ~20% strain, but

with the stress ratio at a much higher level than that reached by the conventional specimens.

As such, the short specimens probably reached a critical state at a shear strain of about 20%.

However, the critical state could not be observed in short CID samples with the current setup

if it does exist.

3. The “critical state” line for the CIU short specimens has a gradient of about 2.60 and

is evidently much steeper than that for conventional long specimens reported by Chin (2006)

as well as in this study, which has a gradient of about 1.12 in the same stress range. Scanning

electron micrographs of short specimens after testing do not show the high degree of particle

break-up seen within the shear bands of conventional long specimens of Chin’s (2006) study

and this study.

4. In general, the isotropic compression curve of the remoulded treated marine clay

depends mainly on the cement content and lies between the compression curves of the

untreated marine clay and the corresponding intact treated marine clay. The isotropic

332
CHAPTER 7

compression index of remoulded treated marine clay is much lower than that of intact (i.e. not

remoulded) treated marine clay; the trend being more obvious when the cement content is low.

When the cement content is low (e.g. 10%~20%), the isotropic compression index of the

remoulded treated marine clay is slightly lower than that of untreated marine clay. On the

other hand, when the cement content is high (e.g., 30%~50%), it’s slightly higher than that of

untreated marine clay. It was found that the curing period, total water content and curing

stress all have little effect on the isotropic compression index of the remoulded treated clay.

5. The experiment results showed that the undrained stress path of remoulded cement-

treated marine clay has a similar shape as those of remoulded marine clay, which is well-

described by the modified Cam Clay. The deviator stress-strain curve is much ductile than

that of intact treated marine clay and similar to that of remoulded marine clay. The excess

pore pressure changes very little after the peak point. The peak value usually occurs at about

6%~8% of axial strain when the confining stress is lower than 250kPa, after which very small

fluctuation can be seen. For confining stress higher than 500kPa, the peak value usually

occurs at about 10%~12% of axial strain, followed by shear band formation. It was found that

the preconsolidation pressure effect on the undrained triaxial behavior increases with cement

content. When cement content is low (e.g. 10%), the preconsolidation pressure effect is very

small and the friction coefficient is almost the same. For higher cement content, the difference

appears to be more significant. However, curing stress and total water content does not

significantly affect undrained shearing behavior and critical state of remoulded cement-treated

marine clay. Curing period also has little effect on the undrained triaxial behavior regardless

of cement content and preconsolidation pressure. For the range of the specimens tested (i.e.

10% to 50% cement content), the friction coefficient of remoulded cement-treated marine

clay lies between 1.50 and 1.77 for preconsolidation pressure of between 44kPa and 70kPa.

6. In this study, the remoulded cement-treated marine clay was considered as a reference

“remoulded” state towards which the microstructure of the cement-treated clay will evolve

with continual shearing. The ultimate state of remoulded cement-treated specimen seems to

be only dependent on the cement content, thereby making it a convenient reference state.

333
CHAPTER 7

7.1.3 Constitutive frame work of cement-treated marine clay

1. Two empirical yield functions for cement-treated marine clay were proposed. The first

yield function is mainly based on original modified Cam-clay, with elliptical yield locus

shape passing through the origin of the stress plane p’-q and change in stress ratio on the top

of the yield locus. Another is adopted from McDowell’s (2000) yield function, with yield

locus passing through the origin of the stress plane and change in stress ratio on the top of the

yield locus. In both yield functions, there are two hardening parameters. One parameter is

responsible for volumetric hardening while another parameter is responsible for

destructuration or kinematic softening. Both parameters are assumed to be only related with

plastic volumetric strain. Consequently, the yield locus expands with the increase in the

plastic volumetric strain and flattens with destructruation due to volumetric plastic strain. By

using two hardening parameters, the yield locus and its evolution can also be obtained

approximately. However, the inherent cohesion or tensile strength could not be considered in

the empirical framework; this clearly deviates from observations which indicate that

cemented soil has a tensile strength.

2. A theoretical yield function was also derived based on an energy equation which was

modified from that of the modified Cam Clay. This follows from experimental observations

that continued shearing led to an elliptical yield locus. The modification involves introducing

a true cohesion parameter into the modified Cam Clay energy equation. This allows the

energy equation to take into account interlocking, friction and cohesion. The new yield

function has three parameters, i.e., true cohesion, friction coefficient of remoulded cement-

treated marine clay, and mean effective stress under isotropic condition of treated marine clay.

3. The proposed yield function fits well with the observed primary yield loci for cement-

treated marine clay specimens with different mix proportion and cured under different

confining pressure for different curing time period, using appropriate values of friction

coefficient and true cohesion. The proposed yield function also fits reasonably well with

observed yield loci at various stages of shearing after initial yielding. For a given mix, curing

334
CHAPTER 7

period and curing stress, the value of the friction coefficient is constant while the cohesion is

decreased (much as hardening, or softening parameter) and the pre-consolidation pressure is

increased.

4. In the theoretical framework, the friction coefficient is taken to be a material constant

and can be obtained from corresponding remoulded cement-treated soil. It was found that the

isotropic compression curves of cement-treated marine clay and remoulded cement-treated

marine clay are similar to that of natural structured soil reported by previous researchers.

Therefore, using an approach similar to Liu and Carter’s (2002), the pre-consolidation

pressure can be determined from isotropic compression curves of remoulded and intact

cement-treated marine clay specimens. The initial cohesion can be easily deduced from

unconfined compression tests, thereby allowing the primary yield locus of cement-treated

marine clay to be readily determined from conventional tests. The degradation of the cohesion

parameter can be expressed as power function of plastic work. By decreasing the cohesion

parameter after yielding, changes in the yield locus after primary yielding can be reproduced

reasonably well.

