You are on page 1of 190

CHARACTERIZATION AND MODELING OF CEMENT-

TREATED SOIL COLUMN USED AS CANTILEVER

EARTH RETAINING STRUCTURE

SAW AY LEE

NATIONAL UNIVERSITY OF SINGAPORE

2014
CHARACTERIZATION AND MODELING OF CEMENT-

TREATED SOIL COLUMN USED AS CANTILEVER

EARTH RETAINING STRUCTURE

SAW AY LEE

(B. Eng. (Hons.), UTM; M. Sc., NUS)

A THESIS SUBMITTED

FOR THE DEGREE OF DOCTOR OF PHILOSOPHY

DEPARTMENT OF CIVIL AND ENVIRONMENTAL

ENGINEERING

NATIONAL UNIVERSITY OF SINGAPORE

2014
DECLARATION

I hereby declare that the thesis is my original work and it has been written by me in its

entirety. I have duly acknowledged all the sources of information which have been used in the

thesis.

This thesis has also not been submitted for any degree in any university previously.

_________________

Saw Ay Lee

3 January 2014
Acknowledgement

First of all, I am grateful to The Almighty God for leading me through this research. Without

His guidance this dissertation would not have been possible.

I would like to express the deepest appreciation to my supervisors, Professor Leung Chun Fai

and Associate Professor Tan Siew Ann for their patient guidance, encouragement and useful

critiques of this research work. I am extremely blessed by their unconditional support. I also

thank the research scholarship as well as facilities provided by the National University of

Singapore to make this research a reality.

Special thanks to Mr. Ang Beng Oon and Mr. Foo Hee Ann who patiently helped me in my

laboratory tasks. I am also grateful to Muhammad Faizal and Dr. Xiao Huawen for their

willingness to share the material and equipment during the time I conducted my experimental

works. A special thank you is also extended to Mr. Ann Kee Tong and Mr. Edward Lim for

their help in offering me the resources. Grateful acknowledgement is expressed to Dr.

Namikawa for enlightening me on the tension-softening behavior of cement-treated soil. I

would also like to thank the fellow colleagues, in particular Hartono, Kok Shien, Yang Yu,

Zongrui, Junhui, Xuguang, Sun Jie, Sandi for their company through this journey.

Also, I would like to thank my good friends for their support and continuing belief in me. Last

but not the least, an honorable mention goes to my family for always being there for me.

i
Table of Contents
Acknowledgement ...................................................................................................................... i
Table of Contents ……………………………………………………………………………...ii
Summary .................................................................................................................................... v
List of Tables ........................................................................................................................... vii
List of Figures........................................................................................................................... ix
List of Symbols ...................................................................................................................... xvii
Abbreviations.......................................................................................................................... xxi
Chapter 1 Introduction
1.1 Background ................................................................................................................ 1
1.2 Issues Related to the Use of Cement-treated Soil Columns ...................................... 2
1.3 Objective and Scope of Study.................................................................................... 4
1.4 Structure of Thesis ..................................................................................................... 5
Chapter 2 Literature Review
2.1 Introduction ............................................................................................................... 9
2.2 General Aspects of Cement-treated Soil.................................................................... 9
2.2.1 Physical Properties of Cement-Treated Soil .................................................... 11
2.2.2 Mechanical Properties of Cement-Treated Soil ............................................... 13
2.3 Existing design approaches...................................................................................... 18
2.4 Numerical Modeling on Behavior of Cement-treated Soil ...................................... 23
2.4.1 2-D and 3-D Finite Element Analysis.............................................................. 24
2.4.2 Constitutive Models for Cement-treated Soil .................................................. 26
2.5 Summary .................................................................................................................. 29
Chapter 3 Fracture Behavior of Cement-treated Singapore Marine Clay
3.1 Introduction ............................................................................................................. 46
3.2 Properties of Base Materials .................................................................................... 47
3.2.1 Untreated Marine Clay .................................................................................... 47
3.2.2 Ordinary Portland Cement ............................................................................... 48
3.3 Sample Preparation Procedure ................................................................................. 48
3.4 Testing Procedure and Apparatus ............................................................................ 50
3.4.1 Uniaxial Compression Test.............................................................................. 51
3.4.2 Split Tension Test ............................................................................................ 51
3.4.3 Three-point Bending Notched Beam Test ....................................................... 52
3.5 Compressive Fracture Behavior .............................................................................. 53
3.6 Tensile Fracture Behavior........................................................................................ 55
3.6.1 Split Tensile Strength ...................................................................................... 56
3.6.2 Fracture Energy, Gf .......................................................................................... 58

ii
3.6.3 Tension-softening Relationship ....................................................................... 62
3.6.4 Parametric Studies ........................................................................................... 67
3.6.5 Outstanding Issues ........................................................................................... 70
3.7 Summary .................................................................................................................. 74
Chapter 4 Constitutive Model for Cement-treated Soil
4.1 Introduction ............................................................................................................. 94
4.2 Finite Element Method ............................................................................................ 94
4.3 Constitutive Models for Cement-treated Soils......................................................... 96
4.3.1 Elastic Perfectly-Plastic Tresca Model ............................................................ 96
4.3.2 Isotropic Model................................................................................................ 99
4.3.3 Concrete Damage Plastic Model.................................................................... 100
4.4 Evaluation of Constitutive Model Prediction for Behavior of Cement-treated
Toyoura Sand..................................................................................................................... 105
4.4.1 Drained Triaxial Compression Test by Namikawa (2006) ............................ 106
4.4.2 Direct Tension Test by Koseki et al. (2005) .................................................. 109
4.4.3 Three-point Bending Notched Beam Test by Namikawa (2006) .................. 112
4.5 Evaluation of Constitutive Model Prediction for Behavior of Cement-treated
Singapore Marine Clay ...................................................................................................... 116
4.5.1 Uniaxial Compression Test (UCT) ................................................................ 117
4.5.2 Three-point Bending Notched Beam Test ..................................................... 119
4.6 Conclusions ........................................................................................................... 121
Chapter 5 Numerical Approaches in Simulating Cemented Soil Mass
5.1 Introduction ........................................................................................................... 138
5.2 Weighted Average Simulation (WAS) and Real Allocation Simulation (RAS)
Approaches ........................................................................................................................ 139
5.3 Evaluation of WAS and RAS Approaches in Simulating Cement-treated Soil Mass
…………………………………………………………………………………….141
5.3.1 Test description.............................................................................................. 141
5.3.2 Numerical Analyses ....................................................................................... 142
5.3.3 Results and Discussion .................................................................................. 147
5.4 Study of Cement-treated Ground Improvement Pattern ........................................ 151
5.4.1 Hypothetical Cases ........................................................................................ 153
5.4.2 Numerical Analyses ....................................................................................... 155
5.4.3 Results and Discussions ................................................................................. 161
5.5 Conclusion ............................................................................................................. 167
Chapter 6 Field Studies
6.1 Introduction ........................................................................................................... 192

iii
6.2 Field Case Study 1: Lateral Load Test by Babasaki et al. (1997).......................... 193
6.2.1 Test Description ............................................................................................. 193
6.2.2 Numerical Analyses ....................................................................................... 195
6.2.3 Results and Discussions ................................................................................. 198
6.3 Field Case Study 2: Waterway Construction in Northeastern Singapore .............. 201
6.3.1 Characteristics of Site .................................................................................... 201
6.3.2 Construction Method ..................................................................................... 202
6.3.3 Numerical Analyses ....................................................................................... 204
6.3.4 Results and Discussions ................................................................................. 208
6.4 Field Case Study 3: Basement Construction in Central Singapore........................ 210
6.4.1 Characteristics of Site .................................................................................... 211
6.4.2 Construction Method ..................................................................................... 212
6.4.3 Numerical Analyses ....................................................................................... 213
6.4.4 Results and Discussions ................................................................................. 215
6.5 Conclusion ............................................................................................................. 217
Chapter 7 Conclusions
7.1 Summary of Findings ............................................................................................ 238
7.2 Recommendations for Further Study ..................................................................... 242
References …………………………………………………………………………...…….243

iv
Summary

Owing to presence of soft soil which covers at least one quarter of land area of Singapore, it is

often necessary to improve the soft soil for various construction purposes particularly for

excavation works. One such ground improvement technique is to improve the soft soil with

cement to increase the in-situ strength and stiffness. The treatment may be conducted at great

depth as embedded struts to support deep excavation or at shallower depth not far below

ground level to support an open excavation work. However, relatively little studies had been

performed to study the behavior of cement-treated wall in an open excavation. This research

covers experimental studies to investigate behavior of cement-treated marine clay and

numerical studies to examine the behavior of cement-treated soil columns used as a retaining

system in an open cut excavation.

The first objective of this study is to understand the material properties of cement-treated

Singapore marine clay in terms of compression and tension behavior. A series of samples

with different mix proportions and curing periods was tested by different means in the

laboratory. The experimental results show that the material strength increases rapidly to a

peak value and then decreases abruptly to a small value upon further straining. The tensile

strength of this material is found to be 11% of its unconfined compressive strength in the

cement content tested in the present study. This material becomes brittle when 20% of cement

content is added and cured for 14 days. The post-peak softening of the treated clay was

derived based on fracture mechanics concept. Numerical calibration analyses were carried out

to evaluate the appropriateness of three available constitutive models based on laboratory test

results and published data. The calibration results show that the concrete damage plasticity

(CDP) model is superior to Tresca and isotropic models in simulating the behaviors of

cement-treated soil in compression, tension and bending.

v
This thesis further examines the significance of modeling the cement-treated soil column

configurations in finite element analysis. This is of importance to account for the localized

overlap areas of treated soil columns and the interaction between treated and untreated soil. In

practice, cement-treated soil mass with certain columnar shaped treated soil are often

analyzed using weighted average simulation (WAS) approach with properties that are

averaged over the treated area. In this thesis, the limitations of generalizing the properties

based on comparison with the results of a published laboratory test were discussed. The

analysis results show that the shortcomings of this approach can be overcome by the real

allocation simulation (RAS) approach with CDP model.

The established numerical approach was then adopted to study hypothetical cases of a vertical

cut with three different ground improvement geometrical patterns. The study demonstrates

that tensile damage in the cement-treated soil columns is the trigger for failure. The ability of

this recommended numerical simulation method – RAS with CDP model is examined by

back-analyzing three field case studies on cement-treated soil columns failure. The numerical

approach provides a fair prediction of the ground response and failure pattern compared to

field observations.

vi
List of Tables

Table 2.1 Summary of selected E-qu relationships for cement-treated clay.


Table 2.2 Summary of some t – qu relationships of cemented soil.
Table 3.1 Basic properties of Singapore marine clay from Marine Bouvelard site.
Table 3.2 Chemical composition and physical properties of Ordinary Portland Cement
provided by supplier.
Table 3.3 Mixture proportions and tests conducted for cement-treated Singapore marine
clay in the present study.
Table 3.4 Summary of fracture energies, Gf from BNT tests.
Table 3.5 Nominal size for materials used in the present study.
Table 3.6 Mixture proportions and experimental tests with addition of filter sand in clay.
Table 3.7 Chemical compositions of filter sand provided by supplier.
Table 3.8 Summary fracture energy Gf for cemented clay and cemented clay+sand.
Table 4.1 Mixing proportions of cement-treated Toyoura sand.
Table 4.2 Design parameters for Tresca model.
Table 4.3 Design parameters for Isotropic model.
Table 4.4 Design parameters for CDP model.
Table 4.5 Summary results of three-point bending notched beam test by Namikawa (2006).
Table 4.6 Tensile design parameters for CDP model in Case T2.
Table 4.7 Mixing proportions of cement-treated Singapore marine clay.
Table 4.8 Calibration parameters for cemented Singapore marine clay using CDP model.
Table 4.9 Damage variables used in CDP model for cemented Singapore marine clay Test
B1.
Table 5.1 Tresca design parameters for WAS approach in Test 2.
Table 5.2 CDP design parameters of lime-cement-treated soil columns for RAS approach
in Test 2.
Table 5.3 CDP damage variables for lime-cement-treated soil columns in Test 2.
Table 5.4 Summary table of hypothetical cases.
Table 5.5 Design parameters for Mohr Coulomb and Tresca models in hypothetical cases.
Table 5.6 CDP model design parameters for cement-treated soil columns used in
hypothetical cases.
Table 5.7 CDP model damage variables for cement-treated soil columns used in
hypothetical cases.
Table 6.1 Model parameters for steel plate, concrete cap and untreated soil (reproduced
from Namikawa et al., 2008).

vii
Table 6.2 Summary of case analyses based on the in-situ strength of cemented soil column.
Table 6.3 Model parameters for CDP model of cemented column.
Table 6.4 Summary table of analysis cases for case study 2.
Table 6.5 Design parameters for Mohr Coulomb and Tresca models used in case study 2.
Table 6.6 Design parameters for case study 3.

viii
List of Figures

Figure 1.1 Cement-treated soil columns used in Lexington, Virgina, USA to facilitate an
open excavation (after Ruffing et al. (2012).
Figure 1.2 Failure case of adopting cement-treated soil columns as a retaining system for
an open excavation in Northeastern Singapore.
Figure 1.3 Examples of soil improvement pattern.
Figure 2.1 Change of water content by in-situ cement treatment in Tokyo Port (after
Kawasaki et al., 1978).
Figure 2.2 Effect of curing time and cement content on final water content of treated clay
(after Kamruzzaman, 2002).
Figure 2.3 Change of density for in-situ cement treatment (after Japan Cement Association,
1994).
Figure 2.4 Effect of cement content on the permeability of cement-treated clay (after
Kawasaki et al., 1981a).
Figure 2.5 Comparison of stress-strain curve obtained by the conventional method
(external strain) and local transducer approach (local strain).
Figure 2.6 Variation of initial modulus at small strain with confining pressure (after
Shibuya et al., 1992).
Figure 2.7 Effect of cement content on stress-strain behavior of treated clay by unconfined
compression test (after Kamruzzaman, 2002).
Figure 2.8 Comparison of stress-strain curves for (a) with and without curing stress; (b)
CIU and UCT test (after Chin, 2006).
Figure 2.9 Peak and residual strengths of cemented soil under different confining
pressures (after Cement Deep Mixing Association of Japan, 1994).
Figure 2.10 Relationship between DST and UCT results (after Saitoh et al., 1980).
Figure 2.11 Effect of cement content and curing time on effective shear strength parameters
of treated clay (after Kamruzzaman, 2002).
Figure 2.12 Effective stress path of triaxial test for treated soil (after Kamruzzaman, 2002).
Figure 2.13 Stress-strain curves of direct tension test. Left: light cemented sand by Das and
Dass, 1995; Right: cemented Toyoura sand by Koseki et al. (2005).
Figure 2.14 Tension-softening relation for cemented Toyoura sand (after Namikawa 2006).
Left: BNT setup; Right: Tensile stress-opening crack displacement relation.
Figure 2.15 Arrangement of typical improvement pattern for columnar treated soil (after
CDIT, 2002).
Figure 2.16 Schematic for computing the area replacement ratio, (after CDIT, 2002).

ix
Figure 2.17 Mobilized shear strength of treated column and soil (after CDIT, 2002).
Figure 2.18 Failure modes of single columns suggested by Kivelo (1998).
Figure 2.19 Failure modes of treated soil columns proposed by Kitazume (2008).
Figure 2.20 Estimated tilting pattern of treated soil columns in collapse failure mode (after
Kitazume, 2008).
Figure 2.21 Pressure distribution diagram used for moment equilibrium check in collapse
failure mode (after Kitazume, 2008).
Figure 2.22 Induced tensile stress in the treated column (after Kitazume, 2008)
Figure 2.23 Pressure distribution diagram used for moment equilibrium check in bending
failure mode (after Kitazume, 2008).
Figure 2.24 Hypothetical excavation model to compare RAS and WAS by Ou and Wu
(1996).
Figure 2.25 3-D FE model of an embedded improved soil raft using RAS approach (after
Yang, 2009).
Figure 2.26 Comparison of test data and calibrated stress-strain curve with isotropic model
in Abaqus by Tjahyono (2011).
Figure 2.27 Numerical calibrated model for axial stress-strain results of uniaxial
compression tests on lime-cement columns using CDP model (after Larsson et
al., 2012).
Figure 3.1 Soil-cement and water-cement ratios from previous studies (reproduced from
Lee et al., 2005) and the present study.
Figure 3.2 Working range of cement-clay mixes from previous studies (reproduced from
Lee et al., 2005) and mix proportion for the present study.
Figure 3.3 Left: Experimental test setup for uniaxial compression test; Right:
Experimental test setup for split tension test.
Figure 3.4 Experimental test setup for three-point bending notched beam test.
Figure 3.5 The UCT stress-strain results of cement-treated Singapore marine clay for 5
samples with Aw = 30% at 28 days curing period.
Figure 3.6 The UCT stress-strain curves of cement-treated Singapore marine clay at 14
days curing period for various cement contents.
Figure 3.7 The UCT stress-strain curves of cement-treated Singapore marine clay at 28
days curing period for various cement contents.
Figure 3.8 The UCT stress-strain curves of cement-treated Singapore marine clay with
Aw=20% at three curing periods.
Figure 3.9 Failure modes of specimen after uniaxial compression test.
Figure 3.10 Unconfined compressive strength of cement-treated Singapore marine clay for
various cement contents and curing periods.

x
Figure 3.11 Variation of split tensile strength versus cement content for three curing
periods for cement-treated Singapore marine clay.
Figure 3.12 Typical tested specimens of cement-treated Singapore marine clay from split
tension test. Left: Front view of testing specimen when failure occurred; Right:
Specimen split into two halves.
Figure 3.13 Relationship between unconfined compressive strength and split tensile
strength for the present cement-treated Singapore marine clay.
Figure 3.14 A stress-strain example from direct tension test for concrete.
Figure 3.15 a) Deformation properties of the material outside the fracture zone: - ; b)
Deformation properties of the fracture zone: - c.

Figure 3.16 Schematic load-deflection curve from a BNT and the corresponding complete
curve when the specimen weight is taken into account.
Figure 3.17 Specimen dimension used in three-point bending notched beam test.
Figure 3.18 BNT test results for Aw = 30% at 14 days curing period: (Left) Load-deflection
curves and (Right) Load-crack mouth opening displacement curves.
Figure 3.19 BNT load-deflection curves of cement-treated Singapore marine clay at 14
days curing period.
Figure 3.20 BNT load-crack mouth opening displacement curves of cement-treated
Singapore marine clay at 14 days curing period.
Figure 3.21 BNT load-deflection relationship of cement-treated Singapore marine clay at
28 days curing period.
Figure 3.22 BNT load-crack mouth opening displacement relationship of cement-treated
Singapore marine clay at 28 days curing period.
Figure 3.23 Fracture zone and the hypothesis stress distribution (modified from Petersson,
1981) in front of the notch tip.
Figure 3.24 J-integral contours around crack tip (left) and a typical tension-softening curve
(right) (after Li and Ward, 1989).
Figure 3.25 Li’s J-integral method (after Rokugo et al., 1989).
Figure 3.26 F - v and c0 - v curves for modified J-integral method by Rokugo et al.
(1989).
Figure 3.27 Fictitious crack width distribution in the modified J-integral method by Uchida
et al. (1991).
Figure 3.28 Tensile stress-crack mouth opening displacement relationship of cement-
treated Singapore marine clay for Test A1 (Aw = 25%).
Figure 3.29 Tensile stress-crack opening displacement relationship of cement-treated
Singapore marine clay for Tests A1 and B1.

xi
Figure 3.30 BNT results for specimen notch widths 0.7 mm and 2.5 mm. Left: Load-
deflection curves; Right: Load-crack mouth opening displacement curves.
Figure 3.31 BNT results for specimen size db = 50 mm and db = 40 mm. Left: Load-
deflection curves; Right: Load-crack mouth opening displacement curves.
Figure 3.32 Stability criterion in a load-deflection test corresponding to the stiffness of
testing machine (after Hillerborg, 1989).
Figure 3.33 UCT stress-strain curves of cement-treated Singapore marine clay with and
without sand (at 28 days curing periods unless otherwise stated).
Figure 3.34 BNT load-deflection curves of cement-treated Singapore marine clay with and
without sand (at 28 days curing periods unless otherwise stated).
Figure 3.35 BNT load-crack mouth opening displacement curves of cement-treated
Singapore marine clay with and without sand (at 28 days curing periods unless
otherwise stated).
Figure 4.1 Typical monotonic stress-strain behavior assumed in Tresca model.
Figure 4.2 Tresca failure criterion in a 3-D stress space.
Figure 4.3 Isotropic model: evolution of the yield surface in 2-D (left) and 3-D (right)
principal stress space.
Figure 4.4 Two ways of modeling crack in finite element analysis (after Pankaj, 1990).
Figure 4.5 Yield surfaces for CDP model in the deviatoric plane, corresponding to
different values of (modified from Abaqus 6.11, 2011).
Figure 4.6 Yield surface for CDP model in plane stress (after Abaqus 6.11, 2011).
Figure 4.7 Definition of cracking strain ̃ used in tension data (after Abaqus 6.11, 2011).
Figure 4.8 Definition of compressive inelastic (or crushing) strain ̃ used in compression
data (after Abaqus 6.11, 2011).
Figure 4.9 Drained triaxial compression test for cemented Toyoura sand by Namikawa
(2006).
Figure 4.10 Loading and boundary conditions simulated in finite element models.
Figure 4.11 Calibrated stress-strain curves for three constitutive models and laboratory
measurement for drained triaxial compression test of cemented Toyoura sand.
Figure 4.12 Direct tension test for cemented Toyoura sand by Koseki et al. (2005).
Figure 4.13 Classical and tension truncated for Tresca criteria (after Antão et al., 2007).
Figure 4.14 Projection of Rankine surface in deviatoric plane.
Figure 4.15 Post-failure stress-fracture energy curve simulated by CDP model (after
Abaqus 6.11, 2011).
Figure 4.16 Calibrated tensile stress-strain curves for three constitutive models and
laboratory measurement for direct tension test of cement-treated Toyoura sand.

xii
Figure 4.17 Three-point bending notched beam test for cement-treated Toyoura sand by
Namikawa (2006).
Figure 4.18 Load-deflection curves of three-point bending notched beam test results for
cement-treated Toyoura sand by Namikawa (2006).
Figure 4.19 Tension-softening relation for cement-treated Toyoura sand by Namikawa
(2006).
Figure 4.20 Finite element meshes for simulating the three-point bending notched beam test:
(a) coarser mesh with 5,353 elements; and (b) finer mesh with 9,866 elements.
Figure 4.21 Load-deflection curves for two different mesh sizes in simulating the three-
point bending notched beam test for cement-treated Toyoura sand.
Figure 4.22 FEM input tension-softening relation in Case T2 for cement-treated Toyoura
sand.
Figure 4.23 Calibrated load-deflection curves and laboratory measurement for three-point
bending notched beam test of cement-treated Toyoura sand.
Figure 4.24 Tensile damage process happened around the notch tip observed in FE analysis
Case T2.
Figure 4.25 Schematic experiment setup and stress-strain curves for UCT of cement-treated
Singapore marine clay.
Figure 4.26 FEM model and stress-strain curves for UCT of cement-treated Singapore
marine clay.
Figure 4.27 Schematic diagram of three-point bending notched beam test B1 of cemented
Singapore marine clay (unit: mm).
Figure 4.28 Load-deflection curves of three-point bending notched beam test B1 for
cement-treated Singapore marine clay.
Figure 4.29 Finite element mesh for simulating the three-point bending notched beam Test
B1.
Figure 4.30 Calibrated load-deflection curves and laboratory measurement for three-point
bending notched beam Test B1 for cement-treated Singapore marine clay.
Figure 4.31 Post-test damage observed in numerical (Top) and laboratory (Bottom) for
three-point bending notched beam Test B1.
Figure 5.1 Schematic of lime-cement columns in shear box test (after Larsson, 1999).
Figure 5.2 (a) Mobilized shear stresses-lateral deformations for Tests 1 and 2; and (b)
damage pattern of the lime-cement-treated soil columns observed in Test 2
(after Larsson et al., 2012).
Figure 5.3 UCT results for lime-cement-treated columns in Test 2 (after Larsson et al.,
2012).

xiii
Figure 5.4 Finite element model for Test 1. Left: Finite element mesh; Right: Boundary
and loading conditions.
Figure 5.5 Finite element meshes for Refined Mass (RM) model in WAS approach.
Figure 5.6 Finite element meshes for RAS approach.
Figure 5.7 Design parameters based on UCT results for lime-cement-treated columns in
Test 2.
Figure 5.8 Two configurations of treated mass for WAS approach in Test 2.
Figure 5.9 Calibrated analysis result of mobilized shear stress and lateral deformation for
Test 1 without treated soil columns.
Figure 5.10a Analysis result of mobilized shear stress and lateral deformation for Test 2 with
WAS approach.
Figure 5.10b Yielding zone appeared in the models for Test 2 with WAS approach.
Figure 5.11 Analysis result of mobilized shear stress and lateral deformation for Test 2 with
WAS and RAS approaches.
Figure 5.12 Schematic failure in treated soil columns for RAS approaches using Tresca
model.
Figure 5.13 Schematic failure in treated soil columns for RAS approaches using CDP
model.
Figure 5.14 Plan view of improvement patterns: a) Grid type (GD); b) Tangential buttress
type (TN); and Double-wall type (DW).
Figure 5.15 Two configurations of WAS approach for grid type ground improvement
pattern.
Figure 5.16 Typical boundary condition and geometry for hypothetical case.
Figure 5.17 Finite element mesh for vertical cut without improvement.
Figure 5.18 Finite element mesh for Whole Mass (WM) model in WAS approach.
Figure 5.19 Finite element mesh for Refined Mass (RM) model in WAS approach.
Figure 5.20 Finite element mesh for grid type (GD) improvement pattern, L1.
Figure 5.21 Finite element mesh for tangential buttress type (TN) improvement pattern, L2.
Figure 5.22 Finite element mesh for double-wall type (DW) improvement pattern, L3.
Figure 5.23 Progressive ground movement, U (in m) for vertical cut without soil
improvement (Case Li).
Figure 5.24 Progressive development of plastic strain (PE) for vertical cut without soil
improvement (Case Li).
Figure 5.25 Ground response for vertical cut with grid type soil improvement pattern with
WAS approach (Left: Case La; Right: Case Lb).
Figure 5.26 Plastic strain (PE) for vertical cut with grid type soil improvement pattern with
WAS approach (Left: Case La; Right: Case Lb).

xiv
Figure 5.27 Ground response for GD type with RAS approach: Case L1-1.
Figure 5.28 Ground response for GD type with RAS approach: Case L1-2.
Figure 5.29 A close-up plastic strain developed in soil and column for Case L1-2.
Figure 5.30 Ground response for GD type with RAS approach: Case L1-3.
Figure 5.31 A close-up plastic strain developed in soil and damage occurred in cemented
column for Case L1-3.
Figure 5.32 Progressive tensile crack development in cemented column for Case L1-3.
Figure 5.33 Ground response for TN type with RAS approach: Case L2-1.
Figure 5.34 Ground response for TN type with RAS approach: Case L2-2.
Figure 5.35 A close-up plastic strain developed in soil and column for Case L2-2.
Figure 5.36 Ground response for TN type with RAS approach: Case L2-3.
Figure 5.37 A close-up plastic strain developed in soil and damage occurred in cemented
column for Case L2-3.
Figure 5.38 Progressive tensile crack development in cemented column for Case L2-3.
Figure 5.39 Ground response for DW type with RAS approach: Case L3-1.
Figure 5.40 Ground response for DW type with RAS approach: Case L3-2.
Figure 5.41 A close-up plastic strain developed in soil and column for Case L3-2.
Figure 5.42 Ground response for DW type with RAS approach: Case L3-3.
Figure 5.43 A close-up plastic strain developed in soil and damage occurred in cemented
column for Case L3-3.
Figure 5.44 Progressive tensile crack development in cemented column for Case L3-3.
Figure 6.1 (a) Soil profile and (b) schematic of test setup (after Namikawa et al., 2008) for
case study 1.
Figure 6.2 Field measurement data. (a) Load-displacement at applied point (after
Namikawa et al., 2008); (b) Post-test damage at columns (after Babasaki et al.,
1997).
Figure 6.3 Laboratory test results for cemented soil. Left: UCT results; Right: Split tensile
strength results (reproduced from Namikawa et al., 2008).
Figure 6.4 Finite element model for lateral load test of cemented soil column in case study
1.
Figure 6.5 The range of qu and st used in finite element analyses for case study 1.
Figure 6.6 FE model with boundary and loading conditions for case study 1.
Figure 6.7 Processed field data for lateral load-displacement at top of the cemented soil
column in field.
Figure 6.8 Comparison of lateral load-displacement relations for field test and analysis
results.

xv
Figure 6.9 Lateral load-displacement relations for Case 2 and Case 3 with tensile strength
varies from 0.1qu to 0.2qu.
Figure 6.10 Damage at cemented column observed in numerical analysis Case 3b: (a)
compressive damage; (b) tensile damage.
Figure 6.11 Lateral load-displacement relations modeled by classic Tresca and CDP models.
Figure 6.12 Yielding zone at cemented column observed in classical Tresca model analysis.
Figure 6.13 Soil profile sketch for case study 2.
Figure 6.14 Proposed construction supported by diaphragm wall and cement-treated soil for
case study 2.
Figure 6.15 On-site excavation profile for case study 2.
Figure 6.16 Collapse of front row cement-treated soil columns for waterway construction in
northeastern Singapore.
Figure 6.17 Typical geometry and boundary condition for case study 2.
Figure 6.18 Finite element mesh for Case C-W (case study 2).
Figure 6.19 Finite element mesh for Case C-R (case study 2).
Figure 6.20 Ground response for Case C-W. Left: Ground movement (m); Right: Plastic
strain.
Figure 6.21 Ground response for Case C-R1. Left: Ground movement (m); Right: Plastic
strain.
Figure 6.22 Ground response for Case C-R2. Left: Ground movement (m); Right: Plastic
strain.
Figure 6.23 Ground movement and tensile damage (dt) at cement-treated soil columns for
Case C-R3.
Figure 6.24 Soil profile and proposed construction method for case study 3.
Figure 6.25 On-site excavation profile and visible cracks at the cement-treated soil columns
appeared immediate after the excavation for case study 3.
Figure 6.26 Localized collapse of cement-treated soil columns for basement construction in
central Singapore (case study 3).
Figure 6.27 Geometry and boundary condition for case study 3.
Figure 6.28 Finite element mesh for case study 3.
Figure 6.29 Numerical analysis results for case study 3.
Figure 6.30 Numerical comparison of crack development in cement-treated soil columns
with steel I-beam (left) and without steel I-beam (right).
Figure 6.31 Comparison of field failure observation and predicted damage by numerical
analysis for case study 3.

xvi
List of Symbols

Sectional area of treated column


Alig Area of the ligament
Al2O3 Aluminium oxide
Aw Cement content
Crack length
Ligament depth
Area of replacement ratio
Difference of notch lengths
b Beam width
Ca Calcium ion
CaO Calcium oxide
Cu Undrained shear strength
Cw Water content
c Cohesion
D Diameter of sample
Elasticity matrix
Degraded elasticity matrix
d Scalar stiffness degradation variable
db Beam depth
dc Compression damage variable
dt Tension damage variable
E Young’s modulus
E0 Initial elastic Young’s modulus
E50 Secant modulus of elasticity
F Load
Fe2O3 Iron oxide
Yield function
G Shear modulus
Gf Tensile facture energy
Gfs Shear fracture energy
Gs Specific gravity
g Gravitational acceleration
Plastic potential function

xvii
First principal stress invariant
J-integral / Potential energy
Rate of energy absorption in the cohesive zone
Second stress invariant
Third stress invariant
K2O Potassium oxide
Ratio of on the TM to that on the CM at initial yield
k Stiffness of the machine in BNT
L Length of sample
LL Liquid limit
MgO Magnesia
MnO Manganese oxide
Equivalent parameter index
Specimen mass
m In-situ moisture content
Na2O Sodium oxide
n Notch width
P2O5 Phosphorus pentoxide
PI Plasticity index
PL Plasticity limit
Property of treated column
Composite property
Property of soil
Effective mean stress
q Von Mises equivalent stress
Effective von Mises equivalent stress
qu UCS
SiO2 Silica
SO3 Sulphuric anhydride
SrO Strontium oxide
s Span
s:c:w Soil : cement : water
Deviatoric stress
TiO2 Titanium dioxide
Potential energy
; Energy

xviii
Energy
w/c Water to cement ratio
ZrO2 Zirconia
Shear strain
Deformation
δc0 Crack mouth opening displacement
δc Crack opening displacement
δcr Maximum crack opening displacement
ij Kronecker delta
δ0 Stroke of the machine in BNT
δv Vertical deflection
Strain
Mean strain
Elastic strain
Plastic strain
̃ Equivalent plastic strain
̃ Compression crushing strain
̃ Tensile cracking strain
b Total unit weight
Strain mobilization factor
Lode angle
ν Poisson ratio
Bulk density
Stress
Compressive strength in the biaxial state
Compressive strength in the uniaxial state
Normal stress
st Split tensile strength
dt Direct tensile strength
t Tensile strength
Uniaxial tensile stress at failure
Effective stress
Algebraically maximum eigenvalue of
Shear strength for treated column
Composite shear strength along the sliding surface
Shear strength for soil

xix
Friction angle
Dilation angle
Total deformation
Eccentricity
. Macauley bracket

xx
Abbreviations

BNT Three-point Bending Notched Beam Test


CASH Calcium aluminium silicate hydrate
CAX4R 4-node Bilinear Axisymmetric Quadrilateral Element
CDIT Coastal Development of Institute Technology, Japan
CDP Concrete Damage Plasticity
CID Isotropically Consolidated Undrained Triaxial Compression Test
CIU Isotropically Consolidated Drained Triaxial Compression Test
CM Compressive Meridian
CPS3 Linear Triangular Element
CPS4R Linear Quadrilateral Element
CSH Calcium silicate hydrate
C3D8R 8-node Linear Brick Element, with Reduced Integration
DW Double-wall Type
FEM Finite Element Method
GD Grid Type
HITEC US High Innovative Technology Evaluation Center
LDT Local Displacement Transducer
LOI Loss on Ignition
OPC Ordinary Portland Cement
RAS Real Allocation Simulation
RILEM French acronym for the International Union of Laboratories and Experts in
Construction Materials. Systems and Structures.
RM Refined Mass Model
STT Split Tension Test
TM Tensile Meridian
TN Tangential Buttress Type
UCT Uniaxial Compression Test
VERT Vertical Earth Reinforced Technology
WAS Weighted Average Simulation
WM Whole Mass Model

xxi
Chapter 1 Introduction

1.1 Background

To effectively utilize very limited land in urban area, underground space is commonly

exploited for development. Various construction methods are employed to facilitate

underground construction in different ground conditions. For ground where bedrock is

encountered at shallow depth, installation of conventional temporary retaining wall such as

sheet pile or soldier pile requires pre-boring process to anchor the wall into bedrock. Pre-

boring work increases time and cost; moreover, the fixity at the bedrock level might induce

large bending moment which compromises the capacity of sheet pile or soldier pile wall.

Construction cost increases tremendously if higher capacity wall such as contiguous bored

pile or diaphragm wall is adopted to facilitate this shallow depth of excavation.

One quarter of Singapore is covered by sedimentary deposit known locally as the Kallang

Formation (Tan et al., 2002; Pitts, 1992), so the development in this formation is inevitable.

This formation consists of deposits of marine, alluvial, littoral and estuarine origins (Lee et al.,

2005). Marine clay is the main constituent of this formation and the thickness is usually

between 10 m to 15 m but in some instance, it can be more than 40 m (Tan et al., 2003). Deep

excavation in this ground condition is always facilitated with robust system such as

contiguous bored pile wall, steel tubular pipe wall or diaphragm wall with a layer of

embedded treated soil together with strutting support (Gaba, 1990; Hsi and Yu, 2005; Shirlaw

et al., 2005). However, this system is too expensive for a small scale development involving

shallow depth of excavation. In such a thick soft soil, ensuring the toe stability of

conventional retaining wall and controlling maximum wall deflection at the base of

excavation (Tanaka, 1994) are always a challenge. As a solution, the designers tend to form

an embedded treated soil layer just beneath the excavation level to increase the stability as

well as provide lateral support to the wall. Nevertheless, for large area development with

1
shallow excavation depth, construction of this embedded treated soil becomes costly and less

practical.

In civil engineering, it is always the particular challenging problems that prompt the adoption

of innovative solutions to optimize the construction efficiency. To overcome aforementioned

ground conditions, column-shaped cement-treated soil offers viable alternative to

conventional earth retaining system. The column-shaped cemented soil can be formed by

deep cement mixing method or jet grouting method. It can serve as a rigid gravity structure

when constructed in continuous overlapping column. Sometimes, different layouts of column

installations are used to achieve the desired effect by utilizing space optimization purpose.

There are some documented successful cases and limited published failure cases. One of the

successful cases was located in downtown Lexington, Virgina, USA (Ruffing et al., 2012)

where the site is underlain by limestone and calcareous shale bedrock similar to the

subsurface condition discussed earlier. Cement-treated soil columns of diameter 2.4 m and

length 8.5 m were constructed to facilitate the 4.9 m deep open excavation. The cross section

of the retaining wall and site photographs are presented in Figure 1.1. On the other hand,

Haque and Bryant (2011) observed substantial soil body movement behind a cement-treated

soil retaining wall in Irving-Las Colinas in Texas, USA. Failure of cement-treated soil wall

did happen in the field but such cases are rarely published. Figure 1.2 shows failure happened

in Northeastern Singapore where cement-treated soil columns were adopted as retaining

system to facilitate an open excavation. The front row of cement-treated soil columns

collapsed when excavation reached 7 m depth.

