You are on page 1of 188

BEHAVIOUR OF SOLDIER PILES AND TIMBER LAGGING

SUPPORT SYSTEMS

HONG SZE HAN


B.Eng.(Hons.), NUS

A THESIS SUBMITTED

FOR THE DEGREE OF DOCTOR OF PHILOSOPHY

DEPARTMENT OF CIVIL ENGINEERING

NATIONAL UNIVERSITY OF SINGAPORE

2002
...Let us run with perseverance the race that is set before us, looking to Jesus the
pioneer and perfecter of our faith...

- Hebrews 12:1

ii
ABSTRACT

ABSTRACT

This thesis discusses the significance of three-dimensional (3D) effects arising

from arching of soil behind retaining wall on wall deflection and ground surface

settlement. Results from literature survey are not applicable to soldier piles retaining

system owing to the added complexities of soil arching, movement between piles and

anisotropy of lagging. A customised version of the CRISP 90 finite element program,

with additional beams and reduced-integration elements incorporated, was used in the

back-analysis. This is necessary in order to account for the local three-dimensional

effects arising from the interaction of the soldier piles and the intervening soils.

Strutted excavations, including soldier-piled excavations, are often analysed

using two-dimensional (2D) finite element (FE) analyses with properties that are

averaged over a certain span of the wall. In this thesis, the effects of “smearing” the

stiffness of the soldier piles and timber laggings into an equivalent uniform stiffness

are examined, based on comparison between the results of 2D and 3D analyses. The

ability of the 3D analyses to model the flexural behaviour of the soldier piles and

timber laggings is established by comparing the flexural behaviour of various FE beam

representations to the corresponding theoretical solutions, followed by a reality check

with an actual case study. Finally, the results of 2D and 3D analyses on an idealized

soldier piled excavation are compared. The findings show that modelling errors can

arise in several ways. Firstly, a 2D analysis tends to over-represent the coupling to pile

to the soil below excavation level. Secondly, the deflection of the timber lagging,

which is usually larger than that of the soldier piles, is often underestimated. For these

reasons, the overall volume of ground loss is, in reality, larger than those given by a

2D analysis. Thirdly, a 2D analysis cannot replicate the swelling, and therefore


iii
ABSTRACT

softening, of the soil face just behind the timber lagging. Increasing the inter-pile

spacing will tend to accentuate the effects of these modelling errors.

Some results of a comparative study into the effects of different constitutive

models in simulating soldier-piled excavations in medium to stiff soils will be

discussed. The field data, which were used as reference in this comparative study,

came from the excavation for the Serangoon station of the North-East Line (NEL)

contract 704. The subsoil at the site is generally residual Bukit Timah Granite. The

temporary retaining structures for the excavation consisted of driven soldier piles and

timber lagging supported by steel struts. The methods of modelling retaining system,

boundary and groundwater conditions and excavation sequence are demonstrated. The

models compared in this thesis include Mohr Coulomb, modified Cam-clay and

hyperbolic Cam-clay. The results of the study indicates that deflection of the soldier

pile may not necessarily be an accurate reflection of the soil face movement behind the

timber lagging, and thereby the ground loss. Furthermore, modelling the soil with a

hyperbolic Cam-clay model results in smaller initial deflection and better agreement

with progressive field measurement compared to the linear elastic Mohr Coulomb

criterion or modified Cam-clay models.

Keywords: deep excavation, finite element method, 3D, soldier piles, piles spacing,
timber lagging, arching effect, Mohr Coulomb, modified Cam-clay,
hyperbolic Cam-clay, beam element
iv
ACKNOWLEDGEMENT

ACKNOWLEDGEMENTS

The Author is especially grateful to Professor Yong Kwet Yew for introducing

him to this fascinating and active field of research and to Associate Professor Lee Fook

Hou for his many most valuable and critical comments during their long discussions.

The Author would like to acknowledge the research scholarship and scholarship

augmentation provided by the National University of Singapore and Econ Piling Pte.

Ltd. respectively. Thanks are due to Econ Piling Pte. Ltd. for their provision of site

instrumentation and monitoring data. The equipment and software support provided

by the Centre for Soft Ground Engineering at the National University of Singapore is

also gratefully acknowledged.

The Author also remembers most stimulating and valuable discussions with

Chan Swee Huat, Chee Kay Hyang, Gu Qian, How You Chuan, Lim Ken Chai, Wong

Kwok Yong and numerous other people which he would like to acknowledge

collectively in order not to forget anyone. Most of all, he would like to thank all of his

family for their love and understanding throughout his entire life and for their

continuing belief in him.

v
TABLE OF CONTENTS

TABLE OF CONTENTS

ABSTRACT iii

ACKNOWLEDGEMENTS v

TABLE OF CONTENTS vi

LIST OF TABLES viii

LIST OF FIGURES ix

LIST OF SYMBOLS xvi

CHAPTER 1 INTRODUCTION 1

1.1 BACKGROUND 1
1.2 SOLDIER PILE WALLS 2
1.3 DEFORMATIONS 4
1.4 OBJECTIVE 5
1.5 SCOPE OF WORK 5

CHAPTER 2 LITERATURE REVIEW 8

2.1 DESIGN APPROACHES FOR RETAINING SYSTEMS 8


2.2 VARIOUS METHODS FOR EVALUATING RETAINING SYSTEMS 9
2.2.1 EMPIRICAL APPROACHES 9
2.2.2 1-D BEAM AND SPRING MODELS 17
2.2.3 2D FINITE ELEMENT ANALYSIS 18
2.2.4 3D FINITE ELEMENT ANALYSIS 24
2.3 SOIL ARCHING 31
2.3.1 DESIGN PROCEDURES FOR SOLDIER PILE WALL 31
2.3.2 RESEARCH ON ARCHING EFFECT 32
2.4 REVIEW SYNOPSIS 35

CHAPTER 3 IDEALISED ANALYSIS OF PILE-SOIL INTERACTION 37

3.1 MODELLING OF SOLDIER PILES IN 2D AND 3D ANALYSES 38


3.2 COMPARISON WITH CASE HISTORY 47

vi
TABLE OF CONTENTS

3.3 3D AND 2D COMPARISONS 54


3.3.1 IDEALISED EXCAVATION 54
3.3.2 WALL DEFLECTION 56
3.3.3 STRESS CHANGES 59
3.4 SYNOPSIS ON 3D AND 2D ANALYSES 65

CHAPTER 4 GEOTECHNICAL STUDY OF SERANGOON MRT SITE 66

4.1 SERANGOON MRT SITE GEOLOGY 66


4.2 RETAINING SYSTEM 76
4.3 CONSTRUCTION SEQUENCE 78

CHAPTER 5 FEM STUDY OF SERANGOON MRT SITE 81

5.1 GEOMETRY OF FINITE ELEMENT MODEL 81


5.2 MODELLING OF EXCAVATION SEQUENCE 84
5.3 PARAMETRIC STUDIES FOR VARIOUS MODELS 85
5.3.1 ELASTIC-PERFECTLY PLASTIC MOHR COULOMB 86
5.3.2 MODIFIED CAM-CLAY 103
5.3.3 HYPERBOLIC CAM-CLAY 123
5.4 SUMMARY OF FINDINGS 142

CHAPTER 6 CONCLUSION AND RECOMMENDATIONS 144

6.1 CONCLUDING REMARKS 144


6.2 RECOMMENDATIONS FOR FUTURE WORK 145

REFERENCES 147

APPENDICES 159

APPENDIX A TRANSFORMATION MATRIX FOR 3D BEAM 159


APPENDIX B EVALUATING BENDING MOMENTS FROM LINEAR-STRAIN
BRICK ELEMENTS 163
APPENDIX C STRESS REVERSAL IN NON-LINEAR SOIL MODEL 166
APPENDIX D PROPERTIES OF TIMBER LAGGING (CHUDNOFF, M., 1984) 168

vii
LIST OF TABLES

LIST OF TABLES

Table 2.1: Empirical approaches contribution by various researchers. 10

Table 2.2: 2D finite element analyses contribution by various researchers. 19

Table 2.3: 3D finite element analyses contribution by various researchers. 25

Table 3.1: Idealised soil and soldier pile properties for beam and brick soldier piles study. 45

Table 3.2: Idealised soil and soldier pile properties for O’Rourke’s case study. 49

Table 3.3: Idealised soil and soldier pile properties for idealised study. 55

Table 3.4: Summary of effective stress analyses for the idealised excavation at OCR = 3 and

×10-8m/s.
k = 1× 55

Table 4.1: Soil properties of Serangoon MRT site. 75

Table 4.2: Struts details for Serangoon MRT retaining system. 76

Table 5.1: Soil properties of Serangoon MRT site. 86

Table 5.2: Soil properties for Layers 1, 2 and 3 for Mohr Coulomb analyses. 89

Table 5.3: Comparison of strut forces at various levels between daily average instrumented

and FEM result for MC2, MC3, MC4 and MC5. 100

Table 5.4: Soil properties for Layer 1 and Layer 2 for modified Cam-clay. 109

Table 5.5: Comparison of strut forces at various levels between instrumented and FEM result

for the best-fit deflection profile for MCC1. 121

Table 5.6: Modification to 3.5m of topsoil to Mohr Coulomb properties. 122

Table 5.7: Soil properties for Layer 1 and Layer 2 for hyperbolic Cam-clay. 129

Table 5.8: Comparison of strut forces at various levels between instrumented and FEM result

for the best-fit deflection profile for HCC4. 135

viii
LIST OF FIGURES

LIST OF FIGURES

Figure 1.1: Components of soldier piles and timber lagging systems. 2

Figure 2.1: Observed settlements behind excavations (after Peck, 1969). 10

Figure 2.2: Summary of settlements adjacent to strutted excavation in Washington, D.C. (after

O’Rourke et al., 1976). 11

Figure 2.3: Summary of settlements adjacent to strutted excavation in Chicago (after

O’Rourke et al., 1976). 12

Figure 2.4: Effects of wall stiffness and support spacing on lateral wall movements (after

Goldberg et al., 1976). 13

Figure 2.5: Calculation of embedment depth of sheet pile wall in relatively uniform competent

soil conditions (after Goldberg et al., 1976) 14

Figure 2.6: Relationship between maximum settlement and stability number for different

batters (after Clough & Denby, 1977). 15

Figure 2.7: Relationship between factor of safety against basal heave and non-dimensional

maximum lateral wall movement for case history data (after Clough et al., 1979). 16

Figure 2.8: Relationship between maximum ground settlements and maximum lateral wall

movements for case history data (after Mana & Clough, 1981). 17

Figure 3.1: Comparison of deflection profiles under distributed load for 2D 43m-cantilever

beam. (a) 32-bit precision. (b) 64-bit precision. 41

Figure 3.2: Comparison of bending moments under distributed load for 2D 43m-cantilever

beam. (a) 32-bit precision. (b) 64-bit precision. 41

Figure 3.3: Effect of aspect ratios for 2D 43m-cantilever beam in 64-bit precision using

RIQUAD elements. (a) Deflection. (b) Bending moment. 42

Figure 3.4: Comparison of deflection profiles under distributed load for 3D 43m-cantilever

beam. (a) 32-bit precision. (b) 64-bit precision. 42

Figure 3.5: Comparison of bending moments under distributed load for 3D 43m-cantilever

beam. (a) 32-bit precision. (b) 64-bit precision. 43

Figure 3.6: Locations of beams and linear strain bricks for modelling soldier piles in the

finite element model of an idealistic excavation. (a) Plan view of beam element

as soldier pile in section A-A. (b) Plan view of RIBRICK element as soldier pile

ix
LIST OF FIGURES

in section A-A. (c) Cross-sectional view in x-y plane of the FEM model. (d) Soil

layers, struts and excavation levels. 44

Figure 3.7: (a) Deflection profile of soldier pile and soil at the center between adjacent piles.

(b) Bending moment profile of soldier piles. 46

Figure 3.8: Plan view of lateral deflection retaining system 9m depth after final excavation

stage. 46

Figure 3.9: Lateral stress contours for at 9.3m below ground level for excavation to 11.5m.

(a) Beam soldier pile elements. (b) RIBRICK soldier pile elements. 47

Figure 3.10: Plan view of retaining system after O’Rourke (1975). 48

Figure 3.11: Finite element mesh used for reality check with O’Rourke (1975). 48

Figure 3.12: Comparison of a series of 2D and 3D analyses from O’Rourke (1975). (a) Soldier

pile deflection. (b) Bending moments. 50

Figure 3.13: Deflection of soldier pile and soil at mid-span with respect to excavation sequence.

(a) Excavation depth at 6.1m. (b) Excavation depth at 11.9m. (c) Excavation

depth at 18.3m. 51

Figure 3.14: Effect of ‘smearing’ on rotational fixity below excavation level. 52

Figure 3.15: Modelling the S-shaped deflection profile of the soldier pile in 3D. 53

Figure 3.16: Construction sequence for idealized excavation from stage A through G. 56

Figure 3.17: Comparison of wall deflection between 2D and different pile spacing for cases

without preloading at final stage of construction. 57

Figure 3.18: Comparison of wall deflection between 2D and different pile spacing for cases

with preloading at final stage of construction. 58

Figure 3.19: Comparison of effects of preloading on wall deflection for 3D-4 at 4.5m below

ground level with stages B, C and D shown in Figure 3.16. 58

Figure 3.20: Location of stress paths plotted for soil behind first level of strut at 4.22m below

ground level in x-z plane. 59

Figure 3.21: Stress path plot for 2D-1 without preload. Stages A to G are shown in Figure 3.16.

Units of stresses in kPa. 60

Figure 3.22: Stress path plot for 2D-2 with preload. Stages A to G are shown in Figure 3.16.

Units of stresses in kPa. 61

x
LIST OF FIGURES

Figure 3.23: Stress path plots for 3D-1 and 3D-2 at point I without preload. Stages A to G are

shown in Figure 3.16. Units of stresses in kPa. 62

Figure 3.24: Stress path plots for 3D-1 and 3D-2 at point II without preload. Stages A to G are

shown in Figure 3.16. Units of stresses in kPa. 63

Figure 3.25: Stress paths plots for 3D-3 and 3D-4 at point I with preload. Stages A to G are

shown in Figure 3.16. Units of stresses in kPa. 63

Figure 3.26: Stress path plots for 3D-3 and 3D-4 at point II with preload. Stages A to G are

shown in Figure 3.16. Units of stresses in kPa. 64

Figure 4.1: Singapore north-east MRT line with location plan. 66

Figure 4.2: Soil profile of Serangoon MRT site at section A-A. 69

Figure 4.3: Soil profile of Serangoon MRT site at section B-B. 69

Figure 4.4: Plan view of Serangoon MRT station. 70

Figure 4.5: Variation of physical properties at Serangoon MRT site. (a) Bulk unit weight.

(b) Atterberg Limits. 71

Figure 4.6: Plasticity chart of G4 soil at Serangoon MRT station. 71

Figure 4.7: Grain size distribution of G4 soil at Serangoon MRT station. 72

Figure 4.8: Variation of soil properties with depth at Serangoon MRT site. (a) SPT values.

(b) Undrained Young’s modulus. (c) Undrained shear strength. (d) Effective

cohesion. (e) Effective friction angle. (f) Over-consolidation ratio. 74

Figure 4.9: Variation of soil permeability with depth at Serangoon MRT site. 75

Figure 4.10: Soldier pile and timber lagging retaining system of Serangoon station. 77

Figure 4.11: Plan view of retaining system for Serangoon MRT station. 77

Figure 4.12: Cross sectional view of retaining system for Serangoon MRT station. 78

Figure 5.1: Finite element model for Serangoon MRT site excavation. 82

Figure 5.2: Plan view of Serangoon MRT station indicating the section B-B’ modelled. 83

Figure 5.3: Excavation with strut and timber lagging installation sequence. 85

Figure 5.4: Assumed piecewise linear variation of E’ with depth for different soil layers. 87

Figure 5.5: Drained Poisson’s ratio versus plasticity index for several lightly overconsolidated

soils after Wroth (1975). 88

Figure 5.6: Plan view of FEM model of retaining system. 89

xi
LIST OF FIGURES

Figure 5.7: Final deflection profile of soldier pile and timber lagging for MC1, MC2. Locations

I and II for each case are defined in Figure 5.6. 90

Figure 5.8: Final deflection profile of soldier pile and timber lagging for MC2, MC3 and ELAS.

Locations I and II for each case are defined in Figure 5.6. 91

Figure 5.9: Apparent Mohr Coulomb yield envelopes at the limits of the effective stress range

of interest. (a) At 6m below EGL. (b) At 12m below EGL. 92

Figure 5.10: Final deflection profile of soldier pile and timber lagging for MC2, MC4 and MC5.

Locations I and II for each case are defined in Figure 5.6. 93

Figure 5.11: Stress path plots for MC2, MC4 and MC5 at 12.3m below ground. (a) Location I.

(b) Location II. Locations I and II for each case are defined in Figure 5.6. Units

of stresses in kPa. 94

Figure 5.12: 2D final deflection profile of soldier pile and timber lagging for MC2, MC4 and

MC5. 95

Figure 5.13: Final deflection profile of soldier pile and timber lagging for MC2, MC6 and MC7.

Locations I and II for each case are defined in Figure 5.6. 96

Figure 5.14: Stress path plots for MC2, MC6 and MC7 at 12.3m below ground. (a) Location I.

(b) Location II. Locations I and II for each case are defined in Figure 5.6. Units

of stresses in kPa. 98

Figure 5.15: Deflection profiles at various stages of excavation for MC5. (a) Stage B. (b) Stage

C. (c) Stage D. (d) Stage E. (e) Stage F. (f) Stage G. (g) Stage H. (h) Stage I. 101

Figure 5.16: Deflection profile of soldier pile for MC5. (a) Before installation of first level

struts. (b) After installation of forth level struts. 102

Figure 5.17: Ground settlement behind Location I of the retaining wall for MC2, MC3, MC4

and MC5. 103

Figure 5.18: Critical state friction angle estimated from triaxial compression tests. (a) At 6m

below EGL. (b) At 12m below EGL. 105

Figure 5.19: Variation of deformability indices with depth at Serangoon MRT site. (a)

Compression index. (b) Swelling index. 106

Figure 5.20: Comparison of laboratory’s odeometer test with MCC soil models. 107

xii
LIST OF FIGURES

Figure 5.21: Axisymmetric FEM undrained (CU) triaxial test model with 9-integration points

eight noded quadrilateral elements. (a) Plane view. (b) Isometric view. 109

Figure 5.22: Comparison of laboratory’s CU triaxial tests with MCC soil models. (a) Depth of

sample is 11m. (b) Depth of sample is 20m. 110

Figure 5.23: Final deflection profile of soldier pile and timber lagging for MCC1 and MCC2. 111

Figure 5.24: Yield zones behind retaining wall at 12.3m below ground at instance when first

yield was detected in MC5 (Between stage C and D of Figure 5.3). (a) MC2.

(b) MC4. (c) MC5. 112

Figure 5.25: Stress path plots for MCC1 and MCC2 at 12.3m below ground. (a) Location I.

(b) Location II. Locations I and II for each case are defined in Figure 5.6. Units

of stresses in kPa. 114

Figure 5.26: Final deflection profile of soldier pile and timber lagging for MCC1, MCC3 and

MCC4. 115

Figure 5.27: Final deflection profile of soldier pile and timber lagging for MCC4 and MCC5. 116

Figure 5.28: Final deflection profile of soldier pile and timber lagging for MCC1, MCC6 and

MCC7. 117

Figure 5.29: Stress path plots for MCC1, MCC6 and MCC7 at 12.3m below ground. (a) Location

I. (b) Location II. Locations I and II for each case are defined in Figure 5.6.

Units of stresses in kPa. 118

Figure 5.30: Deflection profiles at various stages of excavation for MCC1-M. (a) Stage B.

(b) Stage C. (c) Stage D. (d) Stage E. (e) Stage F. (f) Stage G. (g) Stage H.

(h) Stage I. 120

Figure 5.31: Deflection profile of soldier pile for MCC1. (a) Before installation of first level

struts. (b) After installation of forth level struts. 121

Figure 5.32: Ground settlement behind Location I of the retaining wall for MCC1, MCC2,

MCC3, MCC4 and MCC5. 122

Figure 5.33: 2D and 3D ground settlement behind Location I of the retaining wall for

MCC1-M. 123

Figure 5.34: Parameters for G0. (a) Coefficient m. (b) Coefficient n. (Viggiani & Atkinson,

1995) 126

xiii
LIST OF FIGURES

Figure 5.35: Variation of K’ and G with volumetric strain and deviator strain respectively in a

simulated FEM drained triaxial test. 128

Figure 5.36: Variation of K’ and G with mean effective stress, p’, and deviator stress, q,

respectively in a simulated FEM drained triaxial test. 128

Figure 5.37: Comparison of laboratory’s triaxial tests with HCC soil models. (a) Depth of

sample is 11m. (b) Depth of sample is 20m. 129

Figure 5.38: Final deflection profile of soldier pile and timber lagging for HCC1 and HCC2. 130

Figure 5.39: Final deflection profile of soldier pile and timber lagging for HCC1 and HCC3. 131

Figure 5.40: Final deflection profile of soldier pile and timber lagging for HCC1 and HCC4. 132

Figure 5.41: Deflection profile of soldier pile for HCC4. (a) Before installation of first level

struts. (b) After installation of forth level struts. 133

Figure 5.42: Deflection profiles at various stages of excavation for HCC4-M. (a) Stage B.

(b) Stage C. (c) Stage D. (d) Stage E. (e) Stage F. (f) Stage G. (g) Stage H.

(h) Stage I. 134

Figure 5.43: Ground settlement behind Location I of the retaining wall for HCC1, HCC2,

HCC3 and HCC4. 136

Figure 5.44: Normalised ground settlement behind Location I of the retaining wall for MCC1,

MCC2, MCC3, MCC4 and MCC5. 136

Figure 5.45: Normalised ground settlement behind Location I of the retaining wall for HCC1,

HCC2, HCC3 and HCC4. 137

Figure 5.46: 2D vs 3D wall deflections behind Location I of the retaining wall for MCC1 and

HCC4. 138

Figure 5.47: 2D vs 3D ground settlements behind Location I of the retaining wall for MCC1

and HCC4. 138

Figure 5.48: Comparison of effect of kingpost modelling on 2D and 3D soldier pile deflection

for HCC4-M. 139

Figure 5.49: Comparison of effect of kingpost modelling on 2D and 3D soldier pile deflection

for MCC1-M. 140

Figure 5.50: 2D and 3D displacement vectors behind Location I of the retaining wall for

HCC4-M. 141

xiv
LIST OF FIGURES

Figure 5.51: 2D and 3D ground settlement behind Location I of the retaining wall for

HCC4-M. 141

Figure A-1: Rotation transformation of axes for a 3D beam element 162

Figure B-1: A full-integration 8-noded quadrilateral element forming part of a beam. 163

Figure B-2: A full-integration 20-noded linear strain brick element forming part of a beam. 164

Figure C-1: Comparison of FEM modelling of stress reversal with laboratory test results done

by Stallebrass (1990) and Dasari (1996). 166

xv
LIST OF SYMBOLS

LIST OF SYMBOLS

Upper case

A cross-sectional area of soldier pile

A pore pressure coefficient

B soldier pile width

C undrained cohesion

C’ effective cohesion

Cc compression index

Cs swelling index

Cub undrained cohesion at base of excavation

E Young’s modulus

Eh Young’s modulus in the horizontal direction

Eu undrained Young’s modulus

Ev Young’s modulus in the vertical direction

G shear modulus

G0 initial shear modulus

Ghv shear modulus representing the coupling between the horizontal and vertical

directions

H excavation depth

I second moment of area of cross section

Ixx second moment of area of cross section about x axis

Iyy second moment of area of cross section about y axis

K’ effective bulk modulus

K0 coefficient of earth pressure at rest

Ka coefficient of active earth pressure

Kmin’ minimum effective bulk modulus

Knc coefficient of earth pressure at rest in normally consolidated soil

xvi
LIST OF SYMBOLS

Kp coefficient of passive earth pressure

LL liquidity index

M critical state frictional constant

N number of blows on sampling spoon during performance of standard

penetration test

OCR overconsolidation ratio

Pa active earth pressure

PI plasticity index

Pp passive earth pressure

Pt apparent earth pressure

S∞ tangential stiffness at very large strain

S0 initial tangential stiffness

Lower case

c cohesion

c’ effective cohesion

cc compression index

cs swelling index

e void ratio

ecs void ratio on the critical state line for p’=1

k coefficient of permeability

kx coefficient of permeability in the x-direction

ky coefficient of permeability in the y-direction

m overconsolidation ratio, OCR, exponent

n effective stress, p’, exponent

p’ mean effective stress

pc ’ preconsolidation pressure

q deviator stress
xvii
LIST OF SYMBOLS

q̂ modified deviator stress

qf deviator stress at failure

su undrained shear strength

ln a Naperian (natural) logarithm of a

Greek symbols

∆hmax maximum wall deflection

∆vmax maximum ground settlement

εs deviator or shear strain

εv volumetric strain

εy deviator or shear strain at yielding

γs bulk unit weight of soil

γw unit weight of water

κ slope of swelling line in e-ln p’ space

κ0 slope of swelling line in e-ln p’ space for isotropic loading

κi slope of swelling line in e-ln p’ space for one-dimensional loading

λ slope of compression line in e-ln p’ space

λ0 slope of compression line in e-ln p’ space for isotropic loading

λi slope of compression line in e-ln p’ space for one-dimensional loading

υ specific volume

φ angle of friction

φ’ effective angle of friction

φcs critical state angle of friction

ν Poisson’s ratio

ν’ drained Poisson’s ratio

νhh Poisson’s ratio representing the coupling between the horizontal directions

xviii
LIST OF SYMBOLS

νhv Poisson’s ratio representing the coupling between the horizontal and vertical

directions

σh’ effective horizontal stress

σv’ effective vertical stress

σx’ effective stress in x-direction

σy’ effective stress in y-direction

xix
INTRODUCTION

CHAPTER 1 INTRODUCTION

1.1 Background

In 1930’s, an entire section of the Berlin subway collapsed because the bracing

was designed for earth pressures calculated in accordance with an inappropriate theory

(Terzaghi et al., 1996). On the other hand, ground movements damaging to a sensitive

monumental structure occurred adjacent to a deep cut in Washington, D. C., even

though the bracing experienced no structural distress (O’Rourke et al., 1976). The

retaining systems used in the above cases were a combination of soldier piles, lagging,

walers and struts. Similar retaining systems were used in excavations for Munich and

New York subways (Terzaghi & Peck, 1948), excavations for the 12th and 19th Street

Stations of the BART system (Armento, 1972), the Downtown Seattle Transit Project

(Borst et al., 1990), eight project sites in Chicago Area (Gill & Lukas, 1990) and the

Bad Creek pumped storage facility near Salem, South Carolina (Pearlman & Wolosick,

1990).

In Singapore, soldier pile walls are used mainly for excavations in residual soils

and where water drawdown is not a problem. They are less popular than sheet pile

walls, although soldier piles installed in predrilled holes are often used to overcome the

problem of inadequate penetration of sheet piles in areas where rock is encountered at

shallow depths below excavation. In hard driving conditions caused by boulders or

other obstructions, soldier piles can be predrilled if necessary, or relocated to avoid the

obstruction, which gives them an advantage over sheet pile walls. Predrilling is often

used in urban areas to avoid the noise and vibrations of pile driving. Soldier pile walls

were used in basement excavations of Boulevard Hotel and Four Season Hotel along

1
INTRODUCTION

Orchard Boulevard. Serangoon and Woodleigh MRT stations along the north-east line

were excavated using soldier pile and timber lagging support system (Coutts and

Wang, 2000). The retaining system for Harbour Front MRT station consists of soldier

pile wall scheme with sheet pile lagging in the upper soil and shotcrete lagging in the

weathered rock (Chen et al., 2000).

1.2 Soldier pile walls

Soil Arch Formation Soldier Pile

Timber Lagging

Waler

King Post

Strut

Figure 1.1: Components of soldier piles and timber lagging systems.

Soldier piles with timber laggings have been used extensively as an excavation

support system, particularly in stiff soil conditions and where ground water ingress into

the excavated area is not problematic (e.g. GCO, 1990 Tomlinson, 1995; O’Rourke,

1975). Soldier pile walls have two basic components, soldier piles (vertical

component) and lagging (horizontal component) as indicated in Figure 1.1. Soldier

2
INTRODUCTION

piles provide intermittent vertical support and are installed before excavation

commences.

The behaviour of discrete piles in the retaining wall gives rise to complexities in

design. Due to their relative rigidity compared to the lagging, the piles provide the

primary support to the retained soil as a result of the arching effect which will be

described in details in later section. The arching of soil behind the soldier piles is a

local 3D effect (Figure 1.1) that cannot be analysed using 1D or 2D analyses. Spacing

of the piles is chosen to suit the arching ability of the soil and the proximity of any

structures sensitive to settlement. A spacing of 2 to 3 metres is commonly used in

strong soils, where no sensitive structures are present. The spacing is reduced to 1 to 2

metres in weaker soils or near sensitive buildings. The separation between two soldier

piles allows the soil to move between them especially below the excavation level in the

absence of lagging.

The lagging serves as a secondary support to the soil face and prevents

progressive deterioration of the soil arching between the piles. It is often installed in

lifts of 1 to 1.5 metres, depending on the soil being supported and on the convenience

of working.

Overconsolidated clays, all soils above the water table if they have at least some

cohesion and homogeneous, free-draining soils that can be effectively dewatered

provide suitable conditions for the use of soldier pile walls. Advantages of soldier pile

walls are (1) soldier-piles and timber lagging are easy to handle, (2) low initial cost

and (3) can be driven or augured. Furthermore, since the soldier piles are not

contiguous, much fewer soldier piles are often needed to be driven in comparison to

3
INTRODUCTION

sheet piles, thereby yielding significant savings in time and cost of installation and thus

allowing excavation to commence with a minimum of lead time.

The design approaches on discrete pile retaining wall systems are laid out in

numerous code of practices, e.g. BS8002 (1994), GCO (1990), Trada (1990) and

NAVFAC (1982). The details will be reviewed in Chapter 2. The design of retaining

systems using soldier piles and timber lagging for deep excavation are becoming more

demanding because of the increasing depth for which they are designed. Soldier piles

retaining systems are used extensively for excavation in residual soil owing to its low

construction cost as compared with diaphragm wall and bored pile retaining systems.

1.3 Deformations

In the frontispiece to his PhD thesis, Professor John Burland wrote (Burland,

1967):

“Stress is a philosophical concept - deformation is the physical reality.”

Engineers are familiar with concepts and calculations involving stresses in

materials but what really matters – what the public sees – is movement and

deformation which lead to damage and danger. Designers are particularly interested in

making reliable predictions of the magnitudes of movements in the surrounding soil;

and then estimating the effects of these movements on adjacent structures and

facilities. In principle, these predictions can be achieved using powerful numerical

methods such as finite element analyses.