7.2 Recommendations for future study

1. The pre-yield behavior of the cement-treated soil including small-strain nonlinear

behavior was not covered in this study. The stiffness of cement-treated marine clay and its

dependence on various factors such as cement content, total water content, curing stress and

curing period should be investigated under triaxial loading condition. The change in the

stiffness during destructuration should also be studied. The incorporation of local strain

measurement in conventional triaxial test system would be able to meet with these

requirements although it may be more difficult under triaxial compression condition than

under unconfined compression condition. The knowledge about stiffness of cement-treated

marine clay would be useful to practical engineering and numerical analysis.

2. The basic assumption in this study is that all specimens are isotropic. The specimens

were isotropically reconsolidated under specified pressure before shearing. Although this is

335
CHAPTER 7

suitable for laboratory study, it’s different from the field situation of cement-treated soil,

which is cured and loaded under anisotropic condition. The effect of anisotropy on behavior

of cement-treated marine clay could be investigated by curing and testing the specimens

under anisotropic condition, provided that the curing and testing system can be designed

properly.

3. All conventional long specimens in this study and in Chin’s (2006) study showed

post-peak softening behavior and shear band formation in CIU and CID test. As noted by

Chin (2006), once the shear band is formed, the soil specimen is not uniform anymore, and

macroscopic changes in shear strain, specific volume and pore pressure do not reflect those

which occur within the shear band. This is a common bifurcation problem encountered by

investigators in strain softening soils which would benefit for further study. However, this

may require special technique to setup the specimen and measurements.

4. The proposed model can replicate the primary yielding locus and the evolution of

post-yield locus reasonable well, using appropriate parameters. The parameters of the

proposed model can be easily obtained from conventional test. However, it should be pointed

out that this is not fully complete constitutive model. Some issues are needed to investigate

furthermore to develop a full model. For instance, the recompression line under compression

space is needed to study in detail. Moreover, more validation work is necessary to asses the

capability of the proposed model to replicate the stress-strain behavior of cement-treated

marine clay specimens with different mix proportions, curing conditions and triaxial loading

conditions.

336
REFERENCES

Adachi, T., and Oka, F.(1995). “An elasto-plastic constitutive model for soft rock with
strain softening”. Int. J. Numer. and Anal. Methods in Geomech., 19(4), 233-247.

Adachi, T., and Oka, F.(1993). “An elasto-viscoplastic constitutive model for soft rock
with strain softening”. Proc., Int. Conf. on Hard Soils-Soft Rocks, 327-333.

Abdulla Ali A. and Kiousis Panos D. (1997). “Behavior of cemented sands-I. Testing”.
Int. J. Numer. Analyt. Methods Geomech., Vol. 21, 533-547.

Abdulla Ali A. and Kiousis Panos D. (1997). “Behavior of cemented sands-II.


Modelling ”. Int. J. Numer. Analyt. Methods Geomech., Vol. 21, 549-568.

Ahnberg, H., Ljungkrantz, C. and Holmqvist, L. (1995). “Deep Stabilization of Different


Types of Soft Soils”. Proc. 11th ECSMFE, Copenhagen, Vol. 7:7.167-7.172.

Airey D. W. (1993). “Triaxial testing of naturally cemented carbonate soil”. Journal of


Geotechnical Engineering, ASCE, Vol-119, No. 9, 1379-1398.

Al-Tabbaa, A. and Wood, D. M.(1989). “An experimentally based bubble model for
caly”. In Numerical models in geomechanics, Niagara Falls(edited by S.Pietruszczak
and G.N. Pande), Rotterdam, 91-99.

Anagnostopoulos, A., Kalteziotis, N., Tsiambaos, G.K. and Kavvadas, M. (1991).


“Geotechnical properties of the Corinth Canal marls”. Geotechnical and Geological
Engineering, Vol.9, 1-26.

Asaoka A., Nakano M. and Noda T.(2000). “Superloading yield surface concept for
highly structured soil behavior”. Soils and Foundations, Vol.40(2), 99-110.

Assarson, k., Broms, B., Granhom, S. and Paus, K.(1974). “Deep foundations and ground
improvement schemes”. proceedings on Geotextiles, Geomembranes and Other
Geosynthetics in Ground Improvement, Bangkok, 161-173.

Atkinson, J.H. and Richardson, D.(1987). “The effect of local drainage in shear zones on
the undrained strength of overconsolidated clay”. Geotechnique 37, No.3, 393-403.

337
REFERENCES

Atkinson, J.H. (1990). “Discussion of Session 1.2. Proceedings of the Conference on


Engineering Geology of Weak Rock”. Leeds, Cripps, J.C., Coulthard, J.M., Culshaw,
M.G., Forster, A.,Hencher, S.R. and Moon, C.F. (eds), Balkema, Rotterdam, 150.

Azman, Md., Amin Md., Ismail, B. and Hyde. A.F.L.(1994). “Deep foundations and
ground improvement schemes”. Proceedings on Geotexiles, Geomembranes and
Other Geosynthetics in Ground Improvement, Bangkok, 161-173.

Babasaki, R., Suzuki, K., Saitoh, S., Suzuki, Y., and Tokitoh, K.(1991). “Construction
and testing of deep foundation improvement using the deep cement mixing method”.
Deep Foundation Improvements: Dsign, Construction and Testing. ASTMSTP 1089,
M. I. Esrig and R.C. Bachus, eds. ASTM, Philadephia, 224-233.

Balasubramaniam, A.S., Kamruzzaman, A.H.M., Uddin, K., Lin, D.G., Phienwij, N. and
Bergado, D.T. (1998). “Chemical Stabilization of Bangkok Clay with Cement, Lime
th
and Flyash Additives”. Proc. 13 SEAGC, Taipei, 253-258.

Barden L. and McDermott Richard J. W. (1965). “ Use of free ends in triaxial testing of
clays”. Journal of the soil mechanics and foundation division, ASCE, Vol-91, No.
SM6, 1-23.

Barksdale, R. D. & Blight, G. E. (1997). “Compressibility and settlement of residual


soils”. In Mechanics of residual soils, (G. E. Blight ed.), Rotterdam: A. A. Balkema, 95–
154.