1.2 Issues Related to the Use of Cement-treated Soil Columns

To optimize construction cost and efficiency, different layouts of column configurations

(refer to samples shown in Figure 1.3) instead of 100% ground improvement are adopted in

2
design. Owing to the short history of employing cement-treated soil columns as earth

retaining system, its basic design has yet to be firmly established. In current design practice,

engineers often regard it as a gravity type structure. As such, two stability analyses involving

external and internal wall stability are evaluated. For external stability analysis, three failure

modes of the treated soil mass: sliding, overturning, and bearing capacity are examined. For

group columns, it is treated as composite material assuming no failure occurred within the

treated soil columns and the untreated soil. However, averaging the properties for these two

materials with very different behavior indeed violates the underlying mechanism. In general,

a volume ratio to average the properties is adopted and a simple elastic-perfectly plastic

constitutive model is employed to simulate the averaged behavior. This can result in errors

and put the design in risk as the treated soil behaves as a quasi-brittle material (Kamruzzaman,

2002; Das and Dass, 1995) while the behavior of untreated soil is usually ductile.

In addition, the behavior of group columns in the field is unlikely to be that of a composite

material assumed in the design, as the bonding between cemented soil columns and between

cemented soil column (tensile strength) and soil (cohesive strength) might be weaker than the

original ground. Weak bonding may cause separation of columns especially the front columns

along the intended excavation line from the rest of the row resulting in toppling of these

columns, making the composite mass assumption invalid. The tendency of bending failure

also reveals the limitation on current internal stability assessment where focus has been drawn

on shear failure of the columns and overlooks the tensile forces in the columns. It is believed

that the disregard of the above considerations in the design results in a good number of

failures of the cement-treated soil column retaining system in open excavations.

In numerical analyses, the cement-treated soil behavior is very often simulated by a simplistic

constitutive model. Up to date, elastic-perfectly plastic model is used due to simplicity and

parameters that can be readily obtained. However, the nature of this treated soil columns is

more complex where the strength reaches its peak at a very small strain followed by a sudden

3
reduction in post-peak stress to a very low residual value. Hence, adopting simplistic model

may not truly simulate the treated soil behavior. Nevertheless, rather than using the most

advanced model, Lee (2008) highlighted that it is more important to have a better

understanding of the soil behavior in a specified problem. This allows one to choose models

and parameters which adequately reflect the important behavioral aspects of the soil for a

particular problem. In spite of this, very few studies were carried out to evaluate the effect of

constitutive model in simulating the cement-treated soil columns behavior for open

excavation problems.

1.3 Objective and Scope of Study

Employing cement-treated soil columns as an alternative for conventional earth retaining

system to facilitate a shallow depth excavation can be more effective in terms of time and cost

(Ruffing et al., 2012). However, failure of such retaining system may occur if the behavior of

this material is not well understood. As a result, this study aims to investigate the mechanism

of cement-treated soil columns used as earth retaining system in an open excavation. The

scope of study is as follows:

a) To carry out laboratory studies to investigate the compressive and tensile behaviors of

cement-treated marine clay including interpretation of post-peak softening of the

treated clay.

b) To evaluate the effect of different constitutive models (namely elastic-perfectly plastic

Tresca model, isotropic model and concrete damage plasticity model) in simulating the

behavior of cement-treated soil. The results are calibrated against those data obtained

from different types of laboratory tests conducted on cement-treated sand and cement-

treated clay.

c) To simulate the behavior of cement-treated soil columns in open excavation using

numerical approaches. These include evaluating the effectiveness of the commonly

4
used weighted average simulation approach and modeling the treated mass based on the

actual column layout.

d) To compare the field observations from case studies with the proposed numerical

simulations. These field studies would also serve to complement the understanding

developed in items (a), (b) and (c).

e) Finally, issues on cement-treated soil columns in an open excavation that are important

but overlooked in design practice would be addressed.

1.4 Structure of Thesis

The outline of this thesis after this introductory chapter is presented as follows:

a) Chapter 2 reviews the general aspects of cement-treated soil where the changes in

physical and mechanical properties are discussed. The current design procedures and

methods for cement-treated soil columns used in an open excavation are also evaluated.

Existing numerical studies of cement-treated soil are reviewed.

b) Chapter 3 presents the experiment work to establish the understanding of fracture

behavior of cement-treated marine clay in compression and tension. At first, the

experiment set-up and methodology are presented followed by results and analyses.

Tensile fracture energy and tension-softening relation of cement-treated marine clay are

interpreted from the experimental results. Parametric studies and outstanding issues

regarding the laboratory studies are discussed.

c) Chapter 4 firstly describes the constitutive models (namely elastic-perfectly plastic

Tresca, isotropic and concrete damage plasticity models) followed by numerical

calibrations against published data and laboratory results presented in Chapter 3. Three

types of laboratory tests are involved to evaluate the accuracy of model prediction for

cement-treated soil when subject to compression, tension and bending. Calibrations are

made for cement-treated Toyoura sand and cement-treated marine clay.

5
d) Chapter 5 examines the effectiveness of commonly used weighted average simulation

approach in simulating the treated mass as a composite material. Back-analysis on a

laboratory test using this approach is compared against laboratory measurements.

Limitations of this common practice are addressed and overcome with the real

allocation simulation approach incorporating the findings from Chapters 3 and 4.

Thereafter, hypothetical cases of a vertical cut with different ground improvement

geometrical patterns are conducted to further examine the behavior of cement-treated

soil mass. Through this study, the failure mechanisms for cement-treated soil columns

in different layouts are illustrated.

e) Chapter 6 presents the back-analyses of large-scale field case studies. Comparisons

between the field observations and the predicted cement-treated soil columns behaviors

using the proposed numerical approach in the present study are made.

f) Chapter 7 concludes the findings of the present study and proposes recommendations

for future studies.

6
Chapter 2 Literature Review

2.1 Introduction

With increasing popularity in using cement-treated soil in construction works, extensive

studies had been conducted by various researchers to understand its behavior. Previous

studies on the general aspects of cement-treated soil are first reviewed in this chapter. The

behavior of cement-treated soil through experimental studies is thoroughly reviewed to

establish the current state of art in the behavior of cement-treated soil, including its physical

and mechanical properties. These two properties decide the mix proportion and the

performance of the final product; hence it is crucial to understand them well. Existing design

approaches for assessing cement-treated soil in open excavation are then discussed. CDIT

(2002) highlighted the importance of adopting numerical analysis to examine cement-treated

soil problems involving their current states of stress-strain. As such, previous numerical

studies for cement-treated soil are reviewed herein.

2.2 General Aspects of Cement-treated Soil

Cement-treated soil is a product of mixing cement to in-situ soil through different mechanical

methods to increase the strength and stiffness by bonding the particles together. The most

popular methods are jet grouting method and deep cement mixing method. The difference

between these two methods is that the former relies on high pressure to erode and mix the soil

with cement or cement slurry through rotation while the latter uses blades to cut and mix the

soil with cement. The product quality formed by the latter is more uniform and promising.

Mechanical properties of cement treated soils are affected by many factors (Babasaki et al.,

1996). The in-situ strength of treated soils varies widely depending on the level of mixing as

in the field, it is difficult to mix cement based agent well with soil. The mixing performance

can be affected by the efficiency of the equipment used, mixing process and ground condition.

9
Therefore, the inherent heterogeneity of cemented soil has been evaluated in laboratory as

well as in the field by some researches such as Larsson et al. (2005a); Larsson et al. (2005b)

and Chen et al. (2011).

The increase in strength of cement-treated soil is mainly attributed to the three reactions

happening in the process of mixing cement with soil, namely: hydration process, ion

exchange (flocculation), and pozzolanic reaction (Diamond and Kinter, 1965; Assarson et al.,

1974). In the initial stage of mixing, water in the mixture is consumed by cement and

dissociates a product known as calcium hydroxide. Dissociation of calcium hydroxide

increases the electrolytic concentration as well as the pH of the pore water, and dissolves the

SiO2 and Al2O3 (pozzolan) from the clay particles, which then leads to ion exchange,

flocculation and pozzolanic reactions. Dissociation of calcium ion Ca2+ in the pore water will

replace the weaker cations on the surface of clay particles. This then leads to decrease in the

distance between the diffused double layer and hence the small particles of clay flocculate

and coagulate into larger sizes. The dissolved dissociated Ca2+ ions react with the dissolved

pozzolan from the clay particle’s surface to form hydrated gels, resulting in the combination

of soil particles. Hydration reaction dominates the early stage of curing process while the

pozzolanic reaction is more significant in prolonged curing period. Hence, effective friction

angle for the treated soil increases more predominantly at early curing period during the

formation of particle interlocking in the soil-cement skeleton (Kamruzzaman, 2002). Increase

component of pozzolanic reaction products contributes to the increase of strength parameter

c’. Details of engineering properties of cement-treated soils are discussed in the following

sections.

10
2.2.1 Physical Properties of Cement-Treated Soil

In addition to the desired improved properties, it is essential to evaluate the overall properties

change resulted from the chemical reaction, so that the performance of the cement-treated soil

can then be estimated.

a) Change in water content

During the mixing process, the cement mineral reacts with water to produce cement hydration

products (CDIT, 2002), thus it is expected that some amount of water in the soil will be

consumed beyond the amount of water in the mixture slurry itself. It was experimentally

tested that the consumed water to form the cementitious products will not be expelled by re-

heating the products to 105oC (Kamruzzaman, 2002). The water content in the soil will be

much more reduced in dry mixing method compared to wet mixing method. Figure 2.1 shows

the water content after cement treatment decreased about 20% from the original as reported in

Tokyo Port clay field. It was noted that the mixing water-cement ratio, w/c was 0.6

(Kawasaki et al., 1978). Kamruzzaman (2002)’s study showed that majority of the decrease in

water content takes place within the first 7 days of curing (Figure 2.2). It is revealed in some

studies (Uddin et al., 1997; Kamruzzaman, 2002) that higher percentage of cement content

induces a greater reduction in water content.

b) Change in density

There is a concern regarding the effect of cement inclusion into the in-situ soil will induce

density change since the water content is decreased. For the cement treatment in dry form, the

wet density of the treated soil, increases by about 3% to 15% recorded by CDIT (2002).

Conversely, the density change due to slurry form treatment is negligible irrespective of water

to cement ratio, w/c. Figure 2.3 shows the change in density for in-situ cement treatment in

11
dry and slurry forms reported by Japan Cement Association (1994). Wen (2005) reported

similar density with original Singapore marine clay is obtained for jet grout mixing.

c) Change in permeability

Although most of the time, increase in strength and stiffness are the main purposes of mixing

cement to in-situ soil, reduce the permeability to form a cut-off wall where problems of

groundwater seepage is a concern (Yu et al., 1999) is also one of the purposes. The

permeability coefficient of the treated soil is highly influenced by the distribution of pore

sizes inside the soil mix (Porbaha et al., 2000). Generally, laboratory permeability tests

showed that the soil-cement mixture is in the order of 10-10 m/s and is not influenced by the

confining pressure and seepage pressure in the specified range (Yu et al., 1999). With the

increase of vertical load, the coefficient of permeability increases markedly when the

confining pressure is low and changes slightly or even decreases when the confining pressure

is high.

The effect of cement content on the permeability of cement-treated clay was studied by

Kawasaki et al. (1981a) as shown in Figure 2.4. It was found that the permeability reduced

with the increment of cement content from 10% to 20%. The reduction could be due to the

pozzolanic cement substances, which blocked the pores in the soil cement matrix (Broderic

and Daniel, 1990).

However, it is known that the permeability of soil in field can be higher than those obtained in

laboratory test (Tavenas et al., 1986), thus field test is more important in understanding the

actual mechanism. In Shen (1998)’s study, finite element method (FEM) back-analysis

showed that the predicted consolidation degree with elapsed time agreed well with the

measured value if the improved column was assigned with high permeability of 10-5 m/s

which was in the order of 4 to horizontal permeability of surrounding clay. This was contrary

12
to the studies as mentioned before and it was claimed that the existence of fracture cracks

resulted in higher permeability. The expansion pressure (pressure acting on the wall of a deep

mixing column results from the injection of chemical admixtures under working pressure)

increases with the injected volume of cement slurry. According to Shen (1998), with a mixed

cement content of 2% to 4%, the expanding pressure will be greater than the minimum

hydraulic fracturing pressure increment, which indicates that the surrounding clay is

conveniently fractured during deep mixing column installation. The existence of the shearing

force caused by mixing blades in deep mixing makes the surrounding clay more easily to be

fractured.

Lorenzo and Bergado (2006) acknowledged that the product of cement-treated is a porous but

strong matrix and exhibits higher permeability. Similar opinion shared with Suzuki et al.

(1981) where due to cement hydration, increase in pH of the pore fluid and Ca++ ion on clay

surface causes shrinkage of the diffused double layer leading to flocculation and hence

resulting an increase in permeability. This also coincides with the observation by

Kamruzzaman (2002) for 7-day permeability for cement-treated Singapore marine clay in

laboratory test. The permeability of the cemented clay was obtained using falling head

permeability test conducted in a modified oedometer consolidation cell.

2.2.2 Mechanical Properties of Cement-Treated Soil

a) Modulus of elasticity, E

The Young’s modulus is defined as the ratio of the uniaxial stress over the uniaxial strain in

the range of stress in elastic regime. Generally the unconfined compression test (UCT) is

preferred owing to its easy handling, inexpensive and quick confirmation of in-situ strength.

The slope of the stress-strain curve at any point to the origin of the axis is called the secant

modulus, Esecant. Tan et al. (2002) reported that the Esecant for cement-treated soil for the whole

regime of stress-strain curve is nonlinear and decreases with strain increment. The adopted

13
Young’s modulus is the tangent modulus of the initial, linear portion of stress-strain curve.

The initial slope is indicated as E0 and secant modulus at 50% of ultimate strength is denoted

as E50. Table 2.1 summarizes selected E-qu relationships suggested for cement-treated clay.

Table 2.1: Summary of selected E-qu relationships for cement-treated clay.

In-situ improved/
References E-qu In-situ soil type
lab improved
Lab improved & In-situ
Lee et al., 1998 E0 = 80 – 200qu # Marine clay
improved
Kamruzzaman, 2002 E0 = 490qu * Marine clay Lab improved
E50 = 350 – 800qu *
Tan et al., 2002 Marine clay Lab improved
E50 = 150 – 400qu #
Lee et al., 2005 E0 = 80 – 140qu * Marine clay Lab improved
Wen, 2005 E = 200qu Marine clay In-situ improved
Wong and Goh, 2006 E = 100qu Marine clay In-situ improved
Lorenzo and Bergado,
E50 = 150qu Bangkok Clay Lab improved
2006
E0 = Initial elastic Young’s modulus; E50 = Secant modulus of elasticity (at 50% of ultimate strength);
qu = unconfined compression strength; * denotes local strain measurement method; # denotes
conventional strain measurement method.

Nevertheless, there were researchers (Jardine et al., 1984; Goto et al., 1991; Tatsuoka et al.,

1996; Kohata et al., 1997; Tan et al., 2002) highlighted the limitation of measuring the axial

strain using conventional method. In the case of unconfined or triaxial compression test, the

axial strain in conventional method is usually obtained from the axial displacement of the

loading piston or the specimen cap. Goto et al. (1991) identified the sources of error involved

in the external axial strain measurement as: (1) system compliance due to the deflection of top

cap, load cell, cell, piston etc; (2) tilting of the specimen; (3) bending error on bottom and top

of the specimen; (4) strain non-uniformity of the specimen caused by end restraints, leading to

bulging of the specimen; and (5) strain localization of the specimen. As shown in Figure 2.5,

the conventional method tends to underestimate the stiffness compared to the local transducer

approach.

14
On the other hand, there were studies (Consoli et al., 1996; Hosoya et al., 1997; Yin and Lai,

1998) revealed that the stiffness and strength obtained from triaxial compression test depend

on confining pressure. Typically, the strength achieved a higher value in triaxial compression

test with increase in confining pressure. However, these tests were carried out using

conventional method of strain measurement. According to the studies by Tatsuoka et al.

(1997), the change in initial elastic modulus at small strain (Emax) corresponding to the

increase in effective confining pressure is negligible when local axial strain measurement

approach is adopted. Figure 2.6 shows that the modulus at small strain (Emax) is not a function

of confining pressure (Shibuya et al., 1992). Study by Chin (2006) in comparing the stress-

strain behavior from both CIU and UCT tests using conventional method for specimens cured

under loaded-drained condition showed that the initial elastic modulus is very similar for both

tests. Consoli et al. (1996) also found that when tensile strength increases to certain value, the

ratio of triaxial deviatoric stress over the unconfined compression stress approaches unity for

any given confining pressure.

b) Strength

In general, the strength is imparted to a soil by virtue of: i) cohesive forces (termed as ‘c’)

between particles; ii) frictional resistances ( ) due to particles sliding against one another, or

moving from interlocked positions. In the past, the cement-treated soil strength has been

measured through unconfined compression test (UCT), consolidated isotropic undrained test

(CIU) and consolidated isotropic drained test (CID). On top of these, the tensile strength has

been studied by uniaxial direct tension test or most of the time split tension test (STT) is

preferred due to easiness.

From the UCT results, ductile behavior is manifested for very low cement content.

Conversely, the treated soil becomes more brittle at higher cement content. Figure 2.7 shows

that the treated soil deviatoric stress increases abruptly associated with a very small strain

15
followed by a sudden reduction in post-peak stress to a very low residual value. According to

Chin (2006)’s result shown in Figure 2.8a, the CIU test showed that the difference in

deviatoric stress increases significantly with increase in curing stress and resulting in the

material turning brittle. The curing stress is the isotropically loaded stress applied on the

specimen during curing period. Comparisons of stress-strain behavior from both CIU and

UCT tests for specimens under loaded-drained condition, the maximum deviatoric stress is

identical to each other but the post-peak behavior varies (Figure 2.8b). Although both exhibit

post peak softening but with the existence of confining stress, the residual strength is about 70%

to 80% of the peak strength for confining stress beyond 100 kPa. Figure 2.9 (Cement Deep

Mixing Association of Japan, 1994) shows the residual and peak strengths under different

confining pressures: in CIU the residual strength is 30% to 60% lower than the peak strength;

while in CID the residual is 60% to 80%.

Estimation of shear strength of cement-treated soil from direct shear tests is not common. So

far the available published data was by Saitoh et al. (1980) as shown in Figure 2.10 for low

strength, the shear strength can be roughly represented by half of the unconfined compressive

strength. It is worth noting that in these tests, the normal stress during shearing was zero and

is therefore more appropriate in representing soil-cement at shallow depth.

The increase in strength of cemented soil is attributed to hydration and pozzolanic reactions.

Hydration reaction dominates the early stage of curing while pozzolanic reaction is more

significant at prolonged curing period. Hence, effective friction angle for the treated soil

increases more predominantly at early curing period during the formation of particle

interlocking in the soil-cement skeleton. Increased component of pozzolanic reaction products

contributes to the increase in strength parameter c’. Figure 2.11 shows the effect of cement

content and curing time on effective strength parameters of treated clay by Kamruzzaman

(2002). Effective stress path of the cement-treated soil in triaxial test (Figure 2.12) revealed

16
the existence of deviatoric stress (q) beyond the original critical state of untreated soil as

evidence of cementation bond takes place during the shearing state.

It is well known that addition of cement helps to improve the compressibility and strength of

in-situ soil. Conversely, it results in a brittle material that is weak in tension. As a result,

designer often ignores the tensile strength in design; consequently the tensile behavior of

cement-treated soil is always neglected in the study related to cemented soil. Unixial tension

test also known as direct tension test is the direct method of determining the tensile strength

but in most cases it was determined by split tension test. Very often, the tensile strength is

correlated to the key parameter qu as illustrated in Table 2.2. Nevertheless, these empirical

correlations should be applied with care, as they differed from soil to soil.

Table 2.2: Summary of some t – qu relationships of cemented soil.

References Relationship
Porbaha et al., 2000 st = 0.1 – 0.15 qu
dt = 0.1 qu
Saitoh et al., 1996
st = 0.1 – 0.3 qu
Tanaka and Terashi, 1986 t = 0.15 qu
Fang et al., 1994 t = 0.05 to 0.2 qu
Xiao, 2009 st = 0.127 qu

st = Split tensile strength; dt = Direct tensile strength; t = Tensile strength (without test method
information); qu = unconfined compressive strength.

Owing to split test being an indirect tensile strength test, it provides no possibility of strain

measurement according to Das and Dass (1995). This aggravates the current limited

information on the tensile behavior of cement-treated soil. The available published direct

tension test results on sand are presented in Figure 2.13. Direct tension test result reveals the

similar stress-strain curve trend with unconfined compression test where an abruptly decrease

in strength after peak strength is observed. Addition of cement in the untreated soil has

changed the material from ductile to brittle manner. Once the treated material experiences the

17
maximum load, the material cannot maintain the maximum load under further straining. This

post-peak strength reduction in further straining is known as softening behavior. To the best

knowledge of this author, so far the tension-softening behavior of cement-treated soil was

only investigated by Namikawa and Koseki (2006). The study was conducted for Toyoura

sand with three-point bending notched beam test (BNT). From the test, the tension-softening

relation of the cemented Toyoura sand (Figure 2.14) was interpreted using the energy balance

approach.

c) Poisson’s ratio, ν

When a material is compressed in one direction, it usually tends to expand in the other two

directions perpendicular to the direction of compression. The ratio between these two

quantities is named as Poisson’s ratio. CDIT (2002) suggested a range of 0.25 to 0.45

irrespective of the unconfined compressive strength, qu of the cement-treated soil. The

measured Poisson’s ratio for cement-treated clay by Fang and Yu (1998) varied from 0.13 to

0.24. A value of 0.167 was evaluated from the triaxial drained compression test on cemented

Toyoura sand by Namikawa (2006). Larsson et al. (2012) proposed a value of 0.15 in their

study.

2.3 Existing design approaches

The early application of cement-treated soil columns is mainly used in foundation of

structures to increase bearing capacity and reduce settlement. On top of these main criteria of

improvement, the stability of whole system is indeed affected by the layout of improvement

pattern. Therefore, the design considerations based on limit equilibrium approach so far have

been focused on the stability assessment of the improved ground. There are several

improvement geometrical patterns: a) block-type, b) wall-type, c) lattice-type, d) group

column-type, and e) columns-in-contact type as illustrated in Figure 2.15. Since there is no

18
design guidance for cement-treated soil columns used in open excavation, designers often

refer to the design concept which has been developed for the use of cement-treated soil

columns in embankment foundation. In examining the performance of cement-treated soil

columns in an open excavation case, the evaluation of bearing capacity and settlement are

excluded since there is negligible vertical loading acting on the columns as in embankment

case unless construction surcharge is applied.

For block-, wall-, and lattice-type improvement where an extensive area of ground is

improved, it is considered as a rigid gravity-type structure (CDIT, 2002). Two types of

stability analyses are evaluated in current practice, i.e. external stability and internal stability.

For external stability analyses, three failure modes of the treated soil mass: sliding,

overturning, and bearing capacity are examined. On the other hand, in assessing the internal

stability analyses, the allowable strengths of treated soil after discount off the reliability

coefficient of overlapping and scattered strength should be higher than the induced stresses.

The induced stresses in the improved ground are determined based on elastic theory. In

addition, it is a necessity to check extrusion failure for the untreated soil remaining between

treated soil-walls since it is subjected to unbalanced active and passive earth pressure.

In certain applications where the improvement pattern has been optimized by adopting group

column-type, the improvement area is not a rigid body but is considered as a sort of

composite ground with an average properties of treated soil columns and surrounding

untreated soil (Public Works Research CenterCenter, 1999). The average property of the

improved mass is defined by

= + (1 ) (2.1)

where

= composite property, i.e. strength/stiffness,

19
= area replacement ratio,

= property of treated column, and

= property of soil.

The improvement ratio, , represents the percentage of the sectional area of the treated soil

column (Figure 2.6) to the ground occupied by the soil column, which can be computed by

= (2.2)

where

= sectional area of treated column,

d1 = intervals between columns, and

d2 = intervals between columns.

The first stability assessment in the design procedure for group column-type treatment is slip

circle analysis. Eq. (2.1) can be rewritten to consider the shear strength for the treated mass:

= + (1 ) (2.3)

where

= composite shear strength along the sliding surface,

= shear strength for treated soil column, and

= shear strength for soil.

On the other hand, the shear strength for column and soil in Eq. (2.3) can be represented as

= + tan (2.4)

where

= cohesion intercept,

20
= normal stress on the failure plane, and

= friction angle.

Nevertheless, the stability of ground improvement is often analyzed in short term (undrained

analysis) because the strengths of treated soil columns and the soil between columns increase

over time (Navin, 2005; Shen, 1998). Generally the undrained friction angle, ,
for soft

soil is found close to zero so practitioners usually neglect it in design. Furthermore, the

normal stress on failure plane is expected to be small for open excavation involves shallow

depth of improvement. As a result, Eq. (2.3) can be simplified as

, = , + (1 ) , or (2.5)

,
, = + (1 ) , (2.6)

where

, = unconfined compressive strength of treated soil columns, and

, = undrained shear strength of soil.

It is well known that the mobilized strains for soft soil and treated soil column are varied.

Hence, a factor, η, is applied to account for the difference in strain at failure when averaging

the properties for treated soil column and surrounding soil (CDIT, 2002). The basis for

evaluating η is shown in Figure 2.17. Alternatively, Tatsuoka and Kobayashi (1983)’s

experimental test result showed that the residual treated soil column strength is approximately

65% to 90% of unconfined compressive strength. Kitazume et al. (2000) adopted a residual

treated soil column strength value of 80% in analyzing their centrifuge results. In slip circle

analysis, shear failure is assumed as the failure mode of treated soil columns and no failure

occurred in the surrounding soil.

21
The second step of assessment is sliding failure analysis where the treated soil columns and

surrounding soil are assumed to behave as a whole. Nevertheless, some researchers reported

several failure modes that have not been covered in existing design practice. For instance, for

external stability assessment, Kitazume et al. (2000) showed that a collapse failure pattern

where the treated soil column tilts like dominos at the base (Figure 2.19a), is less stable than

sliding failure pattern. For internal stability assessment, some studies (Miyake et al., 1991;

Karastanev. et al., 1997; Hashizume et al., 1998; Kitazume et al., 1996) showed various

failure modes: shear, bending and tensile failure modes corresponding to the ground and

external loading and the locations of each column. Owing to the average fracture energy for

shear failure, Gfs, is approximately 15 times greater than the fracture energy for tensile failure,

Gf (Namikawa and Koseki, 2006), the existing design approach based on shear strength might

not be on the safe side during internal stability assessment (Kitazume and Maruyama, 2007).

Kivelo (1998) and Broms (2004) proposed that for group column-type improved ground,

several failure modes of treated soil columns as illustrated in Figure 2.18 should be taken into

account. However, assumption of the failure plane is made prior to the design checking and

thus affecting the accuracy of the assesment.

In addition, through observation from experimental tests, Kitazume (2008) proposed 3 other

failure modes: collapse failure mode in external stability assessment; shear failure and

bending failure modes in internal stability assessment as shown in Figure 2.19.

(i) Collapse failure mode

In collapse failure mode, it is assumed that unbalance occurred between the

active and pressure earth pressures acting at the edge of treated area. As a result,

the treated area deforms and tilts at the edge of treated area shown in Figure 2.20.

The active and passive pressures distribution diagram is illustrated in Figure 2.21.

Moment equilibrium at the base of the treated columns is checked to evaluate the

possible occurrence of this collapse failure mode.

22
(ii) Shear failure mode

In addition to the shear slip circle plane, there is a tendency where the treated soil

columns and the intervening soils shear together at a horizontal plane as shown in

Figure 2.19b. Full mobilization of shear strengths in treated soil columns is

assumed in the calculation. The estimation of shear failure plane location is based

on the load equilibrium of active and passive earth pressures acting on the edge

boundaries of the treated area.

(iii) Bending failure mode

To assess bending failure, all the treated soil columns are assumed to fail

simultaneously in bending and the treated area above the failure plane is assumed

to deform as a simple shear. Moreover, the treated soil columns are considered to

fail when the induced tensile stress reaches the ultimate tensile strength (Figure

2.22). The moment equilibrium at the assumed failure plane is checked to assess

the bending failure. The pressure distribution diagram is presented in Figure 2.23.

Nevertheless, it should be kept in mind that the above-mentioned assessments in this section

are based on limit equilibrium approach. In the design procedure of examining embankment

stability by CDIT (2002), it is proposed that finite element analysis should be conducted to

study the lateral displacements and stresses developed in the treated ground for case with

more complicated improvement pattern. Therefore, numerical model of cement-treated soil in

previous studies are discussed in the following section.

2.4 Numerical Modeling on Behavior of Cement-treated Soil

Numerical model is often applied to study the problem behavior over time through simulating

the construction procedure in steps. FEM, has proven to be a powerful tool used in analysing

most of the complex geotechnical engineering problems. It is superior over other techniques

23
such as simplified beam-and-spring and empirical charts as less assumption is needed. Studies

regarding the numerical model of cement-treated soil behavior have been well documented.

Following section discusses some publications related to the prediction or back-analysis of

the cement-treated soil performance in case studies. Some studies focused on the constitutive

model used to simulate the behavior of cement-treated soil are also evaluated herein.

2.4.1 2-D and 3-D Finite Element Analysis

Two-dimensional (2-D) model has been widely used in the last century to perform the

numerical analyses in solving geotechnical problems. Plane strain or plane stress condition is

assumed in the 2-D model. Despite knowing that the treated ground is formed from hundreds

or thousands of cement-treated columns, lacking of high performance computational tool has

limited the modeling of individual single column in the finite element analysis. As a result,

similar to limit equilibrium approach, weighted average simulation (WAS) method is adopted

to average the properties of the treated mass in the model.

The equation to calculate the mass properties is similar to Eq. (2.1). Khoo et al. (1997)

adopted this method to estimate the properties of improved soil berm for an excavation in

Singapore. The area replacement ratio in this project was 25%, while an average of shear

strength and stiffness were computed based on Eq. (2.1) to predict the performance of the

retaining system and surrounding ground. The finite element prediction of lateral wall

deflection was close to field measurement. However, the predicted ground settlement was half

of the field measurement. This indicates that some feature might be lacking in the analysis.

Some researchers introduced an empirical parameter in Eq. (2.1) when fine-tuning between

the prediction and field performance. For instance, a reduction factor of 0.5 was applied by

Hsieh et al. (2003) on the strength of jet grouted soil-cement columns when performing a

FLAC back-analysis to study the diaphragm wall behavior for an excavation in Southern part

24
of Taiwan. On the other hand, Ou and Wu (1996) performed a 3-D finite element analysis to

estimate overall material properties of the treated soil mass for column-type of soil

improvement. In the study, real allocation simulation (RAS) approach was first employed to

model a hypothetical excavation problem. At this stage, the treated material properties were

assigned to the mesh elements corresponding to the treated soil columns zones; while the soil

material properties were specified for the untreated zones. Subsequently, the hypothetical

excavation problem was analyzed with weighted average simulation (WAS) where the

material of the treated zone was regarded as a composite ground mass (Figure 2.24). By

comparing the wall deflection computed from the RAS and WAS methods, the following

formula was proposed to evaluate the overall material properties of the composite ground:

= + (1 ) (2.7)

where an equivalent parameter index, was added as compared to Eq. (2.1). When equals

to 1, Eq. (2.7) can be reduced to Eq. (2.1). Various sets of equivalent material parameters

correspond to different values for the composite ground were used in WAS analyses and

compared to RAS results. In the RAS approach, the treated soil columns were distributed

uniformly across the excavation zone. The value ranged from 1.5 to 5.5 for area

improvement ratio of 10% to 75% for treated soil strength qu between 1 MPa and 3 MPa. On

top of these 3-D models, they suggested that plane strain analysis tends to overestimate the

wall deformation. After some trial-and-error analyses, they proposed a reduction factor 0.88

to the value for Eq. (2.7) to be used in a plane strain model. Through this approach, Ou et

al. (1996) managed to model a wall deflection in Taipei city that is close to field observation

using 2-D plane strain analysis as well as 3-D analysis.

Some researchers found that averaging the properties using volume ratio in Eqs. (2.1 – 2.7)

were too crude. As discussed in earlier section, the stress-strain behavior for cement-treated

soil is different from the untreated soil. Some other researchers proposed a two-phase mixture

model to evaluate the stress-strain relationship of mixture. Studies were also carried out on

25
the estimation of elastic moduli of the mixture. An upper bound solution was proposed by

assuming all the elements of the mixture are subjected to the same uniform strain. On the

other hand, lower bound solution was proposed by assuming that all the mixture elements are

subjected to a uniform stress equal to the applied stress. Omine et al. (1999) proposed a

weighted averaging method for evaluating the stress-strain relationship of two-phase mixtures

with random spherical inclusions (inclusion here denotes as treated soil column) based on the

evaluation of stress distribution. In their approach, a stress distribution tensor was introduced

and the value depended on the shape of the inclusion. Linear elastic-perfectly plastic materials

were assumed for the treated soil and original soil. Omine et al. (1999) validated the proposed

method in estimating the coefficient of subgrade reaction and bearing capacities of the

improved masses through a series of model tests.

Nevertheless, the empirical factor is a crude approach where underlying mechanism is not

well addressed. Moreover, the factor is different for site condition which limits its use. Yang

(2009) conducted a series of 3-D finite element analyses using RAS approach to examine the

mobilized mass properties of an embedded improved soil raft in the lateral direction (Figure

2.25). By modeling each individual improved soil column, various influencing factors such as

geometry arrangement, layering within soil-cement columns, overlapping and non-perfect

treatment were able to be addressed. Through RAS approach, stress distribution between

treated and untreated soil around missing soil columns were well defined. This study shed the

light into a deeper understanding of the underlying mechanism that cannot be addressed by 2-

D model as well as WAS approach.

2.4.2 Constitutive Models for Cement-treated Soil

Constitutive model is used to simulate the mechanical behavior of engineering materials using

various mathematical equations. There are different theories available in modeling soil

behavior. Very often, an elastic-perfectly plastic model is used by the practitioners (Khoo et

26
al., 1997; Goh, 2003; Hsieh et al., 2003; Zhang, 2004; Yang, 2009) to simulate the behavior

of cement-treated soil in a finite element analysis. Among all, Mohr-Coulomb model, a

combination of Hooke’s law and the generalized form of Coulomb’s failure criterion is

adopted. However, when undrained shear strength value is used and undrained friction angle

is taken as zero, then the Mohr-Coulomb criterion will reduce to the Tresca criterion. Tresca

model has been widely used in back-analyzing the field performances. For instance, it was

applied by Hsieh et al. (2003) in his back-analysis for a six-level basements excavation in

Kaohsiung, Taiwan. The 7%-area replacement ratio treated mass was simulated by an elastic-

perfectly plastic Tresca model in a finite difference program, FLAC. It was claimed that the

discrepancy between the numerical analysis and field results was due to earlier unidentified

factors such as the preloading effect by the jet grouting process. In addition, this model was

also used by other researchers such as Goh (2003) and Zhang (2004) in their finite element

simulations for centrifuge tests involving cement-treated soil. Yang (2009) also adopted this

model in his 3-D study of mobilized mass properties of embedded improved soil raft as

reported earlier.

Another popular soil model, the Duncan-Chang model also known as the Hyperbolic model,

is often adopted as well. This model is categorized as a non-linear soil model where the soil

stiffness can be formulated as a stress-dependent parameter using a power law formulation.

The distinction is made between a primary loading stiffness and an unloading and reloading

stiffness. This better description of the non-linear and stress-dependent stiffness has made it

preferred over the Tresca and Mohr-Coulomb models. The failure criterion of this model

follows Mohr-Coulomb criterion but it is not able to incorporate dilatation effect. Ou et al.

(1996) adopted this model to capture the behavior of treated mass in their back-analyses for a

three-level basement excavation in Taipei. The improved ground was installed next to the

diaphragm wall with an area improvement ratio of 8%. However, Ou et al. (1996) observed

that the stress-strain curve exhibited nearly linear at pre-failure stage, thus the stiffness

27
modulus exponent, equal to zero was used. Failure ratio was taken as 0.5 to take into account

the brittle stress-strain behavior of the treated soil.

As evoked in many studies that the cement-treated soil exhibits brittle manner, some

researchers tried to capture this essential feature into their analyses. Wong and Goh (2006)

suggested that for jet grouting slabs in very deep excavations, the stresses experienced in the

slab may exceed the strength, thus modeling the post-failure behavior becomes important.

The collapsed section of cut-and-cover tunnel at Nicoll Highway was illustrated to showcase

the effect of strain softening. Analysis that considered strain softening effect by reducing 80%

strength yielded a better strut force prediction compared to field measurement. However, the

way Wong and Goh (2006) modeled the material softening was rather crude. In the analysis,

Tresca model was adopted and the program execution was forced to stop when extensive

yielding was noticed in the improved layer. Subsequently, the strength and modulus of the

improved layer were reduced manually and the execution was then resumed.

