You are on page 1of 254

EFFECT OF A CORNER IN A

THREE-DIMENSIONAL EXCAVATION

LOH CHANG KAAN


(B.ENG. (HONS.) UTM, M.ENG. (NUS))

A THESIS SUBMITTED

FOR THE DEGREE OF DOCTOR OF PHILOSOPHY

DEPARTMENT OF CIVIL ENGINEERING

NATIONAL UNIVERSITY OF SINGAPORE

2003
Dedicated to,

my mother, wife, and two children for their

……….. Continual support, encouragement, and

..........................Endless understanding
Acknowledgements

ACKNOWLEDGEMENTS

Firstly, I would like to extend my deepest gratitude to my first supervisor, Associate

Professor T. S. Tan for his concern and kindness, his continual effort on guiding,

checking and providing ideas throughout my research work. Secondly, I would like to

thank my second supervisor, Associate Professor F.H. Lee, who has provided me with

some most critical ideas and approaches on handling daunting problems that emerged

along the study. In addition, his encouragement is deeply appreciated. Both of my

supervisors’ effort on initiating this research project and shaping up the framework of

this study is acknowledged.

One of the biggest problems on pursuing three-dimensional (3D) study is the handling

of tremendous volume of data generated, compared to 2D study. It is very easy to be

drawn into a whirlpool of data, lost the direction of study, and become exhausted along

the way. It is my friend and colleague, Dr. J. Wang that helps me validate and

prioritize the data, and assists to form a clearer focus of study. His support is

invaluable for the completion of this thesis. Gratitude also extends to my senior and

friend, Dr. T.G. Ng, for his critical proofreading of the thesis, often late into the night.

I would also like to thanks all the laboratory technologists in the Geotechnical

Laboratory, especially Mr. C.Y. Wong, for their help in the experimental works.

Acknowledgement also extends to all my fellow research engineers and scholars for

sharing their experience and stimulating discussion.

i
Table of Contents

TABLE OF CONTENTS

Page

Title Page
Acknowledgements i
Table Of Contents ii
List Of Figures vi
List Of Tables xv
Summary xvi

Nomenclature xix

CHAPTER 1 INTRODUCTION 1

1.1 Three Dimensional Behaviour Of Deep Excavation 1


1.2 Current Understanding On 3D Behaviour Of Deep Excavation 2
1.3 Approach To A Better Understanding On Corner Effect In Deep 2
Excavation
1.4 Necessities Of Centrifuge Model Test 4
1.5 Objectives Of The Study 5

CHAPTER 2 LITERATURE REVIEW 8

2.1 Introduction 8
2.2 Literature Review On Centrifuge Modelling On Deep Excavation 11
2.2.1 Various Methods To Model Excavation In Centrifuge 12
2.2.1.1 Increasing-g Method 12
2.1.1.2 Excavate and Spin Method 13

2.2.1.3 Heavy Liquid Method 13


2.2.1.4 2D In-Flight Excavation Test In The Centrifuge 15
2.2.1.5 Tunnel Excavation Method 17
2.2.2 The Challenge Of 3D In-Flight Excavation Test 18
2.3 Literature Review On 3D Excavation Behaviour 19
2.3.1 General Appreciation Of 3D Excavation Behaviour 19

ii
Table of Contents

2.3.2 Quantification Of 3D Corner Effect 22

2.3.3 Influence Range Of 3D Corner Effect 25

2.3.4 Other Parameters Which Might Influence Corner Effect 29

2.4 Summary 30

CHAPTER 3 DEVELOPMENT OF 3D IN-FLIGHT EXCAVATOR AND 32

CENTRIFUGE MODEL TESTS

3.1 Design And Development Of 3D In-Flight Excavator 32


3.2 Experimental Details 35
3.2.1 The Model Retaining Wall 35
3.2.2 Soil Sample Preparation 37
3.2.3 Centrifuge Model Container Preparation 40
3.2.4 Instrumentation And Excavator Set-Up 41
3.2.5 Excavation Procedures 43
3.3 The Conditions Of In-Flight Excavation Tests 44
3.3.1 Area Modelled In The Experiments 44
3.3.2 Fixities And Water Table Boundary Conditions 45
3.4 Methodology Of Test 46
3.4.1 Parametric Study On The Effect Of Wall Stiffness 46
3.4.2 Parametric Study On The Effect Of Capping Beam 46
3.4.3 Parametric Study On Influence Of Geotechnical Effect To 47
3D Corner Effect

CHAPTER 4 EXPERIMENTAL RESULTS AND DISCUSSIONS 48

4.1 Introduction 48
4.1.1 Initial Conditions Of Tests 48
4.1.2 Soil Characterization 49
4.1.3 Prototype Modelled and Other Consideration 51
4.2 General Characteristics Of 3D Excavation Behaviour 52
4.2.1 Characteristics Of Surface Settlement And Wall Deflection 53

iii
Table of Contents

4.2.2 Characteristics Of Lateral Earth Pressure On Retaining Wall 59


4.3 2D Versus 3D Excavation Behaviour 62
4.4 Retaining Wall Thickness Effect In A Corner Excavation 67
4.5 Presence Of A Capping Beam In A 3D Excavation 70
4.6 Effect Of Soil Strength In 3D Excavation 72
4.7 Summary 74