Baudet B. and Stallebrass S. (2004). “A constitutive model for structured clays”.


Geotechnique 54, No.4, 269-278.

Bazant Z.P.,Planas J.(1998). “Fracture and Size in Concrete and Other Quasi-brittle
Materials”. Boca Raton FL:CRC Press,291-291.

Bergado, D.T., Anderson, L.R., Miura, N. and Balasubramaniam, A.S. (1996). “Soft
Ground Improvement in Lowland and Other Environments”. ASCE press, New York.

Bergado, D. T., Ruenkrairergsa, T., Taesiri, Y., and Balasubramainam, A. S.(1999).


“Deep soil mixing to reduce embankment settlement”. Ground Improvement J., 3(3),
145-162.

338
REFERENCES

Black, D.K. and Lee, K.L. (1973). “Saturating laboratory samples by back pressure”. J.
Soil Mechanics & Foundation Division , ASCE, Vol. 99(SM1), 9484, 75-93.

Bressani, L.A. (1990). “Experimental properties of bonded soils”. PhD thesis, University
of London.

Broderic, G. P. and Daniel, D. E. (1990). “Stabilizing compacted clays against chemical


attack”. Journal of Geotechnical Engineering, ASCE, 116(10), 1549-1567.

Broms, B. B., 1984. “ Stablisation of soft clay with lime columns”. Proceedings, Seminar
on Soil Improvement and Construction Techniques in Soft Ground, NanYang
Technological Institute, Sigapore.

Bruce, D. A., Bruce, M. E. C. and DiMillio, A. F.(1998). “Deep mixing method: a global
perspective”. Geotechnical Special Publication No.81. Soil Improvement for Big Digs,
Proc. Sessions of Geo-Congree 98, Boston, 1-26.

BS1377 (1990). “British Standard Mehtods of Test for Soils for Civil Engineering
Purposes”. British Standard Institution, London.

BS1881 (1983). “Testing concrete-Part 117: Method for determination of tensile splitting
strength”. British Standard Institution, London.

Burland, J.B. (1990). “On the compressibility and shear strength of natural clays”.
Geotechnique, Vol. 29(3), 329-378.

Burland, J.B., Rampello, S., Georgiannou, V.N., and Calabresi, G. (1996). “A Laboratory
Study of the Strength of Four Stiff Clays’. Geotechnique, Vol. 46(3), 491-514.

Callisto L. and Rampello S.(2004). “ An interpretation of structural degradation for three


natural clays”. Can. Geotech. J.Vol-41, 392-407.

Callisto L. and Calabresi G. (1998). “Mechanical behavior of a natural soft clay”.


Getochnique 48, No.4, 495-513.

Cecconi, M., Viggiani, G. & Rampello, S. (1998). “An experimental investigation of the
mechanical behaviour of a pyroclastic soft rock”. Proc. 2nd Int. Symp. on Geotech.

339
REFERENCES

Engng of Hard Soils–Soft Rocks, Naples 1, 473–482.

Chazallon, C., and Hicher, P. Y.(1995). “An elasto-plastic model with damage for bonded
geomaterials”. Numerical Models in Geomechanics NUMOG-5 Conf., G.N. Pande
and S.Pietruszezak, eds., Balkema, Rotterdam,. The Netherlands, 21-26.

Chazallon, C., and Hicher, P. Y.(1998). “A constitutive model coupling elastoplasticity


and damage for cohesive-frictional materials”. Mechanics Cohesive-Frictional
Materials 3, 41-63.

Chandler, R.J.(2000). “Clay sediments in depositional basins: the geotechnical cycle”. Q.


J. Engng Geol. Hydrogeol, Vol.33, 7-39.

Chia, B.H., and Tan, T. S.(1993). “The use of jet grouting in the construction of drains in
soft soils”. Innovation in infrastructure development: Proc., 11th Conf. of ASEAN Fed.
of Eng. Organisations, 20-26.

Chiu C. F., Zhu W. and Zhang C.L. (2008). “Yielding and shear behaviour of cement-
treated dredged materials” . Engineering Geology, Vol. 103, 1-12.

Chua Tong Seng (1990). “ Some considerations in undrained triaxial testing of very soft
clay”. Master thesis, National University of Singapore, Singapore.

Chew S. H., Kamruzzaman, and Lee F. H. (2004). “Physicochemical and Engineering


behavior of cement treated clays”. Journal of Geotechnical and Geoenvironmental
Engineering, ASCE, 696-705.

Chin, K.G., Lee, F.H., and Dasari, G.R. (2004). “Effects of Curing Stress on Mechanical
Properties of Cement-Treated Soft Marine Clay”. Proc. Int. Symp. on Engineering
Practice and Performance of Soft Deposits, Osaka, 217-222.

Chin Kheng Ghee (2006). “constitutive behavior of cement treated marine clay”. PhD
thesis, National University of Singapore, Singapore.

COI (2005). “Report of the Committee of Inquiry into the Incident at the MRT Circle
Line Worksite that led to the Collapse of the Nicoll Highway on 20 April 2004”.
Ministry of Manpower, Singapore.

Consoli N. C., Rotta G. V. and Prietto P. D. M.(2000). “ Influence of curing under stress
on the triaxial response of cemented soils”. Geotechnique 50, No.1, 99-105.

340
REFERENCES

Consoli N. C., Rotta G. V. and Prietto P. D. M.(2006). “ Yielding-compressibility-


strength relationship for an artificially cemented soil under stress”. Geotechnique 56,
No.1, 69-72.

Coop M. R. and Atkinson J. H. (1993). “ The mechanics of cemented carbonate sands”.


Geotechnique 43, No.1, 53-67.

Cotecchia, F., and Chandler, R.J. (1997). “ The Influence of Structure on the Pre-Failure
Behaviour of a Natural Clay”. Geotechnique, Vol. 47(3), 523-544.

Crooks, J. H. A. and Graham, J.(1976). “Geotechnical properties of the Belfast estuarine


deposits”. Geotechnique 30, No. 2, 293-315.