A simple isotropic model in Abaqus software which allows the yield strength to vary with

respect to the equivalent plastic strain was adopted by Tjahyono (2011). The stress-strain

curve was distinguished into four zones, i.e. pre-peak zone, peak zone, post-rupture zone and

residual zone. The initial pre-peak zone followed Hooke’s law of linear isotropic elasticity

while other three zones were in plastic. Figure 2.26 shows the calibrated stress-strain curve

compared with test data by Tjahyono (2011).

Charbit (2009) and Larsson et al. (2012) simulated the lime-cement columns in a shear box

test using the concrete damage plasticity (CDP) model. Concrete damage plasticity model is

able to capture the degradation in strength and stiffness both in compression and tension.

Figure 2.27 shows the stress reduction simulation of uniaxial compression tests for lime-

cement columns by CDP model. This model was initially invented for quasi-brittle materials

such as concrete. Nevertheless, there are several similarities between cement-treated soil and

28
concrete such as brittle behavior and the tensile strength is much lower as compared to its

compressive strength. Using this model, the fracture patterns of the lime-cement columns in

the shear box test were well simulated.

A decade ago, several new sophisticated models (Kasama et al., 2000; Liu and Carter, 2002;

Lee et al., 2004; Xiao, 2009) were proposed to simulate the cement-treated soil behavior.

Generally, these models made efforts to capture the influence of cementation and describe the

shear behavior, yielding, softening, failure criteria as well as dilation effect. Despite these,

little attention had been paid to study the behavior under tensile stresses. Namikawa and

Mihira (2007) proposed an elasto-plastic model that can describe the strain-hardening and

strain-softening responses under tensile and shear failure modes. However, the more

sophisticated a soil model is, the more parameters have to be used on the basis of laboratory

test data. Most of aforementioned advanced models are calibrated based on substantial

laboratory data. For geotechnical engineering applications, available field soil data is always

limited. As a result, some parameters have to be assumed or estimated and this can be argued

whether such models can deliver in practice the accuracy that they pretend (Brinkgreve, 2005).

2.5 Summary

Cement-treated soil is a popular ground improvement method used to increase the strength

and stiffness of in-situ weak soil. It is illustrated in the literature reviews that extensive

amount of research works had been carried out to establish the understanding of this material.

Nevertheless, there are still issues related to the behavior of cement-treated soils that remain

unresolved.

For the past three decades, substantial efforts have been done in laboratory studies (Uddin et

al., 1997; Porbaha et al., 2000; Kamruzzaman, 2002; Chin, 2006; Xiao, 2009) to understand

the properties of cement-treated soil. The factors that influence the strength and stress-strain

29
characteristics were identified. The characterization of the properties of cement-treated soil

has often been presented in terms of strength, deformation, modulus and permeability

characteristics. In many studies, it has been revealed that the soil became brittle by adding

small amount of cement. The material stress reaches peak under a smaller strain and then

decreases abruptly to a much smaller post-peak value. There were disagreements on the

permeability of the cement-treated soil where some studies showed an increase in

permeability against original soil while others showed a decrease. Focus has been put on the

understanding of compressive and shear strengths, but little attention has been paid to the

tensile strength of cement-treated soil. Conventional uniaxial tension test is the direct yet

most appropriate approach to obtain the stress-strain relationship of the cement-treated soil

under tension. However, the execution of this test is very challenging and very limited data

are available so far. Split tension test is often preferred as it is easier to perform and the

results are more consistent (Porbaha et al., 2000). Nevertheless, this test is an indirect test

where the strain cannot be obtained. Moreover, it is not possible to maintain stable loading

after peak stress in direct tension test and split tension test. As such, Namikawa (2002)

conducted three-point bending notched beam test by using fracture energy concept to study

the post-peak tensile behavior of cemented Toyoura sand. His study showed that softening

occurs after the material reached peak stress.

In existing design practice for internal stability assessment, focus has been placed on the shear

failure of cement-treated soil columns using slip circle analysis. However, studies by Larsson

(1999) and Kitazume and Maruyama (2007) both revealed that cement-treated soil columns

could fail in bending and tensile modes. Namikawa and Koseki (2006) compared the shear

fracture and tensile fracture for cement-treated sand. Their study showed that the ratio of

average fracture energy for the tensile failure, Gf, to shear failure, Gfs, is approximately 1:15.

This was due to the breakage area of cementation in the shear failure is much larger than that

in the tensile fracture. In other words, the tendency for cement-treated soil columns to fail in

tension is much higher compared to shear mode. This is supported by Larsson (1999)’s

30
finding where in a shear box test, the lime-cement columns failed in bending due to limited

tensile strength rather than shearing.

FEM has been proven to be a powerful tool in solving geotechnical problems. It is able to

simulate the construction sequence as well as the soil behavior with appropriate constitutive

soil models; thus realistic assumptions are used. In the early days, 2-D finite element analysis

is widely used; the treated ground is generalized to a single material in the analysis. In

common practice, weighted average simulation (WAS) approach is adopted to average the

properties between the treated and untreated soils in term of volume ratio. However, this

WAS approach is highly empirical and sometimes a coefficient is incorporated based on field

observations. In pace with the growth of computer technology, there is an increase in 3-D

FEM modeling prediction in recent years. Ou and Wu (1996) and Yang (2009) simulated the

treated soil columns in 3-D model using real allocation simulation (RAS) approach. Although

such analyses are very complex and require significant computing resources, this is the way

ahead to establish the understanding of the mechanism involved in the treated ground. Using

simplistic averaging two different materials into a single material may overlook the real

underlying mechanism.

The accuracy of prediction by numerical analyses highly depends on the ability of selected

constitutive model to simulate the mechanical behavior of soil. Linear elastic-perfectly plastic

Tresca model is often used by practicing engineers as well as some researchers due to its

simplicity and less parameters needed. However, it is generally too crude to capture essential

features of the treated soil behavior, i.e. brittleness. On the other hand, a few constitutive

models (Kasama et al., 2000; Liu and Carter, 2002; Lee et al., 2004; Xiao, 2009) have been

proposed by researchers according to their laboratory data calibration on cement-treated soils.

Amongst these researchers, efforts have been made to describe the yielding, softening, failure

criteria, shearing and dilation behavior of the treated soil. Very little or no attention has been

paid to the tensile behavior. Nevertheless, aforementioned advanced models require more

31
parameters on the basis of soil investigation data. As a result, the use of above models is

limited to research laboratory data calibration and yet to be applied to field problem

simulations. Namikawa and Mihira (2007) proposed an elasto-plastic model that can describe

strain-hardening and strain-softening responses under tensile and shear failure of treated soil.

As field soil data are always limited, certain parameters have to be estimated if adopting

aforementioned advanced models thus putting the accuracy of the model in doubt.

Although the use of cement-treated soil is common in practice and extensive studies have

been conducted, it is evident that the knowledge of tensile behavior of this material and the

numerical model to address it have not been understood. Furthermore, studies also showed

that this material might fail in tension rather than in shearing as assumed in existing design

practice. Therefore, there is a need to address the key issues raised in this chapter in order to

better understand the behavior of cement-treated soil under tension. As a result, this study is

hoped to overcome the gap by establish the tensile behavior of cement-treated soil through

experimental work; calibrate the numerical constitutive models based on laboratory tests;

back-analysis laboratory test to propose a modeling criteria; and to verify the proposed model

using field cases.

32
Chapter 3 Fracture Behavior of Cement-treated Singapore
Marine Clay

3.1 Introduction

As reviewed in Chapter 2, it is important to understand the fracture behavior of cement-

treated soil. Considering the differences in clay mineralogy and geological conditions, there is

a need to establish the characteristics of local cement-treated clay particularly on its tensile

resistance. As its compression fracture behavior has been well studied by many researchers

(Kamruzzaman, 2002; Chin, 2006; Xiao, 2009), the present study aims to fill the gap in the

understanding of tension fracture behavior of local cement-treated clay. This chapter presents

the test program to examine the fracture behavior of cement-treated Singapore Marine Clay in

tension. The corresponding compression behavior is also determined for comparison.

First of all, the base materials followed by the mix proportion of the materials used in the tests

are presented. Next, the sample preparation and the relevant tests are discussed. Three types

of test, i.e. uniaxial compression test (UCT), split tension test (STT) and three-point bending

notched beam test (BNT) were conducted in the present study to evaluate the fracture

behavior of cement-treated Singapore marine clay. The observed fracture behavior for

cement-treated Singapore marine clay both in compression and tension are presented and

evaluated in depth. Parametric studies were performed to investigate the effects of specimen

size and notch width in three-point bending notched beam test. Subsequently, issues involving

stability performance of three-point bending notched beam test and sand content in cemented

clay were investigated. Finally, a summary is presented to conclude the experimental findings

of this chapter.

46
3.2 Properties of Base Materials

3.2.1 Untreated Marine Clay

The Singapore marine clay typically composes of a sedimentary deposit which is known

locally as the Kallang Formation. This deposit is widely distributed and covers about one

quarter of the total land surface of the island (Pitts, 1992). This formation usually consists of

two layers separated by a stiff desiccated intermediate layer (Chong, 2002). The basic

properties of Singapore marine clay have been well studied (Tan, 1983; Yong et al., 1990;

Chong, 2002; Low, 2004). The upper marine clay is highly plastic with reported liquid limit

and plasticity index typically ranging from 75 to 120 and 50 to 80, respectively. The bulk unit

weight is between 14 kN/m3 to 16 kN/m3 and the natural moisture content is about 60% to

110%. On the other hand, the liquid limit and plasticity index of lower marine clay is reported

to range from 60 to 90 and 39 to 55, respectively. The bulk unit weight for lower marine clay

is slightly higher ranging from 15 kN/m3 to 17.5 kN/m3. The natural moisture content in this

layer is lower compared to upper marine clay which is approximately 47% to 70 %. In-situ

marine clay from Marina Boulevard, Southern part of Singapore Island was collected as

untreated soil in the present study. Table 3.1 presents the basic properties of the untreated

marine clay used in this study.

Table 3.1: Basic properties of Singapore marine clay from Marine Bouvelard site.

Properties Values Properties Value


Liquid Limit, LL (%) 60 Grain Size Distribution:
Plastic Limit, PL (%) 26 Sand (%) 11
Plasticity Index, PI (%) 34 Silt (%) 68
In-situ Moisture Content, m 55 Clay (%) 21
(%)
Specific gravity, Gs 2.57 Total Unit weight, b (kN/m3) 16.7

47
3.2.2 Ordinary Portland Cement

A common type of cement, Ordinary Portland Cement (OPC), is used in this study to treat the

Singapore marine clay. The chemical composition and physical properties of OPC provided

by the supplier are shown in Table 3.2. The cement powder was stored by packing in smaller

quantities using zip-locked plastic bags to ensure airtight in preventing hydration of the

cement powder. The cement content (Aw) used in this study is defined as the ratio of dry

weight of cement powder to the dry weight of soil particles and is expressed as a percentage.

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. The aforementioned

definitions follow as those used by Xiao (2009).

Table 3.2: Chemical composition and physical properties of Ordinary Portland Cement
provided by supplier.

Chemical compositions Value Physical Properties Values


Lime Saturation, Factors (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

3.3 Sample Preparation Procedure

The main purpose of this chapter is to investigate the fracture behavior of cement-treated

Singapore marine clay. Thus a commonly adopted range of cement content, and curing period

48
in practice were examined. Meanwhile, other variables such as water content and curing

condition were set to constant. Cement content in the range from 20% to 40% was selected in

this study; this range being reasonably representative of the cement content used to treat the

weak soil as reported by Lee et al. (2005), see Figure 3.1.

There is always a concern that excessive water in the mixture may lead to “bleeding” or

segregation of cement and soil solids from water. Such condition will occur once water

contact of the mixtures exceeds the bleeding limit, subsequently increases the tendency of

inducing in-homogeneity to the specimen. Figure 3.2 shows the liquid and bleeding limits for

cement-clay mixtures with different admixture proportions for the studies by Chew et al.

(1997) and Lee et al. (2005). In the present study, the total water content (Cw) of 100 % was

maintained throughout all the tests. A total of five mix sets for different proportion of soil-

cement-water are summarized and presented in Table 3.3. By plotting the soil-cement-water

ratio configuration from this study in Figure 3.2, the admixture in the present study falls

within the workable range bounded by liquid and bleeding limits. Hence it is confident to say

that there should not be any bleeding effect in the current adopted admixture proportion.

Table 3.3: Mixture proportions and tests conducted for cement-treated Singapore marine clay
in the present study.
Total
Mix Soil- Water- Cement Curing
water
proportions cement cement content, time Tests
content
(s:c:w) ratio (s:c) ratio (w:c) Aw (%) (days)
(%)
5:1:6 5:1 6:1 20 100 14&28 UCT, STT
UCT, STT,
5:1.25:6.25 5:1.25 6.25:1.25 25 100 14&28
BNT
UCT, STT,
5:1.5:6.5 5:1.5 6.5:1.5 30 100 14&28
BNT
UCT, STT,
5:1.75:6.75 5:1.75 6.75:1.75 35 100 14&28
BNT
5:2:7 5:2 7:2 40 100 28 BNT

Notes: UCT – uniaxial compression test; STT – split tension test; BNT – three-point bending
notched beam test; s – soil, c – cement; and w – water.

49
Untreated marine clay collected from site was first broken down from slumps using a pan

mixer. Meanwhile, any impurity such as shell was removed from the clay. Deionized water

was added to the natural clay to increase its water content. The water content of the clay

sample was identified prior to mixing. By knowing the weight and water content of the

untreated clay, the prescribed amount of cement and water were prepared accordingly. The

clay and cement slurry were then mixed thoroughly in a Hobart mixer using a rotational speed

of approximately 125 rpm for 10 minutes. Halfway through the mixing, it was interrupted for

about 1 minute to scrap the mixture attached to side wall of the mixer bowl in order to ensure

the mixture is uniformly and thoroughly mixed. The cement clay paste was then placed into

the prefabricated molds and submerged into water for curing. The water temperature in the

curing tank was about 24+3oC. A cylindrical PVC split-mold of inner diameter of 50 mm and

100 mm height was used for uniaxial compression test (UCT) specimen. On the other hand,

the split tension test (STT) specimen used the same cylindrical PVC split-mold but the height

was reduced by half to 50 mm. Three-point bending notched beam test (BNT) specimen used

a perspex mold with inner dimensions of 50 mm width x 50 mm depth x 250 mm length. The

contact surfaces between the mold and cement clay paste were greased. To reduce the

trapping of air bubbles in the specimen, the paste was compacted while placing in the mold.

The compaction was done in three layers for 100 mm height and two layers for 50 mm height

by slowly tamping the mold on a supported base (Tan et al., 2002).

3.4 Testing Procedure and Apparatus

After the desired curing time, the specimens were taken out from the curing tank and tested

accordingly. The strength of the treated soil was evaluated through UCT and STT. As the

fracture behavior of a material cannot be determined by its strength parameter, three-point

bending notched beam test was adopted to examine the tension softening behavior of this

treated soil.

50
3.4.1 Uniaxial Compression Test

In the uniaxial compression test (UCT), the specimen was placed in the loading machine on a

base plate. The upper specimen was covered by a cap where the loading shaft was placed on.

A 10 kN capacity load cell was attached to the loading shaft to measure the applied load.

Before starting the loading, the upper cap was adjusted to be in full contact with the specimen

and the loading shaft. The axial deformation of the specimen was monitored by the external

displacement transducer. The test then started by applying a constant axial strain of 1.00

mm/min (1% strain per minute) based on ASTM 5102-09 and BS 1377 - Part 7, (1990). The

experimental test setup is shown in Figure 3.3. The load and deformation values were

recorded to obtain a reasonably complete load-deformation curve. The loading was

terminated when the recorded load decreased or remained constant at 10% to 20% of the peak

load.

3.4.2 Split Tension Test

By definition, the tensile strength shall be obtained by the direct uniaxial tension test.

However, there are always difficulties to achieve a uniform tensile stress across a section of

the specimen, without inducing high localized stress concentrations elsewhere. This usually

happens when the specimen is attached to ordinary clamping grips where the grips give rise to

stress concentrations and multiaxial stresses. Another way of using adhesive to attach the

specimen is also not a good approach. Different properties between the adhesive and

specimen induce shear stresses in the region of the interface, which can cause the fracture to

occur. Therefore, the split tension test (STT) appears to offer a desirable alternative, because

it is much simpler to be executed and hence enhance the test performance consistency.

In the present study, 50 mm by 50 mm cylindrical specimens were used in split tension test.

The test setup is similar to that by Xiao (2009), see Figure 3.3. Displacement controlled

loading is adopted in this test and the loading rate is kept at 0.5 mm/min. Since there is no

51
difference between using loading rate of 0.5 mm/min and 0.005 mm/min as investigated by

Xiao (2009), a faster rate 0.5 mm/min is adopted in the present study.

3.4.3 Three-point Bending Notched Beam Test

One way of quantifying the tensile toughness is by means of fracture energy, Gf. As

recommended by RILEM (French acronym for the international union of laboratories and

experts in construction materials, systems and structures), the fracture energy, Gf can be

obtained through three-point bending notched beam test (BNT). In this test, three test data

were measured, i.e. applied load (F), beam vertical deflection (δv) and crack mouth opening

displacement (δc0). In the present study, the inner dimensions of mold used to prepare the

specimens were 50 mm in height, 50 mm in width and 250 mm in length. After the specified

curing time, the actual dimensions of the specimen were measured and notched at central

bottom by saw-cutting with the help of a jig. The notch depth was equal to half the beam

depth with approximately 2.5 mm wide. Parametric studies (to be further discussed in Section

3.6.4) were conducted to investigate the influence of specimen size and notch width for this

test. Subsequently, two thin aluminum knife edges were glued to the bottom beam surface,

centered across the notch. The shorter sides of the knife edges were placed against the

specimen such that the long sharp edge is farthest from the specimen and spaced 6mm apart

from each other. A clip gauge was then attached to the sharp edges to measure the crack

mouth opening displacement throughout the test. A potentiometer was positioned below the

specimen to measure the beam deflection in vertical direction. The specimen test span is four

times the specimen height which is equal to 200 mm in this case. The experimental test setup

is shown in Figure 3.4. The loading rate is kept at 0.01 mm/min similar to that adopted by

Namikawa (2006).

52
3.5 Compressive Fracture Behavior

Compressive behavior of cement-treated soil is commonly evaluated by uniaxial compression

test (also known as unconfined compression test, UCT) due to its simplicity. According to

Uddin et al. (1997), UCT has been performed to study the strength and strain development of

cement-treated soil for various cement contents and curing times. Although some researchers

have furthered the study involving confining stress, unconfined condition is deemed to be

more representative for cement-treated soil used in an open excavation. In the present study,

the range of cement content commonly adopted in practice is examined. The cement content

varies from 20% to 40%, while the curing period varies from 14 days to 28 days. Figures 3.5

to 3.8 show the stress-strain curves obtained from uniaxial compression tests for cement-

treated Singapore marine clay in the present study. Figure 3.5 shows that the stress-strain

curves for the 5 samples in UCT are reasonably similar including the post-peak region,

thereby illustrating the consistency of the results of the present tests.

In all the tests, the stress-strain curves of the treated samples are generally found to increase

abruptly up to the peak values and then decrease suddenly to small values upon further

straining. Unlike untreated clay which behaves ductile and tends to fail in shearing in UCT,

the addition of cement has transformed the soil into a quasi-brittle material and the failure

mechanism is more complex. Three types of failure modes are observed for UCT in this range

of cement content and curing day. For lower cement content, i.e. Aw = 20% and 25% cured in

14 days, shear and cone failure mode is observed (Figure 3.9a). It is noticed that the damage

occurs in a slanting direction across the specimen and a wedge failure pattern is observed at

the end. For the remaining tests except for Aw = 35% cured in 28 days, cone and split failure

mode as presented in Figure 3.9b is found. In combination to shear failure, the specimen also

experiences compressive crushing with cracks propagate in the vertical direction. For higher

cement content, i.e. Aw = 35% cured in 28 days, vertical compressive crushing is dominant

and columnar damage is observed. As shown in Figure 3.6, even at 14 days of curing period

53
for 20% cement content, quasi-brittle behavior is manifested. The peak load in this case is

maintained for certain period before the strength decreases gently to a very small or

insignificant residual strength. The curvatures of the stress-strain plots change in the post-

peak region; abruption drop is more obvious with increase in cement content and curing

period. In other words, the brittleness of the material increases with increase in cement

content and curing period. The observed strength increase with cement content and curing

period in the present study is consistent with that reported by other studies (Lorenzo and

Bergado, 2006; Kamruzzaman, 2002; Uddin et al., 1997). The brittleness which is indicated

by a sudden decrease of post peak stress is much more pronounced for prolonged curing

period and cement content. For higher strength material, the deviatoric stress increases

abruptly associated with a lower strain.

The unconfined compressive strengths (qu) of the cement-treated Singapore marine clay in the

present study for different cement contents and curing periods are summarized in Figure 3.10.

It is found that a certain required compressive strength can be achieved with increase either in

cement content or curing period. For instance, 28 days strength obtained from addition of 25%

cement content is approximate equal to 14 days strength obtained from addition of 30%

cement content. Hence, in practice, one can increase the cement content to achieve specified

compressive strength when construction time is tight, vice versa cement content can be

reduced to save cost if longer curing period is permitted. The increase in strength with cement

content has been recognized due to cement hydration and pozzolanic reaction that bind

together the clay particles which created a new bonded, stronger matrix of soil. In addition,

owing to pozzolanic reaction that can last for months, the strength of treated soil is expected

to increase with time. Nevertheless, it is observed that the effect of increasing the cement

content is more pronounced in increasing the strength as compared to curing period. The

unconfined compression tests in the present study categorize the cement-treated Singapore

marine clay that added with 20% cement content and above as quasi-brittle materials.

54
3.6 Tensile Fracture Behavior

Tensile strength of a quasi-brittle material is commonly evaluated by using the split tension

test (also known as the Brazilian tension test), as the specimen is easy to be prepared and

tested. However, this indirect tensile strength test provides no possibility of strain

measurement (Das and Dass, 1995). Furthermore, the tensile stress obtained from this test is

estimated from elastic theory (Frocht, 1948) which makes the post-peak behavior not valid.

When the cement-treated soil transforms into a quasi-brittle material, any small crack in a

region with tensile stress would cause a running crack, which might lead to a catastrophic

failure. Although the tensile toughness is an essential property for cementitious material, it

has never been explicitly taken into account in cement-treated soil design. This is due to the

fact that its importance has not been well understood and methods of taking it into account are

lacking. As discussed in Chapter 2 that the cement-treated soil does exhibit tension-softening

behavior, this aspect has not been quantified so far. In fact, there exists no well-established

test method to date. As such, the tension-softening behavior for cement-treated Singapore

marine clay is dealt with in detail herein.

Uniaxial tension test (also known as direct tension test) is a direct method of determining the

softening relation; by measuring the extension of the localization zone, the external load-

displacement relation to determine the post-peak stress-strain relation in the localization zone

can be determined. Nevertheless, there is an inherent problem in executing this test; it is not

possible to obtain a stable condition once the specimen undergoes peak-load (Hillerborg,

1985). As a result, methods using fracture mechanics concept were suggested to determine the

post peak tension-softening relation for quasi-brittle material such as concrete in a realistic

manner (Hillerborg et al., 1976; Li et al., 1987; Uchida et al., 1991; Niwa et al., 1998b). The

concept can be derived from the three-point bending test on a notched beam as recommended

by RILEM. The three-point bend test on a notched beam (also known as single-edge notched

beam) assumes the energy released is consumed by the formation of cracks. As such, fracture

55
energy, Gf, is defined as the energy consumed per unit crack area. This material property

parameter together with the tensile strength provides a comprehensive understanding of the

tensile behavior of a cement-treated soil. On the other hand, the tension-softening ( - δc)

relation interpreted from three-point bending notched beam test (BNT) provides information

on fracture resistance and can be used for numerical simulations of crack formation and

propagation in cement-treated soil. The tension-softening ( - δc) relation is identified as the

functional relationship between the traction acting across a crack plane and the crack opening

distance, in a uniaxial tension specimen that loaded quasi-statically to complete failure (Li et

al., 1987).

3.6.1 Split Tensile Strength

The split tension test is derived from the generation of compression-induced extensional

fracturing. In this test, a circular sample placed between two platens is loaded in compression

producing a nearly uniform tensile stress distribution normal to the loaded (vertical) diametral

plane, inducing the disk to fail in splitting (Rocco et al., 1999). The tensile stress is estimated

from the elastic theory (Frocht, 1948) as:

= (3.1)

where F is the applied load, L is the specimen length and D is the specimen diameter. Along

the horizontal diametral plane, tensile stress decreases from a peak level (Eq. (3.1)) as a

function of distance from the center to the edge of the disk. For quasi-brittle materials such as

concrete and rocks, the tensile splitting strength can be determined using Eq. (3.1) as

suggested by BS1881-117 (1983) and ASTM 3967-08.

56
Split Tensile Strength of Cement-treated Singapore Marine Clay

The split tension test results for the present study are presented in Figure 3.11. The split

tensile strength, , of cement-treated Singapore marine clay increases with increase in

cement content and curing period. Figure 3.12 shows the failed specimens after testing.

During the test, the applied vertical compressive loading at each end of the specimen diameter

induces tensile stresses normal to the vertical. Once the specimen experiences maximum

tensile stresses, fracture initiates perpendicular to the direction of the tensile stress. The

resulting crack then propagates in a plane normal to the tension, and the specimen finally

separates into two halves as shown in Figure 3.12.

Very often, split tensile strength, is correlated to unconfined compressive strength, qu.

Figure 3.13 shows the relationship between split tensile strength and unconfined compressive

strength for cement-treated Singapore marine clay in the present study. It is evident that a

good correlation can be obtained between these two parameters. The split tensile strength can

be correlated to the unconfined compressive strength (Figure 3.13) through the following

relation:

= 0.11 (3.2)

The correlation between unconfined compressive strength and split tensile strength in the

present study is very close to that observed by Xiao (2009) and also lies within the ranges

reported by other studies in Chapter 2. For instance, Porbaha et al. (2000) reported the ratio of

tensile to compressive strength is between 0.1 and 0.15. On the other hand, a ratio of between

0.1 and 0.3 was reported by Namikawa and Koseki (2007) for cement-treated Toyoura sand.

Clough et al. (1981) presented a ratio value of about 0.1 for cemented sand.

57
3.6.2 Fracture Energy, Gf

Petersson (1981) illustrated that the ordinary stress criterion cannot be used in analyzing a

crack problem, thus one has to use fracture mechanics approach. A sample tensile stress-strain

curve from direct tension test for a quasi-brittle material, concrete, is shown in Figure 3.14a.

Experimental result from Heilmann et al. (1969) showed that once the material experienced

maximum stress, the strain crossing the crack increased rapidly while the strain outside the

fracture zone decreased. Meanwhile, it was observed that the fracture zone is a narrow band

across the specimen. Petersson (1981) then modeled the fracture zone with a slit (Figure

3.14b) which is able to transfer stress and the transferring capacity is associated on the width

( c) of the silt. The total deformation ( ) of the specimen is given as:

= +δ (3.3)

where is the strain in the material outside the fracture zone. In other words, the mean strain

( ) of the specimen in Figure 3.14b can be given as:

= +δ / (3.4)

where δ is zero before the tensile strength is reached. After the maximum stress is reached,

the fracture zone width affects the mean strain and consequently the stress-strain curve. As a

result, stress-strain curve is not suitable to be used as a material property. Therefore, a better

way to describe the deformation properties of a material subjected to tensile failure is to use

two relations: one relation between the stress and strain for the material outside the fracture

zone (Figure 3.15a) and the other relation between the stress and the deformation of the

fracture zone (Figure 3.15b).

The almost linear - curve for material outside the fracture zone can be defined by the

Young’s modulus (E) and the tensile strength ( ). The area covered by - c curve in

fracture zone corresponds to the amount of energy that is necessary to create one unit of area

of a crack which is equal to the fracture energy (Gf). The main criterion in measuring Gf is to

58
let the crack propagates in a well-defined distance and then measure the energy consumption

due to the crack propagation. Referring to the method proposed by Hillerborg (1985) and

recommended by RILEM (1985), the fracture energy Gf can be determined by the use of

three-point bending test on a notched beam (BNT).

Figure 3.16 shows a schematic test load, F versus vertical deflection, δv curve obtained from

BNT. The total amount of absorbed energy when the specimen is broken into two halves is

the area below the F – δv curve. Nevertheless, as highlighted by RILEM, the specimen is not

only subjected to the imposed load but also by its own weight. The amount of energy supplied

from the loading force and the weight of the beam must equal to the absorbed energy. The

measured load-deflection data is shown by the solid line. Meanwhile, a hypothetical F – δv

curve denoted by a dashed line is introduced to cater for the correction due to specimen self-

weight. The additional load F0 is the central load which induces central bending moment from

the weight of the specimen that is excluded from the measured load F. F0 is expressed as:

= g (3.5)

where is the specimen mass between the supports and g is the gravitational acceleration.

The total amount of absorbed energy, W can be taken as:

= + + + (3.6)

W0 is the area below the measured load-deformation curve, and W1 is:

= = g (3.7)

in which δ0 is the deformation when F = 0 and the specimen breaks completely. Petersson

(1981) reported that W2 is approximately equal to W1. It was found that typically W3 is so

small (less that 1 – 2% of the total energy) that it can be neglected. As a result, the total

absorbed energy becomes:

59
= + 2 = + g . (3.8)

This fracture energy Gf can be obtained as the amount of energy divided by the area of the

ligament Alig:

( ) ( )
= = ( )
(3.9)

where is the ligament depth, b is the beam width, and db is the beam depth. The specimen

dimension for BNT is illustrated in Figure 3.17.

Fracture Energy of Cement-treated Singapore Marine Clay

The load-deflection (F – δv) and load-crack mouth opening displacement (F – δc0) curves of

the test results for various cement contents at 14 and 28 curing days are presented in Figures

3.19 to 3.22. As illustrated in Figure 3.18, the F – δv and F – δc0 relations are repeatable even

after the peak loads. Figure 3.19 shows the effect of cement content on the mixture is evident

even at 14 curing days period; the maximum applied load, F, that the specimen can resist

increases with increase in cement content. The curve gradient of pre-peak load against

deflection or crack mouth opening displacement becomes steeper with increase in cement

content. In other words, the specimen is getting stiffer and reaches peak load at a smaller

deflection with higher cement content. In comparison to deflection, the crack mouth opening

displacement value is smaller. Meanwhile, the differences between the crack mouth opening

displacement values at peak load for various cement contents are smaller as compared to

deflection. It is also observed that the tail of the curve reduces with increase in cement content.

Similar observations discussed herein are noted for 28 days curing period specimen. However,

it is recorded that the crack mouth opening displacement values at peak loads for various

cement contents are very close to each other in 28 days curing period. In the post-peak load

zone, the load decreases in a steeper gradient for specimen with higher strength. This test

once again manifests the brittle behavior of this material which is consistent to the finding in

60
UCT reported earlier. The term “brittleness” in the present study is defined as the lack of

ductility and the degree of brittleness is reflected in low values of percentage of elongation of

reduction in area (Morley, 1944). Once the brittleness of the material increases with cement

content, it causes the difficulty of maintaining the load stability after peak load. This is

observed in specimens with 28 days curing period for higher cement content greater than 25%,

the load is unstable after 20% reduction of peak load. As a result, the fracture energy (area

covered by the load-deflection curve) cannot be obtained from these instability tests. The test

stability problem will be further investigated and discussed in Section 3.6.5.

The fracture energies calculated from these F – δv curves are summarized in Table 3.4. It is

found that even the material can resist a higher load, the corresponding fracture energy, Gf is

not necessary proportionally higher. As discussed earlier, the material with higher peak load

will decrease sharply at post-peak zone with smaller deflection which reduces the area under

the curve tremendously. In addition, the increment of this peak load is relatively small and

thus contributes less to the overall fracture energy. Smaller δ0 recorded by brittle material

contributes to smaller W1 as well as Wo; and if the increment of peak load is relatively small,

the material although with higher peak load Fmax will experience a smaller value of Gf. This is

shown in the comparison of Gf between Aw = 25% (Test A1) and Aw = 30% (Test B1) for

specimen under 14 days curing period; as well as Gf for Aw = 25% under 14 days (Test A1)

and 28 days (Test A2) curing period.

The fracture process around the notch tip can be illustrated in Figure 3.23 which shows the

hypothesis stress distribution (modified from Petersson, 1981) in front of the notch tip

corresponds to the respective stages indicated in the load-deflection curves. Notched beam

helps the fracture zone to start developing as soon as the specimen becomes loaded due to

stress concentration. A zone of micro-cracks will develop in front of the notch and this

fracture zone reduces the stress concentration. In the beginning at point A, no stress exceeds

61
the tensile capacity. When the maximum load reaches point B, crack (with stress-free surface)

starts to propagate. Even though a crack is formed and propagated, the material is still able to

transfer stress at point C until the crack propagates throughout the entire specimen.

Table 3.4: Summary of fracture energies, Gf from BNT tests.

Curing
Aw (%) Test Fmax (N) Wo (N.m) W1 (N.m) Gf (N/m)
Period
25 A1 12.32 0.002558 0.002795 4.2824
30 B1 16.68 0.002570 0.001871 3.5528
14 days
35 C1 21.48 0.004046 0.001403 4.3598
40 D1 26.12 N.A N.A N.A
25 A2 19.97 0.001999 0.001270 2.6151
30 B2 26.67 N.A N.A N.A
28 days
35 C2 30.49 N.A N.A N.A
40 D2 34.45 N.A N.A N.A

3.6.3 Tension-softening Relationship

It is a more representative way to illustrate the tension-softening behavior of cement-treated

Singapore marine clay through stress-crack opening displacement ( - c) relationship. This

constitutive relationship between the tensile stress transferred across a crack plane and the

separation distance of the crack faces is a material property. As mentioned earlier, the direct

approach of obtaining the tensile stress-crack opening displacement relationship is from

uniaxial direct tension test. However, owing to the challenges discussed earlier, an indirect

approach, energy concept has been adopted to interpret the tension-softening relationship

from BNT. The J-integral theory which represents a way to estimate energy per unit fracture

surface area in a material method was first introduced by Cherepanov (1967) and Rice (1968)

in late 1960s. Later, the experimental method was proposed by Li et al. (1987) and

subsequently improved by Rokugo et al. (1989) and Uchida et al. (1991). In the present study,

the modified J-integral method proposed by Uchida et al. (1991) is adopted to develop the -

62
c relation of cement-treated Singapore marine clay. The basis of this method is discussed in

the following section.

Theoretical Basis of J-integral Technique

The method proposed by Li et al. (1987), is based on J-integral concept shown in Figure 3.24.

The path independent J-integral can be defined as

= Г
[ ] (3.10)

where is a curve around the notch tip, is the strain energy density, is the traction

vector, ui is the displacement vector and is an arc along . Based on cohesive zone model,

Rice (1968) suggested that the general contour can be shrunk into 1 running alongside the

cohesive zone (Figure 3.24) and Eq. (3.10) reduces to

= ( )( / ) (3.11)

where ( ) and ( ) are the normal stress and opening displacement at the point in the

cohesive zone. The upper integral limit L is a measured point ahead of the physical crack

where ( ) = 0. If the crack opening at each point in the cohesive zone increases by an

amount , then the profile of the cohesive zone boundary extends a distance . Eq. (3.11)

may also be expressed as:

( )= ( ) = ( ) (3.12)

where is the crack separation distance. A critical value of Jc = Gf is reached when the crack

opening at physical crack tip reaches maximum opening . Jc may be interpreted as the rate

of energy absorption in the cohesive zone with respect to crack tip propagation.

63
According to Rice (1968), J can also be interpreted as potential energy with respect to crack

advance:

= = (3.13)

where is the potential energy, is the crack length, and the minus sign in Eq. (3.13)

indicates a decrease in potential energy when the crack extends in length.

The basis of Li’s J-integral method of finding the tension-softening ( ) relationship is to

obtain J experimentally using Eq. (3.13) and then substitute it into Eq. (3.12) to find ( ).

Potential energy may be computed simply from a load-deflection curve obtained from the

three-point bending notched beam test. In view that the crack tip position is difficult to locate

accurately, Li et al. (1987) proposed to use two specimens with slightly different notch

lengths. Then, J-integral value can be expressed as a function of δv by dividing the area

between the two load-deflection curves A(δv) by the difference in the ligament area b • :

U
J(δ ) = A(δ ) /[b • a ] (3.14)

where is the difference of notch lengths (= a2 - a1) and b is the width of specimen. The

crack opening is taken as an average of crack mouth opening displacement δc01 and δc02.

Figure 3.25c shows the relation between the J-integral value, J and . By differentiating the J

- curve (similar to Eq. (3.12)), the tension-softening curve can be determined as presented

in Figure 3.25d. However, correction is needed to eliminate the initial ascending part which

may be interpreted as the sum of recoverable deformation and inelastic deformation caused by

micro-cracking prior to the localization of deformation onto the fracture process zone thus

this should not be regarded as part of the - c relationship.

( )= ( )/ (3.15)

64
Modified J-integral Method

(i) Rokugo et al. (1989)

In the method proposed by Li et al. (1987), there is a concern in controlling the experimental

data variation since two specimens are used. Therefore, some modifications had been made

by Rokugo et al. (1989) and Uchida et al. (1991) to obtain the - c relationship from a single

specimen test. Rokugo et al. (1989) proposed a new J-integral method to simplify Li’s

approach. In this method, a virtual specimen with a full notch (without ligament) and no

weight is considered in addition to one loaded notched specimen. The deflection axis in the

actual load-deflection curve is considered as the load-deflection of the virtual specimen.