CHAPTER 5 FUNDAMENTAL BEHAVIOUR OF CORNER EFFECT 76

IN EXCAVATION

5.1 Introduction 76

5.2 Finite Element Analysis 77


5.2.1 Parametric Studies By Varying Excavation Dimensions Of A 80
Corner Excavation
5.2.2 Parametric Studies On The Effect Of Wall Stiffness And The 84
Presence Of Capping Beam To 3D Corner Effect
5.2.3 Summary Of Findings From FEM Studies 85
5.3 Characterisation Of Corner Effects 86
5.3.1 Major Factors Affecting The Corner Effect 86
5.3.2 Characterization Of Structural Factor 88
5.3.2.1 Deformation Due To Excavation Induced Imbalance 88
Load
5.3.2.2 The Effect Of Structural Restraint At Corner 92
5.3.2.3 Corner Effect Due To Structural Restrain At Corner 94
5.3.2.4 The Corner Effect Influence Range 96
5.3.3 Characterization Of Geotechnical Factors 101
5.4 Combined Characterization Of Corner Effect 106
5.4.1 Evaluation Of Corner Effect Hypotheses 107

CHAPTER 6 CONCLUSIONS 113

6.1 Development Of 3D In-Flight Excavator 113


6.2 Summary Of Findings 114

iv
Table of Contents

6.3 Effect Of A Corner In A 3D Excavation 115


6.4 Recommendation For Future Studies 117

TABLES 120

FIGURES 127

REFERENCES 222

v
List of Figures

LIST OF FIGURES

Figure No. Description Page


Fig. 1.1 A schematic diagram showing a typical excavation 127
carried out in the field

Fig. 2.1 Observed settlements behind strutted excavation in 128


Chicago (after O’ Rourke et al., 1976)

Fig. 2.2 Observed settlements behind excavation (after Peck, 128


1969)

Fig. 2.3 Relationship between factor of safety against basal 129


heave and maximum lateral wall movement from case
histories (after Clough et al., 1979)

Fig. 2.4 Relationship between maximum ground settlements 129


and maximum lateral wall movement from case
histories (after Mana & Clough, 1981)

Fig. 2.5 Apparent pressure diagrams for computing strut loads 130
in braced cuts (after Terzaghi et al. 1996)

Fig. 2.6 Distress caused to a buried service by a shallow 131


trenching operation (after Needham and Howe, 1984)

Fig. 2.7 Schematic representation of the “Increasing-g” method 132


to simulate excavation

Fig. 2.8 Schematic representation of the “Heavy Liquid” 132


method to simulate excavation

Fig. 2.9 TIT’s in-flight excavator setup (after Kimura et al. 133
1994)

Fig. 2.10 a) Vertical cut with corner angle ∝ (after Giger 134
and Krizek, 1975)
b) Stability factor Ns as a function of the corner
angle ∝ (after Giger and Krizek, 1975)

Fig. 2.11 Stability Number versus Depth of Excavation Divided 135


by Radius (after Britto and Kusakabe, 1983)

Fig. 2.12 2D Section Used in Trench Excavation Analysis (after 136


De Moor, 1994)

Fig. 2.13 PSR Chart (After Ou et al. 1996) 137

Fig. 2.14 Typical Analysis on a 10m long x 6m deep x 1m wide 138


trench excavation. Ground movement along line e – e

vi
List of Figures

parallel to trench and located 2m below surface:


Comparison between results from three-dimensional
and plane strain analyses (after Nath, 1983)

Fig. 2.15 Variation of maximum wall displacement with the 139


distance from for constant sizes of complementary
wall and various sizes of primary wall, L = Length of
primary wall; B = length of complementary wall.
After Ou et al. (1996)

Fig. 3.1 3D in-flight excavator set-up on centrifuge 140

Fig. 3.2 Schematic diagrams of the 3D in-flight excavator 141

Fig. 3.3 Schematic representation of 3D in-flight excavator set- 142


up on the centrifuge during testing

Fig. 3.4 Indexers/drivers mounted on-board the centrifuge 143

Fig. 3.5 Internal parts of indexers/drivers strengthened by 143


silicone sealant. The indexers/drivers was mounted
near to the centrifuge rotating shaft to minimized
centrifugal force during spinning.

Fig. 3.6 Schematic drawing of a corner of an excavation 144


simulated

Fig. 3.7 2D in-flight excavation test set-up on a narrow 145


centrifuge container

Fig. 3.8 Micro-concrete wall used in the study 146

Fig. 3.9 One of the aluminum alloy wall used in the study. 146

Fig. 3.10 Plan view of the excavation set-up & LVDT set-out for 147
Test 3DK-2

Fig. 3.11 Locations of SGs and TSTs in the experiment: Test 147
3DK-2c

Fig. 3.12 Schematic representation of OCR profiles intended 148

Fig. 3.13 De-airing of pore pressure transducers by boiling 149

Fig. 3.14 Installation of pore pressure transducers into soil 149


sample

Fig. 3.15 a) Completed 3D in-flight excavation test set-up 150


before transporting to centrifuge room
b) 3D in-flight excavation test set-up after

vii
List of Figures

excavation test on centrifuge platform

Fig. 3.16 Excavation in Progress (captured by miniature camera 151


mounted in front of the test sample)

Fig. 4.1 Surface settlement versus elapsed reconsolidation time 152


(Test 3DK-3)

Fig. 4.2 Hyperbolic plot, Elapsed time/Settlement versus 152


Elapsed Time (Test 3DK-3)

Fig. 4.3 Density and undrained shear strength profiles of NC 153


soil used in the experiments

Fig. 4.4 Schematic plan view of model retaining wall edge at 154
container wall face