4
INTRODUCTION

1.4 Objective

The objective of this study is to investigate the mechanism of soil support in a

retaining system consisting of soldier piles and timber lagging.

In particular, the research will attempt to clarify the role of soil arching in

discrete pile retaining walls and to provide an information base for rational design.

The effect of pile spacing and timber lagging stiffness on soil arching will also be

investigated.

1.5 Scope of Work

Chapter 2 reviews the current design procedures and various methods for

evaluating retaining systems. These include the different aspects of retaining wall that

were investigated by researchers using simple beam and spring models to the state of

the art 3D finite element analysis. A definition and behaviour of soil arching and will

also be surveyed.

Chapter 3 had been contributed to journals Computers and Geotechnics entitled

“Three-dimensional pile-soil interaction in soldier-piled excavations” (Hong et al.,

2003). Strutted excavations, including soldier-piled excavations, are often analysed

using two-dimensional (2D) finite element (FE) analyses with properties, which are

averaged over a certain span of the wall. In chapter 3, the effects of “smearing” the

stiffness of the soldier piles and timber laggings into an equivalent uniform stiffness

are examined, based on comparison between the results of 2D and 3D analyses. The

ability of the 3D analyses to model the flexural behaviour of the soldier piles and

timber laggings is established by comparing the flexural behaviour of various FE beam

representations to the corresponding theoretical solutions, followed by a reality check


5
INTRODUCTION

with an actual case study. Finally, the results of 2D and 3D analyses of an idealized

soldier piled excavation are compared. The findings show that modelling errors can

arise in several ways. Firstly, a 2D analysis tends to over-represent the coupling to pile

to the soil below excavation level. Secondly, the deflection of the timber lagging,

which is usually larger than that of the soldier piles, is often underestimated. For this

reasons, the overall volume of ground loss is, in reality, larger than those given by a

2D analysis. Thirdly, a 2D analysis cannot replicate the swelling, and therefore

softening, of the soil face just behind the timber lagging. Increasing the inter-pile

spacing will tend to accentuate the effects of these modelling errors.

Chapter 4 begins by drawing attention to a case study of Serangoon MRT (Mass

Rapid Transit) site. The subsoil at the site consists of residual soil of a local granite

formation known as Bukit Timah Granite. The temporary retaining structures for the

excavation consisted of driven soldier piles and timber lagging supported by steel

struts. In this chapter, the soil properties of Serangoon station’s geological formation

and construction sequence of the retaining system are presented as a prelude to the

subsequent chapter.

Chapter 5 discusses some results of a comparative study into the effects of

different constitutive models in simulating soldier-piled excavations in medium to stiff

soils. The field data, which were used as reference, in this comparative study came

from the excavation for the Serangoon station of the North-East Line (NEL) contract

704. The methods of modelling retaining system, boundary and groundwater

conditions and excavation sequence are described in this thesis. The behaviour in 2D

and 3D for each soil model, Mohr Coulomb criterion, modified Cam-clay and

hyperbolic Cam-clay, is investigated.

6
INTRODUCTION

Chapter 6 concludes the report by emphasising the complexity and importance

of conducting a 3D analysis for soldier pile retaining system. A brief recommendation

for future research work is also given.

7
LITERATURE REVIEW

CHAPTER 2 LITERATURE REVIEW

Reliable predictions of ground deformations are essential in the design process

for (1) assessing the effects of excavation on adjacent facilities; and (2) identifying

sections where special remedial construction measures are required. Available

techniques for estimating wall deflections and soil settlements involve either

interpolation from existing empirical databases or numerical analyses using finite

element methods.

2.1 Design approaches for retaining systems

‘Limit analysis approach’ is often used as a preliminary tool to determine the

component sizes of a retaining structure. In this approach, the equilibrium of the

retaining structure with an assumed failure mechanism or some active earth pressure

distribution is considered to evaluate the restoring forces or moments required.

Comparison of the available restoring forces with the required restoring forces yields a

‘factor of safety’. This is taken to be an indication of the available margin for

diaphragm walls, which are now commonly used in basement construction, gives rise

to several problems.

Firstly, whereas unstrutted walls can be treated as statically determinate

problems once a failure mechanism is assumed for the surrounding soil, the same

cannot be done for strutted walls. For single-strutted walls, using the ‘free earth

support’ or ‘fixed earth support’ assumption can simplify the problem. Multi-strutted

walls, on the other hand, are often difficult to reduce the degree of freedoms to the

point where they can be solved as static problems.

8
LITERATURE REVIEW

Secondly, the requirement to limit ground movement often leads to the usage of

fairly stiff strutting systems. Under such circumstances, ground deformation is often

not large enough to allow full active or passive pressures to be developed around the

wall. Thus, the assumption of active or passive condition usually does not accurately

reflect reality.

Thirdly, even if the factor of safety can be evaluated, this parameter offers no

information on the ground movement that is likely to be incurred. Because of these

limitations, it is often difficult to apply conventional design approaches to multi-

strutted excavations.

2.2 Various methods for evaluating retaining systems

A number of methods are available in practice to evaluate the soundness of

retaining systems. They are empirical approaches based on proposed relations and

charts, beam and springs model and 2 and 3D numerical analyses.

2.2.1 Empirical approaches

Various researchers (e.g. Peck, 1969; Goldberg et al., 1976; O’Rourke et al.,

1976; Clough & Denby, 1977; Clough et al., 1979 and Mana & Clough, 1981) have

proposed empirical relations and charts for the evaluation of strut loads and ground

movements in deep excavations. Table 2.1 summarises some of the empirical

approaches available for retaining system evaluation.

9
LITERATURE REVIEW

Table 2.1: Empirical approaches contribution by various researchers.


Scope Author Retaining system Soil type
Settlement Peck (1969) Soldier pile and lagging -
O’Rourke (1976) Soldier pile and lagging Dense sand and
interbedded stiff clay
Stiffness of support Goldberg (1976) Soldier pile and lagging -
system
Passive soil resistance Goldberg (1976) Soldier pile and lagging -
Effects of passive soil Clough & Denby (1977) Sheet pile Soft to medium clay
buttress
Factor of safety against Clough et al. (1979) Soldier pile and lagging -
basal heave Mana & Clough (1981) Soldier pile and lagging -

Settlement

Ground movements and strut loads in excavations are highly dependent upon

the stress histories, soil properties, ground condition and the support system.

Settlement behind excavations can be estimated based on data obtained from previous

excavations in similar conditions.

Figure 2.1: Observed settlements behind excavations (after Peck, 1969).

Peck (1969) summarised data on settlements behind excavations to produce the

useful chart shown in Figure 2.1. Settlements presented as a percentage of the

excavation depth were plotted against distance from the excavation divided by

10
LITERATURE REVIEW

excavation depth. Three different zones, as defined in Figure 2.1, were identified. It

should be noted that construction technique has a strong influence on movements of

strutted systems.

Figure 2.2: Summary of settlements adjacent to strutted excavation in Washington, D.C. (after
O’Rourke et al., 1976).

O’Rourke et al. (1976) compiled settlement data associated with soldier pile

and lagging excavation in dense sand and interbedded stiff clay of Washington, U.S.A.

(Figure 2.2). Expressed as a percentage of the excavation depth, the surface

settlements were equal to or less than 0.3% near the edge of the cut and 0.05% at a

distance equal to 1.5 times the excavation depth. O’Rourke et al. (1976) carried out a

comparative study of strutted excavations in the soft clay of Chicago using temporary

berms and raking struts. The data are plotted in dimensionless form in Figure 2.3.

Three zones of ground displacement can be distinguished and related to the salient

11
LITERATURE REVIEW

characteristics of construction. These zones approximate to the three zones of

settlement delineated by Peck (1969), with the exception that the widths of the

settlement zones are notably shorter. This indicated that the settlements associated

with the Chicago excavations are confined to areas that are comparatively nearer to the

edges of the excavation.

The charts provided by Peck (1969) and O’Rourke et al. (1976) took into

account the construction and support systems in a general and qualitative manner. It is

doubtful whether these curves, which were compiled from data obtained in North

America, are directly applicable to Singapore soil. The stiffness of the support systems

is not quantitatively accounted for. Peck’s chart also overestimates the settlement due

to the lower stiffness of struts in the 60’s. These empirical tools lack universal

applicability as none of them considered all these factors.

Figure 2.3: Summary of settlements adjacent to strutted excavation in Chicago (after O’Rourke
et al., 1976).

12
LITERATURE REVIEW

Stiffness of support system

The support system stiffness depends on the stiffness of the wall and its

supports, the spacing between supports, and the length of wall embedded below

excavation level. Goldberg et al. (1976) conducted a valuable study into the effect of

wall stiffness and support spacing, and the results are presented in Figure 2.4.

Figure 2.4: Effects of wall stiffness and support spacing on lateral wall movements (after
Goldberg et al., 1976).

Passive soil resistance

Goldberg et al. (1976) provided the method, illustrated in Figure 2.5, for

determining the depth of penetration in ‘competent’ soils, which are capable of

developing adequate passive resistance. The practicality of using this method on

soldier pile walls is doubted, as the piles do not provide adequate contact surface

13
LITERATURE REVIEW

below excavation level and the passive resistance between the soldier piles is non-

existence.

Pt: Apparent earth

Figure 2.5: Calculation of embedment depth of sheet pile wall in relatively uniform competent
soil conditions (after Goldberg et al., 1976)

Effect of passive soil buttresses.

The effectiveness of these in controlling movements in excavation in soft to

medium clay was studied by Clough & Denby (1977). Figure 2.6 shows the

theoretical relationships between settlements behind bermed sheet pile walls and the

γH
stability number, , for the condition of excavation after the berm has been
Cub

removed and the rakers installed.

14
LITERATURE REVIEW

Figure 2.6: Relationship between maximum settlement and stability number for different
batters (after Clough & Denby, 1977).

15
LITERATURE REVIEW

Factor of safety against basal heave

Figure 2.7: Relationship between factor of safety against basal heave and non-dimensional
maximum lateral wall movement for case history data (after Clough et al., 1979).

Clough et al. (1979) and Mana & Clough (1981) reviewed case histories of

sheet pile and soldier pile walls in clays supported primarily by cross-lot struts. The

results are summarised in Figures 2.7 and 2.8, which show that the maximum lateral

movement can be correlated with the factor of safety against basal heave defined by

Terzaghi (1943). The movement increases rapidly below a factor of safety of 1.4 and

1.5, while at higher factors of safety the non-dimensional movements lie within a

narrow range of 0.2% and 0.8%. Moreover, there do not appear to be any significant

differences in lateral movements between sheet pile walls whose tips are embedded in

an underlying stiff layer and those whose tips remain in the moving clay mass. Wall

deflection is sometimes correlated to and used as a control parameter for ground

movement (e.g. Mana & Clough, 1981). For example, Mana & Clough (1981)

reported that the maximum ground settlement ∆vmax typically lies between 0.5 and 1

times the maximum wall deflection ∆hmax for sheet-piled and soldier piled excavations.

This suggests that the ground loss arising from the inward movement of the wall is

closely correlated to that due to ground settlement.


16
LITERATURE REVIEW

Figure 2.8: Relationship between maximum ground settlements and maximum lateral wall
movements for case history data (after Mana & Clough, 1981).

2.2.2 1-D beam and spring models

Multi-strutted wall in a deep excavation is a statically indeterminate structure.

This has led to the development of computer models, which are based on treating the

wall as a beam and the soil around the wall as equivalent generalised springs. An

example is the classical Winkler solution of about 1867, in which the foundation is

considered as a bed of springs.

Hetenyi (1946) developed solutions for a load at any point along a beam. Based

on these equations, a load is applied at the strutted location to simulate the effect of

strutting. The soil stiffness is reflected as the spring constant.

Bowles (1988) also modelled the sheet pile wall as beam elements, but utilised

the concept of subgrade reaction for soil below the excavation level. The matrix

stiffness method of analysis was used to obtain the solution. Pearlman & Wolosick

17
LITERATURE REVIEW

(1990) modelled soldier piles and diaphragm wall using beam on elastic foundation

analysis to study the soil-structure interaction.

An advantage of the beam and spring approach over limit analysis is its ability

to account for structure flexibility and soil stiffness. Thus, the effects of stress

redistribution in soil because of differential structural deflections are accommodated.

Unfortunately, there are some limitations to this approach. There is inherent difficulty

in determining the appropriate spring stiffness of the soil for analysis, as these are not

fundamental soil properties. This method cannot directly simulate construction

sequence, unusual initial soil stresses and development of wall friction. The Winkler

solution did not account for the shear between the soil particles, which is a great

disadvantage to this model. The analytical treatment of the soil continuum as a series

of generalised springs implies that ground movement cannot be directly obtained from

the computation.

2.2.3 2D finite element analysis

A summary of the contribution by various researchers in the area of 2D finite

element simulation of excavation is presented in Table 2.2.

Prediction of excavation wall deformation or ground surface settlement in the

centre section of an excavation has been studied using plane strain finite element

analysis by many researchers (e.g., Clough & Denby, 1977; Clough & Hansen, 1981;

Clark & Wroth, 1984; Borja, 1990; Finno et al., 1991; Whittle et al., 1993). Apart

from wall deflection and ground settlement, other researchers undertook a number of

excavation-related studies.

18
LITERATURE REVIEW

Table 2.2: 2D finite element analyses contribution by various researchers.


Scope of Study Author Retaining system Soil type
Installation effect Finno et al. (1991) Sheet pile wall Chicago clay and
Niagaran limestone
Gunn et al. (1992) Diaphragm wall London clay
De Moor (1994) Diaphragm wall London clay
Ng et al. (1995) Diaphragm wall Gault clay
Effect of wall length Hashash & Whittle Diaphragm wall Boston blue clay
(1996)
Bica & Clayton (1998) Cantilever Wall Sand
(Experimental)
Richards & Powrie Diaphragm wall Kaolin clay
(1998)
Earth pressures Fourie & Potts (1989) Cantilever Wall London clay
Ling et al. (1993) Diaphram Wall Gault clay
Bica & Clayton (1998) Cantilever Wall Sand
(Experimental)
Factor of safety against Mana & Clough (1981) Sheet piles and soldier Clay
basal heave piles
Clough & O’Rourke Sheet piles and soldier Clay
(1990) piles
Effect of water table Hsi & Small (1992) Diaphragm wall Sandy and clayey soil
drawdown
Effect of consolidation Yong et al. (1989) Diaphragm wall Singapore marine clay
in Singapore Lee et al. (1989) Tieback sheet pile wall Singapore marine clay
Parnploy (1990) Diaphragm wall Singapore marine clay
Excavation in residual Lee et al. (1993) Bored pile wall Bukit Timah granite
soils formation
Discrepancies between Lee et al. (1989) Diaphragm wall Singapore marine clay
computed and predicted
results

Installation effect

The effect of wall installation on the lateral stresses in overconsolidated soil

strata and on the behaviour of retaining walls is commonly recognised. It is featured in

many discussions of retaining wall analyses and performance (e.g. Finno et al., 1991;

Gunn et al., 1992; De Moor, 1994 and Ng et al., 1995).

Finno et al. (1991) noted that the computed sheet pile displacements are slightly

greater for the case of no sheet pile installation after initial excavation stage due to the

preloading effect of sheet pile installation. These trends are reversed by the end of

excavation because of the reduction of available passive resistance caused by the

installation procedure.

19
LITERATURE REVIEW

Gunn et al. (1992) found that installation effects do seem to be significant when

walls are propped. This can be understood as the propping action ‘locking in’ the

reduction in lateral stresses associated with wall installation – the soil around the wall

does not have the same freedom to strain and approach the classical stress

distributions. However, the magnitude of installation effects depends strongly on the

initial position of the ground water table. Installation effects are relatively minor when

ground water table is high, and become more significant as it falls below excavation

level.

De Moor (1994) showed that non-dimensional graphs could be produced that

allow the change in lateral stress in the ground adjacent to the wall to be estimated for

any given applied change in lateral stress at the wall excavation boundary, for a range

σh'
of wall panel widths. These estimates may provide more realistic σv ' values than the

in situ conditions often used at present for analysis of diaphragm walls.

Ng et al. (1995) proposed that the stress changes in the ground are dominated

by two distinct mechanisms: horizontal arching and downward load transfer.

Approximate analyses of the three-dimensional effects of diaphragm wall installation

had been carried out using two simplified horizontal and vertical plane strain sections.

The analyses showed the horizontal arching results only in lateral stress redistribution,

whereas downward load transfer can result in substantial stress reductions by shedding

load downwards beneath the toe through the action of shear stresses in the vertical

plane. However, the effects do not appear to be readily quantified. The authors

recognised that techniques used to simulate the effects of wall installation by a general

reduction in horizontal stresses in all soil above the toe level of the wall is less

satisfactory than those that permit stress redistribution.

20
LITERATURE REVIEW

Effect of wall length

Hashash & Whittle (1992) demonstrated that wall length has a minimal effect

on the pre-failure deformations for excavations in deep layers of clay, but does have a

major influence on the location of failure mechanisms within the soil. Base stability is

controlled by the location of the failure mechanism, which is constrained by the wall

length.

Hashash & Whittle (1996) and Bica & Clayton (1998) also observed the trend

of increasing bending moment with depth of embedment. This conflicts with the

design recommendation that maximum bending moment should be evaluated with a

safety factor of 1. This factor must instead be greater than 1 when evaluating the

maximum bending moment.

Richards & Powrie (1998) showed that an increase in embedment depth will

lead to an increase in wall bending moment and a reduction of bottom prop load.

There is no real advantage in terms of limiting ground movements in increasing the

embedment depth without increasing the wall stiffness. The additional movement due

to bending counteracts the reduction in the component of movement due to wall

rotation.

Lateral earth pressures

Fourie & Potts (1989) examined the lateral earth pressures in front and behind

the wall and found that the earth pressures measured near the cantilever wall bottom

correspond to a value of Kp which is significantly lower than just below the excavation

level.

21
LITERATURE REVIEW

Bica & Clayton (1998) showed experimentally that some cantilever wall design

methods that use the same value of passive earth pressure both above and below the

centre of rotation of the wall could be unsafe unless some conservative assumptions

are made.

Ling et al. (1993) studied the reliability of the earth pressure measurements

adjacent to a multi-propped diaphragm wall. It appeared that the earth pressure cells

do not indicate changes in earth pressure with any great accuracy, although the

direction of change is consistent with the derived earth pressures. Back analysis of

earth pressure for final stage of excavation showed stress redistribution (arching) that

was consistent with the changes in earth pressure recorded by instrumentation.

Basal heave

Mana & Clough (1981) and Clough & O’Rourke (1990) showed that

measurement of maximum lateral wall deflection could be correlated with factor of

safety against basal heave, as defined by Terzarghi (1943). The movements increase

rapidly below a factor of safety of 1.4 to 1.5. At higher factor of safety, the non-

dimensional movements lie within a narrow range of 0.2% to 0.8%. Moreover, there

do not appear to be any significant differences in lateral movements between sheet

piles walls whose tips are embedded in an underlying stiff layer and those whose tips

remain in the moving clay mass.

Effects of water table drawdown

Hsi & Small (1992) developed a fully coupled finite element method for

evaluating the deformations and pore pressures in a soil during excavation taking into

account the drawdown of the free surface. This method is implemented in a finite

22
LITERATURE REVIEW

element program, EXCA2 that is able to solve plane strain and axisymmetric

problems. It can be used for elastic or elasto-plastic soils. To ensure the equilibrium

of stresses within each increment of excavation and determine the location of the free

surface, iterations using implicit Euler backward method, are required within each time

step so that equilibrium can be achieved.

Singapore soil

A number of researchers, (e.g. Yong et al., 1989; Lee et al., 1989 and Parnploy,

1990) had analysed excavation in local Singapore Marine Clay. Lee et al. (1993)

investigated the behaviour of excavation in both marine clay and residual soil.

Yong et al. (1989) indicates that an undrained analysis could not account for the

progressive increase in lateral sheet pile movement and build-up of strut loads with

time. The sheet pile movement may be substantially underestimated.

Lee et al. (1989) suggested a design method that uses movement control. It is

based on the idea that reasonable estimates of support system movements can be made,

including effects of soil conditions, structural components and construction process.

The design method is iterative in that trials are needed to adjust the estimated expected

movements to the allowable values.

Parnploy (1990) studied the time-dependent behaviour of excavations in

Singapore marine clay. The author reported that the time-dependent change in ground

movement is associated with the dissipation of negative pore-water pressure. Excess

pressure begins to dissipate immediately and typically continues well beyond the

completion of the excavation, resulting in progressive movements and increase in strut

loads even when no excavation is being carried out. Parnploy (1990) also reported that

23
LITERATURE REVIEW

piles installed prior to the excavation work would increase the stiffness of the soil in

front of the wall and help to reduce the wall deflection.

Lee et al. (1993) proposed two mechanisms for the ground movement namely

dissipation of excess negative pore pressure and drawdown-cum-seepage. For marine

clay formations, pore pressure dissipation is sufficient to account for the observed

ground movement. The low permeability of the soil formation is probably sufficient to

limit the effects of drawdown and seepage over the duration of excavation. On the

other hand, the much higher permeability of granitic residual soil formations ensures

that the first mechanism is completed much more rapidly than the excavation duration.

Observable ground movement is thus likely to be the result of drawdown and seepage.

Among the researchers, only Mana & Clough (1981) and Clough & O’Rourke

(1990) studied the behaviour of soldier piles. They invoked the plane strain

assumption so that problem simplifies drastically to 2D. Owing to arch formation in

the retained soil, anisotropic behaviour of lagging and movement of soil between the

soldier piles, 2D analysis over simplifies the mechanism of basal heave. Plane strain

assumption is not strictly true because the strut loads are discrete point loads, not a line

load as the load redistribution by the walers is not uniform. The plane strain

assumption will lead to unrealistically large lateral stress reductions.

2.2.4 3D finite element analysis

A summary of the contribution by various researchers in the area of 3D finite

element simulation of excavation is presented in Table 2.3.

Three-dimensional effects in excavation were observed in many field

measurements. Dysli & Fontana (1982) observed the corner effects in a well-

24
LITERATURE REVIEW

instrumented excavation and explore it further in a 3D non-linear analysis of an

excavation in clayey soil using the program ADINA on VAX11-780. They failed to

get any meaningful result because of the expensive cost, limitation of computer

hardware and the unavailability of pre/post processor.

Table 2.3: 3D finite element analyses contribution by various researchers.


Scope Author Retaining system Soil type
Installation effect Ng & Yan (1999) Diaphragm wall Gualt clay
Gourvenec & Powrie Diaphragm wall Lias clay
(1999)
Corner effect Giger & Krizek (1975) Unsupported Coulomb yield criterion
(Analytical)
Cardoso (1987) Tieback diaphragm wall Granitic residual soil
Borja et al. (1989) Unsupported Elastoplastic
(Numerical)
Ou & Chiou (1993) Diaphragm wall Silty clay
Matos Fernandes et al. Strutted diaphragm wall Silty clay
(1994)
Ou et al. (1996) Diaphragm wall Silty clay
Ou & Shiau (1998) Diaphragm wall Silty clay
Point load effect Tsui & Clough (1974) Tie-back diaphragm wall -
Matos Fernandes (1986) Strutted diaphragm wall Silty clay
Liu (1995) Diaphragm wall Singapore marine clay
Wall deflection Briaud & Lim (1999) Tieback soldier pile Silty sand
Small strain effect Nasim (1999) Strutted diaphragm wall Singapore marine clay
Arching effect Vermeer et al. (2001) Tieback soldier pile -
Axisymmetric analyses St. John (1975) Unsupported London clay
Britto & Kusakabe Unsupported Clay
(1983)
Simpson (1992) Bored pile London clay

Installation effect

Ng & Yan (1999) confirms the roles of two stress transfer mechanisms: the

horizontal arching and downward load transfer mechanisms during diaphragm wall

installation. A three-dimensional analysis was carried out. The difference in the

computed behaviour between the true and pseudo three-dimensional analyses is that

the downward load transfer and the horizontal arching mechanism were uncoupled in

the pseudo three-dimensional analyses.

Gourvenec & Powrie (1999) clarified the sequence of stress changes and

displacements that takes place during the installation of a diaphragm wall in panels.

25
LITERATURE REVIEW

Their analyses showed that the magnitude and extent of lateral stress reduction near a

diaphragm wall during construction depends on the panel length and are over-predicted

in analyses assuming condition of plane strain. The analyses also showed that three-

dimensional effects tend to reduce lateral soil movements during installation of a

diaphragm wall in panels compared with the plane strain case, but soil movements

increase markedly with panel length at aspect ratios (panel depth : panel length) of less

than three.

Corner effect

Giger & Krizek (1975) analysed the stability of a vertical cut with a variable

corner angle. Subsequently, Cardoso (1987), Borja (1989), Ou & Chiou (1993), Liu

(1995), Ou et al. (1996) and Ou & Shiau (1998) investigated corner effect of a 3D

excavation.

Cardoso (1987) found out that settlements and horizontal deflections are

smaller in the 3D case, especially at the corner, because of the soil arching and mutual

support between intersecting walls. The lateral deflections for 2 and 3D case are

almost the same at midsection, but settlements are overestimated in the 2D case. The

reason being that the intermediate principal stresses in 3D case experience less

reduction than in 2D case during excavation.

Ou & Chiou (1993) reported that for excavations with shorter sides, 3D

analyses might be required to obtain a more realistic prediction of wall behaviour.

The significance of corner constraint was also studied by Matos Fernandes et al.

(1994). The authors also suggested that, near the corners, the structural behaviour is

26
LITERATURE REVIEW

considerably influenced by the stress distribution due to soil arching as well as by the

mutual support between the intersecting walls.

From the above reviews, it was noted that 3D geometrical considerations were

essential in capturing the soil-structure interaction for retaining systems.

Unfortunately, the arching effect afforded by the soldier piles could not be captured in

these literatures.

Point load effect

Tsui & Clough (1974) compared the earth pressure acting on a 0.3m thick

concrete laterally supported by tie backs spaced 1m apart in clay. It was shown that

the assumption of plane strain usually adopted in finite element analysis of strutted and

anchored walls cannot be arbitrarily extrapolated to many practical problems,

particularly those involving soldier pile walls or light sheet pile walls. Diaphragm

walls, however, were shown to approximately satisfy the conditions assumed in a

plane strain analysis. In general, discrepancies increase with wall flexibility, soil

stiffness and tie back spacing.

Matos Fernandes (1986) concluded that 3D effects arising from strut loads are

only significant near the struts and do not play an important role in the behaviour of

the wall.

Liu (1995) simulated the struts in an excavation using springs. The

disadvantage of this simulation is that the stiffness of the springs is a constant

throughout the excavation process. Furthermore, it is difficult to obtain the stiffness of

the raking struts.

27
LITERATURE REVIEW

Wall deflection

Briaud & Lim (1999) studied the influence of various design decision for

tieback walls. One-dimensional beam elements were used to simulate soldier piles and

tendons. The analyses used isotropic shell elements to model the timber lagging. The

authors recommended that the first anchor be installed between 1.2m and 1.5m below

the top of the wall. Deflections accumulated during this period are very difficult to

eliminate by further tieback installation.

Small strain effect

Nasim (1999) showed that non-linear effect is likely to be more dominant away

from the corners whereas, at the corners, the effects may be more important. Other

situations where the non-linear effect is important are in the prediction of the response

of piled and shallow foundations.

Several methods have been used to represent the non-linear behaviour of soils

below the yield locus. The simplest, but not necessarily the best is to use variable

shear modulus, G. For instance, Dasari (1996) developed a “strain-dependent Cam-

clay model” by assuming that

b
G = Bp' n OCR m ε s (2.1)

in the non-linear region, in which B, n, m and b are parameters to be measured from

tests. Similarly, Nasim (1999) assumed that G decreases hyperbolically with εs.

Arching effect

Vermeer et al. (2001) studied both horizontal and vertical arching behind

soldier piled retaining wall using finite elements. Horizontal arching was promoted by
28
LITERATURE REVIEW

means of flexible horizontal lagging timbers and solid contact between the piles and

the surrounding soil. A simple 2D analysis was performed to assess the amount of

arching. As a result, lighter timber laggings could be used and the economy of the

wall construction could be increased. High pre-stress forces in the upper ground

anchors induced vertical arching.

Axisymmetry

St. John (1975) studied the differences between 2 and 3D analyses of square,

unsupported excavation in London Clay using linear elastic finite element analyses.

His results showed that good agreement was obtained between the 2D axisymmetric

and 3D analyses.

Britto & Kusakabe (1983) found that the width of the settlement zone adjacent

to the excavation is approximately equal to 0.4 times the depth of excavation in 2D

axisymmetric analysis. Burland et al. (1979) also suggested that an assumption of

axial symmetry is more appropriate to computations for square excavations than is

plane strain. This assumption is often difficult to apply to excavations that are

irregularly shaped or supported by strutting system. The geometry of the analysis will

induce large compressive hoop stresses in the wall resulting in unrealistically small

deflection.

Simpson (1992), in his study of British Library excavation, reported that the

differences between axisymmetric and plane strain analyses are negligible and

attributed this to the shallow depth to a relatively rigid stratum.

Axisymmetric analyses are simple and intuitively reasonable for unsupported

excavations, it is often difficult to apply to excavations that are irregularly shaped or

29
LITERATURE REVIEW

supported by strutting systems. There is also a possibility that, if the stiffness of the

strutting system is sufficiently high, corner effects may be suppressed to the point that

the excavation behaves in a largely 2D plane strain manner.

Furthermore, if a soldier pile wall supports the excavation, the use of

axisymmetric analysis will require the lateral modulus of the wall to be artificially

reduced in order to avoid large compressive hoop stresses that would render the

computed wall deflection unrealistically small. Finally, it also does not replicate the

lateral flexural action of the walers system.

With the exception of works done by Briaud & Lim (1999) and Vermeer et al.

(2001), most research mentioned are not applicable to soldier piles retaining system

owing to the added complexities of soil arching, movement between piles and

anisotropy of lagging. Matos Fernandes (1986) simulated the soil and wall using

linear elastic elements and the strut with spring. This will hardly reflect the complex

behaviour of the soil that is more likely to be elastoplastic in nature. The work by Ng

et al. (1995) and Ng & Yan (1999) highlighted the importance of using three-

dimensional analyses in the investigation of horizontal arching. Beam elements used

by Briaud & Lim (1999) may not realistically represent the interaction between the

soldier piles and surrounding soil, owing to their inability to replicate the finite cross-

sectional dimensions of real soldier piles. Vermeer et al. (2001) did not investigate the

effect of pile spacing on arching, which will further determine the economy of the

retaining system. The 2D analysis in a horizontal cross section may not give a realistic

view of arching, which occurs in 3 dimensions. Axisymmetric analyses are strictly

applicable to a circular unsupported excavation only. Research with this level of

30
LITERATURE REVIEW

complexity in modelling has never been attempted from the literature survey

conducted.