Figure 3.26 shows the modified load-deflection curve (dashed line) after considering

additional load from half of the specimen weight between two supports. Since the crack

opening for virtual specimen is zero, a fictitious crack width c is taken as half of the crack

mouth opening displacement c0 ( c= c0/2) in actual specimen. In other word, the crack width

is fictitiously assumed to be uniformly distributed throughout the entire ligament length a0.

(ii) Uchida et al. (1991)

The assumption of fictitious crack distribution was later improved by Uchida et al. (1991). As

shown in Figure 3.27, the fictitious crack width is assumed to be distributed along the

rotational axis on top of the notch. Therefore, the fictitious crack width can be expressed as:

( )= / (3.16)

where is the crack mouth opening displacement. Changing variables from to gives

the following equation for energy respect to crack advance:

= ( ( )) = ( ) = ( ) (3.17)

65
in which is the ligament area (b • ) of the specimen. Differentiating Eq. (3.17) against

gives the following

( )= [ ( )+ ( )] (3.18)

Further differentiating Eq. (3.18) against gives the following equation:

( )= ( )= [2 ( )+ ( )] (3.19)

in which (area under load-deflection curve after considering the beam weight) and can

be experimentally obtained from BNT. By replacing with , the tension-softening

relationship can be established by employing Eq. (3.19).

Tension-softening Relationship of Cement-treated Singapore Marine Clay

The tensile stress-crack opening displacement ( t – c) curves of the test results for cement

content, Aw = 25% and 30% at 14 curing days (Tests A1 and B1) are presented in Figure 3.29

and Figure 3.30. For the load-deflection curves presented earlier, unlike long tail softening

observed in concrete test, the descending part is nearly linear for this much brittle material. As

a result, the value is getting close to zero when differentiating it to obtain the stress value. For

instance, some part of the stress value cannot be obtained for Tests C1 to A2. The area

covered by the t – c curve is equal to the fracture energy Gf as represented by the hatch area

in Figure 3.28. It is observed that the tensile stress-crack opening displacement relationship

can be represented by a bilinear form. This is similar to the test results for cement-treated

Toyoura sand and concrete. Nevertheless, it is found that the bilinear relationship is getting

replaced by a linear relationship when brittleness increases as illustrated in Test B1. It is

realized that more efforts are needed to establish the t – c relationship for cement-treated

Singapore marine clay as the interpretation is very sensitive to the curvature of the test data.

66
As such, it is suggested to follow the practice for concrete where the fracture energy Gf or

fracture toughness is used as a parameter to represent the material softening behavior.

3.6.4 Parametric Studies

To the best knowledge of this author, up to date there does not exist any study for tension-

softening study of cement-treated clay. Three-point bending notched beam is a recommended

test to determine the tension-softening behavior for quasi-brittle material such as concrete and

asphalt. Namikawa (2006) successfully applied this test on much weaker material, cement-

treated sand in a smaller dimension compared to concrete and asphalt. Owing to the closer

similarity between cement-treated clay and cement-treated sand compared to concrete and

asphalt, the present test specification for three-point bending notched beam follows that of

Namikawa (2006)’s work. Nevertheless, cement-treated clay is a weaker material compared

to cement-treated Toyoura sand; moreover, the particle size for clay is an order smaller than

Toyoura sand. Therefore, there is a necessity to investigate the influences of specimen size

and notch width on this test for cement-treated Singapore marine clay.

(i) Effect of notch width (n)

Measuring crack mouth opening displacement in a bending test requires the help of pre-

fabricated notch in the specimen. Without the pre-fabricated notch, it is hard to identify the

crack location although in theory it should appear at the center of the specimen where loading

applies. Meanwhile, the fracture zone will not propagate until the tensile strength is reached at

the bottom of the beam. Conversely, for notched beam, the fracture zone due to stress

concentrations starts developing as soon as the loading on the specimen begins (Petersson,

1981). According to the test specification stated by RILEM, the notch depth should be equal

to half the beam depth which is 25 mm for the present specimen size (50 mm x 50 mm x 200

mm). For concrete, Baratta and Underwood (1992) proposed a ratio of narrow notch width to

67
beam depth (n/d) = 0.01 for tests in which a close modeling of the ideal crack compliance is

important. Therefore, for the beam depth of 50 mm in the present study, a notch width of 0.5

mm should be adopted. Based on past experiences, it is recommended to create the notch with

saw-cutting. The reason for this is that creating the notch during casting may cause premature

crack as the strength is still very low. Moreover, splitting action may occur when the mold for

the notch is removed. However, saw cutting a 0.5 mm wide notch is truly a challenge due to

equipment availability to handle this very thin width. As such, two types of notch widths, i.e.

0.7 mm and 2.5 mm are compared herein to investigate the influence of the notch width in the

present study and subsequently to adopt a suitable yet workable notch width. The latter notch

width 2.5 mm was applied by Namikawa (2006) for beam depths ranging from 40 mm to 80

mm.

The load-deflection (F – δv) and the load-crack mouth opening displacement (F – δc0) curves

of the test results for notch widths 0.7 mm and 2.5 mm are presented in Figures 3.30. The test

results for these two notch widths are very similar to each other and recording approximately

the same maximum peak load, deflection as well as crack mouth opening displacement.

Unlike concrete, the test result is affected by the notch width which is believed to be related

to the particle size around the notch tip. For concrete, the material is made of cement paste,

sand and aggregate in which the aggregate size might be bigger or smaller than the notch

width. Thus, the particle size distribution around the notch tip might affect the stress

distribution. For the cement-treated clay in the present study, the particle sizes for clay as well

as cement are relatively small so a more uniform particle size distribution is established

around the notch tip. As such, the test results for notch width 0.7 mm and 2.5 mm are similar.

In view that there is no difference in results between notch widths of 0.7 mm and 2.5 mm for

the test, n = 2.5 mm was adopted for all the tests with specimen size 50 mm x 50 mm x 200

mm due to higher workability.

68
(ii) Effect of specimen size

The specimen size used for three-point bending notched beam test is closely related to the

particle sizes used to prepare the specimen. It is important to ensure that the representative

part of the material is tested and not affected by localized particle distribution. As such, the

recommendation of concrete specimen by RILEM is based on maximum aggregate size. For

maximum particle size between 1 mm to 16 mm, the recommended specimen depth is 100

mm. For Toyoura sand with nominal size approximately 0.2 mm, Namikawa (2006)

conducted three-point bending notched beam tests using specimen depth ranging from 40 mm

to 80 mm. The test results showed no significant specimen size effect on the material fracture

energy, Gf. In the present study, the nominal particle size for Singapore marine clay is

approximately 8 m. Hence, two specimen sizes, i.e. 50 mm x 50 mm x 200 mm and 40 mm

x 40 mm x 160 mm are compared herein to evaluate the influence of specimen size in the

present study and subsequently adopt a suitable specimen size.

The load-deflection curves and load-crack mouth opening displacement curves of the

aforementioned two specimen sizes are presented in Figures 3.31. The trend for these two

specimen sizes is identical with higher load resisted by bigger specimen. Nevertheless, the

resistant stresses after normalizing load to area are similar to both specimens. Specimen size

affects the fracture zone depth as well as the stress distribution within this zone. For a small

specimen, the process zone is comparatively large to the size of the specimen and efficient in

reducing the stress concentrations at the crack tip. Bazant and Kazemi (1991) reported that the

fracture energy, Gf increases as the specimen size increases. For bigger specimen size, the

stress transferring capacity near the notch tip decreases and consequently a crack starts

propagating closer to the maximum load for deeper beams. This is also illustrated in Petersson

(1981)’s study on a range of beam depth from 0.05 m to 1.6 m. Thus it can be seen that the

stress transition zone involves a considerable range of specimen size. It is believed that these

two specimen sizes, i.e. db = 50 mm and 40 mm used in the present study are considered as

small specimens where the 10 mm difference is insignificant compared to the overall

69
transition zone. Therefore, both of the specimens yield similar test results as presented in

Figure 3.31. However, as the size of the specimen increases up to a certain size, the relative

size of the process zone decreases, eventually becoming negligible and Gf would be size

independent when large specimens were used. Therefore, a considerable range of the

specimen size should be studied in detail in future to investigate the specimen size effect on

three-point bending notched beam test for cement-treated marine clay.

Nevertheless, when deciding the specimen size, it should be ensured that the specimen is easy

to handle and the risk of breaking the specimen during handling should be minimized.

Increase in specimen size might induce stability problem; this has been highlighted by

RILEM that the net stress at notched section due to the weight of specimen shall be much less

than the net stress at failure. As such, the choice of the proposed specimen is a compromise

between different requirements, which are often contradictory (Hillerborg, 1985). In the

present study, a beam depth of 50 mm is adopted. Beam depth 50 mm is preferred to 40 mm

as more space is available for placing the transducers underneath the specimen. In addition,

the load that can be resisted by specimen 40 mm is smaller and this might challenge the

handling process as well as the load cell sensitivity. Beam depth greater than 50 mm is not

considered in the present study as there is a concern about the occurrence of stability problem

for increase in specimen size as discussed in the next section.

3.6.5 Outstanding Issues

(i) Stability problem in three-point bending notched beam test

In this section, the stability problems for Tests D1, B2, C2 and D2 are discussed. It is always

a challenge yet important to maintain a stable test in order to obtain a complete load-

deflection and load-crack mouth opening displacement data for further interpretation of the

material fracture mechanism. However, stability problem had occurred in Tests D1, B2, C2

70
and D2 where a sharp increase in beam deflection and crack mouth opening displacement

which is an early sign indicating the specimen will soon break into two halves. The stability

problem is believed to be related to the stiffness of the testing machine as compared to the

brittleness of the tested material.

It is recommended to perform fracture test in deformation control instead of load control

owing to the latter method always produces an unstable fracture when the maximum load is

reached (Petersson, 1981). The deflection, transferred to the specimen can be expressed by

the following

= / (3.20)

where is the stroke of the machine, F is the force, and k is the stiffness of the machine. A

typical load-deflection curve for a specimen is shown in Figure 3.32. Eq. (3.20) corresponds

to a series of straight lines with a constant decline slope at different locations depending on

the stroke . The stability criterion is that the decline sloping line shall be gentler than the

specimen curve. In other words, the stiffness of the machine has to be greater than the decline

tangent curve which can formally be expressed by:

> / . (3.21)

As observed from the present tests, the brittleness of the cement-treated Singapore marine

clay increases with cement content and curing period. The brittleness is manifested when the

descending part of the F - v curve becomes steeper. To certain extent in the present study,

tests D1, B2, C2 and D2 where the steepest slope of the descending part is greater than the

stiffness of the testing machine, unstable fractures appear. In addition, it is also believed that

the specimen self-weight in some degree aggravated the sudden rupture behavior. For a weak

yet brittle material, the specimen self-weight between two supports can attribute to the

acceleration of crack propagation rate at the center loading point.

71
(ii) Effect of sand on behavior of cement-treated Singapore marine clay

Generally for Singapore marine clay at shallow depth, it is not clean but comprises a mixture

of clay and sand. As such, the effect of increasing the sand content in the mixture of cement

and marine clay on the fracture behavior of cement-treated Singapore marine clay is evaluated.

Filter sands in dry weight mix proportional 20% and 40% were added into the marine clay

sample to mix together with cement. Tables 3.5 and 3.6 summarize the nominal sizes for

materials and mixture proportion used in this study respectively.

Table 3.5: Nominal size for materials used in present study.

Material Marine Clay Cement Filter sand


Nominal size ( m) 8 15 200

Table 3.6: Mixture proportions and experimental tests with addition of filter sand in clay.

Soil
Mix Sand Cement Curing
component Total water
proportions content content,Aw time Tests
:sand-clay content (%)
(s:c:w) (%) (%) (days)
ratio
UCT,
5:1.5:6.5 1:4 20 30 100 28
BNT
UCT,
5:1.5:6.5 2:3 40 30 100 28
BNT

Table 3.7: Chemical compositions of filter sand provided by supplier.

Chemical
SiO2 Al2O3 Fe2O3 CaO MgO TiO2
composition
Value 99 0.52 0.15 0.04 <0.01 0.05

Na2O K2O P2O5 MnO SrO ZrO2 LOI


0.01 <0.01 <0.01 <0.01 <0.01 0.04 0.17

Table 3.5 reveals that the nominal size of filter sand is approximately 25 times greater than

the marine clay. Table 3.7 presents the chemical compositions of filter sand obtained from the

72
supplier. Figure 3.34 shows the UCT results for cement-treated clay mixed with filter sand

(C+Sand) compared with only clay (C), as presented in Section 3.5. The test results show that

the compressive strength decreases with increase in sand content for the same water content

and cement content. This is believed to be closely related to the chemical reaction between

sand, clay and cement particles. Kamruzzaman (2002) compared the content of kaolinite,

quartz and illite in untreated and cement-treated Singapore marine clay through X-ray

diffraction analyses and found that kaolinite is preferentially attacked to form cementitious

constituents (calcium silicate hydrate, CSH and calcium aluminium silicate hydrate, CASH)

followed by illite and quartz. As shown in Table 3.7, the quartz content of filter sand is

reported up to 99% while the kaolinite is less than 1%. As a result, the mixture strength in this

study is lower compared to the mixing with pure clay as the mixture strength is proportional

to the cement hydration and pozzolanic reaction. It is also observed that the brittleness

reduces with increase in sand content.

Figures 3.34 and 3.35 show the BNT results for cement-treated clay mixed with filter sand

(C+Sand) compared with cement-treated clay. Unstable test condition was observed with 20%

addition of sand content. Mixture with 40% sand content exhibits less brittle behavior and the

measured fracture energy summarized in Table 3.8 is close to that of Test B1 reported earlier.

For concrete, Issa et al. (2000b) reported that the fracture toughness increases with aggregate

size associated with increase of resistance encountered by the propagating crack. When a

growing crack reaches an aggregate, it either goes around or penetrates through it. Since the

fracture energy of the interface is smaller than the fracture energy for the aggregate, a crack

more often travels around the aggregate. Nonetheless, the particle size for filter sand in the

present study is relatively small so the aforementioned effect is not reflected in the present

tests.

73
Table 3.8: Summary fracture energy Gf for cemented clay and cemented clay+sand.

Curing
Aw (%) Test No. Fmax (N) Wo (N.m) W1 (N.m) Gf (N/m)
Period
28 days 30 Sand40% 13.96 0.001927 0.002003 3.1440
14 days 30 B1 16.68 0.002570 0.001871 3.5528

3.7 Summary

This chapter examines the fracture behavior of cement-treated Singapore marine clay in both

compression and tension. A commonly used range of cement content between 20% to 40%

was examined in this study. The test program involves three types of tests, i.e. uniaxial

compression test (UCT), split tension test and three-point bending notched beam test. The test

results are observed to be repeatable including the post-peak softening demonstrating the

reliability of the test findings.

UCT was conducted to evaluate the behavior of cement-treated Singapore marine clay under

compression loading. Generally the stress increases rapidly to a peak value and then decreases

rapidly to a small value upon further straining. For material with cement content of 20% and

cured for 14 days, the admixture becomes quasi-brittle. The strength associated with

brittleness increases with cement content and curing period; this finding is consistent with the

findings of existing studies reviewed in Chapter 2. For lower cement content cured in 14 days,

shear and cone failure is observed where the specimen experiences damage in a slanting

direction across the specimen followed by wedge failure at the end. Split and cone failure is

observed for the remaining specimens. For higher cement content of 35% cured in 28 days,

vertical compressive crushing is dominant and columnar failure is observed.

Owing to quasi-brittle behavior manifested in UCT, a split tension test was carried out to

evaluate the tensile strength of cement-treated Singapore Marine Clay. Experimental results

show that the split tensile strength increases with cement content and curing period. The split

74
tensile strength can be well correlated with the unconfined compressive strength, st = 0.11qu

in the present study. Nevertheless, as commented by Das and Dass (1995), this indirect tensile

strength test provides no possibility of strain measurement. In addition, this standalone

strength parameter cannot capture the tensile behavior of the cement-treated soil. As a result,

three-point bend test on notched beam (BNT) was conducted to further investigate the post-

peak tensile behavior for cement-treated Singapore marine clay.

The post-peak tensile behavior is well represented by a relation between the stress and

fracture zone deformation. Fracture mechanics concept is adopted in interpreting the tensile

stress-crack displacement relationship for cement-treated soil from BNT. Fracture energy, Gf

is often used as a property to describe the tension-softening behavior of a material.

Experimental results in the present study observe that Gf does not always increase with

cement content and curing period. There is a condition when the load resistance increment is

relatively small compared to a faster rupture manner associated with smaller deflection (more

brittle material), then the obtained Gf will be smaller. Generally the Gf falls within a range of

2.6 N/m to 4.4 N/m in the present study. A bilinear relation is portrayed in the tensile stress-

crack opening displacement relationship. However, the relationship becomes linear when the

material becomes brittle.

Parametric studies were conducted to investigate the effects of notch width and specimen size

on the BNT. Test results show that there is no difference in adopting notch width 0.7 mm and

2.5 mm. It is believed that a uniform particle size distribution is established around the notch

tip eliminating any localized stress concentration due to individual particle. On the other hand,

specimens with 50 mm depth and 40 mm depth also yield identical results. The 10 mm

difference in size is insignificant compared to the overall stress transition zone which usually

involves hundreds of mm.

75
Finally, the outstanding issues related to the stability problem in BNT and sand content in the

cement-treated clay are discussed. The stability problem that occurred in the few tests in the

present study is closely related to the stiffness of the testing machine comparatively to the

brittleness of the specimens. It is observed that the stability problem is more pronounced for a

much brittle specimen. In Singapore, the upper marine clay may comprise a mixture of clay

and sand at shallow depth of the soil layer. Therefore, a certain proportion of filter sand with

nominal size approximately 0.2 mm was added to the mixing of clay and cement. The test

results show that the unconfined compressive strength and brittleness decrease with increase

in sand content. It is believed that the chemical compositions of filter sand are less active to

formed cementitious constituents as compared to clay.

76
Chapter 4 Constitutive Model for Cement-treated Soil

4.1 Introduction

CDIT (2002) recommends the conduct of numerical analysis in assessing the performance of

cement-treated soil in geotechnical design. As reviewed in Chapter 2, there are still gaps in

previous numerical studies on simulating the behavior of cement-treated soil. Some studies

(Lee et al., 2004; Xiao, 2009; Horpibulsuk et al., 2010; Arroyo et al., 2012) focused on post-

peak behavior of cement-treated soil in compression loading. However, cemented soil can be

weakened in tension and fails in bending as reported by several experimental studies

(Babasaki et al., 1997; Larsson, 1999; Kitazume and Maruyama, 2007). Thus, a constitutive

model that is able to capture the post-peak behavior of cement-treated soil in both

compression and tension is needed in analyzing the problem.

In this chapter, three types of constitutive models, i.e. elastic-perfectly plastic model, isotropic

model, and concrete damage plastic (CDP) model to describe the behavior of cement-treated

soil are reviewed. Evaluation on these model predictions are carried out by referring to the

published laboratory data on cement-treated Toyoura sand by Namikawa (2006) and Koseki

et al. (2005), and cement-treated marine clay in Chapter 3. A decent evaluation is conducted

since the calibrated laboratory tests comprise three possible failure modes in terms of

compressive, tensile and bending. Then, a representative constitutive model to simulate the

cemented soil behavior is identified among these three constitutive models. The shortcomings

and strengths of these constitutive models are examined through the calibration tests.

4.2 Finite Element Method

Improvement in computer technology and CAD systems for the past decades has numerical

studies feasible where sophisticated computational problem can be solved with relative ease.

94
Solving a complex three-dimensional (3-D) problem is no longer out of reach when parallel

processing is available today. The parallel processing is supported by a system harnessing the

power of many CPUs to work together at one time. FEM, one of the numerical techniques,

has proven to be a powerful analysis tool in solving geotechnical engineering problems. This

method is empowered with three distinct features. First, a domain is discretized into

geometrically simple subdomains. Second, in each subdomain the material properties and the

governing relationships of the problem are approximated. Thirdly, each individual segment of

the solution should be compatible with its neighbors and through certain inter-element

relationships; all the elements will be assembled to provide the approximate behavior of the

domain after solving it. In solving geotechnical engineering problem, this method is superior

over other techniques such as empirical charts and simplified beam-and-spring as realistic

assumptions are used (Lee, 2008).

A general purpose finite element software, Abaqus/Standard version 10 which employs

implicit integration scheme is adopted to run the analyses in the present study unless

otherwise stated. This software is developed by Dassult Systems/SIMULIA. Its ability to

solve a wide variety of problems and modeling of many complex geometrics and material

behaviors has been well proven. The analysis procedure in this software involves three stages

of activity: pre-processing, processing and post-processing. At the first stage, input file is

created where the model of the physical problem is defined. The input file can be created

using interface Abaqus/CAE or a text editor. The processing stage involves solving the

numerical problem defined in the model. Lastly, the post-processing stage where the analysis

results are generated and reviewed using the Visualization module of Abaqus/CAE or another

post-processor.

95
4.3 Constitutive Models for Cement-treated Soils

In finite element analysis, the soil behavior can be simulated properly according to the

problem when an appropriate constitutive model is adopted. The literature presented in

Chapter 2 discusses the mechanical behavior of cement-treated soil and the constitutive

models used. The choice of constitutive model to be used depends largely on the analysis type

and the range of stress values that the material is likely to experience. Some sophisticated

constitutive models have been proposed lately but they require extensive and high accuracy of

laboratory data for calibration. As a result, the use of these models so far is limited to

laboratory data calibration. There is yet to see any detailed published field case study

employing them. It is a challenge to adopt the minimal parameters from common laboratory

test to establish the soil mechanical behavior in numerical analysis. Among all, three types of

constitutive soil model, i.e. elastic perfectly-plastic Tresca, isotropic and concrete damage

plasticity models are reviewed in this section to evaluate the most representative simulation of

cement-treated soil behavior.

4.3.1 Elastic Perfectly-Plastic Tresca Model

Elastic perfectly-plastic Tresca model is the most common model used by practitioners (Khoo

et al., 1997; Goh, 2003; Hsieh et al., 2003; Zhang, 2004; Yang, 2009) to simulate the behavior

of cement-treated soil in numerical analysis. This model can be used to define a material

which upon reaching certain stress state, undergoes deformation which is irreversible. It is a

path-dependent behavior model. There are two regions in this model known as elastic and

plastic. The elastic region is governed by linear stress-strain relationship following Hooke’s

law in small strain. In this region, strains and deformations are completely recoverable upon

unloading. In elastic region, total strain is completely contributed by elastic strain and is

related to total stress as:

= (4.1)

96
where is the total stress, is the elasticity tensor, and is the total elastic strain. In

matric form, the stress-strain relationship for linear elasticity in isotropic case is given by

1/ v/ / 0 0 0
/ 1/ / 0 0 0
/ / 1/ 0 0 0
= (4.2)
0 0 0 1/ 0 0
0 0 0 0 1/ 0
0 0 0 0 0 1/

where is the shear strain, is the Poisson’s ratio, E is the Young’s modulus, and G is the

shear modulus. The linear stress-strain relationship is governed by E and . The shear

modulus, G, can be expressed as G = E/2(1+ ).

Beyond the elastic region, the material yields and flows irreversibly under loading. In

plasticity, part of the strains and deformations are irrecoverable upon unloading. The

distinction between elastic and plastic response is known as the yield surface and it is one of

the basic ingredients to describe plastic behavior. Further straining causes the stress state to

remain on the yield surface, as the states that lie outside are not permissible. Monotonic

stress-strain behavior of an ideal plasticity assumed in Tresca model is shown in Figure 4.1.

The elastic material follows the same path in loading and unloading while the plastic material

shows a history-dependent unloading.

Yield function

Tresca criterion assumes yielding occurs when the shear stress on any point in a material

reaches a given value in the same plane. The Tresca model is a pressure-independent model

with a hexagonal prism yield surface when plotted in 3-D principal stress-space (Figure 4.2).

The hexagon shape is formed from the intersection of the Tresca yield surface with a

deviatoric plane. The vertices of the hexagon, which are located at triaxial compression and

97
extension points, create singularities of the normal gradient to the yield surface. The Tresca

yield function, is given as:

= =0 (4.3)

where is the second stress invariant which is a measure of the distance between the current

stress state and the hydrostatic axis in the deviatoric plane; is the Lode angle which defines

the orientation of the stress state with respect to the principal stress in this plane; and is the

undrained shear strength of the soil.

Although principal stresses ( , , ) are commonly referred, the isotropic yield function is

conveniently to be represented by three stress invariants and (Nayak and Zienkiewicz,

1972). These invariants can be expressed in terms of stress tensor, as below. is the first

principal invariant of the Cauchy stress:

= + + . (4.4)

The second and third invariants are associated with the deviatoric stresses:

= ; (4.5)

= (4.6)

where deviatoric stress

= /3 ( denotes the Kronecker delta). (4.7)

The lode angle, , can then be expressed as:

= ( / ) ( 30 30 ). (4.8)

In the plastic region, the material behaves differently from elastic region. The term “flow rule”

is often used to describe the plastic behavior. It is well known that ratios of strain increments

98
in the plastic region are not dependent upon the ratios of the stress increments but instead on

the stresses to which the material is subjected. Therefore, a potential function which is a

scalar function of position is adopted to describe the plastic strain increments. The plastic

strain increment vectors are normal to the plastic potential. For Tresca model, the yield

function, , and the plastic potential, , appear to be coincidently associated (i.e. and are

same) with each other which termed as “associated” flow.

This Tresca model involves three parameters, namely the two elastic parameters from

Hooke’s law (E, ) and one parameter for the yield criterion ( ). This simple constitutive

model with minimum design parameters needed is favored among geotechnical engineers.

However, the Tresca yield criterion only considers single-shear stress effect and ignores the

strength difference of materials in tension, compression, normal stress, intermediate principal

stress, intermediate principal shear stress, hydrostatic stress and twin-shear stresses (Yu et al.,

2009).

4.3.2 Isotropic Model

Isotropic model is one of the elastic-plastic models where the elastic region behaves similarly

as the elastic-perfectly plastic model. In contrast to an elastic perfectly-plastic model, the

yield surface size for isotropic model changes uniformly in all directions such that the yield

stress can increase/decrease in all stress directions when plastic straining occurs. Any

combination of stresses inside the initial yield surface is categorized in the elastic region.

Once beyond the initial yield surface, plastic deformation will be encountered. In isotropic

hardening, the center of the yield surface remains fix but the size of the surface increases.

Any stress states inside the new yield surface will experience elastic deformation; new

deformation occurs when the stress state reaches a new surface. The von Mises yield surface

is used to define isotropic yielding.

99
Yield function

()
The yield function, is given as ( ̃ ) and the von Mises equivalent stress:

()
q= 3 = ( ̃ ) (4.9)

() ()
where ( ̃ ) is the yield stress defined as a function of equivalent plastic strain ( ̃ ).

In Abaqus, the yield stress, , can be defined by the user as a tabled function of equivalent

()
plastic strain, ̃ . The yield stress at a given state is simply interpolated from the input data

and it remains constant for plastic strains exceeding the last value given in the data. The

equivalent plastic strain ̃ in Abaqus is defined as:

̃ = ( ) + ( ) + ( ) +2 ( ) + ( ) + ( ) (4.10)

where , , , , and are the plastic strain components. Under a condition of

uniaxial loading in 2-direction, the above expression can be reduced to ε = (the plastic

strain in 2-direction) (Tjahyono, 2011). Associated plastic flow is applied in this model such

that when the material yields, the inelastic deformation is in the normal direction to the yield

surface. The parameters involve in isotropic hardening/softening model in Abaqus are E, ,

and in a tabled function of .

4.3.3 Concrete Damage Plastic Model

Concrete damage plasticity (CDP) model has been widely used to simulate the crack

development and strength degradation due to damage for concrete (Lee and Fenves, 1998;

Rusinowski, 2005; Josephine, 2011; Wahalathantri et al., 2011; Kmiecik and Kaminski, 2011).

Therefore, it is believed that this model can work reasonably well for cement-treated soil

which has been categorized to be a quasi-brittle material in Chapters 2 and 3. The elastic

region follows Hooke’s laws similar to the models that discussed earlier. CDP model assumes

100
that the two main failure mechanisms are tensile cracking and compressive crushing. The

uniaxial stress-strain relation can be employed into stress-equivalent plastic strain curve and

this is generated automatically from the user-input inelastic strain data corresponding to the

relevant stress. The effective tensile and compressive stresses are then computed to determine

the current state of the yield surface to analyze multiaxial load cases. A modified Drucker-

Prager yield criterion is employed in this model to determine failure by both normal and shear

stress. This model was first developed by Lubliner et al. (1989) and further modified by Lee

and Fenves (1998).

Fracture model

The main criterion in fracture description is the crack recognition. There are two ways of

modeling crack in finite element analysis as shown in Figure 4.4. The discrete approach

models cracking as a separation of elements which requires a very fine mesh for accuracy of

prediction. In other words, the model will demand large number of computations and hence

may not be effective when solving large scale problem. Instead, smeared crack approach

shown in Figure 4.4b where the damaged zone coincides with the dimensions of the elements

is preferred. For CDP model in Abaqus, a scalar damage variable for the material properties is

used to model the damaged zone which is believed to be less mesh dependent compared to the

discrete approach.

Material with scalar isotropic damage in CDP model is defined as:

= ( ) (4.13)

where is Cauchy stress tensor, is the strain tensor, and is the degraded elasticity

matrix. The degraded elasticity can also be written as:

= (1 ) (4.14)

101
where is the initial (undamaged) elasticity matrix, and d is the scalar stiffness degradation

variable. The scalar degradation variable is a function of effective stress tensor, , and

equivalent plastic strains, ̃ , as shown in Eq. (4.15). In CDP model, the stiffness is initially

isotropic and the degradation is characterized by two damage variables, and where

subscripts t and c refer to tension and compression, respectively. The damage variable can

take values from zero, meaning no damage, to one, which represents total loss of strength.

= ( , ̃ ) (4.15)

The “effective” stress tensor discussed in CDP model is defined as:

= . (4.16)

Finally, the Cauchy stress tensor can be related to the effective stress tensor through the

scalar degradation parameter (1 – d):

= (1 ) ; (4.17)

when there is no damage (d = 0), the effective stress tensor, is equal to the Cauchy stress

tensor, .

Yield function

In CDP model, the failure mechanisms: cracking (tension) and crushing (compression) are

represented by increasing the value of damage variables. These variables control the

development of the yield surface and the degradation of the elastic stiffness. The yield

function represents a surface in effective stress space which determines the state of damage. A

modified Drucker-Prager yield criterion is incorporated into the CDP model. Drucker-Prager

model is a pressure dependent criterion formed by two stress invariants of the effective stress

tensor; mean stress, , and the von Mises equivalent stress, .

102
= /3 (4.18)

= 3 (4.19)

According to the yield condition proposed by Lubliner et al. (1989) and further modified by

Lee and Fenves (1998), the failure surface shape in the deviator cross section is not in a circle

but governed by parameter, . Owing to the modification, the model can now account for the

different evolution of strength under tension and compression by the damage/softening

variables, ε and ε , which are referred as tensile and compressive equivalent plastic strains

respectively. The yield function, is given as:

= 3 + ̃ ̃ =0 (4.20)

with

( / )
= ; 0 0.5, (4.21)
( / )

( )
̃ = (1 ) (1 + ), and (4.22)
( )

( )
= (4.23)

where is the algebraically maximum eigenvalue of . The Macauley bracket . is

defined by = (| | + ). / is the ratio of the compressive strength in the biaxial

state to the compressive strength in the uniaxial state. Experimental results reported by

Kupfer et al. (1969) suggested a value of 1.16 for concrete material which is also the default

value in Abaqus. On the other hand, ( ̃ ) and ( ̃ ) are the notations of effective tensile

and compressive stress, respectively.

103
appears only in triaxial compression (Lubliner et al., 1989), that is, in stress state with

< 0 . Let TM and CM represent the “tensile meridian” ( > = ) and the

“compressive meridian” ( = > ) on the yield surface, respectively.

( )
= is defined at a given state . (4.24)
( )

is the ratio of the second stress invariant on the tensile meridian to that on the compressive

meridian at initial yield for any given value of the pressure invariant . Values range from

0.64 to 0.8 was suggested for (Lubliner et al., 1989). The Abaqus default value for is

2/3. For comparison, the difference of the yield surfaces in the deviatoric plane is shown in

Figure 4.5.

The yield surface in plane stress is shown in Figure 4.6. The enclosed area of the yield

surface represents the elastic region. Quadrant I in the coordinate system represents the

material subjected to tensile stress in both the and directions. Meanwhile, if the material

is loaded in compression in both directions, the stress state falls within the quadrant III. The

stress states will be either in quadrant II or IV for load cases where a combination of tensile

and compressive forces is applied (e.g. shear).

Flow rule

Another main ingredient for plasticity is the flow potential function which is used to describe

the plastic strain increments. The CDP model assumes non-associated potential plastic flow.

The flow potential, , used for this model is the Drucker-Prager hyperbolic function:

= ( tan ) + tan . (4.25)

Here, is the dilation angle measured in the plane at high confining pressure; is the

uniaxial tensile stress at failure, taken from the user-specified tension stiffening data; and is

104
a parameter, referred to as the eccentricity that expresses the rate of approach of the plastic

potential hyperbola to its asymptote. In the CDP model, the plastic potential surface in the

meridional plane is assumed in the form of a hyperbola. The shape is adjusted through the

parameter . Parameter can be determined as a ratio of tensile strength to compressive

strength (Jankowiak et al., 2005). The CDP model suggests to assume = 0.1. When = 0,

the surface in the meridional plane becomes a straight line similar to the classiccal Drucker-

Prager hypothesis.

Based on the experimental curves for both uniaxial tension and compression, it is

possible to define the relationships between stress–cracking (damage) strain ( ̃ ) in uniaxial

tension and stress–crushing strain ( ̃ ) in the uniaxial compression. These are better

illustrated through Figure 4.7 and Figure 4.8 in which the values in CDP model are

interpreted as the cracking strain ( ̃ ) and the crushing strain ( ̃ ).

4.4 Evaluation of Constitutive Model Prediction for Behavior of

Cement-treated Toyoura Sand

The behavior of cement-treated soil predicted by three aforementioned constitutive models,

namely Tresca model, isotropic model, and CDP model are evaluated in this section. Cement-

treated soil is known as a quasi-brittle material which has distinct strength asymmetry,

meaning that the uniaxial tensile and compressive behaviors are different as discussed in

previous chapters. Despite many laboratory tests conducted, many cases have little or no data

on tensile strength. The available published laboratory data by Namikawa (2006) and Koseki

et al. (2005) for cement-treated Toyoura sand are adopted here to evaluate the model

predictions.

105
Cement-treated Toyoura Sand

As details are reported by the testers (Namikawa, 2006; Koseki et al., 2005), only a brief

description is presented here. Three laboratory tests, i.e. triaxial compression, uniaxial tension,

three-point bending test on notched beam are modeled in this section. The specimen was

modeled as a single element for the first and second examples. Meanwhile, the third test was

simulated as a boundary value problem. The composition of the cement-treated sand used in

the tests is presented in Table 4.1. The admixture was poured and compacted into the

specified mold after mixing thoroughly. The curing time was 7 days and the desired

unconfined compressive strength was approximately 1800 kPa.

Table 4.1: Mixing proportions of cement-treated Toyoura sand.

Material Toyoura sand Portland cement Bentonite Water


Weight (%) 66.3 10 5 18.7

4.4.1 Drained Triaxial Compression Test by Namikawa (2006)

As mentioned earlier there is little tension tests for cement-treated soil, and compression test

result is often the only data available. Typically compression test is specified by the designer

to validate that the achieved strength of the improved soil exceeds the design value. In other

words, it serves as one of the quality control criteria to verify the effectiveness of ground

improvement in the field. Therefore, a representative model should at least be able to simulate

the treated soil behavior revealed in the compression test.

4.4.1.1 Test Description

A prismatic specimen was trimmed to 160 mm in height with a cross section of 60 mm x 80

mm and set up in the pressure cell. After the specimen was saturated, it was isotropically

consolidated to the predetermined confining stress of 49 kPa. Subsequently under drained

106
condition, compression loading was applied in z-direction at a constant strain rate of

0.01 %/min. The load in z-direction was measured by an inner load cell that is free from

friction at the bearing. In addition to the external displacement transducer attached to the

loading shaft, the displacements of the specimen surfaces in y- and z-directions were

measured by local displacement transducers (LDTs). Owing to possible occurrence of

bending error in loading shaft, the strain measured by LDTs was more accurate and thus

adopted in the present numerical calibration. The relationship between the deviatoric stress

and strain together with the schematic diagram of the test setup are shown in Figure 4.9.

4.4.1.2 Numerical Calibration

A 3-D single element was used to simulate the aforementioned test. The adopted element type

was 8-node linear brick (C3D8R). The boundary condition was modeled by roller where the

bottom plane was restricted from moving verticality. The confining pressure was represented

by distributed loads acting around all the element planes except the bottom one. The loading

test was simulated by applying prescribed displacement from top of the model. The loading

and boundary conditions of the finite element model are shown in Figure 4.10a.