Fig. 4.5 Responses of LVDTs installed at difference planes 155


(Test 3DK-2)

Fig. 4.6 Flow chart showing centrifuge tests conducted 156

Fig. 4.7 Locations of LVDTs, SG and TST in the experiment: 157


Test 3DK-2c

Fig. 4.8 Test: 3DK-2c: Surface settlement at various location 158


behind the retaining wall

Fig. 4.9 Surface settlement profiles behind wall, at various 159


section from corner

Fig. 4.10 Surface settlement behind wall: Test 3DK-2c compare 160
2D tests and published data (after Peck 1969)
S = Surface settlement
h = Depth of excavation
D = Distance behind wall

Fig. 4.11 Surface settlement profiles at various distances from 161


corner: at 3.0m and 7.0m behind wall (Test 3DK-2c)

Fig. 4.12 Surface settlement contour behind retaining wall (Test 162
3DK-2c)

Fig. 4.13 Wall deflection profiles at Perspex window (Test 163


3DK-2c)

Fig. 4.14 A typical scraping of 1 layer of soil in 3D in-flight 163


excavation test

Fig. 4.15 Incremental surface settlement of a single scrapping 164

viii
List of Figures

operation. LVDTs at 7m behind wall, at various


distances from corner (x)

Fig. 4.16 Normalized incremental surface settlement at 7m 165


behind wall of a single scraping operation: At various
distances from corner at different excavation stages
(Test 3DK-2c)

Fig. 4.17 The final settlement at each stage of scrapping. 166


x = distance from corner

Fig. 4.18 Lateral earth pressures versus depth of excavation. 167


Total stress transducers (TSTs ) at passive and active
sides, at 8m and 15m from corner

Fig. 4.19 a) Coefficient of total stress lateral earth pressure 168


P
K=
coefficient, γy , versus depth of excavation. P =
measured lateral earth pressure, γ = density of soil (16
kN/m3), y = depth of TST from soil level

P − γy
b) versus depth of excavation. P = measured
2C u
lateral earth pressure, γ = density of soil (16 kN/m3), y
= depth of TST from soil level, Cu = undrained shear
strength at the TST level

Fig. 4.20 Surface settlement profiles behind wall. Test 3DK-2c 169
compare 2DK-1 and 2DK-3

Fig. 4.21 Normalized wall deflection profiles. 3D test compare 170


2D tests

Fig. 4.22 Incremental surface settlement at 7m behind wall: 2D 171


(Test 2DK-3) versus 3D (Tests 3DK-2 and 3DK-2c)
NF2D = Normalization factor for 2D test
NF3D = Normalization factor for 3D test

Fig. 4.23 Incremental wall top displacement: 2D Test (Test 172


2DK-3) versus 3D Test (Tests 3DK-1, 2 and 3), x =
9.1m
NF2D = Normalization factor for 2D test
NF3D = Normalization factor for 3D test

Fig. 4.24 PSR-sett at 7m behind wall versus distance from 173


corner (x) Test 3DK-2c

Fig. 4.25 Surface settlement at 7m behind wall versus depth of 174

ix
List of Figures

excavation: 3D tests with various wall thickness

Fig. 4.26 Lateral wall top displacement (δ) versus depth of 174
excavation: 3D tests with various wall thickness

Fig. 4.27 Lateral displacement at wall top at various distances 175


from corner: 3D tests with various thickness

Fig. 4.28 Surface settlement profiles at 7m behind wall at 175


various distances from corner: 3D tests with various
thickness

Fig. 4.29 Surface settlement contour behind retaining wall after 176
7m-excavation: 3D tests with various wall thickness

Fig. 4.30 Illustration of offsetting surface settlements with 177


settlement at the corner

Fig. 4.31 Sketch illustration of capping beam restrained at 178


corner

Fig. 4.32 Surface settlement: Test 3DK-2c compare Test 3DK-2 179

Fig. 4.33 Wall deflection profiles at the edge of retaining wall 180
(17m from corner): Test 3DK-2c compare 3DK-2

Fig. 4.34 Lateral displacement at wall top at various distances 181


from corner: 3D tests with various thickness

Fig. 4.35 a) Lateral wall top displacement versus distance from 182
corner. Soft NC compares Stiff OC soils

b) Lateral wall top displacement normalized by


dividing with displacement at x = 15.2m (furthest from
corner), at the particular stage of excavation. Soft NC
compares Stiff OC soils

Fig. 4.36 a) Offset-settlement versus distance from corner. 183


Soft NC compares Stiff OC soils
b) Offset-settlement normalized by dividing with
displacement at x = 14.5m (furthest from corner), at
the particular stage of excavation. Soft NC compares
Stiff OC soils

Fig. 5.1 Typical 1000 elements FEM mesh used in the study 184

Fig. 5.2 Wall deflection profiles at the edge (on Perspex 185
window in the experiment)

x
List of Figures

Fig. 5.3 Configuration of excavation case studied by Ou et al. 186


(1996)

Fig. 5.4 Schematic illustration of size of excavation studied 186

Fig. 5.5 Variation of maximum wall displacement with the 187


distance from for constant sizes of complementary
wall and various sizes of primary wall, L = Length of
primary wall; B = length of complementary wall.
After Ou et al. (1996)

Fig. 5.6 Wall top displacement at various distances from 188


corner, FEM analyses with primary wall length (PL) =
17, 25, 65 and 130m

Fig. 5.7 PSR versus distance from corner plot: FEM analyses 189
results. PL = 65 and 130m

Fig. 5.8 Variation of PSR against distance from corner, x 190


Wall with capping beam, with different wall stiffness

Fig. 5.9 Variation of PSR against distance from corner, x 191


Wall without capping beam, with different wall
stiffness

Fig. 5.10 PSR at a section of x = 30m against the depth of 192


excavation. For tests with and without capping beam

Fig. 5.11 Illustration of a simplified corner excavation 193

Fig. 5.12 Schematic representation of the boundary and loadings 194


conditions.