2.3 Soil arching

In addition to the above research, soldier pile retaining systems have added

complications, viz. (1) soil arching and discrete piles, (2) interaction with timber lags.

Little has been published on the arching effect of soil for the design of retaining walls

using discrete piles spaced some distance apart. As defined by the Committee on the

Glossary of Terms and Definitions in Soil Mechanics (1958), arching is the

phenomenon of the transfer of stress from a yielding part of a soil mass to adjoining

less yielding or restrained parts of the mass.

Terzaghi (1936) and Bosscher (1981) used the trap door testing apparatus and

not discrete piles to quantify arching effect and draw parallelism between the two.

Nevertheless, the assumption that arching produces relieving effect is widely used in

practice where the retaining systems consist of discrete piles and flexible timber

lagging. Numerous uses of this assumption on discrete pile retaining wall systems

exist in current geotechnical practice, BS8002 (1994), GCO (1990), Trada (1990) and

NAVFAC (1982).

2.3.1 Design procedures for soldier pile wall

Soldier piles are often designed and analysed as a contiguous wall systems even

though they are, in reality, not so (e.g. Peck, 1969). Nonetheless, designers have long

recognized that the soldier pile retaining system is not only often more flexible than

other systems but the stiffness of the system is non-uniform in that the piles are much

stiffer than the timber lags. To address this characteristic, Trada (1990) and BS8002

31
LITERATURE REVIEW

(1994) recommended a 20% to 25% reduction of Peck’s earth pressure for the soldier

piles and 50% reduction for the timber lagging. It also indicates that relieving effect of

the arching is much reduced with the use of stiffer lagging materials, for example

concrete, due to lower relative flexibility of the concrete. Armento (1972) also

proposed a similar reduction based on measurements from the excavations for the 12th

and 19th Street Stations of the BART system. This is a simplifying assumption for a

hinge at formation level. It makes an allowance for the difficulty of assessing the

degree of fixity at the toe of the pile. Broms (1964), Teng (1975), NAVFAC (1982)

and the Canadian Geotechnical Society (1992) proposed that the soldier pile of width

B be designed to support the passive earth pressure imposed by an extent of ground of

span 3B (i.e. 1.5B on both sides of the pile) below the bottom of excavation. This is

because the failure in soil due to individual pile elements is different from that of

continuous walls for which pressure distributions are derived. All the above has made

it difficult for designers to quantify the relieving effect of arching in soil.

2.3.2 Research on arching effect

If a portion of an otherwise rigid support of a granular mass yields, the

adjoining particles move with respect to the remainder of the granular mass. This

movement is resisted by shearing stresses that reduce the pressure on the yielding

portion of the support while increasing the pressure on the adjacent rigid zone. This

phenomenon is called the “Arching Effect”. “Arching” derives from the Latin archus,

as in archery, for flight of an arrow. Lusher & Hoeg (1964) suggested arching as a

“thrust ring action” in soil surrounding an opening, and noted the existence of self-

supporting soil arches or domes.

32
LITERATURE REVIEW

Terzaghi (1936) studied arching effect using the trap door testing apparatus. In

1943, Terzarghi described arching action in soils as “one of the most universal

phenomena encountered in soils both in fields and in laboratory,” and devoted a

chapter to the subject. However, he did not actually draw an arch, but used the term

qualitatively to explain observed non-hydrostatic pressure distributions of soils against

retaining walls. Getzler et al. (1968) described earth pressure distributions in terms of

arching action but the authors rarely draw an arch.

Krynine (1945) proved by using Mohr’s circle that the stresses along the wall at

the two ends are not the principal stresses but the rotation of the principal stresses.

Handy (1985a) extended this concept and concluded that soil-arching action may be

depicted as a trajectory of minor principal stress that approximates a catenary. The

concept of the dipping downward catenary arch was further discussed by Kingsley

(1989). He used the force equilibrium to prove that the shape of soil arch behind the

wall can be, theoretically, very close to catenary or circular. This gives strong support

to the assumption made by Handy (1985a) that both minor and major principal stresses

are constant along the arch.

Handy (1985b) further that the self-supported arch must assign its continuous

stress trajectory to the major rather than the minor principal stress. The soil arching

actions features the continuous stress trajectory of minor principal stress for downward

catenary type of soil arch and major principal stress for supportive soil arch.

Al-Hussaini & Perry (1978) found that, as a reinforced earth wall was

surcharged to failure, wall pressures increased markedly above Rankine stresses near

the centre of the wall while horizontal stresses in soil only a foot away were much

lower. These changes are indicative of an arching action near the wall.

33
LITERATURE REVIEW

Bosscher (1981) investigated the role of soil arching in sandy slope using the

trap door apparatus. His findings show that arching can occur in both loose and dense

sand. The density of the soil at the slip surface may have contracted or dilated to the

critical void ratio. Therefore, the soil stress transfer capability is the same in the loose

and dense deposit.

Handy (1985a) explained soil arching in terms of rotation of principal stresses

at the wall thereby producing appreciably higher wall pressure than those predicted by

Rankine or Coulomb analyses. The subsequent lateral wall yielding initiates active-

state soil conditions, and reduces pressures on the lower wall. Nakai (1985) also noted

this behaviour behind retaining wall in his 2D analyses.

As mentioned in previous section, Ng et al. (1995) and Ng & Yan (1999)

studied soil arching as one of the mechanism during diaphragm wall installation. The

horizontal arching mechanism transfers lateral stress (via the shear stress, τyx) from the

center to the edge of the panel.

The discussion made by Handy (1985a & b) and Kingsley (1989) is limited to

vertical arching in soil behind the wall. Horizontal arching was not explained.

Investigations on soil arching were done using trap door testing apparatus in

experiments and 2D finite element analysis to explain the appreciably higher wall

pressure than those predicted by Rankine or Coulomb analyses. Arching was never

explained in terms of pressure distribution on the soldier piles and timber lagging in a

3D finite element analysis.

Due to their relative rigidity compared to the lagging, the piles provide the

primary support to the retained soil as a result of the arching effect. The arching of

34
LITERATURE REVIEW

soil behind the soldier piles is a local 3D effect that cannot be analysed using 1D or 2D

analyses.

2.4 Review synopsis

The descriptions so far have shown that numerical analysis can play an

important role as a research tool in studying earth-retaining structures. Its use as a

design tool, however, is somewhat limited in the past owing to reasons stated by

Clough & Tsui (1977), namely (1) satisfaction with conventional methods; (2)

difficulties with older numerical methods; (3) cost; (4) character of the geotechnical

engineer. The last reason can be explained by the fact that most geotechnical

engineers enjoy their work precisely because they often design using considerable

‘engineering judgment’ and less reliance on theoretical tools. In recent years,

engineers have had the possibility to routinely use advance numerical methods in

design. The possibility to do so is largely due to the increased availability of such

programs, combined with the rapid pace of hardware development (Potts &

Zdravković, 2001).

Results from literature survey are not applicable to soldier piles retaining

systems owing to the added complexities of soil arching, movement between piles and

anisotropy of lagging.

Mana & Clough (1981) and Clough & O’Rourke (1990) studied the behaviour

of soldier piles using 2D analysis. Owing to arch formation in the retained soil and

movement of soil between the soldier piles, 2D analysis tends to oversimplify the

mechanism of basal heave. Plane strain assumptions are not strictly true because the

strut loads are discrete point loads, not a line load as the load redistribution by the

35
LITERATURE REVIEW

walers is non-uniform. Matos Fernandes (1986) studied the effect of anchors and

struts applied to a diaphragm wall at discrete points. The interaction between soil and

soldier piles, as discrete elements is still unknown.

36
IDEALISED ANALYSIS OF PILE-SOIL INTERACTION

CHAPTER 3 IDEALISED ANALYSIS OF PILE-SOIL


INTERACTION

Soldier piles with timber lags have been used extensively as an excavation

support system, particularly in stiff soil conditions and where ground water ingress into

the excavated area is not problematic (e.g. GCO, 1990; Tomlinson, 1995; O’Rourke,

1975). The main advantage of using soldier piles is their relatively low cost and ease

of installation, compared to other forms of support systems such as diaphragm walls

and bored piles. Since the soldier piles are not contiguous, fewer soldier piles are

needed in comparison to sheet piles, thereby yielding significant savings in time and

cost of installation and thus allowing excavation to commence with a minimum lead

time.

Soldier piles are often designed and analysed as a contiguous wall systems even

though they are, in reality, not so (e.g. Peck, 1969, Potts & Zdravković, 2001). The

fact that the soldier piles and timber laggings are of vastly different construction and

have very different stiffnesses means that the interaction between the retaining system

and the soil is three-dimensional (3D) in nature. However, most of the finite element

(FE) analyses on soldier pile walls to date are two-dimensional (2D) and based on the

assumption of plane strain (e.g. O’Rourke, 1975; Clough et al., 1972; Tsui and Clough,

1974 and Gomes Correia and Guerra, 1997). The differences in pile and timber

lagging stiffnesses cannot be modelled by these analyses, which necessarily assume a

“smeared” uniform stiffness for the entire span of the retaining wall. Briaud and Lim

(1999) used 3D FE analyses to study tieback soldier pile walls. However, their study

focuses more on the parameters of the tiebacks and depth of soldier pile embedment

rather the effects of different types of analyses. Furthermore, Briaud and Lim (1999)

37
IDEALISED ANALYSIS OF PILE-SOIL INTERACTION

modelled the soldier piles with beam elements. As will be discussed later, beam

elements may not realistically represent the interaction between the soldier piles and

surrounding soil, owing to their inability to replicate the finite cross-sectional

dimensions of real soldier piles.

In this chapter, the effects of “smearing” the stiffness of the soldier piles and

timber laggings into an equivalent uniform stiffness are examined, based on

comparison between the results of 2D and 3D analyses. “Smeared” uniform stiffness

is used for the retaining systems in the 2D analyses. On the other hand, in the 3D

analyses, different elements and material types represent the soldier piles and timber

laggings. In the first part of this chapter, the ability of the 3D analyses to model the

flexural behaviour of the soldier piles and timber laggings is examined. This is

accomplished by comparing the flexural behaviour of various FE beam representations

to the corresponding theoretical solutions. A reality check is then conducted using

O’Rourke’s study on the G Street test section as a reference. Finally, the results of a

comparative study between 2D and 3D analyses on an idealized soldier piled

excavation will be discussed to highlight the effects of using “smeared” uniform

stiffnesses in 2D analyses. All of the analyses reported herein were conducted using

an in-house version of CRISP (Britto and Gunn, 1990), with 3D beam and reduced-

integration brick elements incorporated.

3.1 Modelling of soldier piles in 2D and 3D analyses

In this section, the modelling of soldier piles in 2D and 3D analyses is first

addressed. Soldier piles act essentially in flexure, and it is well-known that 4-noded

quadrilaterals with high aspect ratios do not model flexural behaviour well, the strain

energy due to the transverse shear stress and strain becomes artificially large, which
38
IDEALISED ANALYSIS OF PILE-SOIL INTERACTION

causes shear-locking (e.g. Livesley, 1983; Reddy, 1993). The 8-noded quadrilateral

element suffers from the same problem, albeit to a lesser extent (e.g. Reddy, 1993). It

has been shown (e.g. Zienkiewicz and Taylor, 1988; Day and Potts, 1993) that, by

using reduced-integration eight-noded quadrilaterals, flexural behaviour in 2D can be

readily modelled. The same may presumably be true for the 20-noded brick element.

In order to assess the ability of various elements to represent the soldier pile, the

idealized problem of a long cantilever beam subjected to a triangular load distribution,

the latter representing an idealized earth pressure profile, was studied using various 2D

and 3D representations, as shown in Figures 3.1 to 3.5. The 2D representations were

achieved using beam elements as well as full- and reduced-integration 8-noded

quadrilateral elements, hereafter termed FIQUAD and RIQUAD, respectively. The 3D

representations were achieved using beam element as well as full- and reduced-

integration 20-noded brick elements, hereafter termed FIBRICK and RIBRICK,

respectively. For each type of element, two different cases were studied. In the first

case, one element was used to model the whole length of cantilever beam. In the

second, the cantilever was modelled using 18 elements of equal length. For the

quadrilateral elements, the element aspect ratios for the 1- and 18-element

representations are 1:70.5 and 1:3.9, respectively.

Figure 3.1a compares the deflected shape computed using 32-bit precision with

the theoretical value from Euler-Bernoulli beam theory. As can be seen, all of the 8-

noded quadrilateral representations show significantly smaller deflection than the

theoretical value. The only representations that come close to the theoretical value are

predictably, those using beam elements. Figure 3.2a shows the corresponding bending

moment profiles obtained from the various representations. For beam elements, the

bending moment profile was interpolated from the integration point values using a

39
IDEALISED ANALYSIS OF PILE-SOIL INTERACTION

spline function. For quadrilateral elements, the bending moments were evaluated

using the method suggested by Rahim and Gunn (1997). As can be seen, the

discrepancy in the deflection is also reflected in the bending moment profiles.

Figures 3.1b and 3.2b show the corresponding deflected shapes and bending

moment profiles computed using 64-bit precision. As can be seen, a dramatic

improvement in accuracy is now achieved in deflection and bending moment in the 18-

element quadrilateral representations. On the other hand, both 1-element quadrilateral

representations show little or no improvement in accuracy. In summary, flexural

behaviour of the cantilever is well represented by single and multiple beam elements,

in both 32- and 64-bit precision. The 18-element quadrilateral representations

adequately represent the cantilever behaviour if they are used with 64-bit precision. In

practice, this is unlikely to be a major constraint since 64-bit precision may be needed,

in any case, to capture the large difference in stiffness between soil and pile (Britto &

Gunn, 1990). Between the two 18-element quadrilateral representations tested, the

RIQUAD representation produces a smoother bending moment profile whereas the

FIQUAD representation produces a rather jagged bending moment profile. The 1-

element quadrilateral representations are unable to adequately replicate the deflection

and bending moment profiles, regardless of whether 32- or 64-bit precision is used. In

other words, the representations, which model cantilever flexure well, are the beam

element and the RIQUAD element with 64-bit precision. Figure 3.3 shows the effect

of using different number of RIQUAD elements, and therefore aspect ratios. As can

be seen, reasonably good bending moment representation can be achieved for aspect

ratio less than about 14. Figures 3.4 and 3.5 show the corresponding deflection and

bending moment plots for the 3D representations. As can be seen, the performance

trends of the various 3D representations closely mirror those of their 2D counterparts,

40
IDEALISED ANALYSIS OF PILE-SOIL INTERACTION

with the beam elements being the most reliable and the FIBRICK element being the

least.

45 (a) 45 (b)

40 40

35 35

30 30

Height (m)
Height (m)

25 25

20 20

15 15 Euler-Bernoulli beam theory


Beam (1 element)
Beam (18 elements)
10 10 Types of Quad Elements Aspect Ratio
Full-integration 1:70.5
5 5 Full-integration 1:3.9
Reduced-integration 1:70.5
Reduced-integration 1:3.9
4.3kN
0 0
0 1 2 3 4 0 1 2 3 4
Deflection (m) Deflection (m)
Figure 3.1: Comparison of deflection profiles under distributed load for 2D 43m-cantilever
beam. (a) 32-bit precision. (b) 64-bit precision.

45 45
Euler-Bernoulli beam theory
Beam (1 element)
40 40 Beam (18 elements)
Types of Quad Elements Aspect Ratio
Full-integration 1:70.5
35 35 Full-integration 1:3.9
Reduced-integration 1:70.5
30 30 Reduced-integration 1:3.9
Height (m)

Height (m)

25 4.3kN 25

20 20

15 15

10 10

5 5
(a) (b)
0 0
0 400 800 1200 1600 0 400 800 1200 1600
Bending Moment (kNm) Bending Moment (kNm)
Figure 3.2: Comparison of bending moments under distributed load for 2D 43m-cantilever
beam. (a) 32-bit precision. (b) 64-bit precision.

41
IDEALISED ANALYSIS OF PILE-SOIL INTERACTION

45 (a) 45 (b)

40 40

35 35

30 30

1:70.5

1:35.2

1:14.1
Height (m)

Height (m)

1:7.0

1:3.9
25 25

20 20
Aspect ratio
15 15 1:70.5
1:35.2
1:14.1
10 10 1:7.0
1:3.9

5 4.3kN 5

0 0
0 1 2 3 4 0 400 800 1200 1600
Deflection (m) Bending Moment (kNm)
Figure 3.3: Effect of aspect ratios for 2D 43m-cantilever beam in 64-bit precision using
RIQUAD elements. (a) Deflection. (b) Bending moment.

45 (a) 45 (b)

40 40

35 35

30 30
Height (m)
Height (m)

25 25

20 20

15 15

Euler-Bernoulli beam theory


10 10
Beam (1 element)
Beam (18 elements)
5 4.3kN 5 Types of Brick Elements Aspect Ratio
Full-integration 1:3.9
Reduced-integration 1:3.9
0 0
0 1 2 3 4 0 1 2 3 4
Deflection (m) Deflection (m)

Figure 3.4: Comparison of deflection profiles under distributed load for 3D 43m-cantilever
beam. (a) 32-bit precision. (b) 64-bit precision.

42
IDEALISED ANALYSIS OF PILE-SOIL INTERACTION

45 45

40 40 Euler-Bernoulli beam theory


Beam (1 element)
Beam (18 elements)
35 35 Types of Brick Elements Aspect Ratio
Full-integration 1:3.9
Reduced-integration 1:3.9
30 30

Height (m)
Height (m)

25 25
4.3kN

20 20

15 15

10 10

5 5
(a) (b)
0 0
0 400 800 1200 1600 0 400 800 1200 1600
Bending Moment (kNm) Bending Moment (kNm)

Figure 3.5: Comparison of bending moments under distributed load for 3D 43m-cantilever
beam. (a) 32-bit precision. (b) 64-bit precision.

In soldier-piled excavations, the soldier piles not only act in flexure but also

attract earth pressures from the retained soil. Although this phenomenon is not

modelled, and thus irrelevant, in 2D analyses, it is relevant to a 3D analysis wherein

soldier piles and timber laggings are modelled explicitly. In this section, the ability of

the beam and brick representations of the soldier piles to capture the interaction

between pile and soil in a 3D analysis will be studied using a typical section of an

idealized unstrutted, soldier-piled excavation shown in Figure 3.6. As can be seen, the

chosen section consists of one half-soldier pile and the ground in front of and behind it.

The soldier piles are modelled using 64-bit precision and are spaced at an interval of

six times the width of the soldier pile, center-to-center.

43
IDEALISED ANALYSIS OF PILE-SOIL INTERACTION

A'
(a) (b) A'
Soldier piles Soil layer 1
Soldier piles x x
(brick element) Soil layer 2
(beam element)
z z  Excavation
levels
A A

15m A A' (c) (d)


Soldier piles
0.0m
-4.5m

-7.5m

-11.5m


-24.0m
y

x
-37.0m

100m

Figure 3.6: Locations of beams and linear strain bricks for modelling soldier piles in the finite
element model of an idealistic excavation. (a) Plan view of beam element as soldier pile in
section A-A. (b) Plan view of RIBRICK element as soldier pile in section A-A. (c) Cross-
sectional view in x-y plane of the FEM model. (d) Soil layers, struts and excavation levels.

In this analysis, the outer cross-sectional dimensions and flexural rigidity EI of

the soldier are based on the properties of the 610 mm × 305 mm × 149 kg/m H-pile

(BSI, 1996). The outer cross-sectional dimensions are only simulated in the RIBRICK

elements, not beam element. Nodes on the four vertical faces of the mesh are only

allowed to move in-plane. The bottom face is fixed in all directions. Timber laggings

are not modelled, as the objective of this exercise is to study the ability of various

soldier pile representation to model the pile-soil interaction. Soil Layer 1 is underlain

by Layer 2 at 22.5m depth to model the increasing stiffness of real ground. Each soil

layer is assumed to be elastic and uniform. The soldier pile is socketed 1.5m deep into

Layer 2. A Poisson’s ratio of 0.02 is used for RIBRICK soldier pile elements as pure

44
IDEALISED ANALYSIS OF PILE-SOIL INTERACTION

bending behaviour is only achieved if Poisson’s ratio is zero (MacNeal, 1994). The

properties of the soldier piles and soil are shown in Table 3.1.

Table 3.1: Idealised soil and soldier pile properties for beam and brick soldier piles study.
Soil layer 1 (Depth: 0-22.5m) γs = 19.5 kN/m3
Elastic model E = 50000 kN/m2
ν = 0.3
K0 = 0.94
kx = 1.0×10-9 m/s
ky = 1.0×10-9 m/s
Soil layer 2 (Depth: 22.5-37m) γs = 19.5 kN/m3
Elastic model E = 120000 kN/m2
ν = 0.3
kx = 1.0×10-9 m/s
ky = 1.0×10-9 m/s
Half soldier pile modelled with beam at 6B pile spacing E = 1×108 kN/m2
Ixx = 1.259×10-3 mm4
Iyy = 9.308×10-5 mm4
ν = 0.29
A = 0.019 m2
Half-soldier pile modelled with RIBRICKs at 6B pile spacing E = 4.36×107 kN/m2
(610×152.5mm) ν = 0.02
G = 2.14×107 kN/m2

The soldier pile is assumed in-place at the start of the analysis. The excavation

process is assumed to occur in three lifts with pore pressure dissipation being modelled

using the Biot’s (1941) consolidation formulation. The final depth of excavation is

11.5m.

As shown in Figure 3.7a, the solid and beam elements give similar results and

the differences appear to be minor. The differences are more clearly reflected in

Figure 3.7b, where the bending moment of the solid elements is clearly larger than that

of the beam element, although the differences are still not substantial. This can be

attributed to the finite cross-sectional dimensions of the RIBRICK element, which

enables it to attract a larger proportion of the earth pressure through the arching

process than the beam elements.

45
IDEALISED ANALYSIS OF PILE-SOIL INTERACTION

0 0

5 5

10 10
Depth (m)

Depth (m)
15 15
Soldier pile location
Beam
RIBRICK
Mid-span location
20 20 Beam
RIBRICK

(a) (b)
25 25
0.00 0.02 0.04 0.06 -200 0 200 400
Deflection (m) Bending Moment (kNm)

Figure 3.7: (a) Deflection profile of soldier pile and soil at the center between adjacent piles.
(b) Bending moment profile of soldier piles.

0.037
Deflection (m)

Beam
0.036 RIBRICK

0.035   Beam element


CL
 Brick element
  Mid-span of lagging

0.034
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
Lateral distance from soldier pile (m)
Figure 3.8: Plan view of lateral deflection retaining system 9m depth after final excavation
stage.

Figure 3.8 shows the local deflection of soldier pile and soil face, in plan view,

at 9m depth below ground surface for half-span between adjacent soldier piles. As can

be seen, the differential movement between pile and soil is clearly larger with the beam

representation than the RIBRICK representation of the soldier pile. Furthermore, as

shown in Figures 3.9a and 3.9b, although the development of stress arch in the soil is

evident in both representations, the local stress distribution is very different. The brick

46
IDEALISED ANALYSIS OF PILE-SOIL INTERACTION

representation results in a much higher concentration of stresses in the vicinity of the

soldier pile than the beam representation, owing to its finite dimensions. In addition,

the stress arch spans over a shorter distance but extends deeper into the retained soil in

the case of the brick representation. This allows the soil movement to be better

coupled to the pile movement than the beam representation. Thus, although a beam

element is slightly more accurate in replicating the flexural behaviour of the soldier

pile, local pile-soil interaction is better represented using solid elements. For this

reason, subsequent 3D analyses discussed herein will model the soldier piles using

reduced-integration 20-noded bricks with 64-bit precision, unless otherwise stated.


0.915
0.732
0.549
0.366 (a)
0.183
Span (m)

0.000
0.00 1.00 2.00 3.00
0.915
0.732
0.549
0.366 (b)
0.183
0.000
0.00 1.00 2.00 3.00
Distance from retaining wall (m)
Figure 3.9: Lateral stress contours for at 9.3m below ground level for excavation to 11.5m. (a)
Beam soldier pile elements. (b) RIBRICK soldier pile elements.

3.2 Comparison with case history

In this section, a comparison is made with O’Rourke’s (1975) field

measurement of an 18.3m-deep braced excavation in the G Street test section (Panel

158). Figure 3.10 shows the plan view of the retaining system. The modelled

geometry and excavation sequence follows closely that of O’Rourke’s. Figure 3.11

shows the 2D and 3D FE meshes used in the current analyses. The 2D mesh follows

47
IDEALISED ANALYSIS OF PILE-SOIL INTERACTION

that of O’Rourke’s, whilst the 3D mesh is merely an “extrusion” of the 2D mesh in the

third dimension.

Strut

Reaction Plates

Waler

Soldier Pile

3’-9” 12’-6” 6’-3”

Figure 3.10: Plan view of retaining system after O’Rourke (1975).

Plan view
Upper Brown Sand
Middle Gray Clay
Gray Sands and Interbedded Stiff Clay strut
Lower Orange Sand
Hard Cretaceous Clay
Cretaceous Sands, Gravels, Clays
timber
 Excavation levels soldier pile lagging
0.0m
EE == 48263.3kPa
48300kPa νν == 0.3
CC == 12.4kPa
12.4kPa φ = 33°
33° 4.18m
 7.60m
EE == 13789.5kPa
13800kPa νν = 0.49
CC == 48.3kPa
48.3kPa φ = 0° 10.62m
 
EE == 103421.4kPa
10300kPa νν == 0.3
0.3 14.33m
CC == 20.7kPa
20.7kPa φφ == 33°
33°  
EE == 137895.2kPa
13800kPa νν = 0.3
CC == 10.3kPa
10.3kPa φ = 35°
35°
EE == 241316.5kPa
24100kPa νν == 0.49
0.49
CC == 289.6kPa
289.6kPa φφ == 0°

EE == 344737.9kPa
34500kPa νν == 0.3
0.3
CC == 48.3kPa
48.3kPa φ = 35°
35°
39.40m
Soil properties Cross sectional view
Figure 3.11: Finite element mesh used for reality check with O’Rourke (1975).

Figure 3.11 shows the soil parameters from O’Rourke (1975), which were

adopted in this comparative study. Since O’Rourke only presented total stress

48
IDEALISED ANALYSIS OF PILE-SOIL INTERACTION

parameters, only a total stress analysis was performed. Furthermore, following

O’Rourke, the Drucker-Prager soil model is used for all soil layers.

Table 3.2: Idealised soil and soldier pile properties for O’Rourke’s case study.
Soldier piles E (kPa) ν Ghv (kPa)
5.09×107 0.02 2.5×107
Struts E (kPa) Ixx (m4) Iyy (m4) Area (m2)
Level 1 (36 WF 260) 1×108 0.0072 4.54×10-4 0.0623
Level 2 1×108 1.19×10-6 1.19×10-6 0.0296
Level 3 1×108 0.55×10-6 0.55×10-6 0.0238
Level 4 1×108 1.19×10-6 1.19×10-6 0.0296
Level 5 1×108 1.19×10-6 1.19×10-6 0.0296
Timber lagging Eh (kPa) Ev (kPa) νhh νvh Ghv (kPa)
12×103 1 0.02 3×10-13 0.3676

In the 2D analysis, the soldier pile wall was modelled using beam elements with

a “smeared” wall stiffness derived by scaling the bending stiffness of individual soldier

piles over the distance separating them and neglecting the effects of the timber lags;

this also follows O’Rourke’s (1975) procedure. In the 3D analyses, the soldier piles

and timber lags are modelled using RIBRICK elements, the properties of which are

shown in Table 3.2. The soldier pile properties were derived from the section

properties of the 24 WF 100 H-piles used in the excavation. Less is known about the

timber lags apart from the fact that they were made from oak wood. Back-calculation

using BS8002 (1994) guidelines and AITC (1994) suggest that the required thickness

of the timber lags is likely to be about 100mm, which falls within the common range of

timber lagging thickness (Tomlinson, 1995). Because of the uncertainty over the

thickness of the timber lagging, the analysis was repeated with timber lagging

thickness of 50mm and 100mm. Since timber lagging is normally installed in the form

of long, narrow planks slotted horizontally in between the flanges of the soldier piles

(e.g. Tomlinson, 1995; BS8002, 1994), the flexural rigidity of the timber lagging to

bending in the vertical direction is likely to be very small. To model this behaviour, an

anisotropic elastic material model is used for the timber lagging, with the Young’s

modulus in the vertical direction being set much lower than that in the horizontal

49
IDEALISED ANALYSIS OF PILE-SOIL INTERACTION

direction, Table 3.2. This differs from Briaud and Lim’s (1999) analyses, which used

isotropic shell elements to model the timber lagging. The braces and walers were

modelled using beam elements, with properties adjusted to replicate those in the field.

0 0

5 5

10 10
Depth (m)

15 Depth (m) 15

Field (O’Rourke, 1975)


IN4B
20 IN5B 20
Soldier pile models
RIQUAD
2D beam
(a) RIBRICK (b)
25 25
0.000 0.005 0.010 0.015 0.020 0.025 -300 -100 100 300 500
Deflection (m) Bending moment (kNm)

Figure 3.12: Comparison of a series of 2D and 3D analyses from O’Rourke (1975). (a) Soldier
pile deflection. (b) Bending moments.

Figure 3.12a compares the computed wall deflection to the O’Rourke’s

measured final deflection. As can be seen, all the computed deflections agree

reasonably well with one another and with the field measurements. The same can be

said of the bending moment diagrams in Figure 3.12b. This indicates that the results of

the FE analyses are reasonably realistic. Nonetheless, it is interesting to note that the

3D FE analysis predicts a slightly larger maximum soldier pile deflection than the 2D

analyses; in spite of the fact that the “smeared” stiffness of the soldier pile wall in the

2D analysis does not take into account the contribution from the timber lagging. In

other words, if the 3D analysis is taken as a reasonably accurate representation of

reality, then the wall deflection from the 2D analysis represents an optimistic, rather

50
IDEALISED ANALYSIS OF PILE-SOIL INTERACTION

than average, estimate of the soldier pile and timber lagging deflection. The reason for

this will be examined in greater detail later.

0 0 0

5 5 5
Depth (m)

Depth (m)

Depth (m)
10 10 10

15 Field (O’Rourke, 1975) 15 15


IN4B
IN5B
Soldier pile location
100mm lagging
20 50mm lagging 20 20
Mid-span location
100mm lagging
50mm lagging
25 25 25
0.00 0.01 0.02 0.03 0.00 0.01 0.02 0.03 0.00 0.01 0.02 0.03
Deflection (m) Deflection (m) Deflection (m)
(a) (b) (c)
Figure 3.13: Deflection of soldier pile and soil at mid-span with respect to excavation
sequence. (a) Excavation depth at 6.1m. (b) Excavation depth at 11.9m. (c) Excavation depth
at 18.3m.