Cuccovillos T. and Coop M. R. (1997). “Yielding and pre-failure deformation of


structured sands”. Geotechnique 47, No.1, 69-72.
Cotecchia, F., and Chandler, R.J. (2000). “A General Framework for the Mechanical
Behaviour of Clays”. Geotechnique, Vol. 50(4), 431-447.

Cuccovillos T. and Coop M. R. (1999). “On the mechanics of structured sands”.


Geotechnique 49, No.6, 741-760.

Deschenes, J., Massiera, M. and Tournier, J. (1995). “Testing of a Cement Bentonite Mix
for a Low Permeability Plastic Barrier” . In Dredging, Remediation, and Containment
of Contaminated Sediments, ASTM 1293, 252-270.

Desrues J. and Viggiani G.(2004). “Strain localization in sand: an overview of the


experimental results obtained in Grenoble using stereophotogrammetry”. Int. J.
Numer. Anal. Meth. Geomech., Vol. 28,.279–321.

Desrues J.and Chambon R.(2002). “Shear band analysis and shear moduli calibration”.
International Journal of Solids and Structures, Vol. 39, 3757–3776.

Desrues J.(1998). “Localisation Patterns in Ductile and Brittle Geomaterials”. In Material


Instabilities in Solids, Edited by René de Borst snd Erik van der Giessen, 137-1258.

341
REFERENCES

Dusseault, M.B. and Morgenstern, N.R.(1979). “Locked sands”. J. Engng Geol. 12, 117-
131.

Enami, A., Hibino, S., Takahashi, M., Akiya, K., and Yamada, M. (1986). “ Properties of
Soil Cement Columns Produced by Compact Machine System for Tenocolumn
Method”. Proc. 21st Annual Meeting of JSSMFE, Tokyo, 1987-1990. (in Japanese)

Endo, M. (1976). “Recent Development in Dredged Material Stabilization and Deep


Chemical Mixing in Japan”. Soil and Site Improvement, University of California,
Berkeley, lifelong learning seminar.

Gajo, A. and Muir Wood, D.(2001). “A new approach to anisotropic bounding surface
plasticity: general formation and simulations for natural and reconstituted clays”. Int.
J. Numer. Anal. Methods Geomech, Vol. 18, 314-333.

Gallavresi, F.(1992). “Grouting improvement of foundation soils”. Proc., Grouting, soil


improvement and geosynthetics, ASCE, New York, vol.1, 1-38.

Gens , A., and Nova, R.(1993). “Conceptual bases for a constitutive model for bonded
soils and weak rocks”. Proc., Int. Conf. on Hard Soils-Soft Rocks, 485-494.

Georgiannou V. N. and Burland J. B. (2001). “A laboratory study of post-rupture


strength”. Geotechnique 51, No. 8, 665-675.

Goto, S., Tatsuoka, F., Shibuya, S., Kim, Y.S. and Sato, T. (1991). “A Simple Gauge for
Local Small Strain Measurements in the Laboratory”. Soils & Found., Vol. 31(1),
169-180.

Harleston N. and Wang X. H.(2009). “Laboratory evaluation of cement reinforced and


cement-flyash-gravel (CFG) reinforced concrete pavement sections”. EJGE, Vol.14,
bundle B.

Head, K.H. (1986). “Manual of Soil Laboratory Testing”. ELE International Limited,
Pentech Press, London.

Hirai H., Takahashi M. and Yamada M. (1989). “An elastic-plastic constitutive model for
the behavior for improved sandy soils”. Soils and Foundations, Vol.29(2), 69-84.

342
REFERENCES

Horpibulsuk, S., Miura, N. and Nagaraj, T.S. (2003). “Assessment of Strength


Development in Cement-Admixed High Water Content Clays with Abrams Law as a
Basis”. Geotechnique, Vol. 53(4), 439-444.

Horpibulsuk S., Miura N. and Bergado D. T. (2004). “Undrained shear behavior of


cement admixed clay at high water content”. Journal of Geotechnical and
Geoenvironmental Engineering, ASCE, 1096-1105.

Horpibulsuk S., Liu M. D., Liyanapthirana D. and Suebsuk J. (2009). “Behavior of


cemented clay simulated via the theoretical framework”. Computers and Geotechnics,
in press.

Huang J. T. and Airey D. W. (1998). “ Properties of artificially cemented carbonate sand”.


Journal of Geotechnical and Geoenvironmental Engineering, ASCE, 492-499.

Huat B. B. K. (2006). “Effect of cement admixtures on the engineering properties of


tropical peak soils”. EJGE paper, Vol. 11, Bundle B.

Jardine, R.J.(1985). “Investigations of pile-soil behavior with special reference to the


foundation of offshore structures”. PhD thesis, University of London.

Jardine, R.J., St. John, H.D., Hight, D.W. and Potts, D.M. (1991). “Some practical
applications of a non-linear ground model”. Proceedings of the lOth European
Conference on Soil Mechanics, Florence, Vol. 1, pp. 223-8.

Jardine, R.J. (1992). “Some observations on the kinematic nature of soil stiffness”. Soils
and Foundations, 32, 111-24.

Jardine, R.J. (1995). “One perspective of the pre-failure deformation characteristics of


some geomaterials”. Proceedings of the International Symposium on Pre-Failure
Deformation Characteristics of Geomaterials, Hokkaido, Japan, Shibuya, S., Mitachi,
T. and Miura, S. (eds),Vol. 2, 855-885.

Jefferies, S.A. (1981). “Bentonite-Cement Slurries for Hydraulic Cutoff”. Proc. 10th
ICSMFE, Stockholm, Vol. 1, 435-440.

343
REFERENCES

Kaomon, M. and Bergado, D. T. (1991). “Ground improvement techniques”. Proceedings


of the 9th Asian Regional Conference on Soil Mechanics and Foundation Engineering
Bankok, 2, 526-546.
Kamruzzaman, A.H.M (2002). “Physico-Chemical & Engineering Behaviour of Cement
Treated Singapore Marine Clay”. PhD thesis, National University of Singapore,
Singapore.