Three constitutive models are compared to evaluate the effectiveness in simulating the

compressive behavior of cement-treated Toyoura sand. At the initial elastic stage, the three

models follow Hooke’s law as quantified by parameters Young’s modulus, E = 2400 MPa and

Poisson’s ratio, = 0.167. Beyond this region, the cemented sand behavior is simulated by

Tresca, isotropic and CDP models separately. For Tresca model, yielding occurs when the

shear stress at any point in the material reaches a given value ( ) in the same plane. Single

parameter, undrained shear strength as shown in Table 4.2 is used to define the yield

surface. With zero friction angle, the is taken as half of the deviatoric stress. For isotropic

model, the yield surface size changes (increase/decrease) uniformly in all directions as plastic

straining occurs. The yield stress is defined as a function of equivalent plastic strain as

107
presented in Table 4.3. CDP model is approximately similar to isotropic model where the

stress-strain curve can be calibrated by defining the compressive stress as a function of

inelastic (or crushing) strain in Table 4.4. Since monotonic loading is simulated herein, the

scalar stiffness degradation variable, d that derived from unloading stage is not considered.

The parameters used to define the yield function in CDP model are shown in Table 4.4.

Table 4.2: Design parameters for Tresca model.

Tension
900 kPa 380 kPa

Table 4.3: Design parameters for Isotropic model.

Yield Stress Equivalent Plastic Strain


1200 kPa 0
1800 kPa 0.0015
1800 kPa 0.0032
1500 kPa 0.0053

Table 4.4: Design parameters for CDP model.

Compression Tension
Compressive
Inelastic Strain Tensile Yield Stress Fracture Energy
Yield Stress
1100 kPa 0 380 kPa 9 N/m
1650 kPa 0.0015
1650 kPa 0.0032
1400 kPa 0.0053

Yielding parameters: = 1o; eccentricity = 0.1; b0/ c0 = 1.16; Kc = 0.67. Identifications of


these parameters have been discussed in detail in Section 4.3.3.

108
In the calculation, isotropic loading of 49 kPa was applied. Then, axial compression loading

simulated by prescribed displacement up to 0.7% strain was applied. The axial and lateral

strains at this stage were recorded and plotted against the laboratory measurements.

4.4.1.3 Results and Discussions

The calibration results for three constitutive models are shown in Figure 4.11. As compared

with the laboratory tests result, the elastic-perfectly plastic Tresca model is not able to capture

the material compressive stress-strain behavior after the specimen experiences 0.05% axial

strain. Owing to a single stiffness and perfectly plastic criteria of this model, the hardening as

well as the softening behavior after the specimen undergoes 0.05% axial strain cannot be

traced using this simple model. Tresca model in this case overestimates the strength or in

other words, underestimates the induced strain. As discussed in Chapter 2, some researchers

recommended to apply the strength as residual strength or to apply a reduction factor to it.

Although this precaution can avoid the overestimation and lead to safe simulation, it is too

drastic to apply making the analysis too conservative. Consequently, this simplified and crude

model might defeat the purpose of adopting a constitutive model in finite element to model

the real behavior of the material. Nevertheless, isotropic and CDP models as shown in Figure

4.11 are capable of capturing the material nonlinear behavior and the calibrated results agree

closely with laboratory result.

4.4.2 Direct Tension Test by Koseki et al. (2005)

A good representative constitutive model shall be at least able to capture both compressive

and tensile behavior of the cement-treated soil. As highlighted in Chapter 2 the study of

tensile behavior is very limited but yet important; the evaluation from the tensile loading

point of view in this section is needed to overcome this gap. The direct tension test is briefly

described in the following.

109
4.4.2.1 Test Description

The specimen was cylindrical with a diameter of 50 mm and 140 mm high. In order to ensure

the tensile failure occurred at the center of specimen, the specimen at mid-height was trimmed

to a smaller diameter as shown in Figure 4.12. The vertical strain was measured by two sets of

LDT installed diametrically opposite to each other. The specimen was held in place by the

grab holder on top and bottom. Gypsum was used to plaster the contact faces between the

specimen and the holder. Tensile loading was applied in z-direction at a constant strain rate of

0.005 %/min. The load in z-direction was measured by an inner load cell. An effort to prevent

the bending error of the loading shaft was made by using a universal joint at the connecting

part. The measurement results of the tensile stress and strain together with the schematic

diagram of the test setup are presented in Figure 4.12.

4.4.2.2 Numerical Calibration

The direct tension test was simplified with one quarter model where a single element with

symmetric condition was modeled. The adopted element type was 4-node bilinear

axisymmetric quadrilateral (CAX4R). The axisymmetric condition was modeled by

constraining the tangential degree of freedom on the bottom and left planes from moving. The

loading test was simulated by applying prescribed displacement from top of the model. The

finite element model with loading and boundary conditions are shown in Figure 4.10b.

Likewise, three constitutive models are used to calibrate this direct tension test. The

parameters used to describe the yield criterion of these three constitutive models are presented

in Tables 4.2 to 4.4.

For a classical Tresca model, the failure criterion as shown in Figure 4.2 only considers

single-shear stress effect and ignores the effect of strength difference in tension and

compression (Yu et al., 2009). In other words, the classical Tresca model allows for the

development of any values of tensile stresses shown in Figure 4.13a. This may be adequate to

110
simulate a cohesive soil behavior but surely not appropriate for cement-treated soil which had

manifested tension-softening behavior in previous laboratory tests. Therefore, a tension

truncated Tresca yield criterion is considered in the present study by adopting the tension

cutoff parameter in Table 4.2. This criterion aims to keep the developing tensile stress values

in the material not to exceed the given truncation value. In Abaqus, this can be achieved by

introducing three additional tension cutoff yield surfaces in form of the Rankine type (Figure

4.14). This model assumes the material to have isotropic properties and hence the ultimate

tensile strength is the same in all orientations. The shear and tension yield surfaces intersect in

the tensile domain of the principal space. As a result, the tension yield surfaces “cutoff” the

shear yield surfaces leading to the effect of the material being able to ultimately sustain an

ultimate tensile stresses (Figure 4.13b). Tension truncated Tresca yield criterion is not a new

thing but has been adopted by other researchers such as Antão et al. (2007) to evaluate the

influence of tension cut-off on the stability of anchored concrete soldier-pile walls in clay.

Similar to classical Tresca model, isotropic model allows for the development of any value of

tension stresses. However, it does not incorporate the cutoff function as available for Tresca

model. Hence, the influence of this will be evaluated in the analysis. In CDP model,

distinction between the compressive and tensile behavior simulations is available so the

model can calibrate the material behavior corresponding to the uniaxial laboratory tests. The

post-tensile failure can be invoked by specifying the post-failure stress as a tabular function of

cracked displacement. Alternatively, the fracture energy, Gf can be specified as a material

property where it assumes a linear loss of strength after cracking as shown in Figure 4.15.

Although Namikawa (2006) had demonstrated bilinear loss of strength after cracking for

cement-treated Toyoura sand; linear loss using Gf = 9 N/m is adopted in the present

calibration as the complete post-failure data is not available in this test. Further investigation

of the tension-softening relationship of cement-treated Toyoura sand will be discussed later in

the calibration of three-point bending notched beam test.

111
4.4.2.3 Results and Discussions

The calibration results by three constitutive models and laboratory measurement are shown in

Figure 4.16. As shown, the post-peak stress modeled by tension truncated Tresca model

remains at the input tension cutoff strength 380 kPa. However, in reality the material

experiences post-peak tension softening as illustrated in the test result. Although the isotropic

model could simulate the compression test well in earlier section, the isotropic yield criterion

which allows for the development of any value of tensile stress has overestimated the actual

material strength. This is clearly shown in Figure 4.16 where the predicted tensile stress-strain

curve exceeds the experimental peak tensile strength without any limit. Without the tension

cutoff function as built into Tresca model, the isotropic model is not able to evaluate the

tensile behavior of cement-treated soil. Nonetheless, the above mentioned shortcoming of the

two models can be overcome by CDP model. In CDP model, a linear loss of strength after

cracking is simulated herein. This built-in function makes the CDP model superior to the

other two models in simulating the behavior of cement-treated soil, as illustrated in Figure

4.16.

4.4.3 Three-point Bending Notched Beam Test by Namikawa (2006)

Unlike the embedded cement-treated soil in deep excavation which is constructed in large

cross sectional area but short length, the cement-treated columnar soil used in open

excavation very often is long and subjected to unbalanced lateral pressures. As such, on top of

calibrating the two basis tests (compression and tension), it is of importance to test the ability

of the model in simulating the behavior of cement-treated soil under bending. Since it has

been shown in earlier calibration tests that elastic-perfectly plastic Tresca model and isotropic

model are inappropriate in simulating the cement-treated soil, only CDP model is further

evaluated herein.

112
4.4.3.1 Test Description

Three-point bending notched beam test was conducted by using beam specimen with a central

notch as shown in Figure 4.17a. The specimen was prepared in a mold with 23 cm in height

and an inner cross section of 6 cm x 8 cm. Upon reaching 7 days curing time, the specimen

was trimmed into the desired size and notched with saw. The notch width of 2.5 mm was

made equal to half the beam depth. A photograph of the specimen under loading is shown in

Figure 4.17b. Vertical loading rate of 0.01 mm/min was maintained throughout the test. As

shown in Figure 4.17b, the applied load, beam deflection at the center, and the crack mouth

opening displacement were measured during the loading process. The beam deflection at

center was measured by laser-type displacement transducer. The crack mouth opening

displacements were measured at both faces of the beam by local displacement transducers

(LDTs). The applied load was measured from the load cell attached to the loading shaft. The

load-deflection curves of three specimens are shown in Figure 4.18 and the tension fracture

energies calculated from these curves are summarized in Table 4.5. The interpreted tensile

stress and crack displacement relations are shown in Figure 4.19.

Table 4.5: Summary results of three-point bending notched beam test by Namikawa (2006).

Case Fmax (N) Wo (N.m) Gf (N/m)


Bab-1 35 0.005 9.6
Bab-2 33 0.0059 12
Bab-3 37 0.0049 9.3

4.4.3.2 Numerical Calibration

A 2-D plane stress finite element analysis was conducted to calibrate the laboratory results.

Sensitivity studies were carried to evaluate the influences of element types (linear triangular

element, CPS3 and linear quadrilateral elements, CPS4R) and mesh sizes. The analysis results

showed that the linear quadrilateral element is more appropriate and the mesh size converges

at a total of 5,353 elements as shown in Figure 4.21. Figure 4.20 shows two studied mesh

113
sizes. For coarser mesh which consisted of 5,353 elements, the typical mesh length was 1.5

mm and getting smaller to 0.5 mm around the notch tip where zone of interest was located.

For finer mesh, the typical mesh length was 1 mm getting smaller to 0.5 mm around the notch

tip attributed to a total of 9,866 elements. Figure 4.21 shows a good agreement between the

computed load-deflection curves for these two meshes, indicating that this model can avoid

the issue of the mesh-size dependency associated with the strain localization problem. In the

finite element model, the specimen bottom supports a 160 mm span was modeled as rollers to

restrict the vertical movement. The simulation was run in two steps. First, gravity loading was

applied to establish initial condition. Subsequently, the loading test was simulated by

applying prescribed displacement on top center of the specimen.

The CDP model simulates the post-failure tensile behavior in term of post-failure tensile

stress as a function of cracking strain, ̃ . The cracking strain is defined as the total strain

minus the elastic strain corresponding to the undamaged material; that is ̃ = ,

where = / and = the initial undamaged stiffness. This relation is well

illustrated in Figure 4.7. Alternatively, the brittle behavior can be characterized by a stress-

crack displacement response instead of a stress-strain response. The post-tensile failure can be

invoked by specifying the post-failure stress as a tabled function of cracked displacement. Or

else, the fracture energy Gf (linear loss of strength after cracking) can be specified as a

material property. The parameters used to describe the yield criterion of CDP model are

similar to previous tests in Table 4.4 where Gf = 9 N/m is applied (Case T1). Nevertheless,

owing to the availability of the interpreted tension-softening relation data in this case (Figure

4.19), the tensile stress as a tabled function of crack displacement is summarized in Table 4.6

was applied in Case T2 (see also Figure 4.22).

In CDP model, the damage in tension is represented by the tension damage variable, dt, which

takes values from zero, representing the undamaged material, to one, which represents a total

114
loss of strength. Owing to absence of unloading data, damage variable, cannot be defined

precisely; thus in this case it is simplified by assuming dt = 0 for maximum tensile strength

and dt = 0.9 for 10% of tensile strength.

Table 4.6: Tensile design parameters for CDP model in Case T2.

Tensile Yield Stress Crack Tensile Yield Stress Crack


(kPa) Displacement (mm) (kPa) Displacement (mm)
380 0 84 0.0296
320 0.0040 75 0.0421
245 0.0083 67 0.0545
185 0.0131 44 0.0875
144 0.0176 24 0.1168
105 0.0216 15 0.1296
88 0.0239 1 0.16

4.4.3.3 Results and Discussions

Figure 4.23 shows the calibration results of three-point bending notched beam test for

cement-treated Toyoura sand using CDP model with Gf (Case T1) and tabular function of

stress-crack opening (Case T2) methods. The load-deflection curve simulated by fracture

energy, Gf, in Case T1 shows an abrupt decrease in load at post-failure zone and the curve

ends with a gentle tail. The actual tests showed a steady decrease in load at post-failure zone

with a long and gentle tail end curve. The abrupt decrease in load at post-failure zone in Case

T1 is due to the simulation of linear loss of strength after cracking by the Gf method.

Nevertheless, the calculated total area covered by load-deflection curve of Case T1 is

approximately 0.0049 N.m which is similar to the test value Wo in Table 4.5. This validates

the numerical simulation as well as the hypothetical consideration to take the beam self-

weight effect in calculating the fracture energy, Gf in section 3.6.2. As shown in Figure 4.23,

Case T2 works better in simulating the present three-point bending notched beam tests. This

is because the post-failure stress is specified as a tabular function of crack displacement from

115
the test results and hence eliminates any assumption made enabling to model the behavior of

the material closely. Figure 4.24 presents the damage path shown in the finite element

analysis for Case T2.

4.5 Evaluation of Constitutive Model Prediction for Behavior of

Cement-treated Singapore Marine Clay

Although cement-treated Singapore marine clay exhibits similar behavior as cement-treated

Toyoura sand to a considerable degree, it is necessary to ensure that the selected constitutive

model can also simulate the behavior of cement-treated Singapore marine clay well. In view

of the findings reported earlier, only the CDP model is selected for further evaluation. The

descriptions of the relevant tests have been presented in Chapter 3. Two laboratory tests, i.e.

uniaxial compression test (UCT) and three-point bending test on notched beam (BNT) are

simulated in this section. The specimen was modeled as a single element for the first example

while the second example was simulated as a boundary value problem. The composition of

the cement-treated clay used in the tests is presented in Table 4.7. Of the few mix proportions

in Chapter 3, cement content Aw = 30% at 14 days of curing period was selected for the

numerical calibration herein.

Table 4.7: Mixing proportions of cement-treated Singapore marine clay.

Mix
Soil-cement Water-cement Cement content, Total water
Proportions
ratio (s:c) ratio (w:c) Aw (%) content (%)
(s:c:w)
5 : 1.5 : 6.5 5 : 1.5 6.5 : 1.5 30 100

116
4.5.1 Uniaxial Compression Test (UCT)

4.5.1.1 Test Description

The test description for UCT was explained in Section 3.4.1. The specimen size was 50 mm

in diameter with 100 mm high. The schematic test and stress-strain curve result are presented

in Figure 4.25.

4.5.1.2 Numerical Calibration

A 3-D single element of 1 m x 1 m x 1m dimension was used to simulate the aforementioned

test. The selected element type was 8-node linear brick (C3D8R). The bottom plane was

modeled by roller to restrict from moving vertically. The test was simulated by applying

prescribed displacement from top of the element. The loading and boundary conditions of the

model are shown in Figure 4.26.

In the initial elastic state, the material was modeled according to Hooke’s law governed by

parameters Young’s modulus, E and Poisson’s ratio, . A value of E = 170 MPa,

approximately 212qu was selected. Poisson’s ratio, = 0.2 was adopted as recommended in

literature reviews in Chapter 2. Beyond this region, the stress-strain behavior of cement-

treated clay is simulated by CDP model. In CDP, the compressive stress can be defined as a

tabular function of inelastic (or crushing) strain in Table 4.8. The stress-strain curve can be

defined in hardening state or even beyond ultimate stress, into the strain-softening regime.

Since monotonic loading is simulated herein and the damage appearance in the model is not

considered, the scalar stiffness degradation variable, d is not considered in this case. In the

absence of compressive damage, dc, the model behaves as a plastic model where inelastic

strain ̃ = plastic strain ̃ . The parameters used to define the yield function in CDP model

are shown in Table 4.8.

117
Table 4.8: Calibration parameters for cement-treated Singapore marine clay using CDP model.

Compression Tension
Compressive
Inelastic Strain Tensile Yield Stress Fracture Energy
Yield Stress
500 kPa 0 88 kPa 3.6 N/m
800 kPa 0.002094
800 kPa 0.008294
530 kPa 0.015882
260 kPa 0.031471
150 kPa 0.042118

Yielding parameters: = 1o; eccentricity = 0.1; b0/ c0 = 1.16; Kc = 0.67. Identifications of


these parameters have been discussed in detail in Section 4.3.3.

In the calculation stage, axial compression loading simulated by prescribed displacement up

to 4.2% strain was applied. The axial stresses and strains in this stage were computed and

plotted against the laboratory measurements in Figure 4.26.

4.5.1.3 Results and Discussions

The stress-strain calibration result for UCT of cemented Singapore marine clay together with

laboratory measurements are shown in Figure 4.26. The numerical results agree very well

with the laboratory test results. In addition to the good simulation of compressive behavior of

cement-treated Toyoura sand, the CDP model is proven to work well in simulating the

compressive behavior of cement-treated Singapore marine clay. These two examples have

shown that CDP model enables a proper definition of the compressive mechanisms (inclusive

of strain hardening and softening effects) for cement-treated soil.

118
4.5.2 Three-point Bending Notched Beam Test

4.5.2.1 Test Description

The test description for three-point bending test on notched beam for cement-treated

Singapore marine clay was presented in Section 3.4.3. The specimen size was 50 mm high, 50

mm wide and 250 mm long. The span between two roller supports was 200 mm. The

schematic test and load-beam deflection curve are presented in Figures 4.27 and 4.28,

respectively.

4.5.2.2 Numerical Calibration

A 2-D plane stress finite element analysis was conducted to calibrate the laboratory results.

Sensitivity studies were conducted to evaluate the influence of mesh sizes. The analysis

results showed that the mesh size converged at a total of 3,358 linear quadrilateral (CPS4R)

elements. The finite element mesh that used to simulate the three-point bending notched beam

Test B1 is presented in Figure 4.29. The typical mesh length was 2.5 mm and made smaller to

0.5 mm around the notch tip where the interested zone was located. In the finite element

model, the bottom supports of the specimen were modeled as rollers to restrict the vertical

movements. The simulation was run in two steps. First, gravity loading was applied to

establish initial condition. Subsequently, the loading test was simulated by applying

prescribed displacement on top center of the specimen.

In previous calibration for cement-treated Toyoura sand (Section 4.4.3) where the tension-

softening behavior was better simulated by specifying the post-failure stress as a tabular

function of cracking displacement instead of fracture energy, Gf. This is due to the

assumption of linear strength loss after crack occurrence in the fracture energy, Gf approach

was too crude to represent the bilinear tension-softening behavior as observed from laboratory

tests. Conversely, the strength loss relation was almost linear as observed for cement-treated

119
Singapore marine clay Test B1. Therefore, it is believed that the fracture energy, Gf approach

can perform the simulation for this test. Design parameters, tensile stress = 0.11qu (correlation

obtained from Section 3.6.1) and Gf = 3.6 N/m were adopted. The parameters used to describe

the yield criterion of CDP model are presented in Table 4.8.

In CDP model, the damages in compression and tension are represented by the compression

damage variable, dc and tension damage variable, dt respectively which takes values from zero,

representing an undamaged material, to one which represents a total loss of strength. The

damage variables used in the present analysis are shown in Table 4.9. Specifying these

damage parameters can help to recognize the damage occurrence in the analysis.

Table 4.9: Damage variables used in CDP model for cemented Singapore marine clay Test B1.

Compression Tension
dc Inelastic Strain dt Crack Displacement
0 0 0 0
0 0.008294 0.96 0.08 mm
0.96 0.042118

4.5.2.3 Results and Discussions

The calibrated load-deflection curve (with E = 212qu from UCT at Section 4.5.1) together

with laboratory result are shown in Figure 4.30. It is noticed that the material stiffness is

underestimated in this case. The stiffness parameter, Young’s modulus was calibrated from

UCT in Section 4.5.1 which was based on the external strain measurement. However, it has

been studied by several researchers that the external strain measurement usually overestimates

the measured strain due to bending error at the contact cap. Therefore, a representative

induced strain must be measured by local transducers attached to the specimen. In view of

this, two higher values for stiffness, E = 400qu and E = 800qu by referring to Tan et al. (2002)

were adopted and the analyses results are also plotted in Figure 4.30 for comparison. The

120
calibration by using Young’s modulus E = 800qu yields reasonably good agreement with the

laboratory measurement. The post-test damage in the model can be identified by the tension

damage variable indication as presented in Figure 4.31. The predicted damage position and

pattern coincide with the laboratory observations where the crack propagates from the edge of

notch throughout the specimen.

Unlike the calibration for cement-treated Toyoura sand in Section 4.4.3, the linear loss

strength after cracking using parameter Gf in this case simulate the three-point bending test on

notched beam well. This is because the tension-softening relation as observed in Figure 3.30

for Test B1 is approximately linear, and can be represented by the method using parameter Gf

in CDP model.

4.6 Conclusions

Numerical analysis is a common tool for the analysis of geotechnical problems today. The

qualitative description of material behavior is formed by the representative constitutive model,

whereas the behavior is further quantified by appropriate model parameters. Therefore, the

constitutive model is a key to solve a particular problem correctly as it represents the material

response to any loading conditions in the analysis. As a result, it is a basic requirement that

the adopted constitutive model should be the most appropriate model to simulate the material

for the problem to be analyzed.

In this chapter, three types of constitutive models are reviewed and evaluated. These three

models follow Hooke’s law for elastic region but the behavior differs accordingly beyond the

elastic region. One of them is elastic-perfectly plastic Tresca model which is typically used

by practitioners in simulating cement-treated soil. Owing to its extensive use, the accuracy

presented by this model is examined. This model is favored due to its simplicity with a single-

shear stress consideration in yield criterion. The second constitutive model is isotropic model,

121
one of the elastic-plastic models. In contrast to the Tresca model, the yield surface size for

this model changes (increase/decrease) uniformly in all directions when plastic strain occurs.

The third model is concrete damage plasticity (CDP) which is often used to simulate crack

development and strength degradation due to damage for brittle material such as concrete.

The CDP model assumes that two main failure mechanisms are tensile cracking and

compressive crushing. A modified Drucker-Prager yield criterion is incorporated in this

model to determine failure both by normal and shear stress. The damage degradation is

characterized by the damage variable which take values from zero, meaning no damage, to

unity which represents total damage loss of strength.

A total of five laboratory tests were employed to evaluate the predictions by these three

constitutive models. Three of them are on cement-treated Toyoura sand with reference to the

published data by Namikawa (2006) and Koseki et al. (2005). Remaining two are on cement-

treated marine clay refer to the tests reported in Chapter 3. The three constitutive models were

first calibrated against drained triaxial compression test and direct tension test for cement-

treated Toyoura sand. For drained triaxial compression test simulation, all of them predict

results close to the laboratory measurement except for the Tresca model. The analysis results

show that owing to a single stiffness and perfectly-plastic criteria assumed, the strain

hardening and softening behavior of the material cannot be captured by the Tresca model. In

direct tension test calibration, isotropic model overestimates the strength grossly as the yield

criterion allows for the development of any value of tensile stress without limits. Some degree

of improvement can be achieved using a built-in feature named as tension cutoff. The post-

peak stress simulated by this tension truncated Tresca model is kept at the input strength value

but it is still incapable of modeling the tension-softening behavior of cement-treated soil.

Hence, it can be concluded that Tresca model and isotropic model are inappropriate in

simulating the behavior of cemented soil. Nevertheless, these shortcomings can be overcome

by using CDP model which yields a close prediction to laboratory measurement in both

compression and tension.

122
To further evaluate the use of CDP model in simulating cement-treated soil behavior,

calibration of three-point bending on notched beam test was conducted. Two methods in CDP

model, namely Gf (Case T1) and tabled function of stress-crack displacement (Case T2) to

simulate the tension-softening relation of cemented Toyoura sand were conducted. The load-

deflection curve simulated by the latter method (Case T2) yields a better agreement with

laboratory measurement. Assumption of linear strength loss after cracking in Gf approach

(Case T2) is too simple to simulate the bilinear tension-softening behavior revealed in the

Namikawa (2006)’s study. Conversely, this simple method is applicable for cement-treated

Singapore marine clay Test B1 reported in Chapter 3, which has a linear tensile stress-crack

displacement relation. The predicted damage propagates from the edge of notch throughout

the specimen and agrees well with laboratory observations. Beside cement-treated Toyoura

sand, CDP model has proven to be an appropriate model in simulating the behavior of

cement-treated marine clay through calibration by UCT and three-point bending notched

beam tests.

123
Chapter 5 Numerical Approaches in Simulating Cemented
Soil Mass

5.1 Introduction

Numerical model with finite element method has been widely applied to analyze geotechnical

problems on cement-treated soil. However, owing to advancement of construction techniques,

relevant analytical procedures for relatively new retaining systems formed by cement-treated

soil columns are yet to be examined. In addition, the accuracy of the numerical analysis is

strongly related to the assumptions and simulations made by the designer. In view of the

above, this chapter attempts to investigate the use of various numerical approaches in

simulating cement-treated soil mass in an open excavation.

Owing to limitations of computational tool in the past, cement-treated soil mass is often

designed and analyzed as a composite material even though they are, in reality, not so. It has

been discussed in Chapters 2 and 3 where the behavior of cement-treated soil is different from

that of untreated soil. However, most of the numerical analyses on cement-treated soil

conducted till to date are two-dimensional employing the weighted average simulation (WAS)

approach. In WAS, a “smeared” uniform strength and stiffness for the treated zone is assumed.

This is crude as it does not take the configuration of cement-treated soil columns into

consideration. WAS is preferred by the practitioners because it requires relatively little effort

to model the complex geometry of cement-treated soil columns. This approach is popularly

used to estimate the distribution of vertical load within an embankment by adopting

recommended coefficients (Vogler and Karstunen, 2009; Adams, 2011). However, the

loading condition experienced by cement-treated soil columns in an open excavation is

significant in the lateral direction as compared to vertical loading for an embankment. In

addition, the overlapping columns formed behind an excavation behave as a cantilever wall

unlike the confined condition in embankment cases. Therefore, the assumption made in WAS

138
considering perfect bonding between cement-treated soil columns and untreated soils forming

a composite material must be carefully examined.

This chapter aims to conduct numerical analysis of cement-treated soil columns used in an

open excavation and seek to shed light on its fundamental behavior. Firstly, the commonly

adopted WAS approach is evaluated against the shear box test results by Larsson (1999). In

addition, the analyses and laboratory test results are compared with the results using real

allocation simulation (RAS) approach where each individual treated soil column is modeled

individually with assigned properties. In the RAS approach, interfaces to allow for relative

movement between treated soil columns and soil are considered. Through this back-analysis,

the understanding of modeling criteria for cement-treated soil columns is established.

Thereafter, hypothetical cases of a vertical cut with different ground improvement patterns are

conducted to further examine the behavior of cement-treated soil mass. Through this study,

the mechanisms that were overlooked by WAS approach are highlighted and the failure

mechanism for cement-treated soil columns with different ground improvement geometrical

patterns are discussed.

5.2 Weighted Average Simulation (WAS) and Real Allocation

Simulation (RAS) Approaches

Cement-treated soil is a product that generally has a higher strength and lower compressibility

compared to the native soil. The construction site can be improved by installing overlapping

columnar cement-treated soil. Sometimes, the columns are arranged in certain improvement

pattern which makes the evaluation of these columns in the treated mass difficult in 1- or 2-D

analyses. As a result, the treated mass is often modeled as a composite material using WAS

(SGF, 2000), alternatively known as homogenization method, is used to evaluate the mass

properties in existing design practice when not adopting 3-D analyses. This can create

139
significant saving in computer storage and computation time because finer mesh is not

required for modeling the actual position and dimension of both treated and untreated soil.

WAS approach is applied to treated ground with customized configuration of columnar

treated soil for obtaining an average strength and stiffness of the treated ground in numerical

analysis. This approach assumes complete interaction between the treated soil column and

surrounding untreated soil (Broms and Boman, 1979). No localized failure between the

treated soil column and untreated soil is considered. The average property of the treated mass

is defined by Eq. (2.1). It is known that the composite strength and stiffness of the improved

area is highly empirical. Some researchers such as Ou et al. (1996) and Hsieh et al. (2003)

imposed an empirical factor when adopting WAS approach in their back-analyzed studies.

Nevertheless, these empirical factors were arbitrarily adjusted in such a way that the predicted

retaining wall behavior would be close to the field measurement or predictions from RAS

approach without further addressing the underlying mechanism involved.

In the real allocation simulation (RAS) approach (Ou et al., 1996; Yang, 2009), 3-D analysis

is required and the treated material properties are assigned to the mesh elements

corresponding to the treated soil area. The native soil material properties are specified

separately from the treated soil area. This allows different constitutive models to be applied to

simulate the behavior of these two materials separately. Thus, the strength mobilization of the

treated soil columns and untreated soil can be closely captured in the analysis. In addition, the

configuration of columnar treated soil as well as the interaction between treated soil and

untreated soil can be addressed clearly in the 3-D numerical analysis. In other words, realistic

assumption is made in the real allocation simulation approach. However, this method requires

finer complex meshes and thus a higher performance computer and longer computational time

to facilitate the calculation.

140
5.3 Evaluation of WAS and RAS Approaches in Simulating

Cement-treated Soil Mass

The case-reference study is based on lime-cement columns in shear box test performed by

Larsson (1999). Through back-analysis of this experiment, the groundwork of modeling

criteria for cement-treated soil mass can be established.

5.3.1 Test description

The experimental material consisted of lime-cement-treated soil columns embedded in kaolin

clay in a shear box. The objective of these tests was to investigate the shear resistance of the

treated soil columns with respect to a horizontal slip surface passing through the soft kaolin

clay. Two shear tests were considered in the present study. At first, a shear test (Test 1) was

performed to obtain the parameters of untreated kaolin clay. Then another shear test was

conducted on twelve treated single columns embedded in the clay (known as Test 2 herein) to

evaluate the competence of WAS and RAS approaches. The experimental procedures are

briefly described below and details are given by Larsson and his co-workers (Larsson, 1999;

Charbit, 2009; Larsson et al., 2012).

The shear box dimension was 0.6 m in height and 0.5 m in diameter. Kaolin clay was

prepared with 50 mm thick sand layer on top and below serving as drainage during

consolidation stage. The installation of lime-cement-treated soil columns was by dry deep

mixing where binder in a dry powder form was mixed with the soft clay. The position of

columns in shear box is indicated in Figure 5.1.

141
The shear box was fabricated in such a way that the upper half of the box was free to move

laterally while the lower part was fixed. The shear box test setup is shown schematically in

Figure 5.1. The test was performed by applying a horizontal traction force in increments and

the induced horizontal deformations were recorded. A surface normal pressure of 10 kPa was

applied to the untreated clay in Test 1 and 15 kPa was applied on lime-cement-treated soil

columns in Test 2. The measured mobilized shear stresses against lateral deformations for

these two tests are presented in Figure 5.2a. After Test 2, the lime-cement columns were

examined to observe the damage pattern as shown in Figure 5.2b. Samples collected from the

columns were subjected to unconfined compression tests (UCTs). The UCT results for these

samples are presented in Figure 5.3.

5.3.2 Numerical Analyses

Model Set-up

3-D FEM analyses using finite element program, Abaqus/Standard, were adopted for the

back-analysis of this case study. Eight-noded brick elements, with reduced integration

(C3D8R) were adopted. The influence of element size was investigated for each model and

the optimal mesh size was determined by trial convergence analysis.

Following the tests, a circular shear box was simulated with a single symmetry boundary

condition to optimize the computational effort. As such, the tangential degree of freedom on

the flat vertical face of the half circle model was restricted, but the model is free to move in

other orthogonal directions. Boundary conditions were applied to the circumferential area of

the shear box. The shear box according to the laboratory setup was modeled in two parts. The

lower part was fixed while the upper part could move laterally during shearing. A partition

between the lower and upper parts was allocated to facilitate these differences in boundary

condition. Similar approach of creating partition in the middle of the model was applied by

142
Ziaie Moayed et al. (2012) when simulating direct shear test. 5 mm and 10 mm thick

partitions were examined to evaluate the influence of partition thickness on the shear test

performance. A comparison of analysis result with calculated shear stresses and strains

showed little difference between these two thicknesses. Therefore, a 5 mm thick partition was

used to distinguish the upper and lower halves of the model in subsequent analyses.

a) Test 1

The finite element mesh for Test 1 is shown in Figure 5.4. The analysis result converged

when the element size of 20 mm was used with a total of 9,484 elements for the whole model.

b) Test 2

Test 2 involved twelve single lime-cement-treated soil columns embedded in kaolin clay and

the results were back-analyzed using WAS and RAS approaches. Two layouts were

considered in the WAS approach. First, average properties were applied to the whole model

named as the whole mass (WM) model. As the geometry is the same as Test 1, the same finite

element mesh model (Figure 5.4) for Test 1 was adopted herein. To further improve the

model, a refined strip to localize the column positions was provided and this model was then

known as the refined mass (RM) model. The analysis result converged when the element size

was 15 mm, with a total of 23,020 elements for the whole model as shown in Figure 5.5. For

RAS approach, the treated soil columns were modeled with their positions and dimensions as

in the laboratory test. The analysis result converged when typical 20 mm element sizes were

applied to the model while finer meshes 10 mm were assigned to the treated soil columns. A

total of 45,338 elements was adopted to build the whole model as presented in Figure 5.6.

143
Material Models

There are three types of material involved in these two tests, i.e. kaolin clay, lime-cement-

treated soil columns and sand. Nevertheless, as only UCT data on lime-cement-treated soil

column was available, the design parameters for clay were then calibrated from Test 1. The

clay was simulated using a classical Tresca model with bulk density, = 1500 kg/m3. The

calibrated cohesion for the clay was 4.2 kPa with E = 1 MPa and = 0.45. The top and

bottom sand layers played no significant effect to the overall test performance; thus linear

elastic behavior with E = 100 MPa and = 0.3 was assumed. The bulk density, was taken

as 1800 kg/m3.

For lime-cement-treated soil columns, two types of constitutive models, i.e. classical Tresca

model and concrete damage plasticity model, were used for comparison purpose. The design

parameters were adopted from the UCT data and presented in Figure 5.7.

i) Classical Tresca model

Owing to the adoption of average property between lime-cement-treated soil columns and

clay in the WAS approach, a simple model classical Tresca was applied. Referring to the

existing design practice as discussed earlier in Chapter 2, cohesive strength is taken as half of

the unconfined compressive strength, = qu/2 which is equal to 60 kPa in the present case.

Considering the different area replacement ratios in WM and RM as illustrated in Figure 5.8,

the average properties based on Eq. (2.1) for these materials are tabulated in Table 5.1.

144
Table 5.1: Tresca design parameters for WAS approach in Test 2.

WAS
Material Case Cu (kPa) E (MPa)
configuration (kg/m3)
Clay - - - 1500 4.20 1 0.45
Treated
- - - 1500 60 20 0.15
columns
WM i 0.12 1500 10.90 3.28 0.45
Composite
RM ii 0.32 1500 22 7.08 0.45
Mass
RM ii 0.197 1500 15.19 4.74 0.45

ii) Concrete damage plasticity model

In elastic region, the behavior of lime-cement-treated soil columns were simulated using

Hooke’s law, similar to WAS approach with E = 20 MPa and = 0.15. The input parameters

for compressive behavior were calibrated from the UCT results as shown in Figure 5.7 and

Table 5.2. Compressive damage was defined by the damage variable, dc which takes values

from zero representing undamaged material, to one which represents a total loss of strength.

Little degradation of the stiffness due to compressive stresses in strain softening zone is

assumed in this model. However, this function is crucial for the residual strain in unloading

condition but has little effect on the monotonic loading except for damage indication. The

damage variables were assumed and defined as a tabular function of inelastic strain presented

in Table 5.3. Sensitivity study conducted to investigate the effect of damage variables ranging

from 0.3 to 0.7 revealed no significant difference. As such, dc = 0.5 was applied

corresponding to the residual strength of 70 kPa.

Owing to the absence of tension tests, the tensile strength was assumed as 0.1qu which was

equal to 12 kPa in this case. This was a reasonable assumption as most of the tensile strength

fell in this range as discussed in Chapter 2. The post-failure behavior was referred to

cemented Singapore marine clay where fracture energy, Gf = 5 N/m was adopted to simulate

the linear loss strength after damage had occurred. Although Gf = 5 N/m is referred to the

cement-treated Singapore marine clay (with higher cement content) whose strength was

145
higher than lime-cement-treated soil columns in this test, the brittleness for lime-cement-

treated soil columns was expected to be less so that they may have chance to reach Gf = 5

N/m according to the condition discussed in Chapter 3. In addition, this was a conservative

assumption compared to Larsson et al. (2012). The tensile damage variables are given in

Table 5.3. Typical parameters used to describe the yield criterion of CDP model were similar

to Charbit (2009).