Fig. 5.13 Active and passive earth pressure distribution on 195


cantilever retaining wall

Fig. 5.14 Illustration of corner restrains effect 196

Fig. 5.15 Normalized wall top displacement versus x profiles: 197


FEM 3D excavation modelling compare beam theories

Fig. 5.16 Illustration and derivation of end restrain effect with a 198
Pin-end condition

Fig. 5.17 Illustration and derivation of end restrain effect with a 198
Pin-Moment end condition

Fig. 5.18 Wall top displacement versus distance from corner (x): 199
Test 3DK-1

xi
List of Figures

Fig. 5.19 Wall top displacement versus distance from corner (x): 200
Test 3DK-2

Fig. 5.20 Wall top displacement versus distance from corner (x): 201
Test 3DK-3

Fig. 5.21 Offset-surface settlement versus distance from corner 202


(x): Test 3DK-1

Fig. 5.22 Offset-surface settlement versus distance from corner 203


(x): Test 3DK-2

Fig. 5.23 Offset-surface settlement versus distance from corner 204


(x): Test 3DK-3

Fig. 5.24 Normalized wall top displacement versus x profiles: 205


FEM 3D excavation modelling compare beam theories

Fig. 5.25 a) Lateral bending moment versus distance from 206


corner (x): Test 3DK-2c

b) Theoretical lateral bending moment versus distance


profile

Fig. 5.26 a) Measured Lm for Test 3DK-1 and hypothesis of Lm 207


for tests with no yielding (Tests 3DK-2 and 3DK-3)

b) Hypothesis on the relationship between Lm and wall


area:
Normalized Lm versus Normalized EA, where
E = Elastic modulus of wall, same for all the test (E1 =
E2 = E3)
A = Cross sectional area of the wall, which is total
depth of wall * thickness of wall. Hence:
A3 = 3 * A1 (A of Test 3DK-3 = 3 times of A of Test
3DK-1)
A2 = 2 * A1 (A of Test 3DK-2 = 2 times of A of Test
3DK-1)
Lm1 = Distance from corner when wall of Test 3DK-1
yielded = 10m

Fig. 5.27 Illustration showing the hypothesis of normalization of 208


corner effects to achieved δy

Fig. 5.28 Total lateral earth pressure coefficient at 8m and 15m 209
from corner, at active and passive sides, versus depth
of excavation

Fig. 5.29 Schematic diagram illustrating the variation of earth 210


pressure effect at active side

xii
List of Figures

Fig. 5.30 Lateral earth pressure at the retained soil side versus 211
depth of excavation

Fig. 5.31 Pa ( 3 D ) 212


γh
Pa ( 2 D )
versus Cu for TSTs at 8m and 15m from
corner for Test 3DK-2c

Fig. 5.32 Pa ( 3 D ) 213


γh
Pa ( 2 D )
versus Cu for Test 3DK-2c (Experiment and
FEM)

Fig. 5.33 Relationship between R and distance from corner 213

Fig. 5.34 Wall top displacement versus distance from corner: 214
Test 3DK-1
a) δ before normalization
b) δ normalized with λs
c) δ normalized with λa and λs

Fig. 5.35 Wall top displacement versus distance from corner: 215
Test 3DK-2
a) δ before normalization
b) δ normalized with λs
c) δ normalized with λa and λs

Fig. 5.36 Wall top displacement versus distance from corner: 216
Test 3DK-3
a) δ before normalization
b) δ normalized with λs
c) δ normalized with λa and λs

Fig. 5.37 Offset-surface settlement at 7m behind wall versus x: 217


Test 3DK-1
a) Offset-Sett before normalization
b) Offset-Sett normalized with λs
c) Offset-Sett normalized with λa and λs

Fig. 5.38 Offset-surface settlement at 7m behind wall versus x: 218


Test 3DK-2
a) Offset-Sett before normalization
b) Offset-Sett normalized with λs
c) Offset-Sett normalized with λa and λs

Fig. 5.39 Offset-surface settlement at 7m behind wall versus x: 219


Test 3DK-3
a) Offset-Sett before normalization

xiii
List of Figures

b) Offset-Sett normalized with λs


c) Offset-Sett normalized with λa and λs

Fig. 5.40 Wall top displacement versus depth of excavation: 220


Tests 3DK-1, 3DK-2 and 3DK-3
a) δ before normalization
b) δ normalized with λs
c) δ normalized with λa and λs

Fig. 5.41 Offset-settlement versus depth of excavation: Tests 221


3DK-1, 3DK-2 and 3DK-3
a) Offset-Sett before normalization
b) Offset-Sett normalized with λs
c) Offset-Sett normalized with λa and λs

xiv
List of Tables

List of Tables

Table No. Description Page No.