Figures 3.13a – c show the deflection profiles leading to various excavation

levels. As can be seen, more substantial differences between measured and predicted

wall deflection are present in the earlier stages. This may be attributed to the fact that

soil behaviour is not truly linearly elastic prior to yielding. The lower strain levels,

which are incurred in the early stages, could have allowed higher soil stiffness to be

mobilised than in the final stages. The timber lagging thickness of 50mm and 100mm

leads to nearly the same result, thus indicating that the results are not heavily

contingent upon the assumed thickness of timber lagging. In both cases, the computed

deflection of the timber lagging is significantly larger than that of the soldier pile

deflection. This is to be expected, since the timber lagging is much more flexible than

the soldier piles and thus expected to deflect more. In other words, significant

51
IDEALISED ANALYSIS OF PILE-SOIL INTERACTION

differences can arise between the soldier pile deflection and that of the intervening

soil.

10
Depth (m)

15

20
2D RIQUAD
3D RIBRICK no smearing
3D RIBRICK smeared stiffness
25
0.000 0.005 0.010 0.015 0.020 0.025
Deflection (m)

Figure 3.14: Effect of ‘smearing’ on rotational fixity below excavation level.

Figure 3.13 also shows that, just below excavation level, there is a short

segment of the soldier pile that moves ahead of the soil. This is to be expected; the

soldier pile wall is not a continuous wall below excavation level. The presence of soil

between adjacent soldier piles allows the piles to move relative to the intervening soil;

this movement being accentuated if the soil yields. In 2D analysis, the discontinuous

nature of the wall cannot be modelled since a uniform wall with a reduced “smeared”

stiffness is assumed instead. This suppresses the relative pile-soil movement, thus

exaggerating the degree of rotational fixity afforded by the pile embedment. As shown

in Figure 3.14, using a “smeared” stiffness for the soldier pile and the intervening soil

in a 3D analysis leads to a deflection profile that is virtually identical with that from a

2D analysis, thus indicating that the difference in deflection between 2D and 3D

52
IDEALISED ANALYSIS OF PILE-SOIL INTERACTION

analyses noted earlier arises from the representation of the non-uniform stiffness of the

soldier pile and intervening soil.

10
Depth (m)

Field (O’Rourke, 1975)


IN4B
IN5B
φ = 33°
15 c = 12.41kPa
c = 5.0kPa
φ = 30°
c = 5.0kPa
c = 2.0kPa
20 c = 0.1kPa

25
0.000 0.005 0.010 0.015 0.020 0.025
Deflection (m)
Figure 3.15: Modelling the S-shaped deflection profile of the soldier pile in 3D.

It is interesting to note that the S-profile present in O’Rourke’s field data was

not predicted in the FE analyses. One possible reason for this is the high values of

cohesion and friction angle prescribed by O’Rourke (1975) for the top layers of soil.

O’Rourke reported significant variation in the modulus of the Upper Brown sand from

about 41MPa to about 97MPa. This is suggestive of significant variation in soil

properties with location, which is not considered in the analysis wherein uniform soil

properties are assumed. As shown in Figure 3.15, progressive reduction in c and φ

values does lead to an S-shape soldier pile deflection profile.

53
IDEALISED ANALYSIS OF PILE-SOIL INTERACTION

3.3 3D and 2D comparisons

3.3.1 Idealised excavation

The complexity of the soil profile and multi-level bracing at the G Street

excavation renders an in-depth study of pile-soil interaction difficult. For this reason,

an idealized excavation involving two levels of bracing in a single soil type is studied,

as shown in Figure 3.6. To enhance the long-term stability of the soil face, 3-inch

thick timber laggings were used. As can be seen, the two levels of bracing were set at

4.5m and 7.5m below the original ground level. The two-levels bracing’s

configuration is suggested to study the stress changes due to strutting at different

depths. Boundary conditions are similar to those used in the back-analysis of the G

Street excavation. Initial groundwater table is assumed to be 2m below original

ground level, this being a typical groundwater table depth in Singapore conditions.

The corresponding 2D mesh is not shown since it is merely a vertical section of the 3D

mesh.

Table 3.3 shows the properties of soil and structural components used in the

analyses. The analyses conducted are effective stress analyses and the soil properties

prescribed are representative of granitic sapprolites. Such geologic material occurs

widely in Singapore (Dames and Moore, 1983) and can be described as a silt of

intermediate or high plasticity (BS5930, 1999). Soldier piles have been used as a

retaining system in such soils (e.g. Wong et al., 1997). The modified Cam-clay model

is used to describe the behaviour of the soil skeleton since it allows soils with different

over-consolidation ratios (OCRs) to be modelled while requiring relatively few

parameters compared to other, more complex models. The lack of detailed

54
IDEALISED ANALYSIS OF PILE-SOIL INTERACTION

information about the soil to be modelled does not justify the use of more sophisticated

models in such an idealized study.

Table 3.3: Idealised soil and soldier pile properties for idealised study.
Soil layer 1 (Depth: 0-22.5m) γs = 19.5 kN/m3
Modified Cam-clay cc = 0.1
cs = 0.01
M = 1.2
ecs = 1.47
ν = 0.3
K0 = 0.94
Soil layer 2 (Depth: 22.5-37m) γs = 19.5 kN/m3
Elastic E = 120000 kN/m2
ν = 0.3
Timber lagging at 3in thickness Eh = 16.3×103 kN/m2
Elastic Ev = 1 kN/m2
νhh = 0.02
ν vh= 2.2×10-5
Ghv = 0.37 kN/m2
Soldier pile modelled with 2D beam element E = 1.09×108 kN/m2
Ixx = 1.259×10-3 mm4
ν = 0.29
A = 0.019 m2
Half soldier pile modelled with RIBRICK at 3B pile spacing E = 2.18×107 kN/m2
(610×152.5mm) ν = 0.02
G = 1.07×107 kN/m2
Half soldier pile modelled with RIBRICK at 6B pile spacing E = 4.36×107 kN/m2
(610×152.5mm) ν = 0.02
G = 2.14×107 kN/m2
Strut at 3.75m spacing modelled with beams E = 2.67×107 kN/m2
Level 1 at 165.0kN/m Ixx = 1.259×10-3 mm4
Level 2 at 441.1kN/m ν = 0.29
A = 0.019 m2
Half strut with stiffness equivalent to 3.75m spacing modelled with E = 1.22×107 kN/m2
beams for cases of 3B pile spacing Ixx = 1.259×10-3 mm4
Level 1 at 75.5kN Iyy = 9.308×10-5 mm4
Level 2 at 201.8kN ν = 0.29
A = 0.019 m2
Half strut with stiffness equivalent to 3.75m spacing modelled with E = 2.44×107 kN/m2
beams for cases of 6B pile spacing Ixx = 1.259×10-3 mm4
Level 1 at 151.0kN (30-60%) (Wong et al., 1997) Iyy = 9.308×10-5 mm4
Level 2 at 403.6kN (40-70%) ν = 0.29
A = 0.019 m2
Note: Walers are assumed to be sufficiently stiff and able to distribute the load evenly to all soldier piles and thus not modelled in
the idealized analyses.

Table 3.4: Summary of effective stress analyses for the idealised excavation at OCR = 3 and
k = 1×10-8m/s.
Analysis Geometric representation Preloading
2D-1 Plane strain No
2D-2 Plane strain Yes
3D-1 3B No
3D-2 6B No
3D-3 3B Yes
3D-4 6B Yes
Note: B is the width of the soldier pile.

55
IDEALISED ANALYSIS OF PILE-SOIL INTERACTION

Figure 3.16 shows the excavation and wall installation sequences of the

idealized excavation. For the non-preloaded cases, the preloading stage is simply

omitted in the analyses. As shown in Table 3.4, the effects of preloading and centre-

to-centre inter-pile spacing on the discrepancy between 2D and 3D analyses are

examined.

Excavation stage
A BC DE F G
0
1
2
3 Install strut level 1

4
Depth (m)

5
6 Install strut level 2

7 Preload strut level 1


8
9
10 Preload strut level 2
Lull period
11 (20 days)
12
0 10 20 30 40 50 60 70 80 90 100 110 120
Time (days)
Figure 3.16: Construction sequence for idealized excavation from stage A through G.

3.3.2 Wall deflection

Figure 3.17 compares the 2D and 3D computed wall deflection profiles for the

non-preloaded soldier piles at the final stage of construction. As can be seen, for inter-

pile spacings of 3 and 6 times the soldier pile width B, the 3D analyses predict slightly

larger maximum soldier pile deflection than the 2D analysis, this being consistent with

the trend indicated for the G-Street excavation. The deflection profile at the mid-span

of the timber lagging is more variable. For an inter-pile spacing of 3B, the timber

lagging deflection remains reasonably well coupled to the soldier pile deflection. For

an inter-pile spacing of 6B, the timber lagging deflection becomes much larger than

56
IDEALISED ANALYSIS OF PILE-SOIL INTERACTION

the soldier pile deflection. It should be noted that an inter-pile spacing of 6B is not

unrealistic; in the G-Street excavation (O’Rourke, 1975), the inter-pile spacing was

approximately 6.5B. Comparison of the soldier pile and timber lagging deflection

profiles shows that, above excavation level, the mid-span deflection is larger than the

soldier pile deflection, which is consistent with the fact that the soldier pile is

supporting the soil face through the timber lagging and the soil arch spanning across

adjacent piles. On the other hand, the reverse is the case just below excavation level.

This occurs because, whereas the soil face at the mid-span can adopt rather sharp

curvature, the soldier pile cannot, owing to its much higher flexural stiffness. Thus,

soldier pile deflection can only decrease gradually below excavation level thereby

resulting in higher deflection in the pile than the mid-span soil. This is similar to the

3D FE results on the deflection of pile groups adjacent to surcharge, reported by

Bransby and Springman (1996).

10
Depth (m)

15
2D-1
Soldier pile location
3D-1
3D-2
20 Mid-span location
3D-1
3D-2

25
0.00 0.01 0.02 0.03 0.04 0.05
Deflection (m)
Figure 3.17: Comparison of wall deflection between 2D and different pile spacing for cases
without preloading at final stage of construction.

57
IDEALISED ANALYSIS OF PILE-SOIL INTERACTION

Figure 3.18 shows the wall deflection profiles of the preloaded cases at the final

stage of construction. As can be seen, preloading pushes the soldier pile backwards

into the retained soil, this combining with the excavation below strut level to produce

the S-shape profile that is also present in O’Rourke’s measurement, but not in the

idealised non-preloaded case discussed in the previous section.

10
Depth (m)

15
2D-2
Soldier pile location
3D-3
20 3D-4
Mid-span location
3D-3
3D-4
25
0.00 0.01 0.02 0.03 0.04 0.05
Deflection (m)
Figure 3.18: Comparison of wall deflection between 2D and different pile spacing for cases
with preloading at final stage of construction.

0.030
Stage
B
0.025
Deflection (m)

D
0.020

0.015 C

0.010
0.0 0.4 0.2 0.6 0.8 1.0
Span (m)
Figure 3.19: Comparison of effects of preloading on wall deflection for 3D-4 at 4.5m below
ground level with stages B, C and D shown in Figure 3.16.

58
IDEALISED ANALYSIS OF PILE-SOIL INTERACTION

Figure 3.19 compares the differences in deflection at the excavation face 4.5m

below ground level before and after preloading. As shown in Figure 3.16, in stage B,

excavation just reaches a depth of 4.5m. At this point, the deflection of the soldier pile

is larger than that of the intervening soil, owing to the large flexural stiffness of the

wall, as discussed earlier. In stage C, the process of preloading pushes soldier pile and

the soil rearwards, but with the soldier pile displacing more. Subsequent excavation

(stage D) allows the soil to move forward more than the soldier pile as the latter is

restrained by the strut and waler system.

3.3.3 Stress changes

This interaction between soil and soldier pile is illustrated in Figures 3.21 to

3.26, which show the stress paths for two points in the retained soil at the upper strut

level, which is located at a depth of 4.22m below ground surface. As shown in Figure

3.20, one point (marked I) is located just behind a soldier pile whereas the other

(marked II) is located just behind the mid-span of the timber lagging. Results for a

corresponding point at the same depth from a corresponding 2D analysis are also

plotted for comparison.

(I)

z (II)

Figure 3.20: Location of stress paths plotted for soil behind first level of strut at 4.22m below
ground level in x-z plane.

59
IDEALISED ANALYSIS OF PILE-SOIL INTERACTION

The deviator stress q is strictly a positive quantity. In Figure 3.21, a modified

quantity q̂ is instead plotted, such that

q̂ =
(σ '
y − σ 'x )q (3.1)
σ − σ 'x
'
y

where σx’ and σy’ are the horizontal and vertical effective stresses

The use of q̂ in place of q allows an active stress state (σy’ > σx’) to be distinguished

from a passive stress state (σx’ > σy’).

120

100

80
Initi
al y
ield
60 locu
G s
D,E,F
40
Deviator stress, q^

20

0 A

-20

-40 B,C
-60

-80

-100 CS
L
-120

0 10 20 30 40 50 60 70 80 90 100 110 120


Mean effective stress, p'
Figure 3.21: Stress path plot for 2D-1 without preload. Stages A to G are shown in Figure
3.16. Units of stresses in kPa.

As Figure 3.21 shows, for the 2D analysis, initial excavation leads to an

increase in mean effective stress due to the drag from the soil flow above the

monitored point. However, this quickly fades away as the formation level approaches

the monitored level [end point B] and lateral effective stress starts to decrease.

60
IDEALISED ANALYSIS OF PILE-SOIL INTERACTION

Meanwhile, q continues to increase as shear stresses build up in the retained soil.

Further excavation below the formation level leads to a reversal in the sign of q̂ [end

point D] as the lateral stress falls below the vertical stress, leading to an active stress

state. Comparison of Figures 3.21 and 3.22 shows that preloading causes the stress

state to remain in a passive, rather than active, state throughout the entire construction

sequence, thereby eliminating the passive-active reversal predicted for the non-

preloaded case.

120

100

80
Init
ia l yie
60 ld l
oc u
s
40
Deviator stress, q^

20

0 A

-20
C
-40 B D
E
F,G
-60

-80

-100 CS
L
-120

0 10 20 30 40 50 60 70 80 90 100 110 120


Mean effective stress, p'
Figure 3.22: Stress path plot for 2D-2 with preload. Stages A to G are shown in Figure 3.16.
Units of stresses in kPa.

As Figures 3.23 and 3.24 show, the initial stress changes behind the soldier pile

(point I) and the timber lagging (point II) are similar to that of the 2D prediction in

Figure 3.21. At both locations, initial excavation leads to an increase in mean effective

stress and a passive stress state due to the drag from the soil flow above the monitored

point. As excavation approaches the monitored level, the release in lateral stress due

61
IDEALISED ANALYSIS OF PILE-SOIL INTERACTION

to soil removal in front of the wall leads to an active stress state [end points B and C].

However, further excavation below the first strut level produces very different stress

paths at the two monitored locations. As shown in Figure 3.23, the build-up of lateral

stress at the base of the soil arch between adjacent soldier piles results in a passive

stress state, which persists up to the end of the construction sequence. On the other

hand, as negative pore pressure in front of the soil arch dissipates, the effective stress

in the soil behind the timber lagging decreases as the soil approaches a state of

incipient active failure at the end of the construction sequence. In physical terms, this

suggests significant softening of the soil behind the timber lagging. These stress path

differences are accentuated as the soldier pile spacing increases from 3B to 6B.

120

100

80
Init
ial
60 yie
ld loc
[B,C] 3B,6B us
40
Deviator stress, q^

20

0 A

-20

-40 [F,G] 3B
[D,E] 3B,6B
-60 [F,G] 6B
-80 Spacing
3B
-100 CS
L 6B
-120

0 10 20 30 40 50 60 70 80 90 100 110 120


Mean effective stress, p'
Figure 3.23: Stress path plots for 3D-1 and 3D-2 at point I without preload. Stages A to G are
shown in Figure 3.16. Units of stresses in kPa.

62
IDEALISED ANALYSIS OF PILE-SOIL INTERACTION

120

100

80 [B,C] 6B
In i
60 tia
[B,C] 3B l yie
ld
l oc
us
40
[D,E,F] 6B
Deviator stress, q^

20
[D,E,F] 3B
0 A

-20

-40

-60

-80 Spacing
3B
-100 CS
L
6B
-120

0 10 20 30 40 50 60 70 80 90 100 110 120


Mean effective stress, p'
Figure 3.24: Stress path plots for 3D-1 and 3D-2 at point II without preload. Stages A to G are
shown in Figure 3.16. Units of stresses in kPa.

120
Spacing
100 3B
6B
80
Init
i al y
60 ie ld l
ocu
s
[B] 6B
40 [B] 3B
Deviator stress, q^

20

0 A

-20
[C] 6B
-40 [C] 3B
[D] 3B [D] 6B
-60 [E,F,G] 3B
[E] 6B
-80
[F,G] 6B
-100 CS
L

-120

0 10 20 30 40 50 60 70 80 90 100 110 120


Mean effective stress, p'
Figure 3.25: Stress paths plots for 3D-3 and 3D-4 at point I with preload. Stages A to G are
shown in Figure 3.16. Units of stresses in kPa.

63
IDEALISED ANALYSIS OF PILE-SOIL INTERACTION

120

100

80 [B,C] 6B
Ini
60 tia
[B,C] 3B ly
iel
d loc
40 us
[D,E,F] 6B
Deviator stress, q^

20
[D,E,F] 3B
0 A

-20

-40

-60

-80 Spacing
3B
-100
CS
L
6B
-120

0 10 20 30 40 50 60 70 80 90 100 110 120


Mean effective stress, p'
Figure 3.26: Stress path plots for 3D-3 and 3D-4 at point II with preload. Stages A to G are
shown in Figure 3.16. Units of stresses in kPa.

Comparison of Figures 3.25 and 3.26 with Figures 3.23 and 3.24 shows that

preloading does not have significant effect that it has on the 2D analysis. The only

significance changes that arise from preloading occur behind the soldier pile and

consist of an earlier reversal back to a passive stress state [before end point C] and a

slightly larger mean effective stress at the end of the construction sequence. Both of

these are directly attributable to the preload. There is no significant difference at

location II.

Comparison of Figure 3.21 with Figures 3.23 and 3.24 as well as Figure 3.22

with Figures 3.25 and 3.26 shows that, at any given end point, the stress state predicted

by the 2D analysis appears to lie somewhere between those of locations I and II in the

3D analysis. This is not surprising in view of the fact the retaining effect of the soldier

pile and timber lagging is “smeared” in a 2D analysis. The large changes in the 2D

64
IDEALISED ANALYSIS OF PILE-SOIL INTERACTION

stress paths between Figures 3.21 and 3.22 appears to be due to the fact that “smeared”

stress path of the 2D analysis is affected by the small variations in the strongly

divergent stress states of the two locations.

3.4 Synopsis on 3D and 2D analyses

Strutted excavations, including soldier-piled excavations, are often analysed

using 2D FE analyses with properties that are averaged over a certain span of the wall.

The foregoing discussion shows that when such an approach is applied to soldier-piled

excavations, modelling errors can arise in several ways. Firstly, by modelling the

discrete soldier piles as a wall, a 2D analysis tends to over-predict the coupling of pile

to the soil below excavation level. In instances where the pile is not embedded into

stiff soil, this modelling error can give rise to discernible underestimation of soldier

pile deflection. Secondly, the deflection of the timber lagging, which is usually larger

than that of the soldier piles, is often underestimated. Thirdly, a 2D analysis cannot

replicate the differences in stress paths taken by the retained soil at different locations

at the same depth. In particular, it cannot model the softening of the soil face just

behind the timber lagging. In instances where the timber lagging is insufficient or its

delayed installation, this softening can give rise to local collapse, which may

exacerbate the ground loss and therefore ground movement. Increasing the inter-pile

spacing will tend to accentuate the effects of these modelling errors.

65
GEOTECHNICAL STUDY OF SERANGOON MRT SITE

CHAPTER 4 GEOTECHNICAL STUDY OF SERANGOON MRT


SITE

The North-East Line is a new mass rapid transit line that will run from the

World Trade Centre in the south of the Singapore island towards the North-East. As

shown is Figure 4.1, it passes through the Central Business District, and following the

corridors of Serangoon Road and Upper Serangoon Road up to the new towns of

Hougang, Sengkang and Punggol. Outram Park and Dhoby Ghaut stations are two

existing stations where the new North-East Line will criss-cross an existing line. The

line is approximately 20 kilometres (km) long and it will be built primarily

underground. It will consist of 16 stations and one depot when fully completed.

Figure 4.1: Singapore north-east MRT line with location plan.

4.1 Serangoon MRT site geology

The Serangoon Mass Rapid Transit (MRT) station will be located underground

at the junction of Upper Serangoon Road and Serangoon Central. In order to study in

detail the behaviour of soldier piles and timber lagging retaining system, field

66
GEOTECHNICAL STUDY OF SERANGOON MRT SITE

measurements have been performed throughout the project. Inclinometers were

installed around the perimeter of Serangoon station. Strain gauges were also installed

at each layer of struts. The depth of excavation reaches 26m below existing ground

level. The support system employed is strutted soldier piles and timber lagging.

The Serangoon MRT station is situated on a ridge of Bukit Timah Granite

Formation, which refers to an entire suite of igneous rocks, principally granite,

adamellite and granodiorite (PWD, 1976). The Central Singapore Granite is a granitic

intrusion which occupies an area in the centre of Singapore island extending some 8

km in the northerly direction and 7 km in the westerly direction where it forms hills

and valleys of both high and low relief. In general, the granite revealed in the MRT

projects had been of grades V and IV. Some limited volumes of grades III, II and I

had been found. The gradation from V to II can be very sudden in the weathering

profile. Weathering of granite is primarily due to chemical action involving the

breakdown of its primary minerals. The resultant soil is mainly sandy and/or clayey

silt. Figures 4.2 and 4.3 show the typical geological profiles at Serangoon station. The

two profiles do not register abrupt changes across the sections. The layout of

boreholes sunk at the Serangoon station site are shown in Figure 4.4.

The Dames & Moore’s report (1983) classified the Bukit Timah formation into

four sublayers:

a) G1: refers to fresh to slightly weathered Bukit Timah Formation

(Weathering Grades I and II);

b) G2: refers to moderately to highly weathered Bukit Timah Formation

(Weathering Grades III and IV);

67
GEOTECHNICAL STUDY OF SERANGOON MRT SITE

c) G3: refers to bouldery soil with boulders exceeding 400mm average

dimension surrounded by completely weathered rock and residual soil

(Weathering Grades III and IV);

d) G4: refers to completely weathered rock or granitic sapprolites in the

form of residual soil (Weathering Grades V and VI);

Soil investigation results indicate that Serangoon Station is underlain by G4 soil

up to a depth of 60m below existing ground level. The geological formation at

Serangoon station can be further divided into the following five sub-units:

a) Layer 1 (Fill & G4a): Stiff Silty Clay - fill is underlain by reddish

brown to light brown silty clay. The SPT N values range from zero to

15 blows/300mm.

b) Layer 2 (G4b): Medium dense Silty Clay to Clayey Silt - medium

dense whitish pink silty clay to clayey silt, with traces of sub-angular

to sub-rounded, fine medium size gravels. The SPT N values range

from 15 to 30 blows/300mm.

c) Layer 3 (G4c): Medium dense to dense Silty Sand - medium dense to

dense dark green silty sand with some fine quartz gravel. The SPT N

values ranges from 30 to 50 blows/300mm.

d) Layer 4 (G4d): Very dense Silty Sand - a layer of very dense greenish

grey silty sand with angular to sub-angular quartz gravel overlies the

weathered granite material. The SPT N values range from 50 to 100

blows/300mm.

68
GEOTECHNICAL STUDY OF SERANGOON MRT SITE

Figure 4.2: Soil profile of Serangoon MRT site at section A-A.

Figure 4.3: Soil profile of Serangoon MRT site at section B-B.

69
Zone 1 Zone 2 Zone 3 Zone 4 Zone 5

Borehole
Additional borehole
Inclinometer

70
0 5 10 20 30
GEOTECHNICAL STUDY OF SERANGOON MRT SITE

Figure 4.4: Plan view of Serangoon MRT station.


GEOTECHNICAL STUDY OF SERANGOON MRT SITE

e) Layer 5 (G4e): Weathered Granite - weathered granite is pinkish grey

to light grey with dip of joint ranging from 20-60 degrees. The SPT N

values are greater than 100 blows/300mm.

Figure 4.5 presents the variations of bulk unit weight and Atterberg Limits

versus depth of G4 soil at Serangoon MRT site.

0 0
Plastic Limit
Liquid Limit

-10 -10

-20 -20
Depth (m)

Depth (m)

-30 -30

-40 -40

(a) (b)
-50 -50
16 18 20 22 0 20 40 60 80 100
Unit Weight (kN/m3) Atterberg Limits (%)
Figure 4.5: Variation of physical properties at Serangoon MRT site. (a) Bulk unit weight. (b)
Atterberg Limits.

60
LL = 50%

CH
50
e
Plasticity index, PI

40 Lin
CL A

30

MH & OH
20

10
ML MH & OL
0
0 10 20 30 40 50 60 70 80 90 100
Liquid limit, LL
Figure 4.6: Plasticity chart of G4 soil at Serangoon MRT station.

71
GEOTECHNICAL STUDY OF SERANGOON MRT SITE

Figure 4.7: Grain size distribution of G4 soil at Serangoon MRT station.

Figures 4.5 and 4.8 show the variation of some of the basic properties of the G4

soils with depth. The Atterberg limits of this soil put it in the class of medium

plasticity inorganic silt as shown in Figure 4.6. This is consistent with the particle size

distribution data in Figure 4.7, which shows a predominant particle size range varying

from sand to silt. As shown in Figures 4.8a, b and c, the SPT blow counts N,

undrained Young’s modulus and undrained shear strength show a gradually increasing

trend with depth, rather than stepwise fashion, thus indicating that the transition from

one layer to the next is often not abrupt. This is consistent with the c’ values, which do

not show any significant increase that would suggest an increase in the degree of

structure or bonding. The significant scatter in the values of effective cohesion c’ and

friction angle φ’ can be explained by the fact that most of the Mohr Circle plots in the

site investigation reports (OYO, 1996; Ove Arup & Partners, 1997) imply a slightly

curved failure envelope, which has been noted by many researchers; e.g. Roscoe et al.

(1958), Roscoe et al. (1963) and Schofield & Wroth (1968), rather than the straight

72
GEOTECHNICAL STUDY OF SERANGOON MRT SITE

line fit prescribed by the Mohr Coulomb envelope. Thus, the derived value of c’ and

φ’ depends somewhat on how the straight line is fitted to the curve.

The overconsolidation ratio values of G4 soil decrease from 4 near the ground

surface to 1 at depth of 20m. In the site investigation for this project, the OCR was

measured using oedometer test. Chong (1998) showed that sampling disturbance tends

to cause a significant reduction in the values of parameters related to the yield surface

of the soil, which includes the preconsolidation pressure pc’. Thus, it is possible that

the measured values of the OCR represent a lower bound of the actual OCR in the

field. It should also be mentioned that the presence of a preconsolidation pressure

which is higher than existing overburden effective stress does not suggest

overconsolidation in the actual sense of the word, since the soil is derived from

weathering of rock rather than consolidation of deposited specimens. For this reason,

the apparent OCR so measured may be more related to the structure of the soil,

including bonds, rather than actual overconsolidation. Various authors (e.g. Terzaghi

et al., 1996; Balasubramaniam & Brenner, 1981; Brumund et al., 1976; Bjerrum, 1973)

have shown that naturally and artificially cemented soils do possess an elevated

preconsolidation pressure.

73
GEOTECHNICAL STUDY OF SERANGOON MRT SITE

0 0 0

-10 -10 -10

-20 -20 -20


Depth (m)

Depth (m)

Depth (m)
-30 -30 -30

-40 -40 -40

(a) (b) (c)


-50 -50 -50
0 20 40 60 80 100 0 100 200 300 400 500 600 0 100 200 300
SPT (N) E (MPa) su (kPa)

0 0 0

-10 -10 -10

-20 -20 -20


Depth (m)

Depth (m)

Depth (m)

-30 -30 -30

-40 -40 -40

(d) (e) (f)


-50 -50 -50
0 10 20 30 40 0 10 20 30 40 0 1 2 3 4
c' (kPa) φ' (°) OCR
Figure 4.8: Variation of soil properties with depth at Serangoon MRT site. (a) SPT values. (b)
Undrained Young’s modulus. (c) Undrained shear strength. (d) Effective cohesion. (e)
Effective friction angle. (f) Over-consolidation ratio.

Permeability assessment employing the variable head method was conducted

for soil layer at track level (depth ranges from 14-20m). Figure 4.9 summarises the

permeability of G4 soil with depth at the MRT site. At borehole NA124 (RL 101.5m)

and BH11 (RL 85.63m), the permeability of Clayey Silt layer registered is 1.70×10-
7
m/s and 2.81×10-8m/s respectively. At a borehole 1.7km away, the permeability in the

Silty Sand layer (RL 92.6m) is 2.42×10-7m/s. The permeability in the vicinity lies in

74
GEOTECHNICAL STUDY OF SERANGOON MRT SITE

the order of 10-7m/s in the top two layers of soil. The values obtained from the site

represent the upper bound permeability indicated in Dames & Moore (1983) for G4

soil. The average overall permeability recommended by Dames & Moore is 1×10-8m/s

with a lower bound of 4×10-10m/s. Lee et al. (1993) reported the typical permeability

of G4 materials to be between 10-7 and 10-9m/s. The water table exists at 5m below

ground. The lowering of water head monitored ranges from 0.1m/day to 1.1m/day for

all stages of excavation.

-10

-20
Depth (m)

-30

-40

-50
0 5 10 15 20
k ( x10-7 m/s)
Figure 4.9: Variation of soil permeability with depth at Serangoon MRT site.

Table 4.1 summarises the soil properties for the various sub-layer of G4 soil at

Serangoon MRT site.

Table 4.1: Soil properties of Serangoon MRT site.


Sub-layer γs E’ c’ φ’ k OCR
(kN/m3) (MPa) (kPa) (m/s)
G4a 18.5 5.0 + 5.5z 11 27° 1.4×10-7 2-4
G4b 19.0 80.0 + 5.0z 17 30° 2.1×10-7 1-2
G4c 19.5 500.0 + 4.5z 25 38° 9.7×10-8 1
G4d 20.0 2000.0 + 2.2z 30 38° 7.8×10-8 1
G4e 21.0 30000.0 + 2.2z 35 40° 4.0×10-8 1
Note: Young’s moduli are expressed as variation with depth, z.

75
GEOTECHNICAL STUDY OF SERANGOON MRT SITE

4.2 Retaining system

The retaining system used in this project is soldier piles with timber lagging.