Table 5.2: CDP design parameters of lime-cement-treated soil columns for RAS approach in
Test 2.

Compression Tension
Compressive
Inelastic Strain Tensile Yield Stress Fracture Energy
Yield Stress
60 kPa 0 12 kPa 5 N/m
120 kPa 0.005
120 kPa 0.013
70 kPa 0.0265

Plasticity parameters: = 1o; eccentricity = 0.1; b0/ c0 = 1.16; Kc = 0.67. Identifications of


these parameters have been discussed in detail in Section 4.3.3.

Table 5.3: CDP damage variables for lime-cement-treated soil columns in Test 2.

Compression Tension
dc Inelastic Strain dt crack displacement
0 0 0 0
0 0.013 0.99 0.83 mm
0.5 0.0265

Contact Properties

In WAS approach, full contact between lime-cement-treated soil columns and clay was

assumed and thus no interface was needed. However, for RAS approach, considering there

might be a relative movement between treated soil columns and soil, the contact surfaces in

146
RAS approach were assigned with relevant properties. The contact surfaces were

distinguished in normal and tangential directions. Contact pressure-overclosure relationship

was applied to normal surface which referred as “hard” contact in Abaqus. Any contact

pressure can be transmitted between the surfaces but the surfaces will separate if the contact

pressure becomes zero (in tension condition). For tangential direction, the frictional behavior

was assumed with classical Coulomb friction. The Coulomb friction model allows the two

contact surfaces to experience limit frictional stress before they start sliding relative to one

another. The coefficient of friction between the contacting surfaces was taken as 0.5.

Numerical Procedures

The numerical procedures for Test 1 and Test 2 were similar; the differences were the

presence of treated soil columns and the magnitude of vertical load applied on top surface of

the model. The simulation of shear test is done in three steps:

1. Gravity load was applied with all the boundary conditions activated to establish the

initial condition of the material prior to loading.

2. Vertical load (10 kPa for Test 1 and 15 kPa for Test 2) was applied on top surface of

the model.

3. A lateral prescribed displacement was applied to the circumferential area of the

model upper part while the lower part remained fixed.

5.3.3 Results and Discussion

Test 1

The calibrated analysis result of mobilized shear stress and lateral deformation without treated

soil columns in Test 1 is presented in Figure 5.9. Owing to a single stiffness and perfectly

147
plastic criteria considered in Tresca model, the best simulation still lacks a little in fully

capturing the clay response. This effect will be evaluated in the next section. The best

simulation was quantified by parameters E = 1 MPa and = 4.2 kPa.

Test 2

Weighted Average Simulation (WAS)

The analysis results of mobilized shear stress and lateral deformation together with yielding

zone appeared in WM and RM models are presented in Figure 5.10a and Figure 5.10b,

respectively. The fully mobilized shear stress obtained in WM model is 10.90 kPa which is

equal to the input shear strength value. In other words, shear failure mode of the overall

system is predicted herein. Meanwhile, RM model also registers similar ultimate shear stress

as WM model but the degree of strain mobilization is different. This is because in RM model,

stronger and stiffer properties were assigned to treated strip area which results in a differential

mobilization between untreated soil and treated strip. The stress mobilization area in WM

model moves uniformly inward to the core of the mass. In RM model, the stress mobilization

area is also moving inward to the core but is not uniform due to higher resistance by the

treated strip. The observed scenario in RM model is more representative for the performance

of a treated mass which consists of treated and untreated soil with difference in strength and

stiffness. Thus difference in strain mobilization is inevitable. Nevertheless, both of these two

models employing weighted average simulation approach overestimate the mobilized shear

stress tremendously, an indication that such simulations had overlooked some key features

and missed the underlying mechanism in the treated system.

148
Real Allocation Simulation (RAS)

For RAS approach, two analyses were conducted where different constitutive models are

employed to describe the behavior of treated soil columns. Similar to WAS approach, Tresca

model is also applied herein for a fair comparison. As discussed in Chapter 4, the CDP model

is superior to Tresca model in modeling the behavior of cement-treated soil, it is thus adopted

herein. The applications of these two constitutive models in back-analyzing a laboratory test

are evaluated.

Figure 5.11 shows the analysis results of the mobilized shear stress and lateral deformation

from WAS and RAS approaches and compared with laboratory measurements. In RAS

approach, each column is simulated instead of a whole mass; thus the difference in strain

mobilization for treated soil columns and untreated soil is properly captured. The analysis

results showed RAS approach recorded a gentle stress-strain curve compared to the laboratory

measurements. This could be due to the lower stiffness selected for kaolin clay in Test 1.

From Figure 5.11, it is found that the simulations by RAS approach yield a closer prediction

to the laboratory measurement. Unlike the full mobilization of whole mass area in WAS,

localized high stress concentration is observed for the treated columns in RAS approach. This

explains why the mobilized shear stresses recorded by RAS approach are much smaller than

WAS approach.

The schematic failure in treated soil columns for Tresca and CDP models using RAS

approach are presented in Figures 5.12 and 5.13, respectively. Although the mobilized shear

stress by Tresca model predicted closer results to the laboratory measurements, a close-up at

the columns failure locations (Figure 5.12) indicates that this model failed to address the

correct failure mode when compared with laboratory results in Figure 5.2b. Conversely, the

tensile damage shown by CDP model in Figure 5.13 coincides well with the plastic hinges

observed in the laboratory test. In CDP model, the tensile damage is represented by the tensile

149
damage variable, dt where values range from zero, representing undamaged, to one, which

represents total loss of strength. The tensile damage first occurred at the back side of rear

column at approximately 60 mm below the slip plane. Subsequent shearing induces

progressive damage at the same location for other columns as well as above the slip plane at

the front side of the columns (Figure 5.13).

Although this is a shear box test by name, which is shearing the material by inducing a slip

plane, bending failure mode is the dominant failure mechanism (as a result of exceeding

tension capacity) observed for the treated soil columns instead of shear failure mode. This is

supported by the study of Namikawa and Koseki (2006) whereby the average fracture energy

for tensile failure, Gf, of the cement-treated soil is approximately 1/15 times of the fracture

energy for shear failure, Gfs. In other words, this material is found to be failed easily in

tension than in shear. On the other hand, a classical Tresca model which allows the

development of any tensile stress values is not able to capture tensile or bending failure.

Therefore, although the mobilized shear stress predicted by Tresca model using RAS is close

to the laboratory results, the predicted failure mechanism is incorrect. A precise prediction of

the damage is essential for engineer to address the overall system mechanism and hence

reinforce the weak area correctly if needed in order to ensure a stable condition.

Through the studies in this section, it is found that even for a shear box test which purely aims

to fail the material in shear, the treated columns would have failed in bending/tension mode,

and not by direct shearing across an imposed slip plane. Therefore, the safety assessment of

cement-treated soil mass employing shear slip analysis of focusing on shear strength in

existing practice is inadequate and unsafe. Consequently, the weighted average simulation

(WAS) approach which averages the properties of the treated mass in terms of shear strength

and stiffness are illustrated to be too crude and naive in capturing the real behavior of treated

soil columns in the shear box test. Conversely, an appropriate constitutive model together

150
with a proper numerical modeling approach which refer to CDP and RAS, respectively are

able to perform a more realistic simulation.

5.4 Study of Cement-treated Ground Improvement Pattern

Unlike a conventional retaining wall, cement-treated soil columns used as earth retaining

system in an open excavation are constructed using gravity wall concept where the designed

gravity mass comprises soil treated with cement. For instance, gravity mass cemented soil

was used to provide temporary as well as permanent support of a river front wall in Columbus,

Georgia, USA in 1991 (Nicholson et al., 1998); and temporary support for a 4.9-m deep

excavation in downtown Lexington, Virgina, USA (Ruffing et al., 2012). Nevertheless,

improving the large ground area with columnar treated soil involves extensive time and cost.

As a result, there is always a desire to optimize the design with the help of innovative

construction method.

Optimization has been considered when cement-treated soil is used as an embedded “strut” on

the passive side of deep excavation. Instead of improving the whole mass of ground, some

(O'Rourke and O'Donnell, 1997; Uchiyama and Kamon, 1998; Tsuzuki et al., 2000) choose to

treat only part of soil forming a buttress or panel type of ground improvement. The reported

field performance appeared to be satisfactory and complied with design requirement.

Although the effectiveness of this partial soil improvement in performing the excavation has

been recorded, the selection of layout pattern and range still lacks guidance. Liao and Tsai

(1993) attempted to examine the passive resistance of soft soil improved by treated soil in

column and buttress patterns. However, the area if improvement ratios are different thus

limiting the reliability of their conclusion drawn from the observed performances of these

patterns.

151
For cement-treated soil columns used as an earth retaining system, Racansky et al. (2008)

modeled 3-D hypothetical cases to evaluate the performance of buttressed jet grouted

retaining wall with different spacings. Mohr Coulomb model with tension cutoff function was

adopted to simulate the jet grouted retaining wall. However, the selected design parameters

for treated soil were very much higher compared to untreated soil thus the incremental shear

strains only occurred at the untreated soil but not at the treated soil. Nicholson et al. (1998)

suggested that a row of overlapping columns can be formed along the intended line of the

proposed excavation. Then additional treated soil columns can be added in a pattern which

will ensure composite action of the mass of soil encompassed by the cement-treated soil

columns. This “composite gravity wall” is also known as “VERT” (Vertically Earth

Reinforced Technology) (Andromalos et al., 2001). It is a proprietary construction method

developed, designed and constructed by Geo-Con, a US company, to support excavations.

This method is still under development at an experimental stage. In their technical report, it is

suggested to construct a capping beam overtop all the columns to tie them together for load

transfer purpose. Steel beam embedded in the front row columns might be needed as

additional support. Nevertheless, Haque and Bryant (2011) observed a substantial soil body

movement for VERT system constructed in Irving-Las Colinas, Texas, USA.

Thus, it can be seen that the ground improvement pattern approach is workable provided that

proper design consideration has been made. The ground improvement pattern is selected

considering the purpose of construction, cost, site condition and stability assurance. Owing to

difficulties in performing extensive quantitative field test involving cost and time on top of

challenges in instruments and monitoring the performance, the following hypothetical case

study adopting the established numerical modeling criteria in earlier sections would provide a

groundwork to establish significant qualitative behavior of cement-treated soil column used as

earth retaining structure in an open excavation in the field.

152
5.4.1 Hypothetical Cases

A 4-m deep vertical open excavation in weak soil is considered as the hypothetical case study

herein. Considering that the cut might not be stable or induces excessive ground

displacements, an optimize construction method of improving the ground by cement-treated

soil columns is carried out. Instead of 100% replacement ratio wherein all the soil in a

particular area is cement-treated, different patterns of columnar treated soil are installed to

achieve the desired performance. This is done by utilizing spaced or overlapping combined

columns (Topolnicki, 2004) taking advantages of soil arching effect (Broms, 2004). The

geometry for treated area in the present study refers to the recommendation used in VERT. In

the technical evaluation report for VERT by HITEC (US High Innovative Technology

Evaluation Center), the suggested width of gravity wall is 60% to 80% of the wall height and

a minimum embedded toe of 1 to 1.2 m. Therefore, in the present study, the treated width for

composite gravity wall is taken as 4 m between the center of front row columns and the center

of rear row columns; and 2 m embedded depth below excavation formation. The diameter of

overlap cement-treated columns is 1.2 m with spacing 1 m center to center. A total of three

improvement patterns are selected to stabilize the vertical cut are shown in Figure 5.14. They

are named as grid type (GD), tangential buttress type (TN), and double-wall type (DW)

ground improvement pattern.

For grid type improvement pattern (see Figure 5.14a), the treated columns are arranged to

form a block which is intended to act as a composite gravity wall. This pattern was adopted

locally with the help of passive soil berm to facilitate a shallow excavation in weak soil. On

the other hand, tangential buttress type improvement pattern is designated to optimize the

mobilization between treated soil columns and soil. The tangential buttress type concrete wall

has been successfully used to facilitate deep excavation as reported by Hwang and Moh

(2008), Chuah and Tan (2010), and Chen et al. (2011). Therefore, the suitability of adopting

this pattern formed by overlapping columns is worthy to examine. It can be formed by

153
overlapping columns along the excavation line (front row) backed up by tangential rows of

overlapping columns in certain intervals (see Figure 5.14b). It is expected that some soil

arching would take place and reduce the full lateral pressure acting on the front row columns

during excavation. Nevertheless, there is a concern that failures may occur due to separation

of the columns (Broms, 2004). This scenario will happen when the tensile resistance of the

cement-treated soil column is exceeded and the overlap area of column is insufficient. In view

of this, a thicker wall named as double-wall type improvement pattern (see Figure 5.14c) is

taken as the third pattern to be examined in the present study. Compared to the successful

retaining system in Lexington, Virgina, USA (Ruffing et al., 2012) with a bigger column

diameter of 2.4 m, typical local practice uses column from 0.85 m to 1.5 m in diameter.

Hence, it is intended to overlap the columns to form a thicker retaining wall. The efficiency of

these three improvement patterns will be discussed in the next section.

The three selected geometries are modeled using RAS approach where each cement-treated

soil column is simulated according to their positions. Meanwhile, typical modeling approach

WAS is also adopted herein to analyze the hypothetical cases. As such, a comparison between

WAS and RAS approaches can be made and the appropriateness of existing practice of WAS

approach used in simulating the excavation problem can be evaluated. Two models for grid

type pattern were considered in the WAS approach as presented in Figure 5.15. First, average

property of treated soil columns and untreated soil were applied to the whole model, termed

as whole mass (WM) model, is analyzed. The total number of installed cement-treated soil

columns for both grid type and double-wall type were the same but the treated ratio might be

slightly less for double-wall type due to column overlap. Treated ratio in the present WM

model was based on grid type and might be slightly higher compared to tangential buttress

type as the total number of installed cement-treated soil columns for the latter was less.

However, it is expected that this minor difference would have little effect on the failure

mechanism to be further discussed in the following section. The second model in the WAS

154
approach adopted the refined mass (RM) where refined strips were provided to localize

assigned treated columns property.

5.4.2 Numerical Analyses

FEM Stability Assessment

Definition of Factor of Safety

Very often, conventional limit equilibrium method presents the ratio of restoring to driving

moments as factor of safety (FOS). In this method, collapse mode is considered where an

assumption on the failing soil slices is made with consequent implication to the overall

equilibrium assessment. Conversely, finite element method aims to solve the defined problem

and presents the current state of ground responses. Therefore, additional effort on the analysis

procedure is needed to define the FOS. In FEM, FOS can be defined in two ways. First, the

original shear strength parameters are divided in order to bring the analyzed problem to the

failure point (Griffiths and Lane, 1999). Some (Matsui and San, 1992; Dawson et al., 1999;

Zheng and Zhao, 2004) named this as strength reduction technique. This way of definition is

similar to conventional limit equilibrium method. To find the FOS, the finite element program

initiates a systematic search for the value of FOS by reducing the soil strength parameters

gradually until a well-defined failure mechanism is fully developed. The second definition on

the FOS used in FEM is named as overloading (Swan and Seo, 1999; Zheng et al., 2006). The

essence of this method is based on monotonically increasing the gravity loading on the soil

until the analyzed problem become unstable and equilibrium solution satisfying internal and

external forces can no longer be obtained (Swan and Seo, 1999). In this method, the FOS is

simply defined as the ratio of maximum gravity acceleration on the problem (where the global

equilibrium solution exists) over the actual gravitational acceleration (g = 9.81 ms-2).

155
Definition of Failure

Definition of failure for an excavation in FEM can be defined by the occurrence of excessive

bulging of the ground profile, limit of the shear stresses on the potential failure surface or

non-convergence of the solution (Griffiths and Lane, 1999). Among the aforementioned

definitions, non-convergence solution approach can be very tricky. This is because the non-

convergence error might be caused by localized element failure instead of a global behavior.

On the other hand, the application of the second definition is limited and only feasible to

Mohr Coulomb model with parameters c and . It is complicated to limit the shear stresses on

the potential failure surface that involved different types of soil model that the strength might

not be described purely by c and . In view of this, defining the failure as the excessive

bulging of the ground profile beyond acceptable limit is the appropriate way in assessing the

overall stability of cement-treated soil columns used in an open excavation. This is a

reasonable assumption as cement-treated soil column is a brittle material, hence excessive

deformation would cause rupture failure of the columns. Nevertheless, this approach cannot

ensure that local column failure can be prevented from occurring.

Model Set-up

A quasi-static 3-D analysis using finite element program, Abaqus/Explicit was adopted to

study the hypothetical case. Explicit time integration scheme was selected because of

substantial yielding that would occur in the materials resulting in numerical convergence

problem (Larsson et al., 2012). A simulation of static processes was achieved by keeping the

loading rate low, and monitoring the kinetic energy to be not exceeding the internal energy by

more than 5%. In view that the geometry of the model consists of sharp curve between

overlapping columns, so first-order tetrahedral elements (C3D4) were used. This type of

element has 4 nodes and 1 integration point. There is always an argument that this element

156
type is stiffer and will underestimate the deformation of the analyzed problem. However, this

limitation can be overcome with very fine meshes.

In order to optimize the computational time, the problem was simplified as a plane strain

condition by considering a 6 m wide from center of tangential row columns to center of

another tangential row columns. The plane strain condition was simulated by restricting the

movement of these two faces in y-direction. The lateral boundary was allowed for four times

the depth of treated soil column, and was restricted from moving in x-direction. The bottom

boundary was chosen to extend for at least two times the depth of treated column from the toe

of column, and fixed from moving in any direction. The typical boundary condition and

geometry for the hypothetical case is presented in Figure 5.16.

A total of six geometry models for twelve analyses were employed to simulate the problem,

as summarized in Table 5.4. The first geometry mode was to simulate a vertical cut in soil

without soil improvement as shown in Figure 5.17. Element sizes were 0.2 m at the vertical

cut face and gradually increased to 1.5 m at the boundary. In general, the number of elements

used in this model was approximately 113,576. Another two geometry models were adopted

for WAS approach simulating the grid type pattern shown in Figures 5.18 and 5.19. Element

size was also kept as 0.2 m within the treated zone and gradually increased to 1.5 m at the

boundary. WM model was built from a total of 111,894 elements while RM model consisted

of 195,939 elements. The remaining three geometry models were used to simulate the grid

type, tangential buttress type and double-wall type ground improvement pattern with RAS

approach as shown in Figures 5.20 to 5.22. Owing to the complex geometry of the columnar

treated soil, the meshes within 5 m of the treated block were refined. The element size for

columns was 0.2 m and gradually increased to 0.6 m within the 5 m extended area and

subsequently reached 1.5 m at boundary. As such, the total elements for GD model, TN

model and DW model were 256,758, 258,434 and 228,696, respectively.

157
Table 5.4: Summary table of hypothetical cases.

Case Improvement Pattern WAS/RAS Soil Model for Treated Mass


Li No - N.A
La Grid type WAS-WM Classical Tresca
Lb Grid type WAS-RM Classical Tresca
L1-1 Grid type RAS Classical Tresca
L1-2 Grid type RAS Tension truncated Tresca
L1-3 Grid type RAS Concrete damage plasticity
L2-1 Tangential buttress type RAS Classical Tresca
L2-2 Tangential buttress type RAS Tension truncated Tresca
L2-3 Tangential buttress type RAS Concrete damage plasticity
L3-1 Double wall type RAS Classical Tresca
L3-2 Double wall type RAS Tension truncated Tresca
L3-3 Double wall type RAS Concrete damage plasticity

Material Models

In these hypothetical cases, the ground material was assumed as homogenous weak soil. The

bulk density ( ) was taken as 1500 kg/m3 while Young’s modulus (E) and Poisson’s ratio ( )

were taken as 20 MPa and 0.3, respectively. The yield criterion for this material was

simulated by Mohr Coulomb model with c’ = 2 kPa and ’ = 22o. On the other hand, cement-

treated soil columns were simulated by two types of constitutive model, i.e. classical Tresca

model and concrete damage plasticity (CDP) model for comparison between common

practice and numerical criteria suggested in the present study. The design parameters for the

cement-treated soil column were those from laboratory results reported in Chapter 3 with

unconfined compressive strength, qu = 980 kPa. In addition to classical Tresca model, the

cement-treated soil columns were also modeled with tension truncated Tresca model for a

better understanding of the simulation performance. The adopted tension strength, t =

0.11*qu is adopted based on the correlation established in Chapter 3. The design parameters in

these analyses are summarized in Tables 5.5 to 5.7. In common WAS approach, classical

158
Tresca model was adopted to average the properties of treated soil columns across the treated

block similar to earlier section but the difference here was that the untreated properties were

not included. The consequences of ignoring the untreated properties into the average

properties will be discussed later.

Table 5.5: Design parameters for Mohr Coulomb and Tresca models in hypothetical cases.

WAS
Material *c (kPa) others E (MPa)
configuration (kg/m3)
#
Silt - - 1500 2 ’ = 22 o 20 0.30
Treated Tf =
- - 1500 490 164 0.15
columns 107.8kPa
WM 0.498 1500 244 - 81 0.30
Composite
RM 0.176 1500 86 - 28 0.30
Mass
RM 0.873 1500 427 - 143 0.30

Notes: *c = unless otherwise stated; # denotes c’.

Table 5.6: CDP design parameters for cement-treated soil columns used in hypothetical cases.

Compression Tension
Compressive
Inelastic Strain Tensile Yield Stress Fracture Energy
Yield Stress
640 kPa 0 107.8 kPa 5 N/m
900 kPa 0.0005
980 kPa 0.002
980 kPa 0.004
850 kPa 0.0083
410 kPa 0.0165
150 kPa 0.0271
60 kPa 0.035

159
Table 5.7: CDP damage variables for cement-treated soil columns used in hypothetical cases.

Compression Tension

dc Inelastic Strain dt crack displacement

0 0 0 0
0 0.004 0.9 0.09 mm
0.95 0.035

Plasticity parameters: = 1o; eccentricity = 0.1; b0/ c0 = 1.16; Kc = 0.67. Identifications of


these parameters have been discussed in detail in Section 4.3.3.

Contact Properties

In WAS approach, full contact between cement-treated soil columns and in-situ soil was

assumed so no interface between surfaces was modeled. However, as mentioned earlier when

the treated soil mass served as retaining system in excavation, the front row free standing

cantilever treated soil columns does not have the confined condition of embankment cases.

Therefore, there could be a relative movement between treated soil columns and in-situ soil

which must be allowed for in the modeling. In RAS model, the contact surfaces were

distinguished in normal and tangential directions. Contact pressure-overclosure relationship

was applied to normal behavior which referred to as “hard” contact in Abaqus. The contact

surfaces will separate if the contact pressure becomes zero (in tension condition). For

tangential direction, the frictional behavior was modeled with classical Coulomb friction. The

coefficient of friction between the contacting surfaces was taken as 0.5.

Numerical Procedures

In these twelve analyses, sequential loading to simulate the installation of cement-treated soil

columns and excavation were not considered. Since it was an assumed excavation condition

and the main interest in the present study was to evaluate the failure mechanism, the analysis

160
was simplified to one step. For wish-in-place treated soil columns with final excavation

profile model (Figure 5.16), the load-displacement-stress response of the system was

investigated quantitatively by increasing the internal force due to the weight of soil with

application of gravity loading (10 kN/m3). In the beginning of this stage, the boundary

conditions shown in Figure 5.16 were activated.

5.4.3 Results and Discussions

Vertical cut without soil improvement (Case Li)

Figure 5.23 shows the ground response for Case Li where the analysis was manually

terminated halfway due to dislodged soil region relative to the original region. Snitbhan and

Chen (1978) found that it seems to be reasonable to consider bulging or loss of ground as an

instability criterion in the analysis. Figure 5.24 clearly reveals that the soil yields along the

sliding mass. Without any support element, the yielded soil slides towards the excavated face

in a wedge form. This is expected for a vertical cut in weak soil.

Vertical cut with grid type ground improvement pattern (Cases La, Lb, L1)

As earlier analysis reveals that a vertical cut in weak soil is not feasible due to stability issue,

ground improvement becomes necessary. In order to optimize the construction time and cost,

grid type ground improvement pattern is expected to work as a composite gravity wall to

facilitate the vertical cut. The analysis results on grid type pattern using WAS and RAS

approaches are compared.

161
WAS Approach:

Whole Mass (Case La) and Refined Mass (Case Lb) Models

The ground movement for the whole mass (WM) model and the refined mass (RM) model in

WAS approach are presented in Figure 5.25. On the other hand, Figure 5.26 shows plastic

strain developed at the passive ground but not in the treated zone for these two models. In

other words, WAS approach predicts potential failure of insufficient passive resistance in this

hypothetical case. Both analysis results reveal gravity wall behavior as assumed in WAS

modeling. No significant difference in ground response is observed between these two models.

Nevertheless, as illustrated in Section 5.3 where WAS approach might overlook the local

failure mechanism occurring at the columns, RAS approach was thus employed to investigate

the mechanism within the columns.

RAS Approach:

Grid Type Ground Improvement Pattern (Case L1-1, L1-2 and L1-3)

In RAS approach, three constitutive models, namely classical Tresca, tension truncated

Tresca and CDP models, were used to describe the cement-treated columns in Cases LX-1,

LX-2 and LX-3 (X named after the improvement pattern), respectively. The most

representative constitutive model for cement-treated soil was examined in Chapter 4 based on

laboratory tests, but the consequences of adopting the model for excavation problems in the

field are evaluated herein through hypothetical cases. Figure 5.27 shows the ground response

for Case L1-1 where nonuniform ground movement is observed within the grid block.

Arching form of soil movement is observed where the maximum movement appears at the

center of the untreated zone and reduces towards the tangential row treated soil columns. In

addition, some plastic strains develop in the passive ground and some behind rear row

columns, similar to that observed from WAS approach. It is believed that the development of

162
plastic strain is mainly due to the compressive pressure built-up when the ground moves

towards the excavated side. No plastic strain is found in the treated soil columns. However,

by limiting the tensile strength of the cement-treated column in Case L1-2, the ground

movement increase is associated with plastic strain observed in the overlap area between front

row of columns and tangential row of columns, as presented in Figure 5.28. A close-up of the

plastic strain appeared in the soil and treated columns are shown in Figure 5.29. The plastic

strain appeared in the overlapping columns is believed to be caused by reaching the limiting

tensile stress in contrary to that observed in Case L1-1. The plastic strain first occurred at the

top of the overlap columns and propagated downward to half of the column length. This

observation is supported by the concerns raised by Broms (2004) where failures may occur

due to separation of the columns. This will be further evaluated by the analysis result of Case

L1-3.

Case L1-3 adopted CDP model to simulate the behavior of treated column as established in

Chapter 4. Figure 5.30 shows the analysis result of ground movement and plastic strain

developed in Case L1-3. The excessive movement observed at the top of the front row

columns implies that the system is no longer stable according to Snitbhan and Chen (1978). A

close-up of yielding developed in the soil and damage occurred in the columns are presented

in Figure 5.31. It is evident that plastic strain has developed in a wedge envelope for the soil

within the grid block while macro-cracks are observed in the tangential row columns and

front row columns. The progressive crack development can be characterized by the damage

variable, dt, as demonstrated in Figure 5.32. As expected, the soil pressure is acting against

the front columns resulting in a “pulling away action” in the overlap area between the front

row columns and tangential row columns. When the induced stress exceeds the tension

capacity, crack initiates at the top of the overlap area and propagates downward along this

overlap area until the end of cantilever part of the columns. The cracks do not propagate down

to the embedded part of the columns as it is under confined. Subsequent loading causes the

columns in tangential rows to separate from one another thus forming new cracks in the

163
overlap areas between columns. The grid type wall becomes vulnerable when the first crack at

overlap area between the front row columns and tangential row columns continue to

propagate into the embedded part till the toe of the columns. Separation of front row columns

from the rest results in a cantilever condition for front columns which would then bend and

fail with maximum movement recorded at the top level. The failure is revealed with the

horizontal crack propagating across the front columns from the back. Failure of front columns

induces a wedge envelope of soil sliding towards the excavated side, and also cracks on other

remaining columns in tangent row. The failure mechanism reported in Case L1-3 is thus more

representative following to the observations made by Broms (2004) and Kitazume (2008).

Therefore, Case L1-3 yields a better prediction for grid type improvement pattern as

compared to Case L1-1 and Case L1-2.

Tangential Buttress Type Ground Improvement Pattern (Cases L2-1, L2-2 and L2-3)

From the previous case, it is learnt that soil arching occurred between two tangential rows

columns but the rear row columns seem not to contribute much to the overall system.

Therefore, a modification of the previous case was made where the rear row columns was

positioned to form a closer spaced tangential row columns (3 m center to center instead of 6

m in Case L1) (Figure 5.14b). This is named as tangential buttress (TN) type ground

improvement pattern in the present study. The ground movement and plastic strain developed

for Case L2-1 and Case L2-2 are presented in Figures 5.33 and 5.34, respectively. Case L2-2

yields similar result to Case L2-1 as no element in the columns reached yielding state, as

shown in the close-up diagram in Figure 5.35. It is noted that the closer spaced tangential

rows columns work better in facilitating the vertical cut due to more arching effect and

smaller soil pressure acting on the front row columns. Nevertheless, as we learn from

previous experience that the analysis is more convincing with CDP model, the

aforementioned observation is further compared with CDP model in Case L2-3.

164
Figure 5.36 shows the ground response for tangential buttress type ground improvement

pattern in Case L2-3. Extensive ground movement appeared in the front row columns similar

to Case L1-3 indicating this system is not stable. However, with closer spaced of tangential

rows columns, the disturbed zone is reduced compared to Case L1 which reveals soil arching.

A close-up shot in Figure 5.37 shows plastic strains developed within a wedge envelope with

corresponding damage observed in the treated soil columns. Figure 5.38 presents the

progressive crack development in the treated soil columns. Similar to Case L1-3, the cracks

first initiate in the overlap area of front row columns and tangential row columns from the top

to the bottom of the excavation level. Again, it demonstrates the weakest part of the columns

configuration is at the overlap area between the front row and tangential row columns. The

crack development in this system is similar to Case L1-3, ended with the front columns

bending and failing together with tortoise-cracks pattern in the other tangential buttress

columns slanting down towards the front columns.

Double-wall Type Ground Improvement Pattern (Cases L3-1, L3-2 and L3-3)

From the previous two cases, it is learnt that the cracks first initiate in the overlap area

between the front row and tangential row columns, then the columns bend and fail when the

cracks propagate across the whole front row columns at the excavation level. As a result,

modification of the improvement pattern was made by repositioning the rear row columns to

overlap with the front row columns in order to form a double overlapping columns (Figure

5.14c) termed as double-wall (DW) type in the present study. The ground movement and

plastic strains developed in Case L3-1 and Case L3-2 are presented in Figures 5.39 and 5.40,

respectively. Case L3-2 reports a greater ground movement compared to Case L3-1. The

plastic strains did not occur in the treated columns in Case L3-1 but did appear in tangential

row columns in Case L3-2. The plastic strains appeared in the overlap area between second

row columns and tangential row columns are caused by limiting tensile stress. Since Case L3-

165
1 adopts classical Tresca model which can allow for any amount of tensile stress, it is not able

to identify this yielding pattern.

A more reliable performance of double-wall type ground improvement pattern is analyzed by

using CDP model as shown in Figure 5.42. Among these three types of ground improvement

pattern, DW recorded the smallest ground movement with approximately 1/5 of the maximum

value of all cases. As shown in Figure 5.43, a wedge envelope of plastic strain is developed in

the soil behind the second row columns but not as severe as that in Case L1 and Case L2.

Progressive crack development in Figure 5.44 illustrates that the crack initiates at the overlap

area between second row columns and tangential row columns. At the same time, horizontal

cracks form at the back of second row columns near the excavation level. The horizontal

cracks progressively form a connection at the back of second row columns. Simultaneously,

the cracks in overlap area in tangential buttress continue to propagate vertically from top

surface to about two-third of the column length prior to slanting downwards towards the front

column toe. At a later stage, the horizontal cracks cut through the second row columns to

reach the front row. Nevertheless, the front row columns still survive as the cracks do not

propagate across the columns as happened in the previous two cases.

In comparison, numerical simulation results using WAS approach presented by Case La and

Case Lb are unable to capture the right mechanism for this hypothetical case. The classical

Tresca model again has its shortcoming of failing to address the tensile damage. As a result,

the analyses results for different ground improvement geometrical patterns adopting this

model for Case L1-1, Case L2-1 and Case L3-1 could not differentiate the comparative

performance of the 3 systems. Although the tension truncated Tresca model works better in

capturing yielding in tensile, it is still not capable of closely predicting the real mechanism as

what had happened in Case L2. Furthermore, this model tends to under-predict the damage

occurred as it assumes no softening effect. Nevertheless, by adopting the more appropriate

166
model, CDP, the aforementioned shortcomings can be addressed and hence enabling the

prediction of the failure as observed by other researchers (Broms, 2004; Kitazume, 2008).

5.5 Conclusion

In this chapter, the appropriateness of WAS and RAS approaches in simulating cement-

treated soil mass is evaluated. In the present study, two models using WAS approach were

considered, i.e. whole mass (WM) and refined mass (RM). The latter model is an

advancement over the former model by localizing the column positions in a strip manner.

Furthermore, the influences of constitutive model used to simulate cement-treated soil in

several selected cases were also examined. The adopted constitutive models include the

conventional classical Tresca model and CDP model as presented in Chapter 4. The tension

truncated Tresca model with a tension cutoff function incorporated in the classical Tresca

model was also employed for comparison.

Back-analyses of lime-cement columns subjected to shear box test were employed to evaluate

the failure criteria for cement-treated soil mass. Both WM and RM models in WAS approach

using classical Tresca model predicted shear failure with full area mobilization at the slip

surface area. Conversely, RAS approach predicted a localized damage in treated column at

some distance above and below the slip surface. The columns bent and induced tensile

damage, thus the mobilized shear stress was smaller compared to WAS approach but closer to

the laboratory measurements. In addition, the failure mechanism and damage positions

predicted by RAS approach using CDP model agreed well with the laboratory observations.

Nevertheless, incorrect failure mode was predicted by classical Tresca model even when the

mobilized shear stress is closely predicted. Through this analysis, it is found that the cement-

treated columns are likely to fail in bending due to limited tension capacity even for a shear

test having an induced slip plane. This is strongly supported by laboratory data where the

average fracture energy for tensile failure, Gf is approximately 1/15 of the fracture energy for

167
shear failure, Gfs (Namikawa and Koseki, 2006). Therefore, the commonly used WAS

approach which averages the property of the treated mass in term of shear strength is

established to be too simplistic in addressing cement-treated soil mass problems. On the other

hand, CDP is found to be an appropriate model. Together with a proper numerical approach,

RAS, it captures the tension-softening behavior to provide a realistic simulation of the failure

response of cement-treated soil columns.

The above established modeling criteria were then employed to study the hypothetical cases

of three different ground improvement patterns (GN, TN and DW as presented in Figure 5.14)

for a 4-m vertical open excavation in weak soil. The dimensions of treated soil columns

follow the one recommended by VERT. The analysis results showed that WAS approach

predicted failure due to insufficient passive resistance. Gravity wall behavior was revealed in

WM and RM models with no yielding within the treated zone. For RAS approach, there was

no major difference in the ground response and plastic strain among these three different

ground improvement patterns when the treated soil columns were simulated using classical

Tresca model. The relative movement between soil and cemented columns was very small.

These two materials worked together as a block and registered the same yielding as reported

by WAS approach. Nevertheless, once the tensile strength of cemented columns is limited

with tension truncated Tresca model, some plastic strain developed at the overlap area

between the front row columns and the tangential row columns.

Unlike confined condition experienced by cemented columns used in compression shear

loading in embankment, the loading condition experienced by cemented columns in an open

excavation is significant in the lateral direction in bending. With the ability of capturing the

tension-softening relation of cement-treated soil, CDP model predicted crack initiation at

overlap area between the front row columns and tangential rows columns. The crack

propagated down to the bottom of excavation level so the front columns were separated from

other columns, hence inducing excessive ground movements. Separation of columns is one of

168
the major concerns raised by Broms (2004). As the front row columns bent, crack propagation

took place at the bottom of excavation level resulting in toppling of the front row columns.

This predicted bending failure mode agrees well with laboratory observation by Kitazume

(2008). In other words, WAS and RAS approaches that adopted classical Tresca model and

tension truncated Tresca model predicted an incorrect failure mechanism for these

hypothetical cases.

As demonstrated in this study, it is revealed that tensile damage is the trigger for failure

initiation. This explains why the closer spaced tangential buttress in Case L2 (TN type) is less

effective as compared to thicker wall formed by double rows of columns in Case L3 (DW

type). The latter improvement pattern provides a longer crack path that requires more energy

for the horizontal crack to propagate across the columns. The damage process obtained with

RAS-CDP analysis is significantly more meaningful for understanding the failure mechanism.

This provides a fundamentally correct concept for the analysis of cement-treated soil used in

an open excavation. Although the above observations seem to be of common sense, this

phenomenological approach shows in detail the propagation of correct damage location, stress

distribution and the interaction mechanism of cement-treated soil columns and untreated soil.

169
Chapter 6 Field Studies

6.1 Introduction

Cement-treatment of existing weak soil is a preferred excavation method as it does not pose

obstacle for future development at the site. Nevertheless, as discussed in Chapter 2, there is

still a lack of understanding on the behavior of cement-treated soil mass in the field. This is

because it is often not economical to conduct large-scale field tests involving extensive

instruments and monitoring.