Table 2.1 Summary of corner effect influence range 120


reported/proposed

Table 3.1 Speed and stroke of NUS’s 3D in-flight excavator 121

Table 3.2 Properties of Malaysian Kaolin clay 122

Table 4.1 Summary of in-flight excavation tests carried out 123

Table 5.1 Summary of soil profile and soil parameters used for 124
FEM analyses

Table 5.2 Initial stress conditions of FEM analyses 125

Table 5.3 Summary of findings from centrifuge modelling and 126


FEM analyses

xv
Summary

SUMMARY

In reality, all excavations are three-dimensional (3D) in nature. In routine engineering

practices, the complicated 3D problem is often simplified and idealised into much

simpler plane strain two-dimensional (2D) problem. However, it is generally known

that such 2D analyses usually produced conservative design especially for small

basement excavation where the corner-to-corner distance is relatively short. If this

corner effect is considered in the analysis, a more accurate and economical design often

can be achieved. In the current study, the main objective was to gain a better

understanding of how a corner affects an excavation.

The current study focuses on the basic aspects of 3D corner effect, with the simplest

configuration possible. A 3D in-flight excavator for the National University of

Singapore centrifuge was developed specifically for the present study to provide

physical data for in depth mechanistic study. An un-braced excavation around a right

angle corner was modeled using a simple soil profile in the centrifuge. This model is

able to capture the salient features of 3D corner effects. From the study, it was found

that the presence of a capping beam, which usually ignored by practicing engineers in

analysis, would enhance the corner effect. The effect was more pronounced in the

early stage of excavation. Comparative study on excavation tests with stiff highly

overconsolidated and soft normally consolidated samples found that the presence of a

corner would affect the wall deflection and surface settlement more in soft soil than that

in stiff soil. However, this is in terms of absolute magnitude. In terms of relative

ratio of wall displacement or surface settlement with distance from corner as

mechanistic study is concerned, there is no apparent correlation with the soil strength. It

xvi
Summary

was also found that the behaviour of corner effect is insensitive to the depth of

excavation.

From the study, the 3D corner effect measured based on the wall displacement was

expressed as δ 3 D = δ 2 D * λs * λa .

Where δ3D = Wall top displacement within the 3D corner effect influence zone

δ2D = Wall top displacement in 2D condition

λs = Corner effect factor due to corner structural restrain

λa = Corner effect factor due to geotechnical effect

λs is derived from a beam theory. It shows to some extent how the effect of a corner is

propagated. λa is due to the variation of earth pressure along the distance from corner,

γh
which is dependent on stability number, .
Cu

Where γ = total unit weight of the soil

h = excavation depth

Cu = undrained shear strength of the soil

It was also found in the study that if the flexural capacity of the retaining wall is large

enough, the influence range of the corner is given by the point where the plane strain

condition starts. If not, then it is decided by the point when the flexural capacity of the

wall was exceeded. Thus La= min .{Lm , L2 D }.

Where Lm = Length from corner when lateral flexural capacity is exceeded.

L2D = Length from corner when plane strain condition is achieved.

If the lateral flexural capacity is exceeded in the retaining wall, the 3D corner effect

may be expressed as δ 3 D = δ y * λ s * λ a , where δy is the wall displacement when the

xvii
Summary

wall lateral flexural capacity is exceeded. This expression was examined using

centrifuge experimental data. The results show that the equation is able to estimate the

corner effect influence to the displacement at the top of wall, as well as surface

settlement at 7m behind the wall. This shows that the λs and λa are able to explain

major aspects of 3D corner effects in the excavations.

xviii
Nomenclature

NOMENCLATURE

A cross sectional area of the retaining wall

B the width of excavation

D distance behind retaining wall

E the Young’s modulus of the wall

g gravitational force

I second moment of area

L the length of excavation

Lm length from corner when lateral flexural capacity is exceeded

Ll simplified corner effect influence range from corner

L2D length from corner when plane strain condition is first achieved

P measured lateral earth pressure

Pa earth pressure at the active side at the toe of the wall in Equation (5.3). Also

the measured lateral earth pressure at the active side.

Pp passive earth pressure at the passive side at the toe of the wall in Equation

(5.3). Also the measured lateral earth pressure at the passive side.

b Total depth of the wall from ground level to the toe of wall in Equation (5.3)

PL the length of primary wall in an excavation

PSR defined as the ratio of the maximum wall displacement (δ) of a section over

the maximum wall displacement (δ2D) of the section

y depth of TST from soil level

Cu the undrained shear strength of the soil

h excavation depth

hw the height of water table

xix
Nomenclature

Ko coefficient of lateral earth pressure at rest

Ko coefficient of lateral earth pressure at rest, total stress term (include water

pressure)

Ka coefficient of lateral earth pressure at active side, total stress term

(include water pressure)

Kp coefficient of lateral earth pressure at passive side, total stress term

(include water pressure)

K a -active coefficient of lateral earth pressure at active state, total stress term

(include water pressure)

qs the uniform surface loading on the retained soil behind the excavation

γh
S the stability number defined as S =
cu

x distance from corner

δ horizontal wall displacement

δ3D wall top displacement within the 3D corner effect influence zone

δ2D wall top displacement in 2D condition

δy wall top displacement when the wall lateral flexural capacity is exceeded

λs corner effect factor due to corner structural restrain

λa corner effect factor due to geotechnical effect

γ the total unit weight of soil

σa the Rankine active earth pressures

σp the Rankine passive earth pressures

xx
Chapter 1 Introduction

CHAPTER 1

INTRODUCTION

1.1 THREE DIMENSIONAL BEHAVIOUR OF DEEP EXCAVATION

Excavation of soil as part of a major infrastructure construction is a common activity in

most heavily built-up cities. Deep excavation is needed for basement, underground

road, sewerage pipe, drainage facilities and other constructions. A schematic diagram

showing a typical excavation carried out in the field is shown in Fig. 1.1. In reality,

excavations are three-dimensional (3D) in nature. At every plane with the distance

from the corner, due to the variation of geometry and influence from the corner, the

behaviour is different, and would interact with each other. This 3D behaviour is very

complicated. Hence, in most engineering practices, a much simpler plane strain two-

dimensional (2D) problem is often used to approximate the 3D problem.