Figure 4.10 shows the retaining system at the MRT site after installation of first layer

of strut in Zone 2. The soldier piles are 610mm × 305mm × 149kg/m H-pile (BSI,

1996) driven at 1.5m centre to centre shown in Figure 4.11. The struts are located

7.5m apart at Zone 2. The first level of strut is 3.5m below ground level. The spacing

between each level of struts varies from 3-4m as illustrated in Figure 4.12. The raking

struts, 3.5m each, (250mm × 250mm × 64.4kg/m H-piles) are connected to the main

strut, one on each side, inclined at 45° as shown in Figure 4.10. The walers consist of

610mm × 305mm × 149kg/m H-piles. No pre-stressing was applied to the struts. The

timber lagging consists of 3in × 8in (76mm × 203mm) Kempas wood. Table 4.2

summarises the prop system from the excavation retaining system.

Table 4.2: Struts details for Serangoon MRT retaining system.


Level Strut Vertical Spacing Waler Raking Strut
1 H610×305 0.5m – 3.5m to ground level H610×305 H250×250
149kg/m 1.2m – 3.15m to 2nd layer strut 149kg/m 64.4kg/m
2 2H610×305 1.2m – 3.15m to 1st layer strut 2H610×305 H250×250
149kg/m 4.0m to 3rd layer strut 149kg/m 64.4kg/m
3 2H610×305 4.0m to 2nd layer strut 2H610×305 H250×250
149kg/m 4.0m to 4th layer strut 149kg/m 64.4kg/m
4 2H610×305 4.0m to 3rd layer strut 2H610×305 H250×250
149kg/m 4.0m to 5th layer strut 149kg/m 64.4kg/m
5 2H610×305 4.0m to 4th layer strut 2H610×305 H250×250
149kg/m 3.0m to 6th layer strut 149kg/m 64.4kg/m
6 2H610×305 3.0m to 6th layer strut 2H610×305 H250×250
149kg/m 3.5m to excavation level 149kg/m 64.4kg/m

76
GEOTECHNICAL STUDY OF SERANGOON MRT SITE

Figure 4.10: Soldier pile and timber lagging retaining system of Serangoon station.

0 10 20 30

metres

EGL = 123.75m

EL = 106.35m

Figure 4.11: Plan view of retaining system for Serangoon MRT station.

77
GEOTECHNICAL STUDY OF SERANGOON MRT SITE

Depth
H610×305×149kg/
3.5m

2H610×305×149kg/m
6.5m

2H610×305×149kg/m
10.5m

2H610×305×149kg/m 14.5m

H610×305×149kg/m @ 1.5m c/c


2H610×305×149kg/m 18.5m

2H610×305×149kg/m
21.5m

25.0m

42.0m

Figure 4.12: Cross sectional view of retaining system for Serangoon MRT station.

4.3 Construction sequence

The excavation depths are 15m at the concourse and approximately 25m at the

platform. Propped soldier piles retaining walls were adopted to temporarily support

the proposed excavation. Construction followed a “bottom-up” type sequence with

steel frames providing temporary propping support as excavation proceeded

downwards. As the permanent works slabs were casted from the bottom upward the

temporary props were removed sequentially, with the floor slabs being used as

permanent props to the permanent walls.

The soldier piles were firstly driven to the designed level. Upon completion of

the soldier piles installation, excavation commenced cell by cell (zone by zone). The

excavation commenced from two ends of the station toward the centre. On reaching

78
GEOTECHNICAL STUDY OF SERANGOON MRT SITE

the designed level, strutting would follow, while excavation to the same level would

continue in the next cell. Thus, strutting was also performed successively from two

ends of station, i.e. Zones 1 and 5, towards the centre. This cycle was repeated until all

of the zones have been excavated to the designed level, after which excavation to the

next level would be carried out in a similar cycle. Timber lagging was installed after

the exposed soil face of the excavation reaches a depth of 1 to 1.5m. Following the

timber installation, further excavation up to another 1m to 1.5m below the installed

timber lagging was carried out. This cycle was repeated until the full excavation

completed. No back-filling of voids or dewatering was conducted in the retained

ground behind the soldier piles with the exception of Zone 1 in Serangoon Station site,

where problems of water ingress and local collapse of soil were encountered (Lee et

al., 2001).

The construction details are as follows:

a) After the installation of kingposts and soldier piles, excavate from

EGL to a depth of 2m below and install timber lagging;

b) Excavate to 5m below EGL and install timber lagging. Install 1st

layer strut at a depth of 3.5m below EGL;

c) Excavate to 8m below EGL and install timber lagging. Install 2nd

layer strut at a depth of 6.5m below EGL;

d) Excavate to 12m below EGL and install timber lagging. Install 3rd

layer strut at a depth of 10.5m below EGL;

e) Excavate to 16m below EGL and install timber lagging. Install 4th

layer strut at a depth of 14.5m below EGL;

79
GEOTECHNICAL STUDY OF SERANGOON MRT SITE

f) Excavate to 20m below EGL and install timber lagging. Install 5th

layer strut at a depth of 18.5m below EGL;

g) Excavate to 23m below EGL and install timber lagging. Install 6th

layer strut at a depth of 21.5m below EGL;

h) Excavate to 25m below EGL and install timber lagging.

80
FEM STUDY OF SERANGOON MRT SITE

CHAPTER 5 FEM STUDY OF SERANGOON MRT SITE

This chapter discusses the interaction between material and geometric

modelling of soldier piled excavations, using the measured data from Serangoon MRT

site as the reference. The computed behaviour of soldier piled excavations will be

examined using 2D and 3D analyses and with 3 soil models, namely Mohr Coulomb

criterion, modified Cam-clay and a hyperbolic Cam-clay model (Nasim 1999). Fitting

attempts and sensitivity studies will be made on the wall deflection and ground

settlement using the three models. By so doing, the performance of each type of soil

model in finite element analysis will be assessed. Additionally, a set of soil parameters

will be determined for each model that gives good agreement between computed and

measured movements. With these sets of parameters, 2D and 3D analyses will be

conducted to examine the stress changes and wall deflection development during

excavation and its final settlement profile.

5.1 Geometry of finite element model

The finite element model for the Serangoon MRT Station excavation is shown

in Figure 5.1. As shown in Figure 5.2 (section B-B’), the modelled section is located

at 20.4m from the nearest corner. The soil domain was modelled using FIBRICK

elements with pore-pressure degree of freedoms at vertex nodes. RIBRICK elements

without pore-pressure degree of freedoms were used for soldier piles for reasons

highlighted in Chapter 3. The struts and walers were modelled with beam elements.

As discussed in Chapter 3, the timber lagging is modelled using RIBRICK elements

with anisotropic elastic material model. Kingposts installed within the excavated area

are also modelled using beam elements as their cross-sectional dimensions are too

81
FEM STUDY OF SERANGOON MRT SITE

small to be modelled viably using solid elements. As will be shown later, the presence

of kingposts has a significant effect on the computed base heave.

As shown in the plan view of Figure 5.1, only half of the strutted section and

half of the intervening soil domain were modelled to take advantage of symmetry.

Timber lagging
Layer 1
Layer 2
Soldier pile Layer 3
Layer 4
Layer 5

Plan view of section A - A'

A A' Depth
0.0m
Strut level 1 3.5m
Strut level 2 6.5m
Strut level 3 10.5m
Strut level 4 14.5m
Strut level 5 18.5m
Strut level 6 21.5m
25.0m
28.5m
King (KP Toe)
Posts

Soldier 42.0m
piles (SP Toe)

79.0m
Cross sectional view
Figure 5.1: Finite element model for Serangoon MRT site excavation.

The modelling of boundary and groundwater conditions is similar to that

discussed in Chapter 3 and will not be repeated herein. The bottom boundary is

located in weathered granite rock which is likely to undergo much less movement than

the overlying strata (Britto and Gunn, 1990). Preliminary trials indicate that the results

82
Zone 1 Zone 2 Zone 3 Zone 4 Zone 5

Borehole
Additional borehole
Inclinometer

B’

83
0 5 10 20 30

Figure 5.2: Plan view of Serangoon MRT station indicating the section B-B’ modelled.
FEM STUDY OF SERANGOON MRT SITE
FEM STUDY OF SERANGOON MRT SITE

of the analyses are not unduly sensitive to the location of the bottom boundary. This is

consistent with the findings of Potts & Zdravković, (2001), who attributed it to the fact

that soil strength and stiffness usually increase with depth. Britto and Gunn (1990)

suggested that the vertical end boundary be placed at a distance of between 4 to 5

times the final depth of the excavation from the retaining wall. Liu’s (1995) finding

for Marine Clay using modified Cam-clay indicates that beyond a distance 3 times the

depth of excavation, the far-field boundary will have little effect on the settlement. In

the present analyses, the far-field vertical boundary was set at 4 times the depth of

excavation. The water table for the subsequent analyses was set at 5m below ground

surface in accordance with standpipes’ readings given in the soil investigation report.

5.2 Modelling of excavation sequence

As summarised in Figure 5.3, the excavation was simulated in 15 stages. Since

fully coupled consolidation analyses were conducted, the increments also modelled the

dissipation of pore pressure in the soil with time. As shown in Figure 5.3, the time

duration for each stage of construction were also modelled in the analyses. The

analyses were performed with the soldier piles already in place, as the driving of the

piles cannot be simulated by CRISP. The installation of the pile was itself a complex

process and its effects on the in-situ stress conditions could not be modelled into the

analyses. The excavation process was simulated by the removal of appropriate soil

elements from the finite element mesh.

84
FEM STUDY OF SERANGOON MRT SITE

Excavation stage
A B C D E F G H I
0
Install timber lag
Install strut level 1 and timber lag
5
Install strut level 2 and timber lag

10 Install strut level 3


Depth (m)

and timber lag


Install strut level 4
15 and timber lag

Install strut level 5


and timber lag
20
Install strut level 6
and timber lag
Install timber lag
25

0 50 100 150 200 250 300


Time (days)
Figure 5.3: Excavation with strut and timber lagging installation sequence.

5.3 Parametric studies for various models

Table 5.1 shows the soil properties used for each type of material model.

Comparison of Figure 5.1 and Table 5.1 shows that the entire excavation took place

within Layers 1 and 2, which correspond to the fill and medium dense silty clay/clayey

silt, the latter being residual soil derived from the weathering of the granitic rock. It is

thus likely that the modelling of these two soil layers will significantly affect the

computed results. For this reason, three types of soil model, namely the Mohr

Coulomb criterion, modified Cam-clay and strain-dependent hyperbolic Cam-clay

(Nasim, 1999) models, are considered for these two layers.

85
FEM STUDY OF SERANGOON MRT SITE

Table 5.1: Soil properties of Serangoon MRT site.


MC Depth (m) K0 γs (kN/m3) E’ (Mpa) c’ (kPa) φcs’ φp’ ν kx (m/s) ky (m/s)
Layer 1 0.0-11.5 0.9 18.5 2.7z + 9.2 11 27° 33° 0.30 1.4×10-7 1.4×10-7
Layer 2 11.5-22.5 0.9 19.0 3.4z – 8.3 17 30° 36° 0.28 2.1×10-7 2.1×10-7
Layer 3 22.5-37.0 1.0 19.5 10.8z – 172.8 27 - 38° 0.30 9.7×10-8 9.7×10-8
Layer 4 37.0-55.0 1.0 20.0 2.2z + 2000.0 30 - 38° 0.30 7.8×10-8 7.8×10-8
Layer 5 55.0-79.0 1.0 21.0 2.2z + 30000.0 35 - 40° 0.33 4.0×10-8 4.0×10-8
MCC Depth (m) K0 γs (kN/m3) κ λ ecs M ν kx (m/s) ky (m/s)
Layer 1 0.0-11.5 0.9 18.5 0.020 0.087 1.74 1.07 0.30 1.4×10-7 1.4×10-7
Layer 2 11.5-22.5 0.9 19.0 0.021 0.109 1.98 1.2 0.28 2.1×10-7 2.1×10-7
HCC Depth (m) K0 γs (kN/m3) C (kPa) n m qf (kPa) G∞ (kPa)
Layer 1 0.0-11.5 0.9 18.5 155 0.8 0.23 177.5 0
Layer 2 11.5-22.5 0.9 19.0 131 0.8 0.23 369.2 0
Note: Young’s moduli are expressed as variation with depth, z.
MC: Mohr Coulomb criterion
MCC: Modified Cam-clay
HCC: Hyperbolic Cam-clay

Based on the soil investigation reports by OYO (1996), Ove Arup & Partners

(1997) and Dames & Moore (1983), an initial estimate of soil parameters for the three

different models are set out in Table 5.1. As discussed below, the hyperbolic Cam-

clay model requires the use of the modified Cam-clay parameters as shown in Table

5.1, as well as five additional parameters, the value for which are shown in Table 5.1.

At larger depths, that is layers 3 to 5, where the soil consists of silty sand only the

elastic-perfectly plastic Mohr Coulomb criterion model is used owing to the limited

laboratory test data available for these deeper layers. Herewith, the elastic-perfectly

plastic Mohr Coulomb criterion model will be known, in short, as the Mohr Coulomb

model. All the three models assumed associated flow rule.

5.3.1 Elastic-perfectly plastic Mohr Coulomb

The Young’s modulus was estimated from Menard pressuremeter data obtained

in this project (OYO, 1996; Ove Arup & Partners, 1997) as well as geotechnical

reports on similar soil types by Dames & Moore (1983). The variation of the

undrained Young’s modulus with depth has been presented in Chapter 4. As shown in

Figure 5.4, in the Mohr-Coulomb analyses which follows, the variation of E’ with

depth for different soil layers were assumed to be piecewise linear. Following Lambe
86
FEM STUDY OF SERANGOON MRT SITE

and Whitman (1979), the effective Young’s modulus E’ is deduced from Eu using the

expression:

Eu (1 + ν ' )
E' = (5.1)
1 .5

10

Linear regression from


field data used in MC1

-10
Depth (m)

-20

-30

-40

-50
0 100 200 300 400 500
E' (MPa)
Figure 5.4: Assumed piecewise linear variation of E’ with depth for different soil layers.

The other elastic deformability property is the Poisson’s ratio. Relatively little

information is available in the literature on the Poisson’s ratio, ν’, of Singapore

weathered granite. As noted by Kulhawy (1990), this parameter often does not vary

greatly. For granular soils:

ν ' ≈ 0.1 + 0.3φ rel (5.2)

with

φ rel =
(φ '−25 ) 
(0 ≤ φ rel ≤ 1) (5.3)
(45 
− 25  )
87
FEM STUDY OF SERANGOON MRT SITE

in which φrel is the relative friction angle.

For cohesive soils, the effective Poisson’s ratio has been related to plasticity

index (PI) for several lightly overconsolidated soils (Wroth, 1975) as shown in Figure

5.5. For layers G4a and G4b, which consist of soils of clayey nature with plasticity

indices in the region of 25 and 17 respectively, the effective Poisson’s ratio was

estimated from PI using Figure 5.5. For layers G4c, G4d and G4e, where the soil is

sandy, Equations 5.2 and 5.3 are used to determine the Poisson’s ratio.

0.6
Drained Poisson's Ratio, ν'

0.4

0.2

0.0
0 20 40 60 80
Plasticity Index, PI(%)

Figure 5.5: Drained Poisson’s ratio versus plasticity index for several lightly overconsolidated
soils after Wroth (1975).

Wall deflection

Figure 5.6 shows the plan view of the FEM model in the vicinity of the wall.

The wall deflection was computed at two points, the first marked I located at a rear

corner of a soldier pile and the second marked II located behind the mid-span of the

timber lagging.

88
FEM STUDY OF SERANGOON MRT SITE

Table 5.2: Soil properties for Layers 1, 2 and 3 for Mohr Coulomb analyses.
Analyses E’ (MPa) c’ (kPa) φ’ K0
MC1 Layer 1 2.7z + 9.2 11 27° 0.9
Layer 2 3.4z – 8.3 17 30° 0.9
Layer 3 10.8z – 172.8 27 38° 1.0
MC2 Layer 1 5.6z + 0.5 11 27° 0.9
Layer 2 4.9z + 80.0 17 30° 0.9
Layer 3 4.5z + 500.0 27 38° 1.0
MC3 Layer 1 5.6z + 0.5 11 27° 0.9
Layer 2 4.9z + 100.0 17 30° 0.9
Layer 3 4.5z + 500.0 27 38° 1.0
MC4 Layer 1 5.6z + 0.5 5 37° 0.9
Layer 2 4.9z + 80.0 5 40° 0.9
Layer 3 4.5z + 500.0 27 38° 1.0
MC5 Layer 1 5.6z + 0.5 30 27° 0.9
Layer 2 4.9z + 80.0 30 30° 0.9
Layer 3 4.5z + 500.0 27 38° 1.0
MC6 Layer 1 5.6z + 0.5 11 27° 0.7
Layer 2 4.9z + 80.0 17 30° 0.7
Layer 3 4.5z + 500.0 27 38° 0.7
MC7 Layer 1 5.6z + 0.5 11 27° 1.0
Layer 2 4.9z + 80.0 17 30° 1.0
Layer 3 4.5z + 500.0 27 38° 1.0
ELAS Layer 1 5.6z + 0.5 - - -
Layer 2 4.9z + 80.0 - - -
Layer 3 4.5z + 500.0 - - -

Timber lagging

II
I
Timber lagging

Figure 5.6: Plan view of FEM model of retaining system.

Figure 5.7 shows two computed deflection profiles at final excavation level.

Deflection profiles MC1 are computed by adopting a Young’s modulus variation with

depth that is inferred from soil investigation using linear regression for each layer.

Deflection profiles MC2 are computed by adjusting the variation of Young’s modulus

that gives a good fit to the measured wall deflection. Comparison of the parameters

for MC1 and MC2 in Table 5.2 shows that, in order to match the field deflection using

Mohr Coulomb model, the Young’s modulus for the top three layers need to be

89
FEM STUDY OF SERANGOON MRT SITE

increased. A more detailed discussion will be carried out on this observation in the

next section together with modified Cam-clay model. Henceforth, the Mohr Coulomb

properties for MC2 Layer 3 will be used for the rest of the analyses to give a realistic

deflection below final excavation level.

10

15
Depth (m)

20

25

30 Field
MC1-I
35 MC1-II
MC2-I
40 MC2-II

45
0
100 200 300 400
Deflection (mm)
Figure 5.7: Final deflection profile of soldier pile and timber lagging for MC1, MC2.
Locations I and II for each case are defined in Figure 5.6.

Figure 5.8 shows the final deflection profile for two Mohr Coulomb cases

designated MC2 and MC3, and one linear elastic case, designated ELAS using the

same elastic parameters as MC2 (Table 5.2). Comparison of the results from MC2 and

ELAS shows that ELAS significantly underestimated the measured deflection of the

soldier pile. Furthermore, there was no significant relative movement between the

timber lagging and the soldier pile. This indicates that plastic deformation contributes

substantially to the computed wall deflection, as well as the relative movement

between soldier pile and timber lagging, in case MC2. Evidently, plastic deformation

and outward movement of the soil was able to occur at the timber lagging between

90
FEM STUDY OF SERANGOON MRT SITE

soldier piles in the Mohr-Coulomb model, whereas it was not allowed to occur in the

elastic model. This point will be further discussed later.

10

Depth (m) 15

20

25
Field
30 MC2-I
MC2-II
MC3-I
35
MC3-II
ELAS-I
40 ELAS-II

45
0 40 80 120 160
Deflection (mm)
Figure 5.8: Final deflection profile of soldier pile and timber lagging for MC2, MC3 and
ELAS. Locations I and II for each case are defined in Figure 5.6.

To assess the sensitivity of the results to the modulus estimated from Menard

pressuremeter data as well as geotechnical reports on similar soil types by Dames &

Moore (1983), the modulus of Layer 2 is increased in MC3 to investigate the effect of

the Young’s modulus in this study. As presented in Figure 5.8, increasing the Young’s

modulus between the depths of 11.5 and 22.5m reduces the wall deflection within this

depth interval. The difference is quite significant for the timber lagging deflection but

much less so for the soldier pile deflection.

Actual shear failure envelopes may exhibit slight curvature in certain cases

(Lambe & Whitman, 1979) but a straight line is often taken over the stress range of

interest and the shear strength parameters determined for that range (Craig, 1996).

This can result in several combinations of fitted cohesion and angle of friction,

91
FEM STUDY OF SERANGOON MRT SITE

depending upon the manner in which the fitting was conducted. As shown in Figure

5.9, a similar curvature exists for the soil at the Serangoon station. Furthermore,

depending upon the method of fitting the linear envelope to the Mohr Circles, the fitted

cohesion may range from 8kPa to 22kPa.

300
(a) c' = 8 kPa
Shear Stress (kPa)

φp' = 34°
200
c' = 22 kPa
φp' = 28°
100

0 100 200 300 400 500


Effective Stress (kPa)

480
(b) c' = 13 kPa
Shear Stress (kPa)

400 φp' = 36.5°


320
240
160
80
0

0 200 400 600 800


Effective Stress (kPa)
Figure 5.9: Apparent Mohr Coulomb yield envelopes at the limits of the effective stress range
of interest. (a) At 6m below EGL. (b) At 12m below EGL.

Table 5.2 shows the cohesion and friction parameters for MC2, MC4 and MC5,

with similar modulus for each layer. MC4 and MC5 represent the possible Mohr

Coulomb yield envelopes fitted at the two limits of the effective stress range of

interest, in addition to MC2, as illustrated in Figure 5.9a. In the site investigation for

this project, OYO (1996) considered the effective stress range of interest to be between

92
FEM STUDY OF SERANGOON MRT SITE

50 and 600kPa. The difference in effective cohesion between the upper bound and

lower bound of the range seldom exceed 5% of the effective stress range. The

effective friction angle at the two bounds is within the difference of 10°.

10

15
Depth (m)

20

25
Field
30 MC2-I
MC2-II
MC4-I
35
MC4-II
MC5-I
40 MC5-II

45
040 80 120 160
Deflection (mm)
Figure 5.10: Final deflection profile of soldier pile and timber lagging for MC2, MC4 and
MC5. Locations I and II for each case are defined in Figure 5.6.

Figure 5.10 shows the final deflection of the retaining wall for MC2, MC4 and

MC5, all of which have the same values of elastic parameters. As this figure shows, at

the soldier pile, that is location I, MC2 shows larger deflection compared to MC4 and

MC5. On the other hand, at the mid-span of the timber lagging, that is location II,

MC4 shows larger deflection than MC2 and MC5. This can be explained in terms of

the differences in effective cohesion and angle of friction. Figure 5.11a and b show the

effective p’-q stress paths for points I and II in analyses MC2, MC4 and MC5 at 12.3m

below ground. In Figures 5.11, q is treated as a strictly a positive quantity in

accordance with the definition of deviator stress. As can be seen, the effective stress

increases to a much higher level for point I than for point II when excavation

93
FEM STUDY OF SERANGOON MRT SITE

approaches the final level. This is attributable to stress arching across the soldier piles,

which has been discussed in Chapter 3. Thus, for the effective stress range at point I,

150
Final excavation
level

Deviatoric stress, q
100

50

(a)
0
0 50 100 150 200
Mean effective stress, p'
3D - behind soldier pile

200
Deviatoric stress, q

150

Final
100 excavation MC2
level MC4
MC5
MC2 yield locus
50 MC4 yield locus
MC5 yield locus

(b)
0
-100 0 100 200 300
Mean effective stress, p'
3D - behind timber lagging
Figure 5.11: Stress path plots for MC2, MC4 and MC5 at 12.3m below ground. (a) Location I.
(b) Location II. Locations I and II for each case are defined in Figure 5.6. Units of stresses in
kPa.

94
FEM STUDY OF SERANGOON MRT SITE

MC2 has the lowest failure envelope, and the readiness of the soil to yield is

manifested as the largest deflection at the soldier pile. On the other hand, for the

effective stress range at point II, MC4 has the lowest failure envelope. Thus, MC4

shows the largest timber lagging deflection.

10

15
Depth (m)

20

25

30
Field
35 2D-MC2
2D-MC4
40 2D-MC5

45
0
20 40 60 80 100
Deflection (mm)
Figure 5.12: 2D final deflection profile of soldier pile and timber lagging for MC2, MC4 and
MC5.

Figure 5.12 shows the wall deflection for the three cases as computed by 2D

analyses. As can be seen, the results from the 2D analyses show a similar trend as

those of the soldier pile, but not the timber lagging, in the 3D analyses. However, for

each case, the maximum deflection computed by the 2D analysis is slightly higher than

that computed by the 3D analysis. Thus, three-dimensional analyses of soldier piled

excavations highlights the presence of a much larger range of stress excursions over

the ground domain than is manifested in two-dimensional analyses. In order to capture

the appropriate relative behaviour between soldier pile and timber lagging in three-

dimensional analyses, more care is needed in the choice of parameters.

95
FEM STUDY OF SERANGOON MRT SITE

10

15

Depth (m)
20

25

30 MC2-I K0 = 0.9
MC2-II K0 = 0.9
35 MC6-I K0 = 0.7
MC6-II K0 = 0.7
MC7-I K0 = 1.0
40 MC7-II K0 = 1.0

45
0
40 80 120 160
Deflection (mm)
Figure 5.13: Final deflection profile of soldier pile and timber lagging for MC2, MC6 and
MC7. Locations I and II for each case are defined in Figure 5.6.

In the above analyses, the process of soldier pile installation was not modelled.

In reality, it is possible that installation of soldier piles may alter the initial stress state,

which will then be reflected by changes in the K0-values in the vicinity of the soldier

piles. It is likely that the changes in stress state will depend upon the method of soldier

pile installation. For example, Symons & Carder (1993) suggest that total stress

reductions of up to 10% may occur close to bored piles in stiff over-consolidated clay.

Ekström (1989) reported an approximately 20% increase of lateral earth pressure on

the surface a square concrete driven pile from in-situ lateral earth pressure measured

using a dilatometer before pile driving in non-cohesive soil.

In the Serangoon MRT Station excavation, the soldier piles were driven into the

ground. It is therefore reasonable to expect an increase in lateral earth pressure in the

vicinity of the pile. However, this increase is likely to be substantially less than that

arising from driving square piles, since the amount of displacement inflicted on the

96
FEM STUDY OF SERANGOON MRT SITE

surrounding soil due to the H-pile is much less. It is thus reasonable to assume an

upper bound of K0 equals 1.0 due to H-pile driving.

Figure 5.13 compares the variation in final wall deflection profile as the K0-

values changes from 0.7 to 1.0. As discussed above, this range of change is likely to

emcompass the changes in K0-values arising from the installation of soldier piles. In

the analyses, the variation in K0 is effected between the head and toe of the soldier

piles.

As can be seen, there is a slight increase in soldier pile and timber lagging

deflections with the increase in K0-value from 0.7 to 1.0. This is not surprising since

an increase K0 leads to an increase in initial lateral earth pressure. However, as in the

case of the other parameters, the changes are not significant. The fact that the mid-

span deflection is much larger than soldier pile deflection remains unchanged.

Figure 5.14 demonstrates the stress paths for the various K0 conditions. The

trend of stress paths for MC2 Mc6 and MC7 are almost similar. This shows that the

wall deflection behaviour and stress paths of Mohr Coulomb model are not heavily

contingent upon the initial stress state resulting from auguring of soil or the driving of

pile into soil where K0 ranges from 0.7 to 1.0.

97
FEM STUDY OF SERANGOON MRT SITE

150

Deviatoric stress, q
100
Initial excavation

50

(a)
0
0 50 100 150 200
Mean effective stress, p'
3D - behind soldier pile

200

Initial
Deviatoric stress, q

150 excavation

100 MC2 K0 = 0.9


MC6 K0 = 0.7
MC7 K0 = 1.0
yield locus
50

(b)
0
-100 0 100 200 300
Mean effective stress, p'
3D - behind timber lagging
Figure 5.14: Stress path plots for MC2, MC6 and MC7 at 12.3m below ground. (a) Location I.
(b) Location II. Locations I and II for each case are defined in Figure 5.6. Units of stresses in
kPa.

Wall deflection from 2D vs 3D analyses

Figure 5.15 shows the measured and computed wall deflection 2D and 3D

progressive movement of the retaining wall for MC5 from Stages B to I as illustrated

98
FEM STUDY OF SERANGOON MRT SITE

in Figure 5.3 compared to the field results. The maximum deflection shown by the

field measurement is approximately 50mm (0.20%H). This is close to the average of

0.18%H reported by Yoo & Kim (1999) for 145 cases of H-piled excavation. The

FEM soldier pile profile for 2D and 3D matches that of the field well in the initial

stages (Stages B-D). From Stages E-I above excavation level, the 2D deflection

profile takes an average value lies between the soldier pile profile and the timber

lagging profiles. This can be attributed to the averaging effect implied by the smearing

of the soldier piles and timber lagging, which prevents the effects of pile-soil

interaction from being exhibited.

Strut forces

Table 5.3 compares the strut forces obtained from finite element analyses of

MC2, MC3, MC4 and MC5. In all of the cases, large discrepancies between the daily

average field and analysed results are found in the first and fifth struts. Furthermore,

the different combinations of parameters do not affect the strut forces significantly. As

will be shown later, the same discrepancies arose with the use of the Cam Clay models.

This suggests that the discrepancies are unlikely to be due to the choice of parameters

or models. In particular, the computed strut forces for the first level greatly exceeded

the field measurement. Figures 5.16a shows the deflection of the soldier pile at an

early stage of the excavation, before the installation of the first level struts. As can be

seen, FEM analysis overestimates the measured soldier pile deflection. A possible

reason is the presence of a 1 to 2m deep concrete pile cap beam which limited the

movement at the initial stages of excavation and reduced the load transferred to the 1st

level strut at 3.5m below EGL. As Loh et al. (1998) noted from centrifuge model tests,

a stiff beam at the top of the wall can act as a strut to limit wall deflection.