Kamruzzaman A. H. M., Chew S. H., and Lee F. H. (2009). “Structuration and


Destructuration behavior of cement-treated Singapore marine clay”. Journal of
Geotechnical and Geoenvironmental Engineering, ASCE, Vol. 135, No. 4, 573-589.

Kasama K., Ochiai H. and Yasufuku N.(2000). “On the stress-strain behavior of lightly
cemented clay based on an extended critical state concept”. Soils and Foundations,
Vol.40, No.5, 37-47.

Kaushinger, J. L., Perry, E. B., and Hankour, R.(1992). “Jet grouting state of the practice”.
Proc., Grouting, soil improvement and geosynthetics: ASCE, New York, vol. 1, 169-
181.

Kavvadas, M., Anagnostopoulos, A., Leonardos, M. and Karras, B. (1993). “Mechanical


properties of Ptolemais lignite”. Proceedings of the International Symposium on the
Geotechnical Engineering of Hard Soils-Soft Rocks, Athens, Anagnostopoulos A.,
Schlosser, F., Kalteziotis, N. and Frank, R. (eds), Balkema, Rotterdam, Vol. 1, pp.
585-92.

Kavvadas, M., Papadopoulos, B. and Kalteziotis, N. (1994). “Geotechnical properties of


Ptolemais lignite”. Geotechnieal and Geological Engineering, 12, 87-112.

Kavvadas M. (1995). “A plasticity approach to the mechanical behavior of bonded soils”.


Proc. 4th Int. Conf. Computational Plasticity, Barcelona.

Kavvadas M. and Amorosi A.(2000). “A constitutive model for structured soils”.


Geotechnique 50, No. 3, 263-273.

Kawasaki, T., Niina, A., Saitoh, S., Suzuki, Y. and Honjo, Y. (1981). “Deep Mixing
Method using Cement Hardening Agent”. Proc. 10th ICSMFE, Vol. 3, 721-724.

344
REFERENCES

Kezdi, A. (1979). “Stabilized Earth Roads”. Development in Geotechnical Engineering,


Elsevier Scientific, New York.

Kilpatrick, B.L., and Garner, S.J. (1992). “Use of Cement Bentonite for Cut-Off Wall
Construction”. Proc. of Grouting, Soil Improvement, and Geosynthetics, ASCE
Geotechnical Publication 30, 803-815.

Koskinen, M., Karstunen, M. and Wheeler, S. J. (2002). “Modelling destructuration and


anisotropy of a soft natural clay”. Proc. 5th Eur. Conf. Numer. Methods Geotech.
Engng, Paris, 11-20.

Lee Kenneth L. (1978). “End restraint effects on undrained static triaxial strength of
sand”. Journal of Geotechnical Engineering, ASCE, Vol-104, No. GT6, 687-703.

Lagioia R. and Nova R. (1995). “An experimental and theoretical study of the behavior of
a calcarenite in triaxial compression”. Geotechnique 45, No. 4, 663-648.

Lagioia R. and Nova R. (1993). “A constitutive model for soft rocks”. Proc., Int. Conf. on
Hard Soils-Soft Rocks, 625-632.

Leddra, M.J., Jones, M.E. and Goldsmith, A.S. (1990). “Compaction and shear
deformation of a weakly-cemented, high porosity sedimentary rock”. Proceedings of
the Conference on Engineering Geology of Weak Rock, Leeds, Cripps, J.C.,
Coulthard, J.M., Culshaw, M.G., Forster, A., Hencher, S.R. and Moon, C.F. (eds)
Balkema, Rotterdam, pp. 45-54.

Lee F. H., Lee Yong, Chew S. H. and Yong K. Y.(2005). “Strength and modulus of
marine clay-cement mixes”. Journal of Geotechnical and Geoenvironmental
Engineering, ASCE, 178-185.

Lee F.H(2008). "How Useful is Numerical Analysis in Geotechnical Engineering?".


Proceedings of the International Conference on Deep Excavation (ICDE) 2008 , Land
Transport Authority, Singapore.

Lee K., Chan D. and Lam K. (2004). “Constitutive model for cement treated clay in a
critical state frame work”. Soils and Foundations, Vol. 44, No.3, 69-77.

345
REFERENCES

Lee Yeong(1999). “A framework for the design of Jet Grouting Piles in Singapore marine
clay”. Master thesis, National University of Singapore, Singapore.

Leroueil S. and Vaughan P. R. (1990). “The general and congruent effects of structure in
natural soils and weak rocks”. Geotechnique 40, No. 3, 467-488.

Leroueil S. (1998). “Contribution to the round table: peculiar aspects of structured soils”.
Proc. 2nd Symp. Geotech. Hard Soils-Soft Rocks, Napoli 3, 1669-1678.

Liu E. L. and Shen Z.J. (2005). “ Binary medium model for structured soils”. Journal of
Hydraulic Engineering, Vol. 36(4), 1-7. (in Chinese)

Liu M. D. and Carter J. P. (1999). “Virgin compression of structured soils”.


Geotechnique 49, No. 1, 43-57.

Liu M. D. and Carter J. P. (2000). “Modelling the destructuring of soils during virgin
compression”. Geotechnique 50, No. 4, 479-483.

Liu M. D. and Carter J. P. (2002). “A structured Cam Clay model”. Can. Geotech. J.Vol-
39, 1313-1332.

Liu MD, Carter JP, Horpibulsuk S, Liyanapathirana DS.. “Modelling the behaviour
of cemented clay”. In: Geo-Shanghai 2006; 65–72 [Geotechnical special
publication no. 152].

Lagioia, R., and Nova, R.(1995). “An experimental model for structured soils”.
Geotechnique, 50(3), 263-273.

Locat, J., Berube, M.-A., and Choquette, M. (1990). “ Laboratory Investigations on the
Lime Stabilization of Sensitive Clays Shear Strength Development”. Can. Geotech. J.,
Vol. 27, 294-304.