The 2D assumption actually assumed that there are no interactions between adjacent

planes. This assumption is more appropriate for a long trench excavation, such as

excavation for pipe laying and drainage works. In a cofferdam excavation, such as

basement excavation where excavation size is limited, there are inevitable interactions

between adjacent planes with distance from corner. For sections near the mid-span of a

large excavation, the behaviour may be approaching that of a 2D problem. However,

the influence range and under what conditions this is valid is not well established.

In illustration shown in Fig. 1.1, it is apparent that the movement of the retaining wall

and soil at the corner would be smaller due to the 3D effect. This corner restrain would

1
Chapter 1 Introduction

extend to a certain distance away from corner. Hence, if this corner effect is considered

in an analysis, a more accurate and economical design can be achieved.

1.2 CURRENT UNDERSTANDING ON 3D BEHAVIOUR OF DEEP

EXCAVATION

For an excavation supported by retaining wall, the unbalance load due to the removal of

earth would cause movement to occur. This movement is restrained by the retaining

wall and the associated bracing system. This is a classical soil-structure interaction

problem, and many 2D solutions are available. In engineering practice, it is intuitively

recognized that the presence of 3D effect, such as corner of a retaining wall, would

influence the soil-structural interaction predicted by the 2D analysis. For example, it is

generally accepted that the corner of a cofferdam is stiffer than other section far from

corner. It is also generally recognized that the movement of a small size excavation

would be smaller than that of larger excavation. In trenching works, engineers would

usually reduce the length of each panel of excavation to control ground movement.

However, to date, there is no specific reference on quantifying the corner effect, nor any

empirical method or design chart to relate the corner effect to a 2D problem. Hence, to

better make use of the 3D effect, which exists naturally in almost any excavation, to

achieve a more economical design, a better understanding on this effect is necessary.

1.3 APPROACH TO A BETTER UNDERSTANDING ON CORNER EFFECT IN

DEEP EXCAVATION

Currently, there are a number of numerical programs capable of performing 2D and 3D

excavation analyses, such as CRISP, FLAC, ABAQUS, PLAXIS and DIGDIRT.

2
Chapter 1 Introduction

These programs are able to model the soil-structural interaction in a realistic sequence

of operation that follows closely the actual geometry of excavation and construction

sequences. The proper modelling of a 2D problem alone is already not easy task. The

accurate modelling of the initial stress state, boundary conditions, soil behaviour

(model), excavation sequences and the interaction between soil and wall are very

complicated. The extension of such analyses to model a true 3D problem will create

many more uncertainties. Thus, while some studies on 3D analyses of excavation were

conducted using such numerical programs, only broad trends of 3D excavation

behaviour was presented. There is little study on the fundamental aspects of 3D corner

effect in an excavation. This is due mainly to the fact that the confidence level on such

numerical analysis does not justify in-depth mechanics study. The confidence level on

such numerical analyses can definitely be improved by proper calibration of such

analyses with good physical data.

Physical data is usually obtained from instrumented field projects. From the process of

back-analysis of field data using the program and subsequent fine-tuning, the accuracy

of the numerical model can be improved. Though field data represent the most

realistic situation, it itself is complicated and is usually not ideal for calibration. The

complexities of field data arise out of a number of situations which are difficult to

avoid. Major contributing factors are complexity of soil profiles, generally unknown

initial conditions, boundary condition and not well defined excavation procedures.

Hence, field data alone is also not satisfactory for in-depth mechanics study of the

corner effect.

3
Chapter 1 Introduction

Centrifuge is a well accepted tool in geotechnical engineering. Though it has certain

limitations, it also offers many advantages. Chief among these are ensuring correct

physical modelling principles, well defined initial and boundary conditions and well

controlled soil properties. These advantages are particularly useful when mechanism

is being examined. Hence, the centrifuge model test is an option to provide data for

fundamental mechanism study on the 3D effect.

1.4 NECCESITIES OF CENTRIFUGE MODEL TEST

Centrifuge model test have many advantages over field test as they can provide reliable

and repeatable data with well defined boundary condition and known material

properties, which are essential in any mechanism study. More useful is the fact that in

the centrifuge, appropriate parametric study can be conducted to ensure that influence of

specific factor of interest can be isolated to enable a better understanding of how their

effectiveness is being mobilized. Some important factors that can be studied through

parametric studies are:

i) 2D versus 3D effect in an excavation

Centrifuge model test can be conducted on 2D and 3D excavation models.

These tests allow the 3D effects to be highlighted.

ii) Effect of retaining wall stiffness

It is known that the stiffness of the retaining wall plays an important role in

the behaviour of an excavation. By varying the characteristics of the

retaining wall, such as the wall stiffness, such effects on the overall

behaviour of an excavation can be carefully evaluated.

iii) Effects of soil properties

4
Chapter 1 Introduction

The effects of soil properties in an excavation can also be assessed through

parametric studies. This includes varying the over-consolidation ratio,

shear strength and stiffness of the soils.