99
FEM STUDY OF SERANGOON MRT SITE

Table 5.3: Comparison of strut forces at various levels between daily average instrumented and
FEM result for MC2, MC3, MC4 and MC5.
Strut level Instrumented (kN) FEM (kN) Instrumented/FEM × 100%
MC2 MC3 MC4 MC5 MC2 MC3 MC4 MC5
1 173.1 427.0 429.0 439.1 410.4 40.5 40.3 39.4 42.2
2 692.3 822.7 825.7 813.3 824.8 84.1 83.8 85.1 83.9
3 836.5 836.6 834.2 811.9 877.5 100.0 100.0 103.0 95.3
4 865.4 839.2 843.6 828.8 839.6 103.1 102.6 104.4 103.1
5 1015.4 555.9 558.2 556.4 534.8 182.7 181.9 182.5 189.9
6 126.9 179.9 181.4 169.6 204.1 70.6 70.0 74.8 62.2

The discrepancy in the fifth level strut forces are unlikely to be caused by the

ground surface activities. Figure 5.16b shows the wall deflection profile after the

installation of the 4th level strut. As can be seen, the analysis significantly over-

estimates the wall deflection at this stage. In other words, there is, in reality, a large

increase in soldier pile deflection between the installation of the 4th and 5th level struts

which is not captured in the analyses; this explains the under-estimation of the 5th

level strut load. The cause of the large increase in wall deflection is unclear, but one

possibility lies in the errors in modelling of the soil profile. As the previous chapter

shows, the soil conditions at this site are quite variable. Since the nearest borehole to

the matched section is about 10m, the interpolated soil profile may not truly reflect the

actual soil profile at the matching section. Another possible reason is the presence of

groundwater problems at this site. Lee et al. (2001) reported that groundwater

problems experienced at the Serangoon Station excavation appear to cause quite

significant softening of the soil, especially near the final formation level. The matched

section is located roughly 30m from the location where groundwater problems were

experienced. It is therefore plausible that some groundwater-induced softening of the

100
0 0 0 0

5 5 5 5

10 10 10 10

15 15 15 15

20 20 20 20

25 (a) 25 25 25

Depth (m)
Depth (m)
Depth (m)
Depth (m)
30 Field 30 30 30
2D
35 35 35 35
3D-I
40 3D-II 40 40 40
(b) (c) (d)
45 45 45 45
0 20 40 60 80 100 0 20 40 60 80 100 0 20 40 60 80 100 0 20 40 60 80 100
Deflection (mm) Deflection (mm) Deflection (mm) Deflection (mm)

0 0 0 0

5 5 5 5

10 10 10 10

15 15 15 15

20 20 20 20

25 25 25 25

Depth (m)
Depth (m)
Depth (m)
Depth (m)

30 30 30 30

35 35 35 35

40 40 40 40
(e) (f) (g) (h)
45 45 45 45

101
0 20 40 60 80 100 0 20 40 60 80 100 0 20 40 60 80 100 0 20 40 60 80 100
Deflection (mm) Deflection (mm) Deflection (mm) Deflection (mm)
FEM STUDY OF SERANGOON MRT SITE

Figure 5.15: Deflection profiles at various stages of excavation for MC5. (a) Stage B. (b) Stage C. (c) Stage D.
(d) Stage E. (e) Stage F. (f) Stage G. (g) Stage H. (h) Stage I.
FEM STUDY OF SERANGOON MRT SITE

retained soil could have occurred towards the end of the excavation construction which

led to a large increase in wall deflection.

0 0

5 5

10 10

15 15
Depth (m)

Depth (m)
20 20

25 25

30 30
Field
35 35 MC2
MC3
40 40 MC4
MC5
45 45
0 20 40 60 0 20 40 60
Deflection (mm) Deflection (mm)
(a) (b)
Figure 5.16: Deflection profile of soldier pile for MC5. (a) Before installation of first level
struts. (b) After installation of forth level struts.

Ground settlement

Figure 5.17 shows the FEM settlement profile behind the retaining wall plotted

against the field settlement. In order to match the wall deflection, the Young’s

modulus of soil was raised beyond that inferred from field tests thus resulting in a

much stiffer settlement profile. As shown in Figure 5.17, all of the Mohr-Coulomb

cases predicted much smaller settlement than the measured values. Thus, in essence,

while the Mohr-Coulomb model was able to match the wall deflection, and to some

extent, the strut loads by increasing the values of the elastic parameters above the

measured values, the settlement profile was significantly underpredicted. This point

will be further discussed in a later section.

102
FEM STUDY OF SERANGOON MRT SITE

Distance from retaining wall (m)


0 10 20 30 40 50 60 70 80 90

0
Settlement (mm) L215(G)
10
L217(G)
L214(G)
20 Field
MC2
L222(G) MC3
30 L203(G)
MC4
MC5
40
Figure 5.17: Ground settlement behind Location I of the retaining wall for MC2, MC3, MC4
and MC5.

5.3.2 Modified Cam-clay

The Serangoon Station site is underlain by residual soil derived from the

weathering of granitic rock. The material may range from highly fissured rock at

depth to completely weathered residual soil with a clayey and silty constitution.

Although it is not strictly an over-consolidated soil, some aspects of its behaviour do

resemble those of over-consolidated soil (Dames & Moore, 1983). For instance, the

Skempton’s pore pressure parameter A is typically about zero at failure, indicating a

tendency to dilate and generate negative pore pressure. The deviator stress-strain

curve also typically shows peak strength at failure, which is again typical of over-

consolidated soil. Dames & Moore (1983) recommended an equivalent over-

consolidation ratio of about 3 for this residual soil.

In Chapter 3 as well as the last section, it has been demonstrated that the

transfer of stress from soil to soldier pile as a result of arching action may result in a

significant increase in effective stress at localized region of the ground. At such

locations, stresses may be sufficiently high to induce normally consolidated behaviour,

even from over-consolidated soil. In such cases, the use of the Cam-clay model is

103
FEM STUDY OF SERANGOON MRT SITE

advantageous compared to the Mohr-Coulomb model since it allows the transition

from over-consolidated to normally consolidated state to be captured. In addition, the

pile-soil interaction may give rise to high shear stresses in localised regions around the

soldier piles. In such cases, the Cam-clay model has the advantage that it allows

stress-induced dilatancy to cease as the critical state is approached, whereas the Mohr-

Coulomb model does not. Finally, as was demonstrated earlier, there is likely to be

significant stress relief in the soil at the excavation face, especially behind the timber

lagging. As the effective stress reduces, the soil is also likely to soften. In such an

instance, the Cam-clay models the softening process to be modelled through the

reduction in modulus of the soil with effective stress, whereas the Mohr-Coulomb

model does not. In this section, the Serangoon Station excavation is back-analysed

using a modified Cam-clay model.

Determination of modified Cam-clay parameters

The critical state frictional constant, M was evaluated using the relationship:

6 sin φcs '


M= (5.4)
3 − sin φcs '

Figures 5.18a and b shows the critical state friction angles corresponding to the depth

of 6m and 12m for the soil at the Serangoon station. Using typical φcs’ of 27° for soil

Layer 1 and 30° for soil Layer 2 leads to M equals 1.07 and 1.2 respectively.

104
FEM STUDY OF SERANGOON MRT SITE

300
φcs' = 29°

Shear Stress (kPa)


200

100

(a)
0

0 100 200 300 400 500


Mean Effective Stress (kPa)

480
φcs' = 31°
Shear Stress (kPa)

400
320
240
160
80
(b)
0

0 200 400 600 800


Mean Effective Stress (kPa)
Figure 5.18: Critical state friction angle estimated from triaxial compression tests. (a) At 6m
below EGL. (b) At 12m below EGL.

Figures 5.19a and b show the compression and swelling indices respectively for

soil up to 24m below EGL. The compression index, cc, and swelling index, cs, were

estimated from odeometer tests. The cc and cs is related to λ and κ through Equations

5.5 and 5.6 respectively:

cc = λo ln 10 (5.5)

c s = κ o ln 10 (5.6)

105
FEM STUDY OF SERANGOON MRT SITE

0 0

-2 -2

-4 -4

-6 -6

-8 -8

-10 -10
Depth (m)

Depth (m)
-12 -12

-14 -14

-16 -16

-18 -18

-20 -20

-22 -22
(a) (b)
-24 -24
0.00 0.20 0.40 0.60 0.00 0.05 0.10 0.15
cc cs

Figure 5.19: Variation of deformability indices with depth at Serangoon MRT site. (a)
Compression index. (b) Swelling index.

Bolton (1991) suggested λo and κo in one-dimensional (1D) loading and λi and

κi isotropic values may be approximately related through the relations

λi = λ o (5.7)

κ i ≈ 1.5κ o (5.8))

Goh (1995) showed from theory of elasticity, κi vary from 1.4-1.8 times κo.

The values of λ inferred for Layer 1 and Layer 2 are 0.087 and 0.109 respectively.

The corresponding values of κ Layer 1 and Layer 2 are 0.02 and 0.021.

The values of λ inferred for Layer 1 and Layer 2 are 0.087 and 0.109

respectively. The corresponding values of κ Layer 1 and Layer 2 are 0.02 and 0.021.

Another parameter which needs to be specified in CRISP is the critical state

void ratio at mean effective stress of 1kPa, or ecs. In this study, ecs was inferred by

simulating the 1D consolidation problem using CRISP and adjusting the critical state

void ratio until the compression curve agrees with that from laboratory test, as depicted

106
FEM STUDY OF SERANGOON MRT SITE

in Figure 5.20. This iterative fitting process returns a value of 1.7 and 2.0 for the ecs of

Layer 1 and Layer 2 respectively.

1.30

1.20

1.10
Void ratio, e

1.00

0.90

odeometer test
0.80
FEM

0.70

0.60

1.0E+0 1.0E+1 1.0E+2 1.0E+3 1.0E+4


Axial load (Mpa)
(MPa)

Figure 5.20: Comparison of laboratory’s odeometer test with MCC soil models.

Determination of in-situ stress state

Apart from the compression and swelling indices, the initial state of the soil

also needs to be specified. The initial earth pressure coefficient K0 was inferred from

the over-consolidation ratio (OCR) determined from one-dimensional consolidation

test data. Consolidation test data conducted on soil samples obtained from the top of

the residual soil stratum indicate an OCR value of about 4. Dames and Moore (1983)

suggested that 1.5 to 3, which was inferred from the ratio of the undrained Young’s

modulus Eu to the undrained shear strength cu, using Duncan and Buchignani’s (1976)

method. Dames and Moore’s (1983) Eu values were obtained as an average from the

recompression curves in unconsolidated undrained (UU) triaxial tests and from the

initial modulus of pressuremeter tests. Given the uncertainty involved in determining

Eu from UU triaxial and pressuremeter tests as well as that in determining OCR from

107
FEM STUDY OF SERANGOON MRT SITE

consolidation tests, the small discrepancy is not unreasonable. For this study, K0 was

inferred using Mayne & Kulhawy (1982) relation that

K 0 = K nc OCR sin φ ' (5.9)

in which Knc is the earth pressure coefficient of the same soil in a normally

consolidated soil and is determined by Jáky’s (1944) relation

K nc = 1 − sin φ ' (5.10)

For the top 11.5m of residual soil stratum, using Equations 5.9 and 5.10 with OCR

ranging from 3 to 4 and φ’ between 30° and 35° leads to K0-values in the range of 0.8

to 1.0. This is also in agreement with Dames and Moore (1983) who recommended a

value of between 0.8 and 1.0. At depths of between11.5-22.5 m, the OCR ranges from

2-4 with φ’ in the range of 27°-39°. This implies that the K0 ranges from 0.57-1.02,

which is also in reasonable agreement with Dames and Moore’s (1983) recommended

range of values.

Table 5.4 summarised the modified Cam-clay properties used in the analyses.

Analysis MCC1 uses the soil properties as inferred above. The OCR in Layer 1 and 2

of MCC2 is artificially raised to 200% of the inferred value so as to enforce

predominantly elastic behaviour. MCC3 represents the upper bound of κ deduced

from field measurement at 0.040 and 0.047 for Layers 1 and 2 respectively. MCC4

adopted the lower bound of κ at 0.013 and 0.014 for Layers 1 and 2 respectively.

108
FEM STUDY OF SERANGOON MRT SITE

Table 5.4: Soil properties for Layer 1 and Layer 2 for modified Cam-clay.
Analyses κ OCR M (φ) K0
Layers 1 2 1 2 1 2 1 2 3
MCC1 0.020 0.021 4 2 1.07 (27°) 1.2 (30°) 0.9 0.9 1.0
MCC2 0.020 0.021 8 4 1.07 (27°) 1.2 (30°) 0.9 0.9 1.0
MCC3 0.040 0.047 4 2 1.07 (27°) 1.2 (30°) 0.9 0.9 1.0
MCC4 0.013 0.014 4 2 1.07 (27°) 1.2 (30°) 0.9 0.9 1.0
MCC5 0.013 0.014 4 2 1.5 (37°) 1.6 (39°) 0.9 0.9 1.0
MCC6 0.020 0.021 4 2 1.07 (27°) 1.2 (30°) 0.7 0.7 0.7
MCC7 0.020 0.021 4 2 1.07 (27°) 1.2 (30°) 1.0 1.0 1.0

Simulation of triaxial test behaviour

Axis of symmetry
σ1
35.25mm

Plane of symmetry

σ3 σ3
20.38mm

σ1
(a) (b)
Figure 5.21: Axisymmetric FEM undrained (CU) triaxial test model with 9-integration points
eight noded quadrilateral elements. (a) Plane view. (b) Isometric view.

As a precursor to simulating the behaviour of the excavation, a FE simulation of

a consolidation undrained (CU) triaxial test was conducted and the results compared to

the measured stress-strain curve in the same kind of test. As shown in Figure 5.21, an

axisymmetric model with 9-integration-points, eight noded quadrilateral elements was

preloaded isotropically to a level of confining stress that the soil is expected to be

109
FEM STUDY OF SERANGOON MRT SITE

subjected to in the field. Strain-controlled axial loading is then simulated by

progressively increasing the vertical displacement of the top boundary.

200 400
(a) (b)
Deviator stress, q (kPa)

Deviator stress, q (kPa)


150 300

100 200

Triaxial test Triaxial test


MCC1 MCC1
50 100
MCC3 MCC3
MCC4 MCC4

0 0
0.00 0.01 0.02 0.03 0.04 0.00 0.01 0.02 0.03 0.04 0.05
Deviator strain, εs Deviator strain, εs

Figure 5.22: Comparison of laboratory’s CU triaxial tests with MCC soil models. (a) Depth of
sample is 11m. (b) Depth of sample is 20m.

As can be seen from Figure 5.22, the FE model predicts linear elastic right up to

the point of yield whereas the measured data indicate that a curve is more appropriate.

The reason for this is that modified Cam-clay model in CRISP does not model any

shear-induced reduction in shear modulus within the yield surface. Thus, small strain

non-linearity is not modelled. For the sample taken from 11m-depth, the elastic

segments of the stress-strain curves from MCC3 and MCC4 appears to bind the

measured curve. The elastic segment of MCC1 also appears to be a reasonable

“average” of the measured curve. For the sample taken from 20-m depth, the

measured curve appears to imply an “average” stiffness that is closer to MCC3 and

definitely lower than that of MCC1. One possible reason could be due bedding errors

in conventional triaxial test sample. The ‘bedding down’ at the ends of the sample

arises due to local irregularities or voids (Jardine et al., 1984). Therefore, the

externally measured strain will often exceed those experienced locally on the sample.

Such errors will not be reflected in a FEM simulation. Notwithstanding the limitations

110
FEM STUDY OF SERANGOON MRT SITE

of the soil model and modelling process, the predicted stress-strain curves appear to

suggest that the computed range of stress-strain behaviour do represent the measured

behaviour reasonably well. It should also be noted that samples taken from the field

often suffer a certain amount of sampling disturbance which commonly leads to a

reduction in the measured strength and modulus of the sample.

The fill properties tend to be highly variable and little or no investigation was

conducted on the fill. Thus for the preliminary investigation, the properties of the fill

were assumed to be the same as those of Layer 1 of residual soil

Wall deflection

10

15
Depth (m)

20

25

30
Field
35 MCC1-I
MCC1-II
40 MCC2-I
MCC2-II
45
020 40 60 80
Deflection (mm)
Figure 5.23: Final deflection profile of soldier pile and timber lagging for MCC1 and MCC2.

Figure 5.23 shows the final deflection profiles for MCC1 and MCC2. As can

be seen, analyses MCC1 and MCC2 predict very similar deflection profiles. Since the

only difference between the two analyses is the larger initial yield surface assumed in

MCC2, the close similarity of the results imply the soil behaviour is likely to be

111
FEM STUDY OF SERANGOON MRT SITE

predominantly elastic. This is also reflected in Figure 5.25, which shows that the stress

paths plotted did not reach the yield surface.

Comparison of Figure 5.23 with Figures 5.8 and 5.10 also suggests that the

modified Cam-clay model predicts a much smaller difference between soldier pile and

timber lagging deflection than the Mohr-Coulomb model. This can be attributed to the

manner in which the parameters for the Mohr-Coulomb and the modified Cam-clay

models were derived. In the case of the Mohr-Coulomb model, the measured angle of

friction was used to define the yield surface. On the other hand, in the modified Cam-

clay model, the angle of friction was used to define the critical state line. This implies

that, given the same angle of friction on the “dry” side of critical, the modified Cam-

clay yield surface can sustain a higher q/p’ stress ratio than the Mohr-Coulomb yield

surface. Thus, if the excavation behaviour is controlled by “dry” side yielding, the

Mohr-Coulomb model will yield more readily than the modified Cam-clay model.

7 - 9 IPs yielded

Soldier piles 4 - 6 IPs yielded

1 - 3 IPs yielded

No IPs yielded
(a) MC1

(b) MC4 (c) MC5


Figure 5.24: Yield zones behind retaining wall at 12.3m below ground at instance when first
yield was detected in MC5 (Between stage C and D of Figure 5.3). (a) MC2. (b) MC4. (c)
MC5.

112
FEM STUDY OF SERANGOON MRT SITE

The presence of yielded zones in the MC cases discussed is illustrated in Figure

5.24, which shows widespread yielding of the soil just behind the soldier piles and

timber lagging. For this reason, the Young’s modulus of the Mohr-Coulomb has to be

raised in order to fit the measured wall deflection. As mentioned earlier, the modulus

values needed to give a good fit to the measured soldier pile deflection significantly

exceeds those derived by fitting the pressuremeter data in Figure 5.4. As the soil

yields, dilatancy causes an expansion in soil volume, which helps to push the flexible

timber lagging out further. This explains the large predicted differences between

soldier pile and timber lagging deflection.

For the modified Cam-clay model, the Young’s modulus is controlled indirectly

by κ and the Poisson’s ratio ν’. For the values of κ and ν’ used in the MCC-analyses,

the implied E’ ranges from about 9MPa to 28MPa, with an average value of about

18MPa; this is in much better agreement with Dames & Moore’s (1983) data than the

range of values for the Mohr-Coulomb model. As shown in Figures 5.25a and b, the

soil remains predominantly elastic with the modified Cam-clay model. Thus dilatancy

is not present in the MCC-analyses. Given the fact that the 30° friction angle is quite a

commonly used value, it is possible that premature yielding may well be a common

syndrome in many Mohr-Coulomb FE analyses of residual or over-consolidated soils

in excavation problems, where localised stress concentration may be present.

113
FEM STUDY OF SERANGOON MRT SITE

150 MCC1
MCC2
MCC1 YL & CSL
MCC2 YL & CSL

Deviatoric stress, q^
100

CSL

50

(a)
0
0 50 100 150 200
Mean effective stress, p'
3D - behind soldier pile

150
100
^
Deviatoric stress, q

50
0 CSL

-50
-100
-150
-200
(b)
-250
0 50 100 150 200
Mean effective stress, p'
3D - behind timber lagging
Figure 5.25: Stress path plots for MCC1 and MCC2 at 12.3m below ground. (a) Location I.
(b) Location II. Locations I and II for each case are defined in Figure 5.6. Units of stresses in
kPa.

114
FEM STUDY OF SERANGOON MRT SITE

10

15

Depth (m)
20

25
Field
30 MCC1-I
MCC1-II
MCC3-I
35
MCC3-II
MCC4-I
40 MCC4-II

45
040 80 120 160
Deflection (mm)
Figure 5.26: Final deflection profile of soldier pile and timber lagging for MCC1, MCC3 and
MCC4.

Figure 5.26 compares the computed results for analyses MCC1, MCC3 and

MCC4. As can be seen, MCC3 shows a much larger wall deflection than MCC1 and

MCC4, this being attributable to its much higher value of κ. This further underlines

the fact that, in the MCC-analyses, wall deflection is largely controlled by elastic

parameters. On the other hand, as shown in Figure 5.27, changing the critical state

angle of friction (from case MCC4 to MCC5) has a much smaller effect.

Figure 5.27 reflects the deflection profiles for MCC4 and MCC5. The angle

implied in this friction constant, M, is the critical state friction angle. An increase in

the M corresponding to an increase in critical state friction angles from 27° to 37° in

Layer 1 and 30° to 39° in Layer 2 does not change the profile as significantly as

MCC3.

115
FEM STUDY OF SERANGOON MRT SITE

10

15

Depth (m)
20

25

30
Field
35 MCC4-I
MCC4-II
MCC5-I
40 MCC5-II

45
020 40 60 80
Deflection (mm)
Figure 5.27: Final deflection profile of soldier pile and timber lagging for MCC4 and MCC5.

Graham (1992) observed that modified Cam-clay predicts shear strains

relatively poorly, especially when clay is heavily overconsolidated. However, the

above discussion shows that, by using appropriate values for the Cam-clay parameters,

meaningful results can still be obtained, notwithstanding the fact that small strain non-

linearity is not modelled. To further assess the significance of small strain non-

linearity, a hyperbolic Cam-clay model (Nasim, 1999) will be used to model the non-

linearity of small-strain non-linearity of the residual soil in a later section.

116
FEM STUDY OF SERANGOON MRT SITE

10

15

Depth (m)
20

25

30 MCC1-I K0 = 0.9
MCC1-II K0 = 0.9
MCC6-I K0 = 0.7
35
MCC6-II K0 = 0.7
MCC7-I K0 = 1.0
40 MCC7-II K0 = 1.0

45
0
20 40 60 80
Deflection (mm)
Figure 5.28: Final deflection profile of soldier pile and timber lagging for MCC1, MCC6 and
MCC7.

As in the case of Mohr-Coulomb models, Figure 5.28 shows slightly increasing

wall deflections with increasing K0-values. Furthermore, as Figure 5.29 shows the

stress paths behind the soldier pile and timber lagging at 12.3m below ground follows

almost parallel paths from initial points. This shows that the pile-soil interaction

mechanism does not change greatly with different in-situ stress conditions. Thus, as in

the case for the Mohr Coulomb models, the wall deflection behaviour and stress paths

for the Modified Cam-clay models are not heavily contingent upon the initial stress

state of K0 between 0.7 and 1.0 resulting from auguring of soil or the driving of H-pile

into soil respectively.

117
FEM STUDY OF SERANGOON MRT SITE

150

Deviatoric stress, q
100
Initial
excavation

50

(a)
0
0 50 100 150 200
Mean effective stress, p'
3D - behind soldier pile

200
Deviatoric stress, q

150

100 MCC1 K0 = 0.9


MCC6 K0 = 0.7
MCC7 K0 = 1.0

50
Initial
excavation
(b)
0
0 100 200 300
Mean effective stress, p'
3D - behind timber lagging
Figure 5.29: Stress path plots for MCC1, MCC6 and MCC7 at 12.3m below ground. (a)
Location I. (b) Location II. Locations I and II for each case are defined in Figure 5.6. Units
of stresses in kPa.

Wall deflection from 2D and 3D analyses

Figure 5.30 shows the 2D and 3D progressive wall deflection after replacement

of the top 3.5m fill material to Mohr Coulomb’s properties. As observed from the

118
FEM STUDY OF SERANGOON MRT SITE

timber lagging deflection in top 3.5m fill, there is a larger difference between soldier

pile and timber lagging deflection. This is due to excavation behaviour being

controlled by “dry” side yielding as discussed earlier. Otherwise, the soldier pile

deflection did not change drastically after modification of the fill properties even

though it made significant improvement to the ground settlement. The 2D profile

takes an average value between the soldier pile and timber lagging deflections in the

Mohr Coulomb layer. In the modified Cam-clay layer above excavation level, the 2D

takes after the soldier pile deflection profile closely.

Comparison of the computed and measured wall deflections at various stages of

excavations shows that, although the final wall deflection is well-matched. The early-

stage wall deflection is significantly over-predicted. The discrepancy appears to

increase towards the initial stages of the excavation. One possible reason for this

discrepancy is the fact that the modified Cam-clay model does not model the

degradation in shear modulus as shear strain increases. Thus, the shear modulus which

is appropriate for matching the final wall deflection is likely to be too low to match the

initial wall deflection, where the shear strains are lower. For this reason, the initial

wall deflection is over-predicted. The notion will be further examined when the

hyperbolic Cam Clay cases are analysed below.

119
0 0 0 0

5 5 5 5

10 10 10 10

15 15 15 15

20 20 20 20

25 (a) 25 25 25

Depth (m)
Depth (m)
Depth (m)
Depth (m)
30 Field 30 30 30
2D
35 35 35 35
3D-I
40 3D-II 40 40 40
(b) (c) (d)
45 45 45 45
0 20 40 60 80 100 0 20 40 60 80 100 0 20 40 60 80 100 0 20 40 60 80 100
Deflection (mm) Deflection (mm) Deflection (mm) Deflection (mm)

0 0 0 0

5 5 5 5

10 10 10 10

15 15 15 15

20 20 20 20

25 25 25 25

Depth (m)
Depth (m)
Depth (m)
Depth (m)

30 30 30 30

35 35 35 35

40 40 40 40
(e) (f) (g) (h)
45 45 45 45
0 20 40 60 80 100 0 20 40 60 80 100 0 20 40 60 80 100 0 20 40 60 80 100
Deflection (mm) Deflection (mm) Deflection (mm) Deflection (mm)

120
Figure 5.30: Deflection profiles at various stages of excavation for MCC1-M. (a) Stage B. (b) Stage C. (c) Stage D.
FEM STUDY OF SERANGOON MRT SITE

(d) Stage E. (e) Stage F. (f) Stage G. (g) Stage H. (h) Stage I.
FEM STUDY OF SERANGOON MRT SITE

Strut forces

Table 5.5 compares the strut forces obtained from the best-fit finite element

analyses of the MCC1. Similar observation to MC5 can be inferred from Table 5.5

and Figure 5.31 where large discrepancies between the field and analysed results are

found in the first and fifth struts. The reasons are as explained in the previous section.

Table 5.5: Comparison of strut forces at various levels between instrumented and FEM result
for the best-fit deflection profile for MCC1.
Strut level Instrumented (kN) FEM (kN) Instrumented/FEM × 100%
1 173.1 395.9 43.7
2 692.3 777.9 89.0
3 836.5 867.5 96.4
4 865.4 970.0 89.2
5 1015.4 503.2 201.8
6 126.9 182.4 69.6

0 0

5 5

10 10

15 15
Depth (m)

Depth (m)

20 20

25 25

30 30

35 35

40 40 Field
FEM
45 45
0 20 40 60 0 20 40 60
Deflection (mm) Deflection (mm)
(a) (b)
Figure 5.31: Deflection profile of soldier pile for MCC1. (a) Before installation of first level
struts. (b) After installation of forth level struts.

Ground settlement

Figure 5.32 compares the computed and measured settlement profiles behind

the retaining wall. As can be seen, in all cases, the ground settlement is over-estimated

121
FEM STUDY OF SERANGOON MRT SITE

by the FE analyses. Examination of the variation of the settlement with depth indicates

that much of the settlement occurs within the fill layer. Given the values of λ and κ

assumed for the fill layer, the implied modulus is of the order of 2MPa at the mid-

depth of the fill. Given that the fill is very heavily compacted, this value is

unrealistically low. In addition, the modified Cam-clay model implies that, right at the

ground surface, the modulus of the soil is zero, which is also unrealistic. In order to

model the fill more realistically, a Mohr-coulomb model with properties as shown in

Table 5.6 was used for the fill. Note that, in order to forestall early yielding, a

relatively high angle of friction of 35° has been adopted for this layer. As shown in

Figure 5.33, the agreement between the computed profile for MCC1-M and the field

data is now much better.

Table 5.6: Modification to 3.5m of topsoil to Mohr Coulomb properties.


Depth (m) γs (kN/m3) E’ (MPa) c’ (kPa) φ’ ν kx (m/s) ky (m/s)
0.0-3.5 18.5 8.0 0 35° 0.30 1.0×10-5 1.0×10-5

Distance from retaining wall (m)


0 10 20 30 40 50 60 70 80 90

0
L215(G)
L214(G) L217(G)
20
L203(G) L222(G)
40
Settlement (mm)

60

80

100 Field
MCC1
120 MCC2
MCC3
140 MCC4
MCC5
160

180
Figure 5.32: Ground settlement behind Location I of the retaining wall for MCC1, MCC2,
MCC3, MCC4 and MCC5.

122
FEM STUDY OF SERANGOON MRT SITE

Distance from retaining wall (m)


0 10 20 30 40 50 60 70 80 90
0
L215(G)
Settlement (mm)

10
L217(G)
L214(G)
20
L222(G) Field
30 2D
3D
L203(G)
40
Figure 5.33: 2D and 3D ground settlement behind Location I of the retaining wall for MCC1-
M.

5.3.3 Hyperbolic Cam-clay

It is now well-recognised that the shear modulus of most soils tends to decrease

as the shear strain increases (e.g. Simpson et al., 1979; Jardine et al., 1986; Burland,

1989), even before the initiation of large-scale yielding. This phenomenon is evident

also from Figure 5.22, in which the triaxial tests data indicate a progressive decrease in

the tangent modulus of the soil that was not reflected by the highly linear pre-yield

segments of the stress-strain curves prescribed by the modified Cam-clay model. This

non-linear behaviour at relatively small shear strains has been variously represented by

multi-nested surface models (e.g. Stallebrass 1990) and by non-linear hysteretic

models (e.g. Dasari 1996; Nasim, 1999). In this and the following sections, the effect

of small strain non-linearity is examined using Nasim’s (1999) hyperbolic Cam-clay

model as a vehicle. More complicated multi-nested models are not used in this

investigation as the emphasis of this discussion is on the influence of geometric and

material factors on soldier-pile-soil interaction rather than the development of

constitutive models for residual soils. In any case, since the triaxial tests in the soil

investigation were not conducted using small strain measurements, the quality of the

123
FEM STUDY OF SERANGOON MRT SITE

data and the knowledge of the soil derived from it is unlikely of a sufficiently high

quality to justify the use of a more complex model.

Determination of hyperbolic Cam-clay’s parameters

Nasim’s hyperbolic Cam-clay model is based on the concept that the pre-yield

non-linear behaviour of a soil can be represented by a hyperbolic function. The

hyperbolic function used by Nasim (1999) is given by:

(S 0 − S ∞ )ε s (0 < ε
q= + S ∞ε s <εy) (5.11)

1 +
(S 0 − S ∞ ) ⋅ ε s 

s

 qf 
 

where

S0 is the initial tangential stiffness

S∞ is the tangential stiffness at very large strain

εs is the shear strain

εy is the shear strain at yielding

qf is the deviator stress at failure

Assuming Equation 5.12 and differentiating Equation 5.11 with respect to εs leads to a

variation of shear modulus G with shear strain εs that is given by:

dq
S= = 3G (5.12)
dε s

G0 − G∞
G= 2
+ G∞ (5.13)
 3(G0 − G∞ ) ⋅ ε s 
1 + 
 qf 
 

where
124
FEM STUDY OF SERANGOON MRT SITE

G0 is the initial shear modulus

G∞ is the shear modulus at very large strain

This allows the shear modulus G to decrease with increasing shear strain following the

trend of the S-shaped curve which has been observed experimentally (e.g. Jardine et

al., 1986; Dasari 1996). The initial shear modulus, G0, is in turn related to the current

state of soil by the relation:

G0 = Cp ' n OCR m (5.14)

where

C is a constant

n is the effective stress, p’, exponent

m is the OCR exponent

The experimental results reported by Wroth et al. (1979) from both dynamic

and static tests on sand indicated that the value of n varied from approximately 0.5 at

small strain to about 1.0 at large strain.