Locat, J. Tremblay, H., and Leroueil, S. (1996). “ Mechanical and Hydraulic Behaviour
of a Soft Inorganic Clay Treated with Lime”. Can. Geotech. J., Vol. 33, 654-669.

346
REFERENCES

Lorenzo G. A. and Bergado D. T. (2004). “ Fundamental parameters of cement-admixed


clay-new approach”. Journal of Geotechnical and Geoenvironmental Engineering,
ASCE, 1042-1050.

Low H. E., Phoon, K. K., Tan T.S., and Leroueil S.(2008). “Effect of soil microstructure
on the compressibility of natural Singapore marine clay”. Can. Geotech. J.45, 161-
176.

Maccarini, M. (1987). “Laboratory studies of a weakly bonded artificial soil”. PhD thesis,
University of London.

Malandraki, V. and Toll, D.G. (1994). “Yielding of a Weakly Bonded Artificial Soil”.
Proc. Int. Symp. Pre-failure Deformation Characteristics of Geomaterials, Sapporo,
Japan (ed. Shibuya, S., Mitachi, T. and Miura, S.), Rotterdam: Balkema, pp. 315-320.

Malandraki, V. and Toll, D.G. (1996). “The definition of yield for bonded materials”.
Geotechnical and Geological Engineering, Vol.14, 67-82.

Malandraki, V. and Toll, D. G. (2000). “Drained probing triaxial tests on a weakly


bonded artificial soil”. Geotechnique 50, 141-151.

Matsuoka H. and SUN, D. A. (1993). “A constitutive law for frictional and cohesive
materials”. JSCE, No. 463/III-22, 163-172(in Japanese).

Matsuoka H. and SUN, D. A. (1995). “Extension of spatially mobilized plane(SMP) to


frictional and cohesive materials and its application to cemented sands”. Soils and
Foundations, Vol. 35, No. 4, 63-72.

McDowell G. R. (2000). “A family of yield loci based on micro mechanics”. Soils and
Foundations, Vol. 40, No.6, 133-137.

McDowell G. R. and Hau .K. W. (2004). “ A generalized modified cam clay model for
clay and sand incorporating kinematic hardening and bounding surface plasticty”.
Granunlar Matter 6, 11-16.

Mitchell R. J. (1970). “On the yielding and mechanical strength of Leda clays”. Canadian
Geotechnical Journal, vol.7, 297-312.

347
REFERENCES

Mitchell, J.K. (1976). “Fundamentals of Soil Behavior”. Wiley, New York.

Mitchell, J.K. (1993). “Fundamentals of Soil Behavior”. 2nd Ed., Wiley, New York.

Miura N., Horpibulsuk S. and Nagagaj T. S. (2001). “Engineering behavior of cement


stabilized clay at high water content”. Soils and Foundations, Vol. 41, No.5, 33-45.

Mroz, Z., Norris, V.A. and Zienkiewicz, O. C.(1978). “An anisotropic hardening model
for soils and its application to cyclic loading”. Int. J. Num. Anal. Methods
Geomechanics 2, 203-221.
Mroz, Z., Norris, V. A. and Zienkiewicz, O. C.(1979). “Application of an anisotropic
hardening model in the analysis of elasto-plastic deformation of soils”. Geotechnique
29, No.1, 1-34.
Wood, D. M. (1990). “Soil Behaviour and Critical State Soil Mechanics”. Cambridge
University Press.
Muir Wood (1995). “Kinematic hardening model for structured soil”. Numerical Models
in Geomechanics-NUMOG V, Pande and Pietruzczak(eds), Balkema, Rotterdam, 83-
88.
Nagaraj. T.S., Pandain, N.S. and Narasimha Raju, P.S.R.(1998). “Compressibility
behavior of soft cemented soils”. Geotechnqiue, 48, NO.2, 281-287.

Nakamura, M., Matsuzawa, S., and Matsushita, M. (1982). “Study of the Agitation
Mixing of Improvement Agents”. Proc. 17th Japan National Conf. on SMFE, Vol. 2,
2585-2588. (in Japanese)
Namikawa T, Koseki J.(2006). “Experimental determination of softening relations for
cement-treated sand”. Soils and Foundations, Vol.46(4), 491–504.

Namikawa T. and Mihira S. (2007). “Elasto-plastic model for cement-treated sand”. Int. J.
Numer. Anal. Meth. Geomech., Vol. 31, 71-107.

Niina, A., Saitoh, S., Babasaki, R., Tsutsumi, I., and Kawasaki, T. (1977). “Study on
DMM Using Cement Hardening Agent (Part 1)”. Proc. 12th Japan National Conf. on
SMFE, 1325-1328. (in Japanese)

Nishibayashi, K. (1988). “Research on Low Strength Improvement Using the Deep


Mixing Method for the Purpose of Post-Improvement Excavation (Part 2)”. Proc.
23rd Annual Meeting of JSSMFE, Tokyo, 2297-2300. (in Japanese)

348
REFERENCES

Nishibayashi, K., Matsuo, T., Hosoya, Y., Hirai, Y., and Shima, M. (1985).
“Experimental Research into Mixing Devices Used for the Deep Mixing Method of
Soil Improvement (Part 4)”. Proc. 20th Annual Meeting of JSSMFE, Tokyo, 1747-
1750. (in Japanese)

Nova, R. (1988). “Sinfonietta classica: an exercise on classical soil modeling”. Proc.


Symp. Constitutive Eq. for Granular Non-Cohesive Soils, Rotterdam, 501-520.

Oka, F., and Adachi, T.(1985). “An elasto-plastic constitutive equation of geologic
material with memory”. Proc., 5th Int. Conf. on Numer. Methods in Geomechanics,
293-300.

Petchgate K., Voottipruex P. and Suknognkol W. (2000). “ Effect of height and diameter
ratio on strength of cement stabilized soft Bangkok clay”. Geotechnical Engineering
Journal, Vol. 31, No.3, 227-239.