The above advantages of centrifuge model tests are well recognised, but to conduct a

realistic 3D excavation test in the centrifuge is very difficult. Due to the high

centrifugal force, the model set-up must be able to work under very harsh environment

with degree of complexity increase proportionally with the centrifugal force. As the

size of the centrifuge platform is generally not large, the space for mounting the soil

model and its modelling apparatus is limited. Thus, to avoid complex development on

the model set-up, two traditional methods were used to approximately model

excavations. They are:

1) the increasing-g method, and

2) the heavy liquid method.

Detail descriptions of these methods are presented in Chapter 2. The traditional

methods have limitation on modelling the correct soil stress and the development of

earth pressures, especially when soft normally consolidated clay profile is used. In

view of the above shortcoming, researchers in the Tokyo Institute of Technology

developed the in-flight excavator to model the excavation more accurately (Kimura et

al. 1993). However, the in-flight excavator was only designed to model plain strain

excavation problem.

1.5 OBJECTIVES OF THE STUDY

In view of the above, this study is carried out with the following objectives:

1) To develop and conduct 3D in-flight excavation tests in the centrifuge to

provide reliable physical data. The key challenge is to develop suitable

5
Chapter 1 Introduction

experimental apparatus to conduct centrifuge test with pre-defined initial

and boundary conditions and to model realistic excavation sequences. The

development of an in-flight excavation apparatus suitable for 3D excavation

is already a very complicated task. To put in place an in-flight strutting

system is more daunting and was not pursued in this project. Thus in this

study, unstrutted excavation was carried out. In all excavations, the first

stage is usually an unstrutted excavation. If the excavation is in soft soil,

this first stage of excavation is often a critical stage where significant

movement can take place if it is not handled properly. Thus the results

from this study, besides their intrinsic value to provide reliable data for

fundamental study, are also realistic for actual excavation, especially the

first stage.

In the field, most retaining walls, such as sheet pile, soldier pile, contiguous

bored pile or diaphragm walls are constructed individually or by panels,

where the lateral flexural stiffness is much smaller than the vertically

flexural stiffness. This phenomenon may reduce the influence range of

corner effect. However, the panels such as that for sheet piles and

diaphragm wall are usually connected by interlocking slot, which may

transfer some moment between panels. In addition, the panels are

generally bound together by a capping beam, thus further increase its lateral

stiffness. Hence, as the wall in the field is not fully rigid or perfectly

hinged, its lateral stiffness is difficult to assess. In the current study, the

monolithic wall with known stiffness was used. This shall shed light on

the fundamental effect of lateral wall stiffness in excavation, as well as

6
Chapter 1 Introduction

reflect to what extend the lateral wall stiffness affect the response of the

ground.

2) Through centrifuge test results, mechanics study to better understand the 3D

behaviour of an excavation would be carried out. This is important as a

better understanding would allow a greater exploitation of corner effect in

practice, resulting in a more economical design.

3) Subsequently, it is aimed to quantitatively evaluate the salient feature of

corner effects. This includes the establishment on how the 3D effect is

developed, the contributors of such effects and the influence to an

excavation. In many of these excavations, capping beams are provided.

These are seldom evaluated, mainly because in 2D analyses, such an effect

cannot be accounted for. The capping beam effect in an excavation is

assessed in this study.

7
Chapter 2 Literature Review

CHAPTER 2

LITERATURE REVIEW

2.1 INTRODUCTION

The mechanics of an excavation supported by retaining wall is a typical soil-structure

interaction problem. This problem is a major topic in geotechnical engineering and

drawn extensive coverage in geotechnical textbooks (Lambe & Whitman 1979; Das

1990; Craig 1992; Terzaghi et al. 1996, Bowles 1996; Sevenoaks 1996). Fundamental

theories on lateral earth pressures, stability of slip and basal heave, as well as the

theories and designs of mechanically stabilized earth and concrete walls, cantilevered

and anchored sheet pile walls, cellular cofferdams are widely available. However, most

theories described above are for excavation at limit state where the problem is

structurally determinate. For deep excavation with multi-level propped retaining wall

designed to limit ground settlement, especially near built-up area, the resulting wall

movement is often insufficient to allow full active or passive pressure to be mobilised.

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

reality. Hence, traditionally, empirical relations and charts are generally used in design

for multi-level strutted retaining wall excavation system. Some of these charts are

proposed by Peck (1969), Goldberg et al. (1976), O’Rourke et al. (1976), Clough and

Denby (1977), Clough et al. (1979), Mana and Clough (1981) and Terzaghi et al.

(1996). Figs. 2.1 to 2.5 show some of these charts.

Although empirical charts are easy to use and able to provide some estimation of the

ground movement, the accuracy of the estimation is not certain. This is because ground

movement and strut loads in excavation are highly dependent upon ground conditions,

8
Chapter 2 Literature Review

soil properties, construction system and sequence. As the empirical tools do not take

all these factors into account, they are likely only applicable to some specific problems.