Viggiani & Atkinson (1995) carried out bender element tests on reconstituted

samples of a variety of different soils and observed that, for fine-grained soils, both the

indices m and n increase with increasing plasticity index in the manner shown in

Figures 5.34a and b respectively. The values of m reported typically range from 0.2 to

0.3 whereas n typically ranged from 0.5 to 0.9. Using bender element tests, Coop &

Jovičić (1999) also reported similar m and n values for a variety of coarse-grained

soils. They also suggested that, for coarse-grained soils, n could reasonably be

approximated by 0.58, while the value of m depended on the history of over-

125
FEM STUDY OF SERANGOON MRT SITE

consolidation or compaction. In this study, the values of m and n were determined

from Figure 5.34.

0.4 1.0

0.9

0.3
Coefficient, m

Coefficient, n
0.8

0.7
0.2

0.6

0.1 0.5
0 10 20 30 40 50 60 70 0 10 20 30 40 50 60 70
Plasticity index, PI Plasticity index, PI

(a) (b)
Figure 5.34: Parameters for G0. (a) Coefficient m. (b) Coefficient n. (Viggiani & Atkinson,
1995)

Since the hyperbolic Cam-clay model is based purely on empirical relationship,

thus matching with laboratory tests is essential in determining the soil parameters. The

soil parameters of the hyperbolic Cam-clay model were determined by simulating a

consolidated undrained triaxial test using finite elements. It is likely that some errors

will be incurred in the fitting process since conventional triaxial tests with external

strain measurements often tend to return lower soil stiffness than the true values.

Based on the CU triaxial tests conducted on Serangoon residual soil, the values of G0

determined are 9.3MPa and 10.4MPa at sample depths of 11m and 20m, respectively.

These benchmark values in turn allow the value of C to be estimated.

At very large strain, the soil is likely to be destructured. For this reason, the

shear modulus at very large strain, G∞, is assumed to be zero. After commencement of

yielding, the strain-dependent model would follow the behaviour of modified Cam-

clay model.

126
FEM STUDY OF SERANGOON MRT SITE

The bulk modulus K’ is related to G through the following equation:

 2 (1 +ν ' ) 
K'=  ⋅G (5.15)
 3 (1 − 2ν ' ) 

Stallebrass’s test data (1990) showed that K’ will ultimately reach a minimum value

for swelling or if loaded in a constant-q stress path. Since the behaviour of the soil is

assumed to be consistent with that prescribed by the modified Cam-clay at large

strains, K’ is limited to be lower bounded:

υp '
K min ' = (5.16)
κ

where

Kmin’ is effective bulk modulus

υ is the specific volume

p’ is the mean effective stress

Figure 5.35 shows the variation of K’ with volumetric strain and G with deviator strain

from FEM simulated drained triaxial test similar to the mesh in Figure 5.21. Figure

5.36 shows the variation of K’ with mean effective stress and G with deviator stress

from FEM simulated drained triaxial test.

127
FEM STUDY OF SERANGOON MRT SITE

Deviator strain, εv x102


0.01 0.1 1 10
80000 40000

Shear modulus, G (kPa)


Bulk modulus, K (kPa)
60000 30000
Bulk Modulus, K
Shear Modulus, G

40000 20000

20000 10000

0 0
10.1 10
Volumetric strain, εv x103
Figure 5.35: Variation of K’ and G with volumetric strain and deviator strain respectively in a
simulated FEM drained triaxial test.

Deviator stress, q (kPa)


0 50 100 150 200 250 300
80000 40000

Shear modulus, G (kPa)


Bulk modulus, K (kPa)

60000 30000
Bulk Modulus, K
Shear Modulus, G

40000 20000

20000 10000

0 0
180 200 220 240 260 280
Mean effective stress, p' (kPa)
Figure 5.36: Variation of K’ and G with mean effective stress, p’, and deviator stress, q,
respectively in a simulated FEM drained triaxial test.

The parameter qf is taken to be the deviator stress at failure in an undrained

triaxial test of the soil. The value of qf can be determined through the least square

fitting of laboratory test data with the strain-dependent relation (Equation 5.11). This

may not be strictly accurate for all soils, but for residual soils, it may not be a bad

approximation since the in-situ effective stress state is likely to be fairly near to being

isotropic, as indicated by K0 ~ 1.0.

128
FEM STUDY OF SERANGOON MRT SITE

Wall deflection

Table 5.7 shows the various hyperbolic Cam-clay cases studied whilst Figure

5.37 compares the measured stress-strain curve from conventional triaxial test with the

predicted curves of the 4 analyses.

Table 5.7: Soil properties for Layer 1 and Layer 2 for hyperbolic Cam-clay.
Analyses C n qf (kPa)
Layer 1 Layer 2 Layer 1 Layer 2 Layer 1 Layer 2
HCC1 155 140 0.8 0.8 177.5 369.2
HCC2 596 524 0.8 0.8 177.5 369.2
HCC3 155 140 0.6 0.6 177.5 369.2
HCC4 250 140 0.8 0.8 250.0 300.0

The values of the properties used in HCC1 are obtained by back-fitting the the

measured stress-strain curve in CU triaxial test in the laboratory, Figure 5.37. In

HCC2, G0 is increased several times above that measured by conventional triaxial test,

in order to examine the effect of underestimating the initial stiffness, G0, due to errors

in the triaxial test data. HCC3 studies the effect of lowering the parameter n to that

which is representative of a low plasticity soil. Finally, the parameters of case HCC4

are determined by back-fitting the final deflection profile of the soldier pile.

175 300
(a) (b)
150 250
Deviator stress, q (kPa)

Deviator stress, q (kPa)

125
200
100
150
75 Triaxial test Triaxial test
HCC1 100 HCC1
50 HCC2 HCC2
HCC3 HCC3
25 50
HCC4 HCC4

0 0
0.00 0.01 0.02 0.03 0.04 0.00 0.01 0.02 0.03 0.04 0.05
Deviator strain, εs Deviator strain, εs

Figure 5.37: Comparison of laboratory’s triaxial tests with HCC soil models. (a) Depth of
sample is 11m. (b) Depth of sample is 20m.

129
FEM STUDY OF SERANGOON MRT SITE

Figure 5.38 shows computed final wall deflection profiles from analyses HCC1

and HCC2. As can be seen, the measured soldier pile deflection matches the HCC1

results better than HCC2 results. HCC2, with larger initial stiffness, appears to be

overly stiff in the final stages of excavation. This suggests that the stress-strain curve

determined from the triaxial test gives a reasonable representation of the “real” stress-

strain curve in the stress range experienced in this study, for the purpose of prediction

soldier pile deflection. Driving of soldier piles to the required depth could have

resulted in remoulding of the surrounding soil thereby reducing the soil stiffness in the

vicinity. The ‘real’ stress-strain relationship thus approaches that of the ‘disturbed’

sample used in the CU triaxial test.

10

15
Depth (m)

20

25

30
Field
HCC1-I
35 HCC1-II
HCC2-I
40 HCC2-II

45
020 40 60 80
Deflection (mm)
Figure 5.38: Final deflection profile of soldier pile and timber lagging for HCC1 and HCC2.

130
FEM STUDY OF SERANGOON MRT SITE

10

15

Depth (m)
20

25

30
Field
35 HCC1-I
HCC1-II
HCC3-I
40 HCC3-II

45
040 80 120 160
Deflection (mm)
Figure 5.39: Final deflection profile of soldier pile and timber lagging for HCC1 and HCC3.

Figure 5.39 shows the final deflection profiles for HCC1 and HCC3. As can be

seen, HCC3 gives a much larger wall deflection that HCC1 and the measured wall

deflection. This can be readily explained by the lower value of n used in analysis

HCC3, which reduces the initial modulus. Thus, n=0.8 appears to be better describe

the variation of G0 with mean effective stress. This is not surprising since the residual

soil encountered in the Serangoon Station site consists predominantly of fine-grained

soils, with medium-high plasticity. Figure 5.40 compares the deflection profiles for

HCC 1 and HCC4. As can be seen, the results of the two analyses are quite close; the

main difference being that HCC4 computes a smaller soldier pile deflection than

HCC1, in layer 1 of the soil near the ground surface. This is not surprising, given the

higher modulus assigned to the layer 1 soil in HCC4.

131
FEM STUDY OF SERANGOON MRT SITE

10

15

Depth (m)
20

25

30
Field
HCC1-I
35 HCC1-II
HCC4-I
40 HCC4-II

45
020 40 60 80
Deflection (mm)
Figure 5.40: Final deflection profile of soldier pile and timber lagging for HCC1 and HCC4.

Comparison of the performance of the three models in predicting wall

deflection alone shows that the modified Cam-clay and the hyperbolic Cam-clay

behaviour very similarly. With appropriate values of parameters, both models can

reasonably predict the deflection of the soldier pile. Furthermore and in contrast to the

Mohr-Coulomb model, both the Cam-clay models predict relatively small differences

between soldier pile and timber lagging deflection. This is due mainly to the shape of

the Cam-clay yield surface. In other words, the relative performances of the two Cam-

clay models are not readily differentiable by wall deflection at final excavation alone.

132
FEM STUDY OF SERANGOON MRT SITE

0 0

5 5

10 10

15 15
Depth (m)

Depth (m)
20 20

25 25

30 30

35 35

40 40 Field
FEM
45 45
0 20 40 60 0 20 40 60
Deflection (mm) Deflection (mm)
(a) (b)
Figure 5.41: Deflection profile of soldier pile for HCC4. (a) Before installation of first level
struts. (b) After installation of forth level struts.

Some differences in wall deflection prediction may be observed in the early

stages of excavation. Figures 5.41a shows the deflection of the soldier pile at an early

stage of the excavation, before the installation of the first level struts for the case

HCC4. Comparison of Figures 5.41a and 5.31a shows that case HCC4 computes a

smaller wall deflection profile than case MCC1. It is also much closer to the field

result than case MCC1. This supports the earlier hypothesis that the over-prediction of

wall deflection by the modified Cam-clay at the initial stages of excavation arises from

its inability to capture the degradation in shear modulus with shear strain.

133
0 0 0 0

5 5 5 5

10 10 10 10

15 15 15 15

20 20 20 20

25 (a) 25 25 25

Depth (m)
Depth (m)
Depth (m)
Depth (m)
30 Field 30 30 30
2D
35 35 35 35
3D-I
40 3D-II 40 40 40
(b) (c) (d)
45 45 45 45
0 20 40 60 80 100 0 20 40 60 80 100 0 20 40 60 80 100 0 20 40 60 80 100
Deflection (mm) Deflection (mm) Deflection (mm) Deflection (mm)

0 0 0 0

5 5 5 5

10 10 10 10

15 15 15 15

20 20 20 20

25 25 25 25

Depth (m)
Depth (m)
Depth (m)
Depth (m)

30 30 30 30

35 35 35 35

40 40 40 40
(e) (f) (g) (h)
45 45 45 45
0 20 40 60 80 100 0 20 40 60 80 100 0 20 40 60 80 100 0 20 40 60 80 100
Deflection (mm) Deflection (mm) Deflection (mm) Deflection (mm)

134
Figure 5.42: Deflection profiles at various stages of excavation for HCC4-M. (a) Stage B. (b) Stage C. (c) Stage D.
FEM STUDY OF SERANGOON MRT SITE

(d) Stage E. (e) Stage F. (f) Stage G. (g) Stage H. (h) Stage I.
FEM STUDY OF SERANGOON MRT SITE

Development of wall deflection with excavation depth

Figure 5.42 shows the 2D without kingposts and 3D progressive movement of

the retaining wall for HCC4-M from Stage B–I as illustrated in Figure 5.3 compared to

the field results. With almost the same final deflection profile as the MCC, the HCC

soil model, with non-linear stiffness, shows a more gradual development of the soldier

pile wall deflection compared to the former. It was noted the in all soil models, the 2D

deflections below the excavation level are significantly larger than the 3D deflection.

Strut forces

Table 5.8 compares the strut forces obtained from the best-fit finite element

analyses for HCC4. Similar to MC5 and MCC1, large discrepancies between the field

and analysed results are found in the first and fifth struts. The reasons for these

discrepancies have been presented earlier and will not be repeated here.

Table 5.8: Comparison of strut forces at various levels between instrumented and FEM result
for the best-fit deflection profile for HCC4.
Strut level Instrumented (kN) FEM (kN) Instrumented/FEM × 100%
1 173.1 386.7 44.8
2 692.3 746.5 92.7
3 836.5 850.7 98.3
4 865.4 947.6 91.3
5 1015.4 490.8 206.9
6 126.9 166.7 76.1

Ground settlement

Figure 5.43 shows the computed ground settlements behind the retaining wall

after excavation to the final level. As for the modified Cam-clay model, the hyperbolic

Cam-clay model also predicts much larger settlement than the field data. This is also

attributable to the fact that the moduli in the hyperbolic Cam-clay are also a function of

the effective stress. As the effective stress approaches zero near the ground surface, so

does the modulus.


135
FEM STUDY OF SERANGOON MRT SITE

Distance from retaining wall (m)


0 10 20 30 40 50 60 70 80 90

0
L215(G)
L214(G) L217(G)
20
L203(G) L222(G)
Settlement (mm)

40

60
Field
80 HCC1
HCC2
100 HCC3
HCC4

120
Figure 5.43: Ground settlement behind Location I of the retaining wall for HCC1, HCC2,
HCC3 and HCC4.

Distance from retaining wall (m)


0 10 20 30 40 50 60 70 80 90

0.2
Settlement
(mm)

0.4
Settlement
Normalised

0.6 MCC1
MCC2
MCC3
0.8 MCC4
MCC5
1
Figure 5.44: Normalised ground settlement behind Location I of the retaining wall for MCC1,
MCC2, MCC3, MCC4 and MCC5.

Figure 5.44 and 5.45 compare the relative decay in settlements computed by the

modified Cam-clay and the hyperbolic Cam-clay model respectively. In these figures,

all the settlement values have been normalised by their respective maximum so that the

maximum normalised settlement in all cases is 1. As can be seen, the hyperbolic Cam-

clay cases predict a slightly higher rate of attenuation in settlement with distance, after

136
FEM STUDY OF SERANGOON MRT SITE

the point of maximum settlement, than the modified Cam-clay cases. This effect is

well-documented (e.g. Jardine et al., 1986).

Distance from retaining wall (m)


0 10 20 30 40 50 60 70 80 90

0.2
Settlement
(mm)

0.4
Settlement
Normalised

0.6
HCC1
HCC2
0.8 HCC3
HCC4
1
Figure 5.45: Normalised ground settlement behind Location I of the retaining wall for HCC1,
HCC2, HCC3 and HCC4.

Figure 5.46 shows various cases of 2- and 3D analyses which explore the

effects of some other factors on wall deflection. As can be seen, the results are also

slightly dependent upon whether the analysis is 2- or 3D. Given the same parameters,

the 3D analysis predicts a slightly smaller wall deflection than the 2D analysis,

especially below excavation level. This is attributable to the mechanism of soil flow

around the soldier piles, which have been discussed in Chapter 3. As discussed then,

the flexural stiffness of the soldier piles tends to drag the retained soil into the

excavated area, and this effect is accentuated in the 2D analysis which modelled the

soldier piles as a wall with equivalent flexural rigidity. For this reason, the settlement

predicted by the 2D analysis is also slightly larger than that predicted by the 3D

analysis, Figure 5.47.

137
FEM STUDY OF SERANGOON MRT SITE

10

15

Depth (m)
20

25

30

35
MCC1-2D
MCC1-3D
40 HCC4-2D
HCC4-3D
45
20 40 0 60 80
Deflection (mm)
Figure 5.46: 2D vs 3D wall deflections behind Location I of the retaining wall for MCC1 and
HCC4.

Distance from retaining wall (m)


0 10 20 30 40 50 60 70 80 90

20
Settlement (mm)

40

60

80 MCC1-2D
MCC1-3D
100 HCC4-2D
HCC4-3D
120
Figure 5.47: 2D vs 3D ground settlements behind Location I of the retaining wall for MCC1
and HCC4.

One factor which may have an effect on the settlement profile is the kingposts

within the excavation area. Figure 5.48 illustrates the effect of modelling the kingposts

in the analyses. As mentioned earlier, kingposts were modelled using beam elements.

Two 2D analyses were performed; the first without modelling the kingposts and the

138
FEM STUDY OF SERANGOON MRT SITE

secondly modelling the kingposts as a kingpost wall with equivalent axial stiffness. As

can be seen, the wall deflection predicted by the two 2D analyses bracket the

prediction of the 3D analyses (with and without kingposts). If the kingposts were not

modelled, the wall deflection of the 2D analysis is larger than that of the 3D analysis.

This arises from the soldier-pile-soil interaction discussed above. On the other hand, if

the kingposts were modelled, the wall deflection of the 2D analysis is smaller than that

of the 3D analysis. This is because the kingposts had to be modelled as an equivalent

wall in the 2D analysis, which inhibits the flow of soil around the kingposts, thereby

giving rise to a spurious stiffening effect. The effect of the kingpost is particularly

evident in the wall deflection just below excavation level and in the attenuation of

settlement just after the maximum settlement is reached. Similar observation is made

in MCC1-M (Figure 5.49).

10

15
Depth (m)

20

25

30

35
2D without kingpost
2D with kingpost
40 3D without kingpost
3D with kingpost
45
20 0 40 60 80 100
Deflection (mm)
Figure 5.48: Comparison of effect of kingpost modelling on 2D and 3D soldier pile deflection
for HCC4-M.

139
FEM STUDY OF SERANGOON MRT SITE

10

15

Depth (m)
20

25

30

35
2D without kingpost
40 2D with kingpost
3D with kingpost
45
20 0 40 60 80 100
Deflection (mm)
Figure 5.49: Comparison of effect of kingpost modelling on 2D and 3D soldier pile deflection
for MCC1-M.

The above effects are also evident from the displacement field of the retained

soil. Figure 5.50 depicts the displacement vectors for both 2D without kingpost and

3D with kingpost HCC4-M analyses. In 3D analysis, the slightly larger deflection of

timber lagging allows for horizontal movement of soil towards the mid-span which is

not possible in the case of 2D analysis as reflected in Figure 5.42. The continuous

simulation of piles in the 2D causes deeper soil movement immediately behind the

retaining wall. When there are no kingpost to restrain the swelling of soil at the

excavation base, the larger upward movement in the excavation is likely to cause

larger settlement a distance away from the retaining wall reflected in Figure 5.51.

When the kingposts are modelled in the 2D analysis, the settlement beyond 10m from

the retaining wall approaches that of 3D analysis. The additional restraint to the lateral

movement and swelling of soil below the excavation afforded by 2D modelling of

kingposts reduces the ground settlement.

140
FEM STUDY OF SERANGOON MRT SITE

-10
Depth (m)

-20

-30
2D
3D
-40
0 10 20 30 40 50 60
Distance from center of excavation (m)
Figure 5.50: 2D and 3D displacement vectors behind Location I of the retaining wall for
HCC4-M.

Distance from retaining wall (m)


0 10 20 30 40 50 60 70 80 90
0
L215(G)
Settlement (mm)

10
L217(G)
L214(G)
20
Field
30 L222(G) 2D without kingposts
L203(G) 2D with kingposts
3D with kingposts
40
Figure 5.51: 2D and 3D ground settlement behind Location I of the retaining wall for HCC4-
M.

In the immediate zone up to 10m, ground settlements for both 2D analyses are

almost identical but larger than 3D profile. This can be attributed to the larger 2D

deflection profiles than in 3D reflected by Figure 5.47 that affects the near field ground

settlement in Figure 5.50. This shows that the effects of kingposts cannot be properly

simulated in 2D analyses.

141
FEM STUDY OF SERANGOON MRT SITE

5.4 Summary of findings

In the Mohr-Coulomb model, the measured angle of friction was used to define

the yield surface, on the other hand, the modified Cam-clay clay, the angle of friction

was used to define the critical state line. This implies that, on the “dry” side of critical,

the modified Cam-clay yield surface is larger than the Mohr-Coulomb yield surface. If

the excavation behaviour is controlled by “dry” side yielding, the Mohr-Coulomb

model will yield more readily than the modified Cam-clay model. For this reason, the

Young’s modulus of the Mohr-Coulomb has to be raised in order to fit the measured

wall deflection. In this case study, the soil remains predominantly elastic with the

modified Cam-clay model, thus, dilatancy is not observed in the MCC-analyses.

Given the fact that the 30° friction angle is not uncommonly used value, it is possible

that premature yielding may well be a common syndrome in many Mohr-Coulomb FE

analyses of residual or over-consolidated soils in excavation problems.

The modified Cam-clay model predicts linear elastic right up to the point of

yield whereas the measured data indicate that a curve is more appropriate. The reason

for this is that modified Cam-clay model in CRISP does not model any shear-induced

reduction in shear modulus within the yield surface, thereby the stiffer initial tangent

modulus of hyperbolic Cam-clay that is not reflected by the highly linear pre-yield

segments of the stress-strain curves computed by the modified Cam-clay model.

The discussion above shows that all three models used are able to compute wall

deflections which are in reasonable agreement with the measured deflection. This

implies that wall deflection alone may not be sufficient to highlight the differences

between the various models.

142
FEM STUDY OF SERANGOON MRT SITE

In the ground settlement studies, the discrepancy in the implied modulus is

accentuated in both the MCC and HCC models for the shallow soil layers where the

effective stresses are lower. Such discrepancy can be overcome by prescribing the fill

as Mohr Coulomb material. It was also reflected in the modified and hyperbolic Cam-

clays analyses that the attenuation of the settlement profile with distance is more

pronounce in the latter. In the 2D analyses, the kingposts had to be modelled as an

equivalent wall, which inhibits the flow of soil around the kingposts, thereby giving

rise to a spurious stiffening effect.

143
CONCLUSION AND RECOMMENDATIONS

CHAPTER 6 CONCLUSION AND RECOMMENDATIONS

6.1 Concluding remarks

The findings arising from foregoing discussion can be summarised as follows:

1) When 2-D analysis is applied to soldier-piled excavations, modelling errors may

arise in several ways as follows:

a) Firstly, by modelling the discrete soldier piles as a wall, a 2D analysis tends to

over-predict the coupling of pile to the soil below excavation level. In

instances where the pile is not embedded into stiff soil, this modelling error can

give rise to discernible underestimation of soldier pile deflection.

b) Secondly, the deflection of the timber lagging, which is usually larger than that

of the soldier piles, is often underestimated.

c) Thirdly, a 2D analysis cannot replicate the differences in stress paths taken by

the retained soil at different locations at the same depth. In particular, it cannot

model the softening of the soil face just behind the timber lagging. In instances

where the timber lagging is insufficient or its installation, this softening can

give rise to local collapse, which may exacerbate the ground loss and therefore

ground movement. Increasing the inter-pile spacing will tend to accentuate the

effects of these modelling errors.

2) Comparison with field data from the Serangoon MRT site shows that using the

measured angle of friction to define the yield criteria in Mohr Coulomb and critical

state line in modified Cam-clay leads to very different predicted behaviour in the

vicinity in the soldier pile wall in the 3-D analysis, which is attributable to the
144
CONCLUSION AND RECOMMENDATIONS

differences in the “dry-side” yielding behaviour of the two models. These

differences in behaviour between soldier pile and timber lagging is not evident in

2-D analysis, which cannot explicitly predict the behaviour of the timber.

3) Comparison of the modified and hyperbolic Cam-clay models shows that while the

former may be able to fit the final deflection profile if assigned appropriate model

parameters, it tends to over-predict the wall deflection in the early stage of

excavation owing to its inability to model the reduction in the moduli of the soils

with increasing shear strain.

4) The kingposts have a discernible effect in reducing the wall deflection below

excavation level, which in turn affects the far-field settlement characteristics of the

retained soil.

5) In this study, soldier pile installation was not explicitly modelled but the sensitivity

of the analysis to the change of K0-values arising from pile installation was studied.

The range of K0-values varied is likely to encompass that incurred by the

installation of the H-piles. The results of these studies indicate that the broad trend

of behaviour was not altered.

6.2 Recommendations for Future Work

The knowledge gathered from this study pave way for further numerical studies

to assess the effect of piles spacing and pile width on the arching action in the retained

soil. The inter-pile spacing can be gradually increased from a fore-determined value,

with constant equivalent stiffness of soldier piles, until the collapse of retained soil is

observed. Effectiveness of timber lagging in preventing progressive deterioration of

the soil arching between the piles can thus be examined. This will provide guidelines

145
CONCLUSION AND RECOMMENDATIONS

to suitability of pile spacing used in mobilising the arching action in the retained soil

based on 3D stress distribution in finite element analysis.

In reality, unsaturated soils are often encountered. The inability of Crisp90 to

model unsaturated flow is widely recognised (Lee et al., 1993 and Hsi & Small, 1992).

The conductivity relation of a particular soil can be incorporated to modify Biot’s

consolidation theory for analysing the unsaturated flow in residual soil. The

drawdown effect of groundwater table can thus be more reasonably represented.

146
REFERENCES

REFERENCES

AITC (1994). Timber Construction Manual, 4th Edn. American Institute of Timber

Construction. Wiley, New York.

Al-Hussaini, M. and Perry, E.B. (1978). Field Experiment of Reinforced Earth Wall. Journal

of Geotechnical Engineering, Div., ASCE, Vol. 104, No. GT3, Mar., pp. 307-322.

Armento, W.J. (1972). Criteria for Lateral Pressures for Braced Cuts. Proceedings of the

Specialty Conference on Performance of Earth and Earth-Supported Structures. Vol 1,

Part 2, Purdue University, Indiana, pp. 1283-1302.

Balasubramaniam, A.S. and Brenner, R.P. (1981). Consolidation and Settlement of Soft Clay.

Soft Clay Engineering, (Eds. A.S. Balasubramaniam and R.P. Brenner), Elsevier, New

York, pp. 481-566.

Bica, A. V. D. and Clayton, C. R. I. (1998). An Experimental Study of Behaviour of

Embedded Lengths of Cantilever Walls. Geotechnique, Vol. 48, No. 1, pp. 731-745.

Biot, M.A. (1941). General Theory of Three Dimensional Consolidation. Journal of Applied

Physics, Vol. 12, pp. 155-164.

Bjerrum, L. (1973). Problems of Soil Mechanics and Construction on Soft Clays and

Structurally Unstable Soils. Proceedings of the 8th International Conference on Soil

Mechanics and Foundation Engineering, Vol. 3, pp. 111-159.

Bolton, M. (1991). A Guide to Soil Mechanics. Macmillan Education Ltd. London.

Borja, R.I., Lee, S.R. and Seed, R.B. (1989). Numerical Simulation of Excavation in

Elastoplastic Soils. International Journal for Numerical and Analytical Methods in

Geomechanics, Vol. 13, pp. 231-249.

Borst, A.J., Conley, T.L., Russell, D.P. and Boirum, R.N. (1990). Subway Design and

Construction for the Downtown Seattle Transit Project. Proceedings of Design And

Performance of Earth Retaining Structures, ASCE Special Conference, Ithaca, New

York, pp. 510-524.

147
REFERENCES

Bosscher, P.J. (1981). Soil Arching in Sandy Slopes. Ph.D. Thesis, The University of

Michigan, U.S.

Bowles, J.E. (1988). Foundation Analysis and Design. McGraw Hill, New York.

Bransby, M.F. and Springman, S.M. (1996). 3-D Finite Element Modelling of Pile Groups

Adjacent to Surcharge Loads. Computers and Geotechnics, Vol. 19, pp. 301-324.

Briaud J.-L. and Lim, Y (1999). Tieback Walls in Sand: Numerical Simulation and Design

Implications. Journal of Geotechnical and Geoenvironmental Engineering, Vol. 125,

No. 2, pp. 101-110.

Britto, A.M. and Gunn, M.J. (1990). CRISP 90 User's & Programmer’s Guide. Cambridge.

Britto, A.M. and Kusakabe, O. (1983). Stability of Axisymmetric Excavations in Clay. Journal

of Geotechnical Engineering, Div., ASCE, Vol. 109, No. 5, pp. 666-681.

Broms, B.B. (1964). Lateral Resistance of Piles in Cohesive Soils. Journal for Soil Mechanics

and Foundation Engineering, ASCE, Vol. 90, SM2, pp. 27-64.

Brumund, W.F., Jonas, E., and Ladd, C.C. (1976). Estimating In Situ Maximum Past

Preconsolidation Pressure of Saturated Clays from Results of Laboratory

Consolidometer Tests. Special Report 163, Transportation Research Board, pp. 4-12.

BS 5930 (1999). Code of Practice for Site Investigations. British Standards Institutions.

BS 8002 (1994). Code of Practice for Earth Retaining Structures. British Standards

Institutions.

BSI (1996). Steelwork Design Guide to BS 5950: Part 1: 1990, Vol.1, Section Properties,

Member Capacities, 4th Edn. British Standards Institutions.

Burland, J.B. (1967). Deformation of Soft Clay. Ph.D. Thesis, University of Cambridge, U.K.

Burland, J.B. (1989). ‘Small is Beautiful’-the Stiffness of Soils at Small Strains. Canadian

Geotechnical Journal, Vol. 26, pp. 499-516.

Burland, J.B., Simpson, B. and St. John, H.D. (1979). Movements around Excavations in

London Clay. Proceedings of 7th European Conference in Soil Mechanics, Brighton,

Vol. 1, pp. 13-30.

148
REFERENCES

Canadian Geotechnical Society (1992). Canadian Foundation Engineering Manual. Canadian

Geotechnical Society, Technical Committee on Foundations.

Cardoso, A.S. (1987). Soil Nailing Applied to Excavations. Ph.D. Thesis, University of Porto,

Portugal.

Chen, D.C., Osborne, N., Shim, J.D. and Hong, E. (2000). An Observational Approach to the

Monitoring of the Excavation at Harbour Front Station, Singapore. Tunnels and

Underground Structures, (Eds. J. Zhao, J.N. Shirlaw & R. Krishnan), Balkema,

Rotterdam, pp. 465-471.

Chong, P.T. (1998). Characterisation of Singapore Soils. Ph.D. Conversion Report, National

University of Singapore.

Chudnoff, M. (1984). Tropical Timbers of the World, Agriculture Handbook 607, U.S.

Department of Agriculture, Forest Service, Forest Products Laboratory, Madison, WI.