The studies in the preceding chapters established the understanding of the fundamental

behavior of cement-treated soil and identified the appropriate constitutive model, CDP, in

capturing both the compressive and tensile behaviors including its post-peak manners. The

shortcoming of the commonly adopted WAS approach is that it fails to capture the crack

development within the cement-treated soil column. This can be overcome by employing the

RAS approach with CDP model. The results of the hypothetical studies reported in the

previous chapter demonstrated the potential damage of cement-treated soil columns used as

earth retaining structure in an open excavation, i.e. separation of front row columns from the

others, bending of columns and horizontal crack propagation across the entire front row

columns resulting in collapse. In order to further the studies, this chapter aims to present the

limited available field case studies which are back-analyzed using the proposed numerical

approach, RAS-CDP. It is hoped that the behavior of cement-treated soil columns used as

earth retaining structure in an open excavation in the field can be well understood.

A total of three field case studies is back-analyzed in this chapter. First, a lateral load test

conducted by Babasaki et al. (1997) in Japan is studied. This test was solely a lateral load test

on cantilever cemented soil columns without other influence factors such as multiaxial

loading, boundary effect and column configurations. The back-analysis of this uniaxial

192
loading test under a well-defined condition would provide confidence in adopting the

proposed numerical method prior to a more complex excavation problem. Subsequently, two

localized collapse case studies in Singapore, i.e. waterway construction in Northeastern area

and basement construction in Central area, are evaluated. The former involved a complex

overlapping cement-treated soil columns configuration to facilitate an open excavation. The

latter involved a single row of overlapping cement-treated soil columns with reinforced I-

beam at intervals to facilitate the basement open excavation for a residential development.

6.2 Field Case Study 1: Lateral Load Test by Babasaki et al. (1997)

The in-situ lateral load test of cement-treated soil column conducted by Babasaki et al. (1997)

is briefly described herein. The test was conducted to examine the failure mechanism of

cantilever cement-treated soil columns subjected to lateral loading. This condition is close to

an open excavation condition where the overlapping cement-treated columns are formed

along the intended line of the proposed excavation and experienced lateral loading from the

retained soil.

6.2.1 Test Description

Soil Profile

Figure 6.1a shows the soil profile of this test. The first 1.75 m below the ground consists of

soil with debris. In view that the debris might affect the final mixing product, it was replaced

with fill. Beneath the fill is a 2 m thick sandy silt with SPT N value of 2. This is followed by

sand with reported SPT N values between 2 and 4, overlying a silt layer that mixed with loose

sand. Subsequently, 2 m thick gravel with SPT N values between 24 and 48 is encountered

overlying a sandy silt layer with SPT N values between 7 and 11. The groundwater table is

approximately 3.3 m below the ground level.

193
Test Program

The cement-treated soil columns were constructed by two-axis soil mixer with 1 m in

diameter and spaced at 0.8 m center-to-center. The column length was 9 m below the ground

level. Upon reaching the required curing time, the surrounding soil around the columns was

excavated to approximately 5.4 m deep to expose the columns for testing. The experimental

setup with elevations of displacement transducers is shown in Figure 6.1b. The lateral load

and column deflection were measured by load cell and displacement transducers, respectively.

Occurrence of cracks at the tensile surface of treated soil columns was monitored with micro

displacement transducers installed at every 10 cm interval up to 1 m from the excavated

ground level.

The load was applied in cycles approximately 6 months after the construction of columns and

the results are presented in Figure 6.2a. In order to avoid localized compression failure at the

loading point, the column head was protected with cast-in concrete and the loading point was

strengthened with steel plate. The observed damage pattern in the columns is shown in Figure

6.2b. After the test, samples of the cement-treated soil were collected for unconfined

compression and split tension tests to identify the material in-situ strengths. The unconfined

compressive strength, qu of the samples was established to range between 2.8 MPa to 7.6 MPa.

On the other hand, the split tensile strengths ( st) fell within the range of 0.1 to 0.2qu. The

laboratory test results for both tests are presented in Figure 6.3.

194
6.2.2 Numerical Analyses

Model Set-up

3-D FEM analyses using finite element program, Abaqus/Standard, were adopted for back-

analysis of this case study. Owing to the inherent existence of sharp curves at the overlap

column area, tetrahedral elements (C3D4) were employed to address these unique geometries.

A mesh refinement study was conducted and the mesh size was decided when the

convergence of analysis result was achieved. In the present analysis, the analysis result

converged when the element length for all the materials except steel plate was 0.15 m. A finer

mesh size, 0.05 m was adopted for steel plate where the applied load focused. As a result, a

total of 72,878 elements were employed to build the model as presented in Figure 6.4.

Material Models

There were four materials, i.e. untreated soil, cement-treated soil, concrete cap and steel plate,

to be assigned with relevant properties. As this study focuses on cement-treated soil column,

the other materials were simulated with a simple model-linear elastic. This is considered

reasonable as concrete and steel are much stronger than the cemented soil column. In addition,

the reported damage was found at explored cemented soil column and hence the untreated soil

did not play a significant role in this test. The elastic properties for these three materials were

taken from Namikawa et al. (2008) and presented in Table 6.1.

Table 6.1: Model parameters for steel plate, concrete cap and untreated soil (reproduced from
Namikawa et al., 2008).

Steel Plate Concrete Untreated Soil


E (MPa) ν E (MPa) ν E (MPa) ν
210000 0.3 25000 0.167 93 0.3

195
The parameters for cement-treated soil were estimated from the laboratory test results as

presented in Figure 6.3. The initial design value for qu was 3.1 MPa but the laboratory result

showed a range from 2.8 MPa to 7.6 MPa. It is common to see the strength of in-situ cement-

treated soil falls in a range as this material has been recognized by some researchers (Larsson

et al., 2005a; Larsson et al., 2005b and Chen et al., 2011) as a heterogeneous material. As a

result, a range of qu values according to the laboratory data was adopted herein. The st fell

between 0.1qu to 0.2qu. In view of this, a series of analysis was conducted, based on four

different qu values with three different st values correlated to each qu as summarized in Table

6.2.

Table 6.2: Summary of case analyses based on the in-situ strength of cemented soil column.

Case qu t

1 2.8 MPa
a) 0.20qu;
2 3.1 MPa
b) 0.15qu;
3 4.4 MPa
c) 0.10qu
4 4.9 MPa

Case 1 catered for upper bound analysis by adopting the lowest qu = 2.8 MPa. This case was

analyzed for three different tensile strengths, i.e. 0.2qu, 0.15qu and 0.1qu, denoted by symbol

‘a’, ‘b’, and ‘c’, respectively. Case 2 considered initial design value, qu = 3.1 MPa. Similar to

Case 1 and the remaining cases, three different tensile strength values corresponded to qu in

each case were analyzed. Referring to UCT data in Figure 6.3, two qu values appeared further

away from the rest. Therefore, by discarding these outliers, a mean value of the remaining

data, qu = 4.4 MPa, was adopted in Case 3. If incorporating these two outliers, an average

value qu = 4.9 MPa was obtained and adopted in Case 4. The range of these adopted analysis

parameters is better illustrated in Figure 6.5.

196
The behavior of cement-treated soil in this field case study was simulated by CDP model. The

linear elastic parameters followed those given in Namikawa et al., (2008) where = 1700

kg/m3, E = 2000qu, and = 0.167. Owing to absence of tension softening test result, the Gf

was taken as proportional to the laboratory results of Namikawa 2006 (Gf0 = 9 N/m with t0 =

380 kPa). The tension-softening was simulated by a linear strength loss approach with

parameter Gf as illustrated in Section 4.43. Unfortunately, no stress-strain data for

compression test was available. As such, assumption was made by taking the residual

compression strength as 60% of qu when inelastic strain became 0.0005. The analysis result

was used to investigate the effect of parameters used. Typical values for determining the yield

criterion were similar to those discussed in Chapter 3. The CDP model parameters used in the

present analysis are summarized in Table 6.3.

Table 6.3: Model parameters for CDP model of cemented column.

Compression Tension
Compressive
Inelastic Strain Tensile Yield Stress Fracture Energy
Yield Stress
*qu 0 * t Gf = t(Gf0/ t0)

0.6*qu 0.0005
Notes: 1. * refer values shown in Table 6.2.

2. Plasticity parameters: = 1o; eccentricity = 0.1; b0/ c0 = 1.16; Kc = 0.67. Identifications of


these parameters have been discussed in detail in Section 4.3.3.

Contact Properties

Since the concrete cap was cast in-situ, the surface between concrete cap and cemented soil

column was defined as perfect contact modelled by tie constraint in Abaqus. It tied these two

materials together so that there was no relative motion between them. Similar contact

property was applied between the concrete cap and steel plate as well. On the other hand, the

contact surface between the untreated soil and treated soil column might not be fully tied all

197
the time. However, the present analysis concentrated on the exposed part of cemented soil

column and it was believed that the contact behavior in the embedded part would not cause

significant effect to the overall test performance. As such, for computational simplicity, the

contact surface for untreated soil and treated soil column was defined as tie constraint as well.

Numerical Procedures

The boundary condition of the circumference area of embedded untreated soil was restricted

from normal direction movement. On the other hand, the model base was restricted from

moving vertically. Beyond aforementioned surfaces, the leftovers were set free. The boundary

condition for the present model is shown in Figure 6.6. The load test was simulated in two

steps:

1. Initial state condition was established with all the boundary conditions were activated.

2. A lateral prescribed displacement was applied at the center point of steel plate. The

steel plate was positioned at the center of concrete cap.

6.2.3 Results and Discussions

Prior to the comparison between field and analysis results, the raw field data was first

processed. Three cycles of loading were conducted and the results are shown in Figure 6.2a.

Nevertheless, a monotonic loading was simulated in the present finite element analysis. As

such, the unloading data was disregarded and the third cycle displacement was reinitialized

from the residual displacement induced by previous cycles. The processed field data for

lateral load-displacement at top of the cemented soil column is then presented in Figure 6.7.

The resistance load decreased after reaching peak at approximately 2 mm displacement.

Figure 6.8 shows the comparison of field and analysis results for four cases presented in

lateral load-displacement relations. The load resistance and displacement depend on the

198
cemented soil column strength, qu. Figure 6.8 shows that the field results lay between the

predictions by Case 2a and Case 3a. In other words, qu in the range of 3.1 MPa and 4.4 MPa

well reflects the in-situ strength of cemented soil columns. Figure 6.9 presents the analysis

results of these two cases with various tensile strength, i.e. 0.1qu, 0.15qu and 0.2qu denoted by

alphabet ‘a’, ‘b’, and ‘c’, respectively. The problem becomes very complicated when damage

occurred resulting the degradation of material strength and stiffness, thus difficult to obtain a

converging solution. As a result, the analysis terminated after a few post peak readings, even

yet to see major damage and final collapse. In spite of this, the analysis results are similar to

the field results. The lateral load resistance of cemented soil column in these cases increases

linearly to the ultimate value and then decreases upon further displacement.

Through the result comparison made between various tensile strengths, it is revealed that the

ultimate lateral load resistance depends on the tensile strength of the cemented soil column.

Since the tensile strength correlates to qu, this complements the observation made on Figure

6.8. Among the adopted parameters, the analysis result of Case 3b (qu = 4.4 MPa with t =

0.15qu) agrees well with the field measurement. The damage at cemented soil column for

Case 3b is presented in Figure 6.10. No compression crushing (represented by compressive

damage variable, dc) was observed which indicated that the mobilized compressive stress did

not reach the cemented soil column strength, qu. The applied lateral load at the top of the

cantilever column tended to bend the column thus inducing tensile stress at the column face

aligned to the loading face. Once the tensile strength of cemented soil column is exceeded, the

strength and stiffness degradation quantified by tensile damage variable, dt is manifested. As a

result, the resistance load decreases after reaching peak value. The tensile damage variable in

Figure 6.10b represents the physical crack which coincided well with field observation

(Figure 6.2b).

In addition, a conventional approach using classical Tresca model to simulate the cement-

treated soil column was analyzed for comparison purpose. The analysis result in lateral load-

199
displacement relation is plotted together with Case 3b and field measurement as shown in

Figure 6.11. Prior to damage, the CDP model and classical Tresca model yield similar

prediction of load-displacement relations. However, the load resistance in classical Tresca

model increases linearly without limit and surpasses the ultimate load resistance of 75 kN

reported in the field. It is clear that classical Tresca model grossly overestimates the load

resistance. As discussed in Chapters 4 and 5, this is believed to be attributed to the shear

failure mechanism assumed in classical Tresca model which allows for any development of

tensile stress, therefore this model failed to capture the tensile damage in the present case

study.

The yielding predicted by classical Tresca model only occurred when the lateral displacement

was 3.6 mm, see Figure 6.12. It is noted that the yielding locations are at the opposite side of

loading, in contrary to the observation in both field and CDP model. In Case 3b which

adopted CDP model to simulate the behavior of cemented soil column, crack occurred at

approximately 0.3 m above the excavated ground level and propagated across the column

parallel to the loading direction. It is evident that the crack pattern simulated in Case 3b is in

good agreement with field observation. Although classical Tresca model is commonly used in

practice to simulate the behavior of cement-treated soil, it is truly not an appropriate model

for problem involving lateral loading, as demonstrated in this case study.

Through this back-analysis, it is again demonstrated the significance of tension-softening

relation of cement-treated soil when subjected to lateral stress. This study has illustrated the

effectiveness of the proposed numerical approach, RAS-CDP model in providing an accurate

prediction of the behavior of cement-treated soil column subjected to lateral stress.

200
6.3 Field Case Study 2: Waterway Construction in Northeastern

Singapore

In northeastern Singapore, an approximately 4 km waterway was constructed to provide

water-based recreational activities. The excavation depth from existing ground level ranged

from 6 to 15 m. Owing to the space limitation, implementing a gentle slope cut was not

possible in some places. Furthermore, the existing ground of this site was weak hence

required a workable yet economical method to facilitate the construction. Compared to

conventional excavation method using retaining wall and lateral support, in-situ treatment of

the ground was the preferred method. This construction method left no structural element that

might be an obstacle for future development. In addition, improving the in-situ soil strength

and stiffness not only helped in stabilizing temporary excavation but also the permanent slope

stability.

Owing to unforeseen situations, there was a collapse of front row cement-treated soil columns

at a location. This case provided an opportunity to carry out a detailed study to supplement

the earlier findings for local soil conditions. It is hoped that from this field study, the

behaviors of cement-treated soil columns used as retaining system in an open excavation as

observed in hypothetical cases in Chapter 5 can be further evaluated.

6.3.1 Characteristics of Site

Geological Formation

According to the geological map of Singapore, the site is located in old alluvium (OA)

formation. It is a by-product of weathering of granite and low-grade metamorphic rocks (Ni,

2005). Generally, this is a dense material with mixture of sand, gravel, silt and clay.

Nevertheless, alluvial members, i.e. peaty clay (E) and fluvial sand (F1) are identified in

201
certain boreholes from the site investigation work. It is not surprise to encounter these alluvial

members as the site is located next to a river.

Subsoil Profile

At the collapse location, there were two available boreholes, i.e. BH-N13 and BH-A5. From

the soil investigation report, the first 2 to 2.4 m below the ground consists of fill sand mixed

with hard core or concrete fragment. Underneath the fill is the sandy clay with thickness

between 5.5 to 9 m. This material is weak with reported SPT N value of 3 in BH-N13 and

gradually improved to SPT N value of 8 in BH-A5. This is followed by a 3-m thick dark grey

soft peaty clay overlying a 3.3 to 6 m thick clay and silt layer with SPT N value of 3.

Subsequently, a silty sand layer followed by a sandy silt layer can be found with SPT N

values gradually increase with depth. The soil profile sketch is presented in Figure 6.13. The

groundwater table is approximately 1.1 m below the ground level.

6.3.2 Construction Method

In this collapse location, one side of the waterway was needed to be retained at its original

ground level for future development. In addition, to cater for future development, no structure

element was allowed to be leftover beyond the 1-m thick diaphragm wall boundary. In other

words, it was a cantilever wall system as no tie-back anchor or beam was permitted. The

waterway width was approximately 30 to 50 m and the retained ground levels were varied

which made the installation of temporary lateral support became impossible. Hence a

cantilever retaining system was to be built to serve in both temporary and permanent. As

mentioned earlier, cement-treating existing weak ground was adopted to facilitate the

construction work. The treated mass forming a “chair” configuration to provide a self-stable

cantilever retaining system was first proposed. To save time and cost, it was proposed to

202
optimize using grid type ground improvement pattern, see Figure 6.14. The treated ratio,

for cemented soil mass in front the diaphragm wall was 0.58 according to Eq. (2.2). The

cement-treated soil column diameter was 1.1 m and installed at 1 m center to center to form

the grid pattern. The design UCS for cement-treated soil column was 1 MPa.

Construction Sequences

The proposed construction sequence is as follows:

1) Install the diaphragm wall and deep cement mixing columns. Sample the core of in-

situ cement mixing columns for strength verification. Rectification work to be carried

out by compensating the columns that do not comply the strength requirement (UCS

= 1 MPa).

2) Excavate to reduced level 9.5 m (-RL 9.5 m) beneath the general ground level.

3) Excavate to –RL 15 m with slope shown in Figure 6.14.

4) Construct the reinforced concrete drain and backfill to required level.

However, amendments were made for the actual construction. The constructed cement-treated

soil column was 0.85 m in diameter instead of 1.1 m but the same proposed treatment ratio

remained. The column overlap area was kept at 100 mm. The cutoff level of deep cement

mixing columns in front of the diaphragm wall was at –RL 8 m. During the excavation, the

cement-treated soil columns were not trimmed to the proposed level thus creating an

approximately 7 m high cantilever shown in Figure 6.15. In the initial proposed design, the

cantilever length for front row columns was only 2.5 m. After a certain period of excavation,

a vertical crack occurred at the overlap column area followed by the collapse of front row

columns (Figure 6.16). Unfortunately, no core sample was done on the collapsed columns to

further investigate the in-situ strength. Prior to the commencement of excavation, random

203
sampling core test should be conducted to verify the in-situ strength on site. Although the

core test result is not able to be retrieved from the field record as a reference for this study, it

is believed that at least the strength requirement 1 MPa UCS was achieved on site since no

compensated column is identified from the as-built column drawing. Thus, the assumption

made on UCS = 1 MPa for the present back-analysis is a reasonable value.

6.3.3 Numerical Analyses

Model Set-up

In the initial design stage, 2-D finite element analyses with WAS-classical Tresca model was

adopted by others to in evaluate the performance of this earth retaining system. In order to

further investigate the cause of failure, 3-D finite element analyses with RAS approach and

CDP model were conducted in the present study. Quasi-static analyses using Abaqus/Explicit

were employed to solve the problem.

The cement-treated area covered approximately 28 m wide with a 1-m thick diaphragm wall

of 30 m deep. To incorporate boundary extension, the model involved an enormous number

of elements using RAS approach to capture the complex geometry of overlapping cement-

treated soil columns. To overcome this, the boundary of cement-treated area was modeled till

the edge of diaphragm wall (denote as interest zone to be studied in Figure 6.15) as the failure

plane would not across this robust wall. Considering the waterway construction was in a long

stretch, the analysis problem was simplified as a plane strain condition by taking 3 m wide

from center of a tangential row column to center of another tangential row column. The plane

strain condition was simulated by restricting the movement of these two faces in y-direction.

The lateral boundary for excavated side was allowed for two and a half times the excavation

depth, and restricted from moving along x-direction. The bottom boundary was allocated to

204
extend for at least two times the excavation depth, and fixed from moving in any direction.

Typical geometry and boundary condition for the problem is presented in Figure 6.17.

Similar to previous studies, first-order tetrahedral elements (C3D4) were employed to capture

the complex geometry. A total of two geometry models for four analyses were conducted to

study the problem, as summarized in Table 6.4. The first geometry (Case C-W) adopted a

commonly used approach, WAS – classical Tresca model shown in Figure 6.18. Element size

of 0.15 m was adopted along the vertical cut face and gradually increased to 1.2 m at the

untreated boundary and 0.4 m at the treated boundary. As such, the total elements that used to

establish the first geometry were 373,024. Unlike the simplified mesh in WAS model, the

difficulty of meshing the second geometry had increased tremendously when taking the

column configuration into consideration. In view of this, the treated area just in front the

diaphragm wall being the furthest zone from vertical cut face was simplified with rectangular

strip instead of columnar shape. This is to maintain the mesh quality and analysis integrity at

the interested zone. A total of 202,841 elements was used to establish the cement-treated soil

columns while another 240,523 elements were used to model the untreated ground, attributing

to a total of 443,364 elements for this geometry. Figure 6.19 shows the finite element mesh

for second geometry namely, Case C-R in the present study.

Table 6.4: Summary table of analysis cases for case study 2.

Case Total elements WAS/RAS Soil Model for Treated Mass


C-W 373,024 WAS-WM Classical Tresca
C-R1 443,364 RAS Classical Tresca
C-R2 443,364 RAS Tension truncated Tresca
C-R3 443,364 RAS Concrete damage plasticity

205
Material Models

The soil parameters used in the present analyses as deduced from the site investigation report

are summarized in Table 6.5. The yield criterion for the material was simulated by Mohr

Coulomb model described by c’ and ’. On the other hand, cement-treated soil column was

simulated by three types of constitutive model, i.e. classical Tresca, tension truncated Tresca

and concrete damage plasticity models for comparison purpose. With the performance of

these three constitutive models in simulating behavior of cement-treated soil columns in an

open excavation assessed in Section 5.4.3, the reliability of the assessment is further

evaluated using this case study. Owing to the lack of retrieving the stress-strain data for

cement-treated soil column on site, the design parameters were then assumed by taking to the

laboratory results reported in Chapter 3 as the qu values were similar. Tensile strength, t =

0.11*qu was adopted according to the correlation established in Chapter 3. In WAS approach,

Eq. (2.1) was used to average the properties of cement-treated soil columns across the treated

zone without considering the untreated properties. As illustrated in Section 5.4, the exclusion

of untreated properties in computing the average properties for treated zone had insignificant

effect to the overall performance. The design parameters in these analyses are summarized in

Table 6.5. The CDP model design parameters for cement-treated soil columns are similar to

those presented in Tables 5.6 to 5.7.

Contact Properties

Similar to Chapter 5, perfect contact between cement-treated soil columns and clay was

assumed in WAS approach. Conversely, a specific contact property between these two

materials was assigned in RAS model. In RAS model, the contact surfaces were distinguished

in normal and tangential directions. Contact pressure-overclosure relationship was applied to

normal behavior which referred to as “hard” contact in Abaqus. The contact surfaces will

206
separate if the contact pressure becomes zero (in tension condition). For tangential direction,

the frictional behavior was modeled with classical Coulomb friction.

Table 6.5: Design parameters for Mohr Coulomb and Tresca models used in case study 2.

Material (kg/m3) Cohesion (kPa) E (MPa)


Sandy Clay 1800 0 24 15 0.3
Peat 1600 0 20 6 0.3
Clay with Silt 1600 0 24 8 0.3
OA (N<10) 1900 2 30 1.75*N 0.3
OA (10<N<30) 1900 2 33 1.75*N 0.3
OA (30<N<50) 2000 5 33 1.75*N 0.3
Treated Column 1600 490 - 164 0.15
Composite Mass
1600 284 - 95 0.3
( =0.58)

Numerical Procedures

In these four analyses, the numerical procedures were simplified to one step without

considering the installation of cement-treated soil columns and excavation sequence. The

cement-treated soil columns were modeled as wish-in-place together with final excavation

profile then the whole model was subjected to an increment of internal force due to the weight

of soil with application of gravity loading (10 kN/m3). In the beginning of this stage, the

boundary conditions shown in Figure 6.17 were activated. Similar procedure was applied by

Chen and Mizuno (1990) in investigating the load-displacement-stress response of a vertical

slope. This analysis presented the final stability of the excavation thus it might over-predict

the stability of on-going excavation stages.

207
6.3.4 Results and Discussions

The analyses were manually terminated upon excessive ground movements. Analysis result

for Case C-W adopting WAS approach with classical Tresca model is presented in Figure

6.20. The treated zone works as a massive block pushing against the passive retained soil,

thus developing plastic strain (PE) at the passive retained soil and induces ground heave at

excavated level. With subsequent loading, a circular slip envelope of small movement

develops at the cantilever cement-treated block initiating at approximately 2 m above the

excavated level. A passive resistance failure is predicted by Case C-W. However, the above

mentioned scenario was not observed in the field which put the reliability of this numerical

approach in question.

For Cases C-R, it is hoped that with the RAS approach to capture the column configurations

and properties, a closer prediction to field behavior can be achieved. Figures 6.21 to 6.23

show the ground movements and plastic strains developed in Cases C-R1, C-R2 and C-R3,

respectively. It is observed that the untreated soil settles and moves relatively from the treated

soil columns. Both Cases C-R1 and C-R2 predict plastic strains develop at the passive

retained soil similar to Case C-W. Unlike Case C-W predicts small movement in the treated

zone, these two cases report wedge envelope movement slanting upward from the excavated

level. Analysis result of Case C-R1 shows plastic strains develop at the overlap area between

treated soil columns in tangential rows. A more pronounced yielding occurs at these overlap

areas when tensile strength of the cement-treated soil is limited as in Case C-R2. As such, the

ground movement reported in Case C-R2 is much greater than in Case C-R1. No obvious

damage is noted in the front row columns in these two cases.

However, compared to field observation in Figure 6.16, damage occurred near the excavated

level and at the middle part of the front row columns. As such, it can be said that the

predictions by Case C-W, Case C-R1 and Case C-R2 are less precise. Conversely, a more

208
exact prediction is demonstrated by RAS approach with CDP model in Case C-R3 presented

in Figure 6.23. By capturing the tension-softening relation of cement-treated soil column with

CDP model, the analysis illustrates how the tensile crack is formed in the cement-treated soil

columns. The damage first occurs at the overlap area between columns in tangential rows

similar to hypothetical cases in Section 5.4. Subsequent loading induces horizontal cracks

across the columns at the middle part of the cantilever front columns. Simultaneously, the

vertical cracks at overlap area propagate downward and across the front columns horizontally

at excavated level. When these micro cracks join up to become macro cracks, collapse of

front row columns occurs. The damage locations predicted by Case C-R3 agree well with the

site observation.

This field case study has demonstrated that RAS approach using CDP model to incorporate

the tension-softening behavior of cement-treated soil columns is the appropriate way in

simulating the behavior of cement-treated soil columns used as retaining system in an open

excavation. In other words, the observations addressed in hypothetical cases in Section 5.4 for

three different layout patterns using this approach are reliable and can be projected to field

performance.

Nevertheless, it should be noted that the limitation in present study is that the numerical

simulation is catered for the worst condition as without considering the excavation unloading

sequence. Wish-in-place columns with final excavation profile model with application of

gravity loading is checked against final stability. Therefore, during the construction period,

the ground might experience unloading so the performance might not as worse as presented in

the analysis. In view of this, the analysis results were selectively outputted to coincide with

existing site condition. In spite of this, the prediction in Case C-R3 still agrees well with site

observation.

209
Through the back-analysis of this case study and hypothetical case studies presented in

Section 5.4, it is learnt that cracks initiate at the overlap areas between the front row columns

and the tangential rows columns and eventually resulting in separation of columns. This can

be improved by constructing a capping beam overtop the columns to tie them together as

recommended by VERT. However, one should be aware that the loading condition

experienced by cement-treated soil columns used as earth retaining structure in an open

excavation is significant in the lateral direction. Therefore, tying the column configurations

overtop does not ensure the whole system integrity as localized crack may occur at the

exposed front row columns when the induced tensile stress exceeds the capacity as happened

in the present case study. As demonstrated in Section 5.4, a thicker front wall formed by

bigger diameter columns will provide a safer retaining system.

6.4 Field Case Study 3: Basement Construction in Central

Singapore

The third field case study concerns a residential development in central Singapore. In

Singapore, the land is very limited and most development projects need to fully utilize the

available space. As such, a car park is often built underneath the residential building. In this

project, the development covered a L-shaped area with longest span of 200 m and 135 m wide.

The site encountered soft soil at shallow depth. Cement-treated soil columns reinforced with

small steel beam members is adopted as retaining system along the intended excavation line

to facilitate the basement construction work. In certain locations, additional embedded

cement-treated soil berm was formed to provide a stiffer support. Generally, the excavation

depth was approximately 4.5 m.

However, due to over-excavation for a deep pile cap construction, a localized collapse of

cement-treated soil columns was reported at site. In addition to the previous two field case

210
studies, this case with different ground condition and cement-treated soil column

configurations is back-analyzed to supplement the understanding of cement-treated soil

columns used as retaining system in an open excavation.

6.4.1 Characteristics of Site

Geological Formation

According to the geological map of Singapore, the site is located in Bukit Timah granite

formation. The Bukit Timah granite is mainly an acidic igneous rock formed during the

Triassic period. The degree of weathering of the residual soils from this formation is

reasonably uniform, decreasing gradually with increasing depth due to chemical

decomposition under the humid tropical climate. Its dominant, granitic component is grey and

medium to coarse-grained. However, soft clay Kallang soil layer is detected in a few

boreholes from site investigation.

Subsoil Profile

From the borehole next to the collapse location, the first 2 m below the ground is categorized

as fill layer made of clayey silt mixed with concrete hard core. Underneath the fill is a 4.8-m

thick grey soft clay followed by a 4.5-m thick soft dark grey sandy clay. This material is weak

with reported SPT N value of 2 and an effective friction angle of 23o. Subsequently, the Bukit

Timah granite formation is encountered. The highly weathering residual soil, clayey silt of

5.5-m thick with SPT N value of 5 is first encountered. The SPT N value of the residual soil

layer then gradually increases with depth. The soil profile is presented in Figure 6.24. The

groundwater table is approximately 1.2 m below the ground level.

211
6.4.2 Construction Method

The excavation depth is 4.5 m. Owing to the presence of an approximately 9-m thick soft clay,

adopting a conventional retaining wall such as sheetpile wall is too flexible and requires

additional embedded strut support to control the maximum wall deflection at the base of

excavation (Tanaka, 1994). In addition, wall extraction at the end of construction consumes

time and may induce additional ground movement if not well conducted. Soldier pile wall

requires effective lagging to avoid the clay from squeezing into the excavation side. Stiffer

wall such as contiguous bored pile wall or diaphragm wall is too expensive for such shallow

depth of excavation. Hence, it has been decided to adopt cement-treated soil columns as the

retaining system to save cost and time. The retaining wall was formed from overlapping 1.5 m

diameter columns installed at 1.25 m center-to-center. Steel member, I beam with 200 mm x

200 mm in dimension was inserted at every 2.5 m interval to strengthen the retaining system.

To reduce the lateral earth pressure acting on the wall, it was proposed to trim the ground

immediate behind the wall to 2 m lower than the existing level. The proposed construction

method is presented in Figure 6.24.

Construction Sequence

The proposed construction sequence is as follows:

1) Install the deep cement mixing columns and steel members. Sample the core of in-

situ cement mixing columns for strength verification. Rectification work to be carried

out by compensating the columns that do not comply the strength requirement.

2) Trim the ground to reduced level 2 m (-RL 2 m) from the existing ground with slope

gradient 1V : 1.5H.

3) Excavate to –RL 4.5 m as shown in Figure 6.24.

4) Construct the base slab and wall before backfill to ground level.

212
However, in the actual construction, the cutoff level of deep cement mixing columns was at

ground level and the ground was not trimmed to the proposed level. As a result, the cantilever

span for the cement-treated soil columns became 4.5 m instead of 2.5 m in the intended

design. The retaining system became vulnerable when further excavation was carried out to

construct a 3-m deep pile cap without approval from the designer. The on-site excavation

profile is shown in Figure 6.25. As a consequence, the cantilever span is much larger at 7.5 m.

Cracks were observed in the cement-treated soil columns (Figure 6.25) immediately after the

over-excavation and failure was reported in the following day, see Figure 6.26.

6.4.3 Numerical Analyses

Model Set-up

To investigate the cause of failure, 3-D finite element analysis with the proposed numerical

method, RAS approach and CDP model was carried out. Abaqus/Explicit was employed to

solve the problem as it involved substantial yielding in the collapse case.

As the collapse of cement-treated soil columns was due to the localized over-excavation, only

this localized area was considered in the present study. The problem was considered as a 5 m

wide span between the center of two columns and their faces were restricted from moving in

y-direction. The lateral and bottom boundaries were at least two and a half times the

excavation depth, and were restricted from moving in x-direction and fixed in any direction,

respectively. The geometry and boundary condition for the problem is presented in Figure

6.27.

A total of 470,274 tetrahedral (C3D4) elements were employed for this problem. The adopted

element sizes were 0.15 m at the column locations and gradually increased to 1.2 m at

boundaries. The two steel members (I-beam with 200 mm x 200 mm in dimension) were

213
modeled as shell comprising a total of 5,760 linear quadrilateral elements (S4R) embedded

into the columns. Figure 6.28 presents the finite element mesh for the present case study.

Material Models

The soil parameters as derived from the site investigation are summarized in Table 6.6. The

yield criterion for the material was simulated by Mohr Coulomb model described by c’ and ’.

As illustrated in previous case studies where the WAS approach and Tresca model are

incapable in simulating the cement-treated soil columns in excavation problem, thus only the

proposed numerical method – RAS approach with CDP model was considered. The CDP

model parameters for cement-treated columns are similar to Section 5.4, as presented in

Tables 5.6 to 5.7. In terms of contact properties for the surface between cement-treated soil

column and untreated soil, “hard” contact in Abaqus was applied to normal direction while

classical Coulomb friction behavior was applied for tangential direction.

Table 6.6: Design parameters for case study 3.

Material (kg/m3) cohesion (kPa) E (MPa)


Fill 1800 2 30 10 0.3
Marine Clay 1500 - 22 5 0.3
Sandy Clay 1900 - 23 10 0.3
Clayey Silt (N=5) 1800 2 28 10 0.3
Clayey Silt (N=15) 1800 3 28 30 0.3

Yield stress
Material (kg/m3) E (GPa)
(MPa)
Steel I-beam 7800 275 210 0.3

214
Numerical Procedures

Similar to case study 2, the numerical procedures for the present case were simplified to one

step without considering the installation of cement-treated soil columns and steel members as

well as the excavation sequences. The cement-treated soil columns and steel members were

modeled as wish-in-place together with final excavation profile then the whole model was

subjected to an increment of internal force due to the weight of soil with application of

gravity loading (10 kN/m3). In the beginning of this stage, the boundary conditions shown in

Figure 6.27 were activated. This analysis presented the final stability of the excavation thus it

might over-predict the performance during on-going excavation stages.

6.4.4 Results and Discussions

In the field, progressive development of cracks may happen very fast and within the

embedded region. As such, it is not easy to record the whole progress for failure mechanism

investigation. Nevertheless, with the help of appropriate numerical method proposed in the

present study, the understanding of the failure process can be established. Figure 6.29 shows

the numerical analysis result of case study 3 in terms of lateral ground movements and

progressive development of cracks in cement-treated soil columns. The analyses were

manually terminated upon excessive ground movements. It is noted that the ground

movements are similar to those of vertical cut with maximum lateral movement occuring at

certain distance above excavated level. As a result, the cement-treated soil columns wall

bulges and sustains tensile stress at this location at the exposed column face.

Cracks (represented by tensile damage variable, dt) occur when the induced tensile stresses

exceed the cement-treated soil capacity. For this 7.5 m cantilever wall, the cracks appear at

exposed column above the excavation level except the top 2.5 m column remains intact.

Compared to 4-m cut hypothetical cases in Section 5.4, case studies 2 and 3 both show

215
appearance of horizontal cracks in the middle part of the cantilever wall. This indicates that

the induced tensile stress beyond its capacity may happen at the cantilever span instead at the

excavation level, thus depending on the cantilever span and cement-treated soil strength.

Moreover, with the presence of embedded steel beam members in the columns, the crack

locations will shift compared to unreinforced case as demonstrated in case study 3 and

observation by Babasaki et al. (1997). Figure 6.30 shows that for case without steel beam

member, the horizontal cracks join up and continue propagate across the columns. On the

other hand, for case with steel beams, when the horizontal cracks meet steel beams, they tend

to turn around and propagate at easier path that requires less energy. Cross-sections of the

cement-treated soil columns with and without steel beam member shown in Figure 6.30

present the crack development in the cement-treated soil columns. In this case study, the steel

I-beam member size (200 mm x 200 mm) is relatively small compared to the cement-treated

soil column 1.5 m in diameter, thus the cracks continue to propagate across the columns

(especially at the overlap area as the crack path is the shorter) and collapse occurs when these

cracks join up and form major cracks. In other words, the contribution by this small steel

beam member to the overall performance of cement-treated soil column wall is insignificant.

It is noted that there is no compressive damage (represented by damage variable, dc) occur. A

comparison between the field observation and analysis result in Figure 6.31 shows that the

prediction by the present study is reasonable. The damage occurs at the lower part of the

columns and the upper part still remains intact as observed in the field.

From this case study, it is learnt that a considerable size of steel beam member is required to

design a stable retaining wall formed by cement-treated soil columns. Insertion of such steel

beam members in every column is preferable otherwise it becomes the weak point for cracks

to propagate through as lesser energy is needed compared to the reinforced columns. In

addition, the observations from published successful case in downtown Lexington, Virgina,

USA (Ruffing et al., 2012), hypothetical cases in Section 5.4, case studies 2 and 3, show that

the front row columns thickness play the most significant role for this type of cantilever

216
retaining structure. For cement-treated soil columns not reinforced with steel beam members,

it is essential to design a sizable column diameter to resist the induced tensile stress as well as

to provide strong resistance against crack propagation.