Pitts, J. (1992). “Landforms and Geomorphic evolution of the Islands during the
Quaternary”. In: Physical Adjustment in a Changing Landscape: The Singapore Story,
A.Gupta and J.Pitts, eds., Singapore University Press, Singapore, 83-143.

Porbaha, A., Shibuya, S. and Kishida, T. (2000). “State of Art in Deep Mixing
Technology. Part III: Geomaterial Characterization”. Ground Improvement, Vol. 3,
91-110.

Prevost, J. H.(1978). “ Plasticity theory for soil stress-strain behavior”. J. Engng Mech.
Div. ASCE104, No.5, 1177-1194.

Rao, S.N., and Rajasekaran, G. (1996). “Reaction products formed in lime-stabilized


marine clays”. J. Geotech. Engrg., 122(5), 329-336.

Roscoe, K.H., and Burland, J.B. 1968. “On the generalized stress strain behaviour of ‘wet
clay”. In Engineering plasticity. Edited by J. Heyman and F.A. Leckie. Cambridge
University Press, 535–609.

Rotta G. V., Consoli N. C., Prietto P. D. M., Coop M .R. and Graham J.(2003). “Isotropic
yielding in an artificially cemented soil cured under stress”. Geotechnique 53, No. 5,
493-501.

349
REFERENCES

Rouania M. and Wood D. M. (2000). “A kinematic hardening constitutive model for


natural clays with loss of structure”. Geotechnique 50, No. 2, 153-164.

Rowe P. W. and Barden Laing(1964). “ Importance of free ends in triaxial testing”.


Journal of the soil mechanics and foundation division, ASCE, Vol-90, No. SM1, 1-27.

Saada A. S., Bianchin G. F. and Liang L.(1994). “ Cracks, bifurcation and shear bands
propagation in saturated clays”. Geotechnique 45, No. 2, 339-343.

Saitoh, S., Suzuki, Y., Shirai, K. (1985). “Hardening of Soil Improved by Deep Mixing
Method”. Proc. 11th ICSMFE, Vol. 5, 1745-1748.

Saitoh, S. (1988). “Experimental Study of Engineering Properties of Cement Improved


Ground by the Deep Mixing Method”. PhD thesis, Nihon University, Japan. (in
Japanese)

Saitoh, S., Shirai, K., Okumura, R., and Kobayashi, Y. (1990). “ Laboratory Experiments
on the Mixing Methods for Improvement of Sandy Soils”. Proc. 25th Annual Meeting
of JSSMFE, Tokyo, 1955-1958. (in Japanese)

Samieh A. M. and Wong R. C.K.(1997). “Deformation of Athabasca oil sand at low


effective stresses under varying boundary conditions”. Can. Geotech. J., Vol. 34,
985–990.

Schnaid, F., Prietto, P. D. M. and Consoli, N.C.(2001). “Characterization of cemented


sand in triaxial compression”. J. Geotech. Geoenviron. Engng. ASCE 127, No. 10,
857-868.

Shen, S.-L. (1998). “ Behaviour of Deep Mixing /Columns in Composite Clay Ground”.
PhD thesis, Saga University, Japan.

Shen Zhu-jiang (1993a). “A elasto-plastic damage model for cemented clay”. Chinese
Journal of Geotechnical Engineering, Vol.15(3), 21-28. (in Chinese)

Shen Zhu-jiang (1993b). “A nonlinear elastic damage model for structured clay”. Journal
of Nanjing hydraulic research institute, Vol.3, 247-255. (in Chinese)

350
REFERENCES

Shibuya, S., Tatsuoka, F., Teachavorasinskun, S., King, X.J., Abe, F., Kim, Y.S. and Park,
C.S. (1992). “ Elastic Deformation Properties of Geomaterials”. Soils & Found., Vol.
32(3), 26-46.

Smith P. R., Jardine R. J. and Hight D. W. (1992). “The yielding of Bothkennar clay”.
Geotechnique 42, No. 2, 257-274.

Stavridakis E. I. (2006). “Effect of curing time and bentonite content on the quantitative
evaluation of engineering behavior of cement treated clayed mixtures under soaked
conditions”. EJGE paper, Vol. 11, Bundle A.

Sugawara, S., Shigenawa, S., Gotoh, H., and Hosoi, T.(1996). “Large scale jet grouting
for prestrutting in soft clay”. Grouting and deep mixing: Proc. IS Tokyo’96, 2nd Int.
Conf. on Ground Improvement FGeosystems, 353-356.

Sun Dao Jun (2008). “ Effect of a newly developed lignosulphonate superplasticizer on


properties of cement pastes and mortars”. Master thesis, National University of
Singapore, Singapore.

Suzuki, K., Nichikawa, T., Hayashi, J. and Ito, S.(1981). “Approach by zeta potential on
the surface change of hydration of C3S”. Cement and Concrete Research, 11, 759-764.

Suzuki, M., Fujimoto, T., Yamamoto, T., and Okabayashi, S. (2004). “Strength Increase
of Cement-Stabilized Soil due to Initial Consolidation”. Proc. Int. Symp. on
Engineering Practice and Performance of Soft Deposits, Osaka, 211-216.

Taki, O. and Yang, D. (1991). “Soil Cement Mixed Wall Technique”. Geotechnical
Engineering Congress, ASCE, Special Publication, No. 27, 298-309.

Tatsuoka F. and Kobayashi A. (1983). “Triaxial strength characteristics of cement-treated


soft clay”. Proc. 8th ECSMFE, Vol. 8, No. 1, 421-426.

Tatsuoka, F. and Shibuya, S. (1991). “Deformation characteristics of soils and rocks from
feld and laboratory tests”. Proceedings of the Ninth Asian Regional Conference on Soil
Mechanics and Foundation Engineering, Southeast Asian Geotechnical Society, Bangkok,
Vol. 2, 53-114.