Clarke, B.G and Wroth, C.P. (1984). Analyses of the Dunton Green Retaining Wall Base on

Results of Pressuremeter Tests. Geotechnique, Vol. 34, No. 4, pp. 549-562.

Clough, G.W. and Denby, G.M. (1977). Stabilising Berm Design for Temporary Walls in

Clay. Journal of Geotechnical Engineering, Div., ASCE, Vol. 103, No. GT2, pp.75-90.

Clough, G.W. and Hansen, L.A. (1981). Clay Anisotropy and Braced wall Behaviour. Journal

of Geotechnical Engineering, Div., ASCE, Vol. 107, No. 7, pp. 893-913.

Clough, G.W. and O’Rourke, T.D. (1990). Construction-Induced Movements of In-situ Walls.

Proceedings of Design And Performance of Earth Retaining Structures, ASCE Special

Conference, Ithaca, New York, pp. 439-470.

Clough, G.W. and Tsui, Y. (1977). Static Analysis of Earth Retaining Structures. Numerical

Methods in Geotechnical Engineering, (Eds. C.S. Desai & J.T. Christian), McGraw

Hill, New York, pp. 506-527.

Clough, G.W., Hansen, L.A. and Mana, A.I. (1979). Prediction of Support Excavation

Movements Under Marginal Stability Conditions in Clay. Proceedings of 3rd

International Conference on Numerical Methods in Geomechanics, Aachen, Vol. 4, pp.

1485-1502.

149
REFERENCES

Clough, G.W., Weber, P.R. and Lamont, J. (1972). Design and Observation of a Tie-back

Wall. Proceedings of the Specialty Conference on Performance of Earth and Earth-

Supported Structures, Vol. 1, New York, pp. 1367-1389.

Committee on Glossary of Terms and Definitions in Soil Mechanics (1958). Glossary of

Terms and Definitions in Soil Mechanics. Journal of the Soil Mechanics and

Foundations. Div., ASCE, Vol. 84, No. 4, pp. 1-43.

Coop, M.R. and Jovičić, V. (1999). The Influence of State on the Very Small Strain Stiffness

of Soils. 2nd International Symposium for Pre-faliure Deformation of Geomaterials IS-

Torino ’99, Turin, pp 175-181.

Coutts, D.R. and Wang, J. (2000). The Monitoring and Analysis of Results for Two Strutted

Deep Excavations in Singapore using Vibrating Wire Strain Gauges. Tunnels and

Underground Structures, (Eds. J. Zhao, J.N. Shirlaw & R. Krishnan), Balkema,

Rotterdam, pp. 465-471.

Craig, R.F. (1996). Soil Mechanics, 5th Edn. Chapman & Hall. London.

Dames and Moore (1983). Detailed Geotechnical Study Interpretative Report, Submitted to

Singapore Mass Rapid Transit System.

Dasari, G. R. (1996). Modelling of the Variation of Soil Stiffness during Sequential

Construction. Ph. D. Thesis, University of Cambridge, U.K.

Day, R.A. and Potts, D.M. (1993). Modelling Sheet Pile Retaining Walls. Computers and

Geotechnics, Vol. 15, pp. 125-143.

De Moor, E.K. (1994). An Analysis of Bored Pile/Diaphragm Wall Installation Effects.

Geotechnique Vol. 44, No. 2, pp. 341-347.

Duncan, J.M. and Buchignani, A. L. (1976). An Engineering Manual for Settlement Studies.

Geotechnical Engineering Report, University of California, Berkeley.

Dysli, M. and Fontana, A. (1982). Deformations Around the Excavation in Clayey Soil. Proc.

Of the Int. Symposium on Numerical Models in Geomechanics, Zurich, Sept. pp. 634-

642.

150
REFERENCES

Ekström, J. (1989). Field Study of Model Pile Group Behaviour in Non-cohesive Soils –

Influence of Compaction due to Pile Driving. Ph. D. Thesis, Chalmers University of

Technology, Göteborg, Sweden.

Finno, R.J., Harahap, I.S. and Sabatini, P.J. (1991). Analysis of Braced Excavations with

Coupled Finite Element Formulations. Computers and Geotechnics, Vol. 12, pp. 91-

114.

Fourie, A.B. and Potts, D.M. (1989). Comparison of Finite Element and Limit Equilibrium

Analysis for an Embedded cantiliver Retaining Wall. Geotechnique, Vol. 39, No. 1,

pp. 175-188.

GCO (1990). Review of Design Methods for Excavations. Geotech. Control Office, Hong

Kong.

Getzler, Z., Komovnik, A. and Mazurik, A. (1968). Model Study on Arching Above Buried

Structures. Journal of the Soil Mechanics and Foundations. Div., ASCE, Vol. 94, No.

SM5, pp. 1123-1141.

Giger, M.W. and Krizek, R.J. (1975). Stability Analysis of Vertical Cut with Variable Corner

Angle. Soils and Foundations, The Japanese Society of Soil Mechanics and

Foundation Engineering, Vol. 15, No. 2, pp. 63-71.

Gill, S.A. and Lukas, R.G. (1990). Ground Movement Adjacent to Braced Cuts. Proceedings

of Design And Performance of Earth Retaining Structures, ASCE Special Conference,

Ithaca, New York, pp. 471-488.

Goh, S.H. (1995). Propagation and Attenuation of Ground Shock in Dry Soil. M. Eng. Thesis,

National University of Singapore, Singapore.

Goldberg, D.T., Jaworski, W.E. and Gordon, M.D. (1976). Lateral Support Systems and

Underpinning, Federation of Highway Administration. Rep. FHWA-RD-75-128, 3

volumes.

Gomes Correia, A. and Guerra, N.M. da C. (1997). Performance of three Berline-type

Retaining Walls. Proceedings of the 14th International Conference of Soil Mechanics

and Foundation Engineering, Hamburg, Vol. 2, pp. 1297-1300.

151
REFERENCES

Gourvenec, S. M. and Powrie, W. (1999). Three-dimensional Finite-element Analysis of

Diaphragm Wall Installation. Geotechnique, Vol. 49, No. 6, pp. 801-823.

Graham, J. (1992). The Key Material Properties in Geotechnical engineering need to be

measured, instead of being specified. Proceedings of the Wroth Memorial Symposium,

Oxford, pp. 1-18.

Gunn, M.J. Satkunananthan, A. and Clayton, C.R.I. (1992). Finite Element Modelling of

Installation Effects. Proceedings Conference on Retaining Structure by ICE,

Cambridge, pp. 46-55.

Handy, R.L. (1985a). The Arch in Soil Arching. Journal of Geotechnical Engineering, Div.,

ASCE, Vol. 111, No. 3, pp. 302-318.

Handy, R.L. (1985b). Discussion of “The Arch in Soil Arching”: Closure by Richard L.

Handy. Journal of Geotechnical Engineering, Div., ASCE, Vol. 111, No. 4, pp. 275-

277.

Hashash, Y.M.A. and Whittle, A.J. (1992). Analysis of Braced Diaphragm Walls in Deep

Deposits of Clay. Res. Rep. R92-19, Department of Civil Engineering, Massachusetts

Institute of Technology (MIT), Cambridge, Mass.

Hashash, Y.M.A. and Whittle, A.J. (1996). Ground Movement Prediction for Deep

Excavations in Soft Clay. Journal of Geotechnical Engineering, Div., ASCE, Vol. 122,

No. 6, June, pp. 474.-486.

Hentenyi, M. (1946). Beams on Elastic Foundations. The University of Michigan Press, Ann

Arbor, MI.

Hong, S.H., Lee, F.H. and Yong, K.Y. (2003). Three-dimensional Pile-soil Interaction in

Soldier-piled Excavations. Computers and Geotechnics, Vol. 30, pp. 81-107.

Hsi, J.P. and Small, J.C. (1992). Ground Settlements and Drawdown of the Water Table

around an Excavation. Canadian Geotechnical Journal, Vol. 29, pp. 740-754.

Jáky, J. (1944). The Coefficient of Earth Pressure at Rest. Journal of the Society of Hungarian

Architects and Engineers, Oct., Budapest, pp 355-358.

152
REFERENCES

Jardine, R.J., Symes, M.J. and Burland, J.B. (1984). The Measurement of Soil Stiffness in the

Triaxial Apparatus. Geotechnique, Vol. 34, No. 3, pp. 323-340.

Jardine, R.J., Potts, D.M., Fourie, A.B. and Burland, J.B. (1986). Studies of the Influence of

Non-linear Stress-strain Characteristics in Soil-structure Interaction. Geotechnique,

Vol. 36, No. 3, pp. 377-396.

Kingsley, H.W. (1989). Arch in Soil Arching. Journal of Geotechnical Engineering, Div.,

ASCE, Vol. 115, No. 3, March, pp. 415-419.

Krynine, D.P. (1945). Discussion of “Stability and Stiffness of Cellular Cofferdams” by Karl

Terzaghi. Transactions, ASCE, Vol. 110, pp. 1175-1178.

Kulhawy, F.H. (1990). Manual on Estimating Soil Properties for Foundation Design. Electric

Power Research Institute, Report no. EL-6800, Ithaca, New York.

Lambe, T.W. and Whitman, R.V. (1979). Soil Mechanics, SI Version. John Wiley & Sons, Inc.

New York.

Lee, F.H., Tan, T.S. and Yong, K.Y. (1993). Excavation in Residual Soils with High

Permeability. Proceedings of 11th Southeast Asian Geotechnical Conference,

Singapore, pp. 739-744.

Lee, F.H., Tan, T.S. and Wang, X.N. (2001). Groundwater Effects in Soldier-piled Excavation

in Residual Soils. Proceedings of Underground Singapore 2001, Singapore, pp. 249-

260.

Lee, F.H., Yong, K.Y., Lee, S.L. and Toh, C.T. (1989). Finite Element Modelling of Strutted

Excavation. Numerical Models in Geomechanics (NUMOG III), Niagra Falls, Canada,

pp. 577-584.

Lee, S.L., Parnploy, U., Yong, K.Y. and Lee, F.H. (1989). Time-dependent Deformation of

Anchored Excavation in Soft Clay. Symp. On Underground Excavations in Soils and

Rocks, Including Earth Pressure Theories, Buried Structures and Tunnels, Bangkok,

Thailand, pp. 377-384.

153
REFERENCES

Lings, M.L., Nash, D.F.T., Ng, C.W.W. (1993). Reliability of Earth Pressure Measurements

Adjacent to a Multipropped Diaphragm Wall. Proceedings of an International

Conference on Retaining Structures, Cambridge, U.K., pp. 258-269.

Liu, K.X. (1995). Three Dimensional Analyses of Deep Excavation in Soft Clay. M. Eng.

Thesis, National University of Singapore, Singapore.

Livesley, R.K. (1983). Finite Element: An Introduction for Engineers. Cambridge University

Press, Cambridge.

Loh, C. K., Tan T. S. and Lee F. H. (1998). Three-dimensional excavation tests. Proceedings

of International Conference Centrifuge 98, 23-25 September, Tokyo, Japan, pp. 649-

654.

Lusher, U. and Hoeg, K. (1964). The Beneficial Action of the Surrounding Soil on the Load

Carrying Capacity of Buried Tubes. Proceedings of the 1964 Symposium on Soil-

Structure Interaction, Tuscon, Arizona, pp. 393-402.

Macneal, R.H. (1994). Finite Elements: Their Design and Performance. Marcel Dekker, Inc.,

New York.

Mana, A.I. and Clough, G.W. (1981). Prediction of Movements for Braced Cuts in Clay.

Journal of Geotechnical Engineering, Div., ASCE, Vol. 107, pp.759-777.

Massing, G. (1926). Eigenspannungen und Verfesttigung beim Messing. Proceedings of 2nd

International Congress of Applied Mechanics.

Matos Fernandes, M.A. (1986). Three-dimensional Analysis of Flexible Earth Retaining

Structures. Proceedings of Second NUMOG (Eds. G.N. Pande & W.F. van Impe), M.

Jackson & Sons, UK.

Matos Fernandes, M.A., Cardoso, A.J.S.,Trigo J.F.C. and Marques J.M.M.C. (1986). Finite

Element Modelling of Supported Excavations. Soil-Structure Interaction: Numerical

Analysis & Modelling (Ed. J.W. Bull), E & FN Spon.

Mayne, P.W. and Kulhawy, F.H. (1982). Ko – OCR Relationships in Soil. Journal of the

Geotechnical Engineering Division, ASCE, Vol.108, No. GT6, pp. 851-872.

154
REFERENCES

Nakai, T. (1985). Finite Element Computations for Active and Passive Earth Pressure

Problems of Retaining Wall. Soil and Foundations, Vol. 25, No. 3, pp. 98-112.

Nasim, A. S. M. (1999). Numerical Modelling of Soil Profile and Behaviour in Deep

Excavation Analyses. M. Eng. Thesis, National University of Singapore, Singapore.

NAVFAC (1982). Design Manual 7.2: Foundations and Earth Structures, Department of the

Navy, Naval Facilities Engineering Command, Washington D.C.

Ng, C.W.W., Simpson, B., Lings, M.L. and Nash, D.F.T. (1995). An Approximate Analysis of

the Three-dimensional Effects of Diaphragm Wall Installation. Geotechnique, Vol. 45,

No. 3, pp. 497-507.

Ng, C.W.W. and Yan, R. W. M. (1999). Three-dimensional Modelling of a Diaphragm Wall

Construction Sequence. Geotechnique, Vol. 49, No. 6, pp. 825-834.

O’ Rourke, T.D. (1975). A Study of two Braced Excavations in Sands and Interbedded Stiff

Clay. PhD Thesis, University of Illinois, U.S.

O'Rourke, T.D., Cording, E.J. and Boscardin, M. (1976). The Ground Movements Related to

Braced Excavation and their Influence on Adjacent Buildings. U.S. Department of

Transportation, Report no. DOT-TST 76, T-23.

Ou, C.Y. and Chiou, D.C. (1993). Three-Dimensional Finite Element Analysis of Deep

Excavation. 11th Southeast Asian Geotechnical Conference, Singapore, pp. 769-774.

Ou, C.Y. and Shiau, B.Y. (1998). Analysis of the Corner Effect on Excavation Behaviours.

Canadian Geotechnical Journal, Vol. 35, pp. 532-540.

Ou, C.Y., Chiou, D.C. and Wu, T.S. (1996). Three-Dimensional Finite Element Analysis of

Deep Excavations. Journal of Geotechnical Engineering, Div., ASCE, Vol. 122, No. 5,

pp. 337-345.

Ove Arup and Partners, (1997). North East Line Contract 704: Additional Soil Investigation

for Proposed Serangoon Station. Ove Arup and Partners, Singapore.

OYO, (1996). Geotechnical Soil Investigation and Related Works (Serangoon to Kangkar):

Final Report. OYO Corporation Singapore Branch, Singapore.

155
REFERENCES

Parnploy, U. (1990). Time Dependent Behaviour of Excavation Support System in Soft Clay.

Ph. D. Thesis, National University of Singapore, Singapore.

Pearlman, S.L. and Wolosick, J.R. (1990). Anchored Earth Retaining Structures-Design

Procedures and Case Studies. Proceedings of Design And Performance of Earth

Retaining Structures, ASCE Special Conference, Ithaca, New York, pp. 525-539.

Peck, R.B. (1969). Deep Excavations and Tunnelling in Soft Ground. Proceedings of 7th

Conference on Soil Mechanics and Foundation Engineering, Mexico. State of the Art

Report, pp.225-290.

Potts, D. M. and Zdravković, L. (2001). Finite Element Analysis in Geotechnical Engineering:

Application. Thomas Telford, London.

PWD, (1976). Geology of the Republic of Singapore. Public Works Department, Singapore.

Rahim, A. and Gunn, M.J. (1997). Bending CRISP. CRISP NEWS, No.4, Sage Engineering

Ltd.

Reddy, J.N. (1993). An Introduction to the Finte Element Method, 2nd Edn. McGraw Hill, Inc.

New York.

Richards, D.J. and Powrie, W. (1998). Centrifuge Model Tests on Doubly Propped Embedded

retaining Walls in Overconsolidated Kaolin Clay. Geotechnique, Vol. 48, No. 6, pp.

833-846.

Roscoe, K.H., Schofield, A.N. and Thurairajah, A.H., (1963). Yielding of Soils in States

Wetter than Critical, Geotechnique, Vol. 13, pp. 211-240.

Roscoe, K.H., Schofield, A.N. and Wroth, C.P., (1958). On the Yielding of Soils.

Geotechnique, Vol. 8, pp. 22-53.

Schofield, A.N. & Wroth, C.P. (1968). Critical State Soil Mechanics, McGraw Hill, London.

Simpson, B. (1992). Retaining Structures: Displacement and Design. Geotechnique, Vol. 42,

No. 4, pp. 541-576.

Simpson, B., O'Riordan, N.J. and Croft, D.D. (1979). A Computer Model for the Analysis of

Ground Movements in London Clay. Geotechnique, Vol. 29, No. 2, pp. 149-175.

156
REFERENCES

St John, H.J. (1975). Field and Theoretical Studies of the Behaviour of Ground around Deep

Excavations in London Clay. PhD Thesis, University of Cambridge, U.K.

Stallebrass, S.E. (1990). Modelling the Effect of Recent Stress History on the Deformation of

Overconsolidated soils. PhD Thesis, The City University, U.K.

Symons, I.F. and Carder, D.R. (1993). Stress changes in stiff clay caused by the installation of

embedded retaining walls. Retaining Structures (Ed. C.R.I. Clayton), Thomas Telford,

London, pp. 227-236.

Teng, W.C. (1975). Soil Stresses for Design of Drilled Caisson Foundations Subjected to

Lateral Loads. ASCE, Conference on Analysis and Design in Geotechnical

Engineering.

Terzaghi, K. (1936). The Shearing Resistance of Saturated Soils and the Angles between the

Planes of Shear. 1st International Conference of Soil Mechanics and Foundation

Engineering., Cambridge, Vol. 1, pp. 54-56.

Terzaghi, K. (1943). Theoretical Soil Mechanics. John Wiley and Sons, Inc., New York.

Terzaghi, K. and Peck, R.B. (1948). Soil Mechanics in Engineering. Practice. John Wiley and

Sons, Inc., New York.

Terzaghi, K., Peck, R.B. and Mesri, G. (1996). Soil Mechanics in Engineering Practice, 3rd

Edn. John Wiley and Sons, Inc., New York.

Tomlinson, M.J. (1995). Foundation Design and Construction, 6th Edn. English Language

Book Society, London.

Trada (1990). Timber in Excavations. Thomas Telford Ltd, London.

Trautmann, C.H. and Kulhawy, F.H. (1987). CUFAD-A Computer Program for Compression

and Uplift Foundation Analysis and Design. Eletric Power Research Institute, Report

no. EL-4540-CCM, Vol. 16.

Tsui, Y. and Clough, G.W. (1974). Plane Strain Approximations in Finite Element Analyses of

Temporary Walls. Proceedings of Conference on Analysis and Design in Geotechnical

Engineering, University of Texas, Austin, 1, pp173.

157
REFERENCES

Vermeer, P.A., Punlor, A. and Ruse, N. (2001). Arching Effects behind a Soldier Pile Wall.

Computers and Geotechnics, Vol. 28, pp. 379-396.

Viggiani, G. and Atkinson, J.H. (1995). Stiffness of Fine-grained Soil at Very Small Strain.

Geotechnique, Vol. 45, No. 2, pp. 249-265.

Whittle, A.J., Hashash, Y.M.A. and Whitman, R.V. (1993). Analysis of Deep Excavation in

Boston. Journal of Geotechnical Engineering, Div., ASCE, Vol. 119, pp. 69-90.

Wong, I.H., Poh T.Y. and Chuah, H.L. (1997). Performance of Excavation for Depressed

Expressway in Singapore. Journal of Geotechnical & Geoenvironmental Engineering,

Div., ASCE, Vol. 123, No. 7, pp. 617-625.

Wroth, C.P. (1975). In-situ Measurement of Initial Stresses and Deformation Characteristics.

Proceedings of ASCE Specialty Conference on In-situ Measurement of Soil Properties,

Vol. 2, Raleigh, pp. 180-230.

Wroth, C.P., Randolph, M.F., Houlsby, G.T. and Faehy, M. (1979). A Review of the

Engineering Properties of Soils with Particular Reference to the Shear Modulus.

OUEL Report No. 1523/84. University of Oxford.

Yong, K.Y., Lee, F.H., Parnploy, U. and Lee, S.L. (1989). Elasto-Plastic Consolidation

Analysis for Strutted Excavation in Clay. Computer and Geotechnics, Vol. 8, pp. 311-

328.

Yoo, C.S. and Kim, Y.J. (1999). Measured Behaviour of In-situ Walls in Korea. Proceedings

of the 5th International symposium on Field Measurements in Geomechanics,

Singapore, pp. 211-216.

Zienkiewicz, O.C. and Taylor, R.L. (1988). The Finite Element Method, 4th Edn, Vol. 1.

McGraw-Hill, London.

158
APPENDICES

APPENDICES

Appendix A Transformation Matrix for 3D Beam

The stiffness matrix of a three dimensional beam element with reference to the

member axes between the end nodes i and j has been derived in the earlier section.

The global axes are brought to coinside with the local member axes by sequence of

rotation about y, z and x axes respectively. This is referred to as y-z-x transformation.

Figure A-1 shows the three rotations of the global axes. The angle of rotation are

denoted as β, γ and α respectively.

These series of rotations about the y-, z- and x-axes, can be resolved into their

respective vector components through the following transformations,

[T '] = [Tα ][Tγ ][Tβ ] (A.1)

where

 cos β 0 sin β 
[ ] Tβ =  0 1 0 
− sin β 0 cos β 

 cos γ sin γ 0
[ ] Tγ = − sin γ cos γ 0
 0 0 1

1 0 0 

[Tα ] = 0 cosα sin α 
0 − sin α cos α 

thus,

159
APPENDICES

 (cos β cos γ ) (sin γ ) (cos γ sin β ) 



[T '] = − (sin β sin α + cos β sin γ cosα ) (cos γ cosα ) (sin α cos β − sin β sin γ cos α )
 − (sin β cos α − cos β sin γ sin α ) − (cos γ sin α ) (cos β cos α − sin β sin γ sin α )
(A.2)

where

β is the rotation of global y-axis

γ is the rotation of global z-axis

α is the rotation of global x-axis (angle of tilt)

The vector components can be expressed in terms of the direction cosines of the

member, we get:

 
 Cx Cy Cz 
 
 
− C C cos α − C z sin α − C y C z cos α + C x sin α  (A.3)
[T '] =  x y 2 2
2
C x + C z cosα
2
2 2 
Cx + Cz Cx + C z
 
 C x C y sin α − C z cosα 2 2 C y C z sin α + C x cos α 
 − C x + C z sin α 
2 2 2 2
 Cx + Cz Cx + C z 

where

x j − xi y j − yi z j − zi
Cx = , Cy = , Cz =
L L L

and

L = ( x j − xi ) 2 + ( y j − yi ) 2 + ( z j − zi ) 2

160
APPENDICES

When the element lies vertical, i.e., when local x-axis coincides with the global

y-axis, Cx = Cz = 0. This makes some of the terms in the transformation matrix

indeterminate. Thus the above procedure breaks down.

In a general case of a vertical member, β = 0° and γ = 90° or 270°. The rotation

matrix for either case can be written as:

 0 Cy 0 
 
[T 'vert ] = − C y cosα 0 sin α  (A.4)
 C y sin α 0 cosα 

As there are eighteen degrees of freedom for a three-dimensional beam element

the rotation transformation matrix can now be expressed as

[T '] [0] [0] [0] [0] [0] 


 [0] [T '] [0] [0] [0] [0] 

 [0] [0] [T '] [0] [0] [0] 
[T ] =  (A.5)
 [0] [0] [0] [T '] [0] [0] 
 [0] [0] [0] [0] [T '] [0] 
 
 [0] [0] [0] [0] [0] [T ']

161
APPENDICES

ym

α
xm

LCy
γ x
β

2 2
LCz
L Cx + Cz

LCx

z
zm

Figure A-1: Rotation transformation of axes for a 3D beam element

162
APPENDICES

Appendix B Evaluating Bending Moments from Linear-strain Brick

Elements

For quadrilateral models, the bending moments were evaluated using the

method suggested by Rahim and Gunn (1997). For a plane strain quadrilateral element

representing part of a structure, the bending moment in x-y plane in Figure B-1 is

found from the integral:

h2

M = ∫σ
−h 2
x y.dy (B.1)

where h is the depth of the beam and σx is the stress normal to the cross section.

η
+1
I
M
h II -1 +1
ξ

III
y -1
x
Global coordinate system Local coordinate system
Figure B-1: A full-integration 8-noded quadrilateral element forming part of a beam.

For reason associated with numerical integration, it is convenient to work with a

‘normalised’ local coordinate system shown in Figure B-2 based on the following

relationship:

dy h
= (B.2)
dη 2

This leads to

163
APPENDICES

h
y= η (B.3)
2

Using the local coordinates, we obtain

+1
 h2 
M = ∫ σ x  η.dη (B.4)
−1  4 

The above integral can be evaluated using a three-point, one-dimensional Gauss

integration rule. Thus:

 h 2  III
M =  ∑ σ xi × ηi × ϖ i (B.5)
 4  i=I

Similarly for reduced-integration 8-noded quadrilateral element,

 h 2  II
M =  ∑ σ xi × ηi × ϖ i (B.6)
 4  i=I

In the case of 3D, which was considered in analyses of Chapter 3, we obtain the

following integral:

hz 2 h y 2

M= ∫ ∫σ
− hz 2 − h y 2
x y.dydz (B.7)

η
hz +1

I -1

Mzz +1
ξ
hy II
3 ζ -1
y III 2
1
x
z
Global coordinate system Local coordinates system

Figure B-2: A full-integration 20-noded linear strain brick element forming part of a beam.

164
APPENDICES

In local coordinates as defined in Figure B-2, the integral is

 hy 2  hz 
+1 +1
M = ∫ ∫σ x   η.dηdζ (B.8)
 4  2 
−1 −1  

Thus for full-integration 20-noded linear strain brick,

 h y 2 hz  3 III
= 
∑∑
M zz σ xik × η i × ϖ i × ϖ k (B.9)
 8
  k =1 i = I

and reduced-integration 20-noded linear strain brick

 h y 2 hz  2 II
= 
∑∑
M zz σ xik × η i × ϖ i × ϖ k (B.10)
 8
  k =1 i = I

where

h is the dimension of the element in x-, y- or z- direction

η is the normalised local coordinate

ω is the weight at respective local coordinates for the element

165
APPENDICES

Appendix C Stress Reversal in Non-linear Soil Model

200
Dasari (1996)
Stallebrass (1990)
Deviator stress, q (kPa)

150 FEM modelling Dasari (1996)


FEM modelling Stallebrass (1990)

100

50

0
0.000 0.002 0.004 0.006 0.008 0.010
Deviator strain, εs

Figure C-1: Comparison of FEM modelling of stress reversal with laboratory test results done
by Stallebrass (1990) and Dasari (1996).

The stiffness of soil on unloading is not the same as the stiffness in loading at

the same strain level. During unloading, the rate of decay of soil stiffness is slower.

This effect can be seen in a strutted excavation where the soil is first excavated and the

subsequent strutting will cause some stress reversal. Further excavation will again

trigger the stress reversal process. This behaviour can be approximately modelled by

Massing’s rule (1926):

q − qur  ε − ε ur 
= F s  (C.1)
2  2 

From Equation 5.11, assuming S=3G and G∞=0,

 ε − ε ur 
3G0  s 
q − qur  2 
= (C.2)
2 3G  ε − ε ur 
1+ 0  s
q f  2 

166
APPENDICES

 3  3G  ε − ε     ε − ε ur  3G0  1  
 G0 1 +   − 3G0  s
0
 s ur
  2  
1 dq  2  q f  2     2  qf 
=
2 dε s 3G  ε − ε ur 
1+ 0  s
q f  2 

3
G0
= 2 (C.3)
2
 3G0  ε s − ε ur 
1 +  
 q f  2 

Simplify,

dq 3G0
= (C.4)
dε s  3G  ε − ε 
2

1 +
0
 s ur

 qf  2 

Variation of G can be rewritten as:

G0
G= 2
(C.5)
 3G0  ε s − ε ur 
1 +  
 qf  2  

Tests from Stallebrass (1990) and Dasari (1996) are used to verify the rule as

shown in Figure C-1. Figure C-1 shows good approximation of Massing’s rule with

stress-strain behaviour of real soil in laboratory tests.

167
APPENDICES

Appendix D Properties of Timber Lagging (Chudnoff, M., 1984)

STANDARD NAME:

KEMPAS (Medium Hardwood)

BOTANICAL NAME:

Koompassia malaccensis

FAMILY

Leguminosae

DISTRIBUTION

Very abundant in all lowland forests and in the mountains, often growing on moist,

peaty or fresh water swamp soils and on low ridges.

GENERAL DESCRIPTION

Sapwood is well-defined and yellow in colour. Heartwood is pinkish when fresh and

darkening to a bright orange-red or deep brown. Grain is interlocked, often very

interlocked. Texture is coarse and even. Vessels are large and with simple

perforations, few or very few, mostly solitary, but also in radial groups of 2 to 6 and

evenly distributed. Wood parenchyma is abundant and predominantly paratracheal, as

conspicuous aliform type; apotracheal type may occur as narrow terminal bands. Rays

are moderately fine, just visible to the naked eye on the cross- section. Ripple marks

are generally distinct. Concentric bands of included phloem are often present.

PHYSICAL PROPERTIES

Air-Dry Density: 770.00 - 1120.00 kg/m3

Shrinkage Radial: 2.00 %

Shrinkage Tangential: 3.00 %

Seasoning: Seasons fairly slowly with little or no defects

Recommended Kiln Schedule: E

MECHANICAL PROPERTIES

Strength Group: A
168
APPENDICES

Static Bending MOE: 18600.00 N/mm2

Static Bending MOR: 122.00 N/mm2

Compression perpendicular to grain: 7.52 N/mm2

Compression parallel to grain: 65.60 N/mm2

Shear Strength: 12.40 N/mm2

DURABILITY

Moderately durable

TREATABILITY

Easy

WORKING PROPERTIES

Planing: Easy

Planing Finish: Smooth

Boring: Slightly difficult

Boring Finish: Rough

Turning: Slightly difficult

Turning Finish: Rough

Nailing: Poor

USES

Heavy construction, railway sleepers, posts and cross arms (telegraphic and power

transmission), beams, joists, rafters, piling, columns (heavy duty), fender supports,

pallets (permanent, heavy duty), doors, window frames and sills, tool handles (heavy

duty) and marine construction. Untreated the timber is suitable for parquet and strip

flooring, flooring (medium to heavy traffic), panelling and vehicle bodies (framework

and floor boards)

169

You might also like