6.5 Conclusion

This chapter presents three field case studies involving cement-treated soil columns and back-

analyzed by 3-D FEM using RAS approach and CDP model. The different loading and

ground conditions, column strengths and configurations in these field cases have provided

valuable data in studying the behavior of cement-treated soil columns used in an open

excavation.

The back-analysis of a lateral load test conducted by Babasaki et al. (1997) has demonstrated

that the proposed numerical method is able to provide a fair prediction of lateral load

resistance-displacement relation and damage pattern compared to the field measurements.

Conversely, the classical Tresca model overestimates the load resistance and predicts the

failure mechanism incorrectly. It is noted that the heterogeneous property of the cement-

treated soil should be taken into consideration for a more precise prediction.

For a cement-treated mass involving certain column configuration such as that in field case

study 2, conventional WAS approach with Tresca model predicts passive resistance failure

associated with ground heave and no damage occurred in the treated mass which is not

appropriate. RAS approach can present a better simulation provided that an appropriate

constitutive model is used. For instance, RAS approach with classical Tresca model and

tension truncated Tresca model predict yielding at cement-treated mass but not at the front

columns as reported in the field. Only after incorporating the tension-softening behavior of

the cement-treated soil in CDP model, the analysis (Case C-R3) manages to capture the

tensile damage appeared in the front row columns that is consistent with field observations.

217
In this chapter, it is observed that for a long span cantilever front wall formed by overlapping

cement-treated soil columns, the tensile damage may occur at certain distance above the

excavation level and this depends on the cantilever span and the cement-treated soil strength.

Addition of steel members in the cement-treated soil columns may shift the damage position

which is consistent with the test observation by Babasaki et al. (1997). The back-analyses of

field case studies revealed that tensile damage with softening is the trigger for failure

initiation of this retaining system particularly at the front row columns. As such, the

embedded lengths of the cement-treated soil columns seem not to be effective as required in

conventional retaining wall.

The back-analysis results of these three field case studies are found to be generally consistent

with field observations. Moreover, the predicted development of progressive failure in these

field cases coincide with the findings presented earlier for hypothetical cases in Section 5.4.

As such, it is confident to say that the prediction for three different ground improvement

geometrical patterns using the proposed numerical method in Section 5.4 are reliable and can

be projected to field performance.

There is few design implications learnt from the present study. First, for retaining structure

that formed from cement-treated soil columns in certain configuration, the weakest part is at

the overlap area between front row columns and others. Cracks tend to initiate at the overlap

area and cause separation of columns. This can be improved by constructing a capping beam

overtop the columns to tie them together as recommended by VERT. However, for long span

cantilever front wall, tensile cracks may still occur at the cantilever span. Hence, a more

effective way as demonstrated in Section 5.4 is to provide a sizable column diameter to resist

the induced stresses instead of designing other column configurations. Reinforcing the front

row columns with steel beam members is only effective when a considerable member size is

embedded in every column. Otherwise, it will not intercept the crack propagation if the steel

member size is too small compared to the overall column diameter.

218
Chapter 7 Conclusions

7.1 Summary of Findings

The studies presented in preceding chapters aim to provide a better understanding of the

behavior of cement-treated soil columns used as earth retaining structure in an open

excavation. The findings of this study are summarized as follows.

1) Through experiments, the behaviors of cement-treated Singapore marine clay are

established as follows.

a) The compressive strength increases rapidly up to the peak value and then decreases

abruptly to a small value upon further straining. At cement content = 20% with 14

days curing period, the treated material manifests brittle behavior with shear and cone

failure observed in the test sample. The strength associated with brittleness increases

with cement content and curing period. For higher cement content and longer curing

period, vertical compressive crushing is dominant and columnar fracture is observed.

b) Tensile strength is measured by split tension test which can be well correlated with the

unconfined compressive strength as st = 0.11qu for the cement content tested in the

present study. The tensile post-peak behavior is presented by a relation between stress

and fracture zone deformation. Three-point bending notched beam test was employed

to derive the tension-softening relation based on fracture mechanics concept. For the

mix proportion used in the present study, Gf falls within a range of 2.6 N/m to 4.4 N/m.

A bilinear relation is observed in the tensile stress-crack opening displacement

relationship, but the relation becomes linear when the material brittleness increases.

2) The use of three-point bending notched beam test to study the tension-softening relation of

cement-treated clay is a relatively new approach although it has been widely used in

concrete and cement-treated sand by Namikawa (2006). Therefore, parametric studies

238
were conducted to evaluate the influences of test parameters, i.e. notch width and sample

size. Test results show that there is no difference between notch width 0.7 mm and 2.5 mm

due to a uniform particle size distribution around the notch tip. On the other hand, there is

also no difference in the test results for 50 mm and 40 mm sample sizes. The 10 mm

difference in size is insignificant compared to the overall stress transition zone which

usually involved hundreds of mm according to literatures. The stability problem in the

tests is closely related to the stiffness of the testing machine comparing to the brittleness of

the cement-treated clay. It is observed that the stability problem is more pronounced for a

more brittle specimen.

3) The comparison between Tresca, isotropic and concrete damage plasticity (CDP) models

based on calibration of laboratory tests in the present study and published data shows that

the third model is far superior to others in simulating the behavior of cement-treated soil:

a) Classical Tresca model assumes single stiffness and perfectly-plastic failure so it

cannot capture the strain hardening and softening behaviors of cement-treated soil.

Although the model can be improved with a built-in tension cutoff function to limit

the tensile strength, it is still incapable of modeling the strain softening part.

b) Isotropic model predicts close to the uniaxial compression test measurements but not

the tension failure as the yield criterion in this model allows the development of any

value of tensile stress.

c) In CDP model, the user-input uniaxial stress-strain relations can be employed to

determine the current state of the yield surface to analyze multiaxial load cases. As a

result, this model can well capture the hardening and softening behaviors of cement-

treated clay and sand. The model assumes a linear loss of strength after initial

cracking which is fitted well to the behavior of cement-treated clay observed in the

present study. Moreover, the post-failure stress can also be defined as a tabled

function of crack displacement which is demonstrated to predict the bilinear relation

of cement-treated Toyoura sand well.

239
4) When weighted average simulation (WAS) approach is employed to model the cement-

treated soil columns used in an open excavation, several errors can occur. These errors are

identified with real allocation simulation (RAS) approach using CDP model with

comparisons made to laboratory test and field cases.

a) Homogenizing the mass properties only allows for a single constitutive model thus

violates the underlying mechanism for treated soil and untreated soil with very

different characteristics. It is shown in this study that the commonly used Tresca

model fails to capture the tensile damage developed in the cement-treated soil

columns as shear failure of the treated mass is predicted.

b) The front columns that form along the intended line of proposed excavation are

subjected to lateral load and results in a “pulling away action” at overlap areas

between front columns and other tangential columns. Cracks initiate when the

induced tensile stress exceeds the capacity. Therefore, regarding the treated mass as a

composite material will not allow failure to happen within the mass is incorrect.

c) Perfect bonding between treated soil and untreated soil assumed in WAS approach is

not appropriate as the analysis result shows that untreated soil moves relatively apart

from the cement-treated soil columns.

5) Hypothetical studies of a vertical cut improved by three types of ground improvement

geometrical patterns (namely Grid, Tangential buttress and Double wall) show that the

vulnerable part of this retaining system is the front row columns. The rear parallel row

columns in Grid pattern do not contribute much resistance for the retaining system. A

thicker front wall (i.e. double-overlapping rows or bigger diameter of cement-treated soil

column) is more effective than other ground improvement geometrical patterns as the

cracks consume more energy to propagate through a longer path to cause failure. This

finding is not possible to be observed in the WAS approach as it does not take the column

configuration into consideration.

240
6) Back-analyses of field case studies confirmed that tensile damage is the trigger for failure

initiation of cement-treated soil columns used as retaining structure in an open excavation.

The horizontal tensile crack positions that appear in the front columns depend on the

cantilever span and may also appear at the middle span or at the excavation level. Addition

of steel members in the cement-treated soil column will shift the horizontal tensile crack

positions. Nevertheless, the contribution of the steel member to the retaining wall

performance highly depends on the member size.

7) The damage patterns obtained by the proposed numerical method, RAS approach with

CDP model are generally consistent with the trend observed from field case studies. In this

proposed numerical method, each individual treated soil column is modeled individually

with assigned properties described by CDP model. CDP model captures the tension-

softening behavior to provide a realistic simulation of failure response of cement-treated

soil columns. It shows in detail the propagation of correct damage locations, stress

distribution and the interaction mechanism of cement-treated soil columns and untreated

soil. This is significantly meaningful as it provides a fundamentally concept for the

analysis of cement-treated soil columns used as earth retaining system in an open

excavation.

8) There are some design implications that can be learnt from the present study. For retaining

structure with cement-treated soil columns in certain configuration, the weakest part is at

the overlap area between front row columns and others. This can be improved by

constructing a capping beam overtop the columns to tie them together as recommended by

VERT. However, for long span cantilever front row columns, tensile crack may occur at

the cantilever span. This can be overcome with a sizable column diameter to resist the

induced stresses which is superior to other column configurations design. Reinforcing the

front row columns with steel beam members is only effective when a considerable member

size is embedded in every column.

241
7.2 Recommendations for Further Study

This research studies several fundamental characterizations of cement-treated soil columns

used as earth retaining structure in an open excavation. However, some improvements can be

done as follows:

(a) The knowledge gained from the laboratory work in this study paves the way for further

development of an established laboratory test guidelines in determining the tension-

softening relation of cement-treated clay. The currently adopted test guideline of three-

point bending notched beam test is based on RILEM’s recommendation for concrete.

Although concrete and cement-treated clay both behave as quasi-brittle materials; the

particle size, strength and brittleness are quite different from each other. Therefore, it

would be of interest to conduct a detailed study of laboratory test guidelines in

determining the tension-softening relation of cement-treated clay.

(b) Since CDP model is proven to be an appropriate constitutive model in simulating the

behavior of cement-treated soil, the yielding parameters can thus be studied. So far, the

relevant parameters used in the present study are inferred from concrete test data.

(c) To better reflect the actual cement-treated soil columns and ground performance in finite

element analysis, it is recommended that the excavation be modelled in its proper

sequences.

(d) In view that cement-treated soil is so brittle and failure would occur very suddenly, it

would be of interest to study any measurement (e.g. addition of fiber to increase ductility

of the final mixture product or insertion of a comparatively large size of steel member in

the cement-treated soil column) that can prevent the collapse failure as reported in field

case studies 2 and 3.

242
REFERENCES
Abaqus 6.11, A. (2011). Online manual, SIMULIA. Providence, RI, USA.

Adams, T. E. (2011) Stability of levees and fllodwalls supported by deep-mixed shear walls. Virgina
Polytechnic Institute and State University.

Andromalos, K., Hegazy, Y. and Jasperse, B. (2001). "Stabilization of soft soils by soil mixing." Soft
Ground Technology,194-205.

Antão, A. N., Guerra, N. M., Cardoso, A. S. and Fernandes, M. M. (2007). "Earth pressures of soils in
undrained conditions. Application to the stability of flexible retaining walls." 5th International
Workshop Applications of Computational Mechanics in Geotechnical Engineering, Portugal,247-256.

Arroyo, M., Ciantia, M., Castellanza, R., Gens, A. and Nova, R. (2012). "Simulation of cement-
improved clay structures with a bonded elasto-plastic model: A practical approach." Computers and
Geotechnics, 45, 140-150.

Assarson, K., Broms, B., Granhom, S. and Paus, K. (1974). Deep Stabilization of Soft Cohesive Soils.
Sweden, Linden Alimark.

Babasaki, R., Suzuki, Y., Nakama, T. and Yamada, K. (1997). "Study for reinforced earth retaining
wall by deep mixing method." Proceeding 32th Annual Conference of Japanese Geotechnical
Society,1733-1734.

Babasaki, R., Terashi, M., Suzuki, T., Maekawa, A., Kawamura, M. and Fukazawa, E. (1996). "Factors
influencing the strength of improved soils."JGS Technical Committee Report.

Baratta, F. I. and Underwood, J. H. (1992). "Notch dimensions for three-point bend fracture specimens
based on compliance analyses."ARL-TR-35, U.S. Army Research Laboratory

Bazant, Z. and Kazemi, M. (1991). "Size dependence on concrete fracture determined by rilem work-
of-fracture method." International Journal of Fracture, 51, 121-128.

Brinkgreve, R. B. J. (2005). "Selection of soil models and parameters for geotechnical engineering
application." Geotechnical Special Publication, ASCE, 128, 69-98.

Broderic, G. P. and Daniel, D. E. (1990). "Stabilizing compacted clays againts chemical attack."
Journal of Geotechnical Engineering, 116(10), 1549-1567.

Broms, B. B. (2004). Lime and lime/cement columns. Ground Improvement 2nd edition. M. P. M. a. K.
Kirsch, Spon Press: 252-330.

Broms, B. B. and Boman, P. (1979). "Stabilization of soil with lime columns." Ground Engineering,
12(4), 23-32.

243
CDIT (2002). The Deep Mixing Method:Principle, Design and Construction. The Netherlands, A.A.
Balkema.

Center, P. W. R. (1999). Deep mixing method design and execution manual for land works, 92-99.

Charbit, B. (2009) Numerical analysis of laterally loaded lime/cement columns. MSc, KTH Royal
Institute of Technology.

Chen, J., Lee, F. H. and Ng, C. C. (2011). "Statistical analysis for strength variation of deep mixing
columns in Singapore." Geo-Frontiers, Dallas, TX,576-584.

Chen, W. F. and Mizuno, E. (1990). Nonlinear analysis in soil mechanics, theory and implementation,
Elsevier.

Cherepanov, G. P. (1967). "The propagation of cracks in a continuous medium." Journal of Applied


Mathematcis and Mechancis, 31(3), 503-512.

Chew, S. H., Lee, F. H. and Lee, Y. (1997). "Jet Grouting in Singapore Marine Clay." Proc. 3rd Asian
Young Geotechnical Engineers Conference, Singapore,231-238.

Chin, K. G. (2006) Constitutive behavior of cement treated marine clay. Ph.D, National University of
Singapore.

Chong, P. T. (2002) Characterisation of Singapore Lower Marine Clay. Ph.D, National University of
Singapore.

Chuah, S. S. and Tan, S. A. (2010). "Numerical study on a new strut-free counterfort embedded wall in
Singapore." Earth Retention Conference 3, Bellevue, Washington, US,740-747.

Clough, G. W., Sitar, N., Bachus, R. C. and Rad, N. S. (1981). "Gemented sands under static loading."
Journal of Geotechnical Engineering Divison, 107(6), 799-817.

Consoli, N. C., Schnaid, F., Rohlfes Junior, J. A. and Preitto, P. D. M. (1996). "Engineering properties
of residual soil-cement mixtures." Grouting and Deep mixing, The Second International Conference on
Ground Improvement Geosystems, Tokyo,25-30.

Das, B. M. and Dass, R. N. (1995). "Lightly cemented sand in tension and compression." Geotechnical
and Geological Engineering, 13, 169-177.

Dawson, E. M., Roth, W. H. and Drescher, A. (1999). "Slope stability analysis by strength reduction."
Geotechnique, 49(6), 835-840.

Diamond, S. and Kinter, E. B. (1965). "Mechanisms of soil-lime stabilization: an interpretive review.


Highway Research Record." 92, 83-102.

244
Fang, T. S. and Yu, F. J. (1998). "Engineering properties of jet-grouted soilcrete and slime." 2nd
International Conference on Ground Improvement Techniques, Singapore,589-596.

Fang, Y. S., Liao, J. J. and Lin, T. K. (1994). "Mechanical properties of jet grouted soilcrete."
Quarterly Journal of Engineering Geological, 27(3), 257-265.

Frocht, M. M. (1948). Photoelasticity. New York, John Wiley.

Gaba, A. R. (1990). "Jet grouting at newton station, Singapore." Tenth Southeast Asian Geotechnical
Conference, Taipei,77-79.

Goh, T. L. (2003) Stabilisation of an embedded improved soil berm. Ph.D, National University of
Singapore.

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 and Foundations, 31(1), 169-180.

Griffiths, D. V. and Lane, P. A. (1999). "Slope stability analysis by finite elements." Geotechnique,
49(3), 387-403.

Haque, M. A. and Bryant, J. T. (2011). "Failure of VERT wall system: forensic evaluation and lessons
learned." Geo-Frontiers 2011: Advances in Geotechnical Engineering, Dallas, Texas, US

Hashizume, H., Okochi, Y., Dong, J., Horii, N., Toyosawa, Y. and Tamate, S. (1998). "Study on the
behavior of soft ground improved using deep mixing method." Proc. of the International Conference
on Centrifuge 98, 851-856.

Heilmann, H. G., Hilsdorf, H. H. and Finsterwalder, K. (1969). "Festigkeit und Verformung von Beton
unter Zugspannungen." Deutscher Ausschuss fur Stahlbeton, (Heft 203 (in German)).

Hillerborg, A. (1985). "The theoretical basis of a method to determine the fracture energy GF of
concrete." Materials and Structures, 18, 291-296.

Hillerborg, A. (1989). "Stability problems in fracture mechanics testing." International Conference on


Recent Developments in the Fracture of Concrete and Rock, University of Wales, UK,369-378.

Hillerborg, A., Modeer, M. and Petersson, P.-E. (1976). "Analysis of crack formation and crack growth
in concrete by means of fracture mechanics and finite elements." Cement and Concrete Research, 6,
773-782.

Horpibulsuk, S., Liu, M. D., Liyanapathirana, D. S. and Suebsuk, J. (2010). "Behaviour of cemented
clay simulated via the theoretical framework of the structured cam clay model." Computers and
Geotechnics, 37(1), 1-9.

Hosoya, Y., Ogino, T., Nasu, T., Kohata, Y., Hibi, Y. and Makihara, Y. (1997). "JGS TC Report: An
evaluation of the strength of soils improved by DMM." Grouting and Deep mixing, The Second
International Conference on Ground Improvement Geosystems, Tokyo,919-924.

245
Hsi, J. P. and Yu, J. B. Y. (2005). "Jet grout application for excavation in soft marine clay."
Proceedings of the 16th International Conference on Soil Mechanics and Geotechnical Engineering,
Osaka, Balkema,1485-1488.

Hsieh, H. S., Wang, C. C. and Ou, C. Y. (2003). "Use of Jet Grouting to Limit Diaphragm Wall
Displacement of a Deep Excavation." Journal of Geotechnical and Geoenviromental Engineering, 129,
146-157.

Hwang, R. N. and Moh, Z. C. (2008). "Evaluating effectiveness of buttresses and cross walls by
reference envelopes." Journal of GeoEngineering, 3(1), 1-11.

Issa, M. A., Issa, M. A., Islam, M. S. and Chudnovsky, A. (2000b). "Size effects in concrete fracture -
Part II: Analysis of test results." International Journal of Fracture, 102(1), 25-42.

Jankowiak, I., Kakol., W. and Madaj, A. (2005). "Identification of a continuous composite beam
numerical model, based on experimental tests." 7th Conference on Composite Structures, Zielona
Gora,163-178.

Jardine, R. J., Symes, N. J. and Burland, J. B. (1984). "The measurement of soil stiffness in triaxial
apparatus." Geotechnique, 34(3), 323-340.

Josephine, V. C. (2011) Material modelling of reinforced concrete at evelated temperatures. MSc,


Technical University of Denmark.

Kamruzzaman, A. H. M. (2002) Physico-chemical and engineering behaviour of cement treated


Singapore marine clay. Ph.D, National University of Singapore.

Karastanev., D., Kitazume, M., Miyajima, S. and Ikeda, T. (1997). "Bearing capacity of shallow
foundation on column type DMM improved ground." Proc. of the 14th International Conference on
Soil Mechanics and Foundations Engineering, 3, 1621-1624.

Kasama, K., Ochiai, H. and Yasufuku, N. (2000). "On the stress-strain behaviour of lightly cemented
clay based on an extended critical state concept." Soils and Foundations, 40(5), 37-47.

Kawasaki, T., Niina, A., Saitoh, S. and Babasaki, R. (1978). Takenaka Technical Research Report.
Studies on engineering characteristics of cement-base stabilized soil. 19: 144-165.

Kawasaki, T., Niina, A., Saitoh, S., Suzuki, Y. and Honjyo, Y. (1981a). "Deep mixing method using
cement hardening agent." Proceedings of the 10th International Conference on Soil Mechanics and
Foundation Engineering, Stockholm,721-724.

Khoo, K. S., Orihara, K., Egi, F., Arii, T. and Yamamoto, K. (1997). "Excavation with soil berm
improved by jet grout piles." Proceedings of 3YGEC, Singapore,189-196.

Kitazume, M. (2008). "Stability of ground column type DM improved ground under embankment
loading behavior of sheet pile quay wall."Report of the Port and Airport Research Institute. Nagase,
Yokosuka, Japan, Port and Airport Research Institute

246
Kitazume, M., Ikeda, T., Miyajima, S. and Karastanev., D. (1996). "Bearing capacity of improved
ground with column type DMM." Proc. of the 2nd International Conference on Ground Improvement
Geosystems, 1, 503-508.

Kitazume, M. and Maruyama, K. (2007). "Internal stability of group column type deep mixing
improved ground under embankment loading." Soils and Foundations, 47(3), 437-455.

Kitazume, M., Okano, K. and Miyajima, S. (2000). "Centrifuge model tests on failure envelope of
column type deep mixing method improved ground." Soils and Foundations, 40(4), 43-55.

Kivelo, M. (1998) Stabilization of embankments on soft soil with lime/cement columns. Ph.D, Royal
Institute of Technology.

Kmiecik, P. and Kaminski, M. (2011). "Modelling of reinforced concrete structures and composite
structures with concrete strength degradation taken into consideration." Archives of Civil and
Mechanical Engineering, 11(3), 623-636.

Kohata, Y., Muramoto, K., Maekawa, H., Yajima, J. and Babasaki, R. (1997). "JGS TC Report:
Deformation and strength properties of DM cement-treated soils." Grouting and Deep mixing, The
Second International Conference on Ground Improvement Geosystems, Tokyo,905-911.

Koseki, J., Sato, T. and Nishimoto, T. (2005). "Anisotropy in strength characteristics of cement-treated
sand prepared by compaction." Proc. Symp. Cement-treated Soils, JGS,365-368.

Kupfer, H., Hilsdorf, H. K. and Rusch, H. (1969). "Behavior of concrete under biaxial stresses." ACI
Journal, 66(8), 656-666.

Larsson, S. (1999). "Shear box apparatus for modeling chemical stabilized soil - Introductory tests."
Proceedings of the International Conference on Deep Mixing, Best Practice and Recent Advances,
Stockholm,732-785.

Larsson, S., Dahlstrom, M. and Nilsson, B. (2005a). "Uniformity of lime-cement columns for deep
mixing: a field study." Ground Improvement, 9(1), 1-15.

Larsson, S., Dahlstrom, M. and Nilsson, B. (2005b). "A complementary field study on the uniformity
of lime-cement columns for deep mixing." Ground Improvement, 9(2), 67-77.

Larsson, S., Malm, R., Charbit, B. and Ansell, A. (2012). "Finite element modelling of laterally loaded
lime-cement columns using a damage plasticity model." Computers and Geotechnics, 44, 48-57.

Lee, F. H. (2008). "How useful is numerical analysis in geotechnical engineering?" International


Conference on Deep Excavations (ICDE), Singapore

Lee, F. H., Lee, Y., Chew, S. H. and Yong, K. Y. (2005). "Strength and modulus of marine clay-
cement mixes." Journal of Geotechnical and Geoenviromental Engineering, 131(2), 178-186.

247
Lee, J. and Fenves, G. L. (1998). "Plastic-damage model for cyclic loading of concrete structures."
Journal of Engineering Mechanics, 124(8), 892-900.

Lee, K. H., Chan, D. and Lam, K. C. (2004). "Consitutive model for cement treated clay in a critical
state frame work." Soils and Foundations, 44(3), 69-77.

Lee, Y., Yogarajah, I. and Lee, F. H. (1998). "Engineering properties of grouted Marine Clay." 2nd
International Conference on Ground Improvement Techniques, Singapore,267-275.

Li, V. C., Chan, C.-M. and Leung, C. K. Y. (1987). "Experimental determination of the tension-
softening relations for cementitious composites." Cement and Concrete Research, 17, 441-452.

Li, V. C. and Ward, R. J. (1989). "A novel testing technique for post-peak tensile behavior of
cementitious materials." Fracture Toughness and Fracture Energy,183-195.

Liao, H. J. and Tsai, T. L. (1993). "Passive resistance of partially improved soft clayey soil." 11th
Southeast Asian Geotechnical Conference, Singapore,751-756.

Liu, M. D. and Carter, J. P. (2002). "A structured cam clay model." Canadian Geotechnical Journal,
39, 1313-1332.

Lorenzo, G. A. and Bergado, D. T. (2006). "Fundamental characteristics of cement-admixed clay in


deep mixing." Journal of Material in Civil Engineering ASCE, 18(2), 161-174.

Low, H. E. (2004) Compressibility and Undrained Behavior of Natural Singapore Marine Clay: Effect
of Soil Structure. M.Eng, National University of Singapore.

Lubliner, J., Oliver, J., Oller, S. and Onate, E. (1989). "A plastic-damage model for concrete."
International Journal of Solids and Structures, 229-326.

Matsui, T. and San, K. C. (1992). "Finite element slope stability analysis by shear strength reduction
technique." Soils and Foundations, 32(1), 59-70.

Miyake, M., Akamoto, H. and Wada, M. (1991). "Deformation characteristics of ground improved by a
group of treated soil." Proc. of the International Conference on Centrifuge 91, 295-302.

Morley, A. (1944). Strength of materials. London, Longman.

Namikawa, T. (2006) Study on tensile and shear failure characteristics of cement-treated sand. Ph.D,
University of Tokyo.

Namikawa, T. and Koseki, J. (2006). "Experimental determination of softening relations for cement-
treated sand." Soils and Foundations, 46(4), 491-504.

Namikawa, T. and Koseki, J. (2007). "Evaluation of tensile strength of cement-treated sand based on
several types of laboratory tests." Soils and Foundations, 47(4), 657-674.

248
Namikawa, T., Koseki, J. and Suzuki, Y. (2008). "Finite element analysis of a full scale bending test of
cement treated soil column." The 12th International Conference of IACMAG, Goa, India,3635-3641.

Namikawa, T. and Mihira, S. (2007). "Elasto-plastic model for cement-treated sand." International
Journal for Numerical and Analytical Methods in Geomechanics, 31, 71-107.

Navin, M. P. (2005) Stability of Embankments Founded on Soft Soil Improved with Deep-Mixing-
Method Columns. Ph.D, Virginia Polytechnic Institute and State University.

Nayak, G. C. and Zienkiewicz, O. C. (1972). "Convenient form of stress invariants for plasticity."
Journal of the Structural Division. ASCE, 98, 949-954.

Ni, Q. (2005) Engineering Properties of Singapore Old Alluvium. Ph.D, National University of
Singapore.

Nicholson, P. J., Mitchell, J. K., Bahner, E. W. and Moriwaki, Y. (1998). "Design of a soil mixed
composite gravity wall." Soil Improvement for Big Digs, Geotechnical Special Publication, ASCE,
Boston, ASCE,27-40.

Niwa, J., Sumranwanich, T. and Tangtermsirikul, S. (1998b). "New method to determine tension
softening curve of concrete." Proceedings FRAMCOS-3, 347-356.

O'Rourke, T. D. and O'Donnell, C. J. (1997). "Field behavior of excavation stabilized by deep soil
mixing." Journal of Geotehnical and Geoenvironmental Engineering, 123(6), 516-524.

Omine, K., Ochiai, H. and Bolton, M. D. (1999). "Homogenization method for numerical analysis of
improved ground with cement-treated soil columns." Proceedings of International Conference on Dry
Mix Methods for Deep Soil Stabilization, Stockholm, Balkema,161-168.

Ou, C. Y. and Wu, T. S. (1996). "Evaluation of material properties for soil improvement in
excavation." Grouting and Deep Mixing, Rotterdam,545-550.

Ou, C. Y., Wu, T. S. and Hsieh, H. S. (1996). "Analysis of deep excavation with column type of
ground improvement in soft clay." Journal of Geotechnical Engineering, 122(9), 709-716.

Pankaj, P. (1990) Finite element analysis in strain softening and localisation problems. Ph.D,
University of Roorkee.

Petersson, P.-E. (1981). "Crack growth and development of fracture zones in plain concrete and similar
materials."Report TVBM-1006. Division of Building Materials, Lund Institute of Technology, Lund,
Sweden

Pitts, J. (1992). Landforms and Geomorphic Evolution of the Islands during the Quaternary. Singapore,
Singapore University Press.

249
Porbaha, A., Shibuya, S. and Kishida, T. (2000). "State of the art in deep mixing technology. Part III:
geomaterial characterization." Ground Improvement, 4(3), 91-110.

Racansky, V., Schweiger, H. F. and Thurner, R. (2008). "FE-Analysis of the Behaviour of Buttressed
Jet Grouted Retaining Walls." The 12th International Conference of IACMAG, India,3984-3992.

Rice, J. R. (1968). "A path independent integral and the approximate analysis of strain concentration by
notches and cracks." journal of Applied Mechanics, 35, 379-386.

RILEM (1985). "Determination of the fracture energy of mortar and concrete by means of three-point
bend tests on notched beams." Materials and Structures, 18, 285-290.

Rocco, C., Guinea, G. V., Planas, J. and Elices, M. (1999). "Mechanisms of rupture in splitting tests."
ACI Material Journal, 96(1), 52-60.

Rokugo, K., Iwasa, M., Seko, S. and Koyanagi, W. (1989). "Tension softening diagrams of steel fiber
reinforced concrete." Fracture of Concrete and Rock,513-522.

Ruffing, D. G., Sheleheda, M. J. and Schindler, R. M. (2012). "A case study: unreinforced soil mixing
for excavation support and bearing capacity improvement." Grouting and deep mixing: 4th
International Conference on Grouting and Deep Mixing, New Orleans, LA

Rusinowski, P. (2005) Two-way concrete slabs with openings: experiments, finite element analysis and
design. Msc, Lulea University of Technology.

Ryan, C. R. and Jasperse, B. H. (1989). "Deep soil mixing at the Jackson lake dam." ASCE
Geotechnical and Construction Divisions Special Conference

Saitoh, S., Kawasaki, T., Nina, S., Babasaki, R. and Miyata, T. (1980). "Research on DMM using
cementitious agents (part 10) - engineering properties of treated soils." Proc. of the 15th National
Conference of the JSSMFE, 717-720 (in Japanese).

Saitoh, S., Suzuki, Y., Nishioka, S. and Okumura, R. (1996). Required strength of cement improved
ground, Balkema.

SGF (2000). "Lime and lime cement columns."Guide for design, construction and control. Report
2:2000, Linkoping, Swedish Geotechnical Society.(Swedish)

Shen, S. L. (1998) Behaviour of deep mixing columns in soft clay ground. Ph.D, Saga University.

Shibuya, S., Tatsuoka, F., Teachavorasinskun, S., J., K. X., Abe, F., Kim, Y. S. and Park, C. S. (1992).
"Elastic deformation properties of geomaterials." Soils and Foundations, 32(3), 26-46.

Shirlaw, J. N., Tan, T. S. and Wong, K. S. (2005). "Deep excavations in Singapore Marine Clay."
Geotechnical Aspects of Underground Construction in Soft Ground, Proceedings of the 5th
International Symposium TC28, Amsterdam, the Netherlands,13-28.

250
Snitbhan, N. and Chen, W. F. (1978). "Elastic-plastic large deformation analysis of soil slopes."
Computers and Structures, 9, 567-577.

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.

Swan, C. C. and Seo, Y. K. (1999). "Limit state analysis of earthen slopes using dual continuum/FEM
approaches." International Journal for Numerical and Analytical Methods in Geomechanics, 23, 1359-
1371.

Tan, S. L. (1983). "Geotechnical properties and laboratory testing of soft soils in Singapore." Proc. 1st
Int. Seminar on Construction Problems in Soft Soils, Singapore,TSL1-47.

Tan, T. S., Goh, T. L. and Yong, K. Y. (2002). "Properties of Singapore marine clays improved by
cement mixing." Geotechnical Testing Journal, 25(4), 1-12.

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." Characterisation and engineering properies of natural
soils,429-454.

Tanaka, H. (1994). "Behaviour of braced excavation in soft clay and the undrained shear strength for
passive earth pressure." Soils and Foundations, 34(1), 53-64.

Tanaka, H. and Terashi, M. (1986). "Engineering properties of treated soil by DMM in field."Report of
Port and Harbour Research Institute.(Japanese)

Tatsuoka, F. and Kobayashi, A. (1983). "Triaxial strength characteristics of cement treated soft clay."
Proc. of the 8th European Regional Conference on Soil Mechanics and Foundation Engineering, 1,
421-426.

Tatsuoka, F., Kohata, Y., Uchida, K. and Imai, K. (1996). "Deformation and strength characteristics of
cement-treated soils in Trans-Tokyo Bay Highway project." Grouting and Deep mixing, The Second
International Conference on Ground Improvement Geosystems, Tokyo,453-459.

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, 1(1), 37-57.

Tavenas, F., jean, P., Leblond, P. and Leroueil, S. (1986). "In situ measurement of permeability in soft
clays." ASCE Special Conference on Use of In Situ Test in Geotechnical Engineering,
Blacksburg,1034-1048.

Tjahyono, S. (2011) Experimental and numerical modelling of spudcan penetration in stiff clay
overlying soft clay. Ph.D, National University of Singapore.

Topolnicki, M. (2004). In situ soil mixing. Ground Improvement 2nd Edition. M. P. M. a. K. Kirsch,
Spon Press: 331-428.

251
Tsuzuki, M., Nakajima, J., Takayama, S. and Kanekura, T. (2000). "Behavior of cantilever earth
retaining with buttress type stabilisation." Geotechnical Aspects of Underground Construction in Soft
Ground,593-598.

Uchida, Y., Rokugo, K. and Koyanagi, W. (1991). "Determination of tension softening diagrams of
concrete by means of bending tests." Journal of Materials, Concrete Structures and Pavements of
JSCE, 426, 203-212.

Uchiyama, N. and Kamon, M. (1998). "Deformations of buttress type ground improvement and
untreated soil during strutted excavation in clay." 2nd International Conference on Ground
Improvement Techniques, Singapore,505-512.

Uddin, K., Balasubramaniam, A. S. and Bergado, D. T. (1997). "Engineering behavior of cement-


treated Bangkok soft clay." Geotechnical Engineering, SEAGS, 28(1), 89-119.

Vogler, U. and Karstunen, M. (2009). Applicaiton of volume averaging technique in numerical


modeling of deep mixing. London, Taylor & Francis Group.

Wahalathantri, B. L., Thambiratnam, D. P., Chan, T. H. T. and Fawzia, S. (2011). "A material model
for flexural crack simulation in reinforced concrete elements using ABAQUS." In Proceedings of the
First International Conference on Engineering, Designing and Developing the Built Environment for
Sustainable Wellbeing, Brisbane,260-264.

Wen, D. (2005). Use of jet grouting in deep excavations, Elsevier.

Wong, K. S. and Goh, A. T. C. (2006). "Modelling jgp slab in deep excavation analysis." International
Conference on Deep Excavations, Singapore

Wong, K. S. and Goh, A. T. C. (2006). "Modelling JGP Slab in Deep Excavation Analysis."
International Conference on Deep Excavation, Singapore

Xiao, H. W. (2009) Yielding and failure of cement treated soil. Ph.D, National University of Singapore.

Yang, H. B. (2009) Mobilised mass properties of embedded improved soil raft in an excavation. Ph.D,
National University of Singapore.

Yin, J. H. and Lai, C. K. (1998). "Strength and stiffness of Hong Kong marine deposits mixed with
cement." Geotechnical Engineering, SEAGS, 29(1), 29-44.

Yong, K. Y., Karunaratne, G. P. and Lee, S. L. (1990). "Recent devlopments in soft clay engineering in
Singapore." Proc. Kansai Int. Geotech. Forum '90: Comparative Geotechnical Engineering, Osaka,
Japan,1-8.

Yu, M., Xia, G. and Kolupaev, V. A. (2009). "Basic characteristics and development of yield criteria
for geomaterials." Journal of Rock Mechanics and Geotechnial Engineering, 1(1), 71-88.

252
Yu, Y., Pu, J., Ugai, K. and Hara, T. (1999). "A study on the perneability of soil-cement mixture." Soils
and Foundations, 39(5), 145-149.

Zhang, Y. D. (2004) An embedded improved soil berm in an excavation-mechanisms and capacity.


Ph.D, National University of Singapore.

Zheng, H., Tham, L. G. and Liu, D. F. (2006). "On two definitions of the factor of safety commonly
used in the finite element slope stability analysis." Computers and Geotechnics, 33, 188-195.

Zheng, S. Y. and Zhao, Y. R. (2004). "Application of strength reduction FEM in soil and rock slope."
Chinese Journal Rock Mech Eng, 23(19), 3381-3388.

Ziaie Moayed, R., Tamassoki, S. and Izadi, E. (2012). "Numerical modeling of direct shear tests on
sandy clay." International Journal of Civil and Environmental Engineering, 6, 106-110.

253

You might also like