351
REFERENCES

Tatsuoka, F.,Uchida, K., Imai, K., Ouchi, T., and Kohata, Y. (1997). “Properties of
Cement Treated Soil in Trans-Tokyo Bay Highway Project”. Ground Improvement,
Vol. 1(1), 37-57.

Timoshenko and Goodier (1970). “Theory of elasticity”. Third Ed, McGraw-Hill, Japan,
122-123.

Terashi, M. (1997). “Theme Lecture: Deep Mixing Method – Brief State-of-Art”. Proc.
14th ICSMFE, Vol. 4, 2475-2478.

Tan S. T., Goh T. L. and Yong K. Y. (2002). “Properties of Singapore marine clays
improved by cement mixing”. Geotechnical Testing Journal, Vol. 25, No.4, 1-11.

Tan, T.S., Phoon, K.K., Lee, F.H., Tanaka, H., Locat, J., and Chong, P.T. (2003). “A
Characterisation Study of Singapore Lower Marine Clay”. In Characterisation and
Engineering Properties of Natural Soils, Tan et al. (eds.). Swets & Zeitlinger, Lisse, 429-
454.

Tan, S. L. (1983). “Geotechnical properties and Laboratory testing of soft soils in


Singapore”. Proc., 1st International Seminar on Construction Problems in soft soils,
Nanyang Technological Institute, Singapore, 1-47.

Uddin, K., Balasubramaniam A. S. and Bergado D .T. (1997). “Engineering Behavior of


Cement-Treated Bangkok Soft Clay”. Geotech. Eng., Vol. 28(1), 89-119.

Vatsala A., Nova R., and Murthy B. R. Srinivasa(2001). “Elastoplastic model for
cemented soils”. Journal of Geotechnical and Geoenvironmental Engineering, ASCE,
Vol. 127, 679-687.

Vaughan, P.R. (1988). “Characterising the mechanical properties of the in-situ residual
soil”. Proceedings of the 2nd International Conference on Geomechanics in Tropical
Soils, Singapore, Balkema, Rotterdam, Vol. 2, pp. 469-87.

Vaughan, P.R., Maccarini, M. and Mokhtar, S.M. (1988). “Indexing the engineering
properties of residual soil”. Quarterly Journal of Engineering Geology, 21, 61-84.

352
REFERENCES

Vaughan, P.R. (1990b). “Discussion of session 1.1, Proceedings of the Conference on


Engineering Geology of Weak Rock”. Leeds, Cripps, J.C., Coulthard, J.M., Culshaw,
M.G., Forster, A., Hencher, S.R. and Moon, C.F. (eds), Balkema, Rotterdam, p. 111.

Viggiani, G., Rampello, S. and Georgiannou, V.N.(1993). “Experimental analysis of


localization phenomenon in triaxial tests on stiff clays”. Proc. 1st Symp. Geotech.
Hard Soils-Soft Rocks, Athens 2, 849-856.

Wheeler, S. J.(1997). “A rotational hardening elasto-plastic model for clays”. In


Proceedings of the 14th International Conference on Soil Mechanics and Foundation
Engineering, Hamburg, A.A. Balkema, Rotterdam, Vol.1, 431-434.

Wissa, A.E.Z., Ladd, C.C. and Lambe, T.W. (1965). “Effective Sress Strength Parameters
of Stabilized Soils”. Proc. 6th ICSMFE, Montreal, Vol. 1, 412-416.

Wong P. K. K. and Mitchell R. J. (1975). “Yielding and plastic flow of sensitive


cemented clay”. Geotechnique 25, No. 4, 763-782.

Wong R. C. K.(1999). “Mobilized strength components of Athabasca oil sand in triaxial


compression”. Can. Geotech. J., Vol. 36, 718–735.

Wood D. M. (1990). “Soil behavior and critical state soil mechanics”.150-153,


Cambridge University Press.

Wood D. M.(1995b). “Kinematic hardening model for structured soil”. Proceedings of


the international symposium on numerical models in geomechanics, Davos, 83-88.

Wood D. M. (2004). “Experimental inspiration for kinematic hardening soil models”.


Journal of Engineering Mechanics, ASCE, 656-664.

Yin, J.H., and Lai, C.K. (1998). “Strength and Stiffness of Hong Kong Marine Deposits
Mixed with Cement.”. Geotech. Eng., Vol. 29(1), 29-44.

Yoshizawa, H., Okumura, R., Hosoya, Y., Sumi, M., Yamada, T. (1996). “JGS Technical
Committee Report: Factors affecting the quality of treated soil during execution of
DMM”. Proc. 2nd Int. Conf. on Ground Improvement Geosystems, Grouting and
Deep Mixing, Tokyo, Vol. 2, 931-937.

353
REFERENCES

Yu, Y., Pu, J. and Ugai, K. (1997). “Study of Mechanical Properties of Soil Cement
Mixture for a Cutoff Wall”. Soils & Found., Vol. 37(4), 93-103.

Yu, Y.Z., Pu, J.L. and Ugai, K. (1998). “A damage model for soil-cement mixture”. Soils
and Foundations, 38(3), 1-12.

Yu, Y. Z, Pu, J. L., Ugai, K. and Takashi, H.(1999). “A study of the permeability of soil-
cement mixture”. Soils and Foundations, 39(5), 145-149.

Yong, D. M., Hayashi, K., and Chia, B. H.(1996). “Jet grouting for the construction of a
RC canal in soft marine clay”. Grouting and deep mixing: Proc. IS Tokyo’96, 2nd Int.
Conf. on Ground Improvement Geosystems, 375-380.

Yong, K. Y., Karunaratne, G. P., and Lee, S. L.(1990). “Recent developments in soft clay
engineering in Singapore”. Proc. Kansai Int. Geotech. Forum’90: Comparative
Geotechnical Engineering. Osaka, Japan, 1-8.

Zhao X. H., Sun H. and Lo K.W.(2002). “An elastoplastic damage model for soil”.
Geotechnique 52, No.7, 533-536.

354

You might also like