Furthermore, there are growing concern on the impact of excavation induced ground

movement to the surrounding structures and foundations (Burland et al. 1979; Brassinga

and Van Tol, 1991; Finno et al. 1991; Poulos and Chen 1997; Tan et al. 1995). Also,

the awareness on the impact of excavation to buried pipes, high voltage cables and

services are also rapidly increasing (Needham and Howe 1984, Philips 1986, Nath

1993). Fig. 2.6 shows an illustration of distress caused to a buried service by a

trenching operation. Hence, since 1970s, two-dimensional (2D) plane strain finite

element method (FEM) is often used to predict ground movements induced by

excavations owing to its ability to incorporate soil non-linearity, boundary conditions,

simulate construction sequences and model time-dependent consolidation phenomena

(Clough and Mana 1976, Clough and Hansen 1981). It is able to predict the excavation

behaviour more accurately at various stages of excavation for a more proper movement

control around the excavation.

However, it is noted that most of the theories, empirical charts and FEM analyses

described above are devoted for 2D problem. While it is recognized that most

excavation in the field are 3D in nature, there are very little understanding devoted to it.

This is mainly because 3D problems are very complicated, and since 2D analysis is

usually more conservative, it is deemed sufficient in most design. The fact that the

extensive computing resources required to carry out 3D analysis also discourage

engineers venture into 3D analysis. However, due to the more stringent requirements

for movement control placed by authorities, and the increasing awareness of movement

control, there are growing needs for more accurate prediction of excavation-induced

9
Chapter 2 Literature Review

movements especially for excavation adjacent to critical structures. Hence, the need

for 3D analysis arises for more accurate prediction of the soil-structural interaction

problem. In addition, the fast growing development in information technology has

lowered the cost of computing resources drastically and made 3D analysis more

affordable. As pointed out by Lee et al. (1998), it is expected that the saving of

construction cost can easily outweigh the cost of conducting 3D analysis, should the 3D

analysis is conducted properly. Hence, the real challenge is to carry out a proper 3D

analysis, and to achieve this, a better understanding on the 3D behaviour must first be

sought.

Presently, the understanding on the 3D excavation behaviour is very limited. This is

mainly due to the lack of good physical data to calibrate numerical analyses and to

allow in depth mechanics study to be carried out. Often, while field data represent the

most realistic situation, its complex soil profile, ill-defined initial conditions, and

complicated excavation procedures limits its usage for mechanics study. Hence, as

presented in Chapter 1, the first objective of the current study is to develop and carry

out proper centrifuge excavation test to provide a realistic glimpse of the excavation

behaviour around a corner under fairly simple boundary and initial conditions. This

will facilitate the interpretation of results and allow the capture of the salient features of

3D excavation behaviour. The development for centrifuge modelling of excavation

will be discussed in the following section.

Through carefully planed and executed centrifuge experiments, the response of the soil

and retaining wall could be obtained. The mechanics of excavation behaviour around

10
Chapter 2 Literature Review

the corner of an excavation can then be investigated. As such, the literature review in

this study is divided into two main parts, which are:

1) Literature review on centrifuge modelling on deep excavation, and

2) Literature review on 3D excavation behaviour.

2.2 LITERATURE REVIEW ON CENTRIFUGE MODELLING ON DEEP

EXCAVATION

Centrifuge is a well accepted tool in modelling and solving geotechnical problems as

evident by the number of papers presented in international centrifuge conferences

namely Centrifuge ’88, Centrifuge ’91, Centrifuge ’94 and centrifuge ’98. Though it

has certain limitations, it also offers many advantages; chief among these are ensuring

correct physical modelling principles, well defined initial and boundary conditions and

well controlled soil properties. Many literatures on the background and application on

centrifuge modelling tests are available (Craig and Rowe, 1981, Craig, 1984; Schofield,

1980, 1988; Kimura 1988; Scotts 1988; Taylor 1994; Leung et al. 1991).

In the centrifuge, the soil sample and other model set-up are spun at a high angular

velocity to exert a centrifugal force to the model that is equivalent to N-time the

gravitational force, g. In typical applications, the centrifuge is spun at Ng-level ranging

from 50 to 200-g. In the National University of Singapore’s (NUS) Centrifuge (Lee et

al. 1991, 1992), to simulate a 100-g field, the centrifuge is spun at about 350 revolutions

per minute. At this speed, all the testing apparatus and control system must be mounted

onto the centrifuge platform to be spun together with the soil sample, and test carried

out in-flight. Such apparatus is given a prefix “in-flight”. The operations of the

apparatus must be remotely controlled, from a control room via a slip rings system.

11
Chapter 2 Literature Review

Due to the high centrifugal force, the model container must be sufficiently strong and

stiff to prevent failure or excessive deformation during spinning. In addition, the in-

flight modelling set-up, in this case, the apparatus to model excavation, must be able to

work under its amplified self-weight and high working load. As the size of the

centrifuge platform is generally not large, in the NUS centrifuge, 0.7m wide x 0.75m

long, the space for the mounting of a soil model and its modelling apparatus is limited.

Thus, to model excavation realistically in the centrifuge is difficult. Many researchers

have devoted extensive effort in the development of devices to model excavation in the

centrifuge. Some of these innovative development and methods to model 2D-

excavation (excavation to study the plane strain effect only) in the centrifuge are

presented below.

2.2.1 Various Methods to Model Excavation in Centrifuge

2.2.1.1 Increasing-g Method

The increasing-g method is one of the pioneering methods to model excavation (Lyndon

and Schofield 1970, Toyosawa et al. 1994). In this method, the soil in the excavation

area is first removed in a 1-g environment. The model is then mounted on a centrifuge

and subjected to increasing centrifugal acceleration to generate imbalance load, mimic

the condition of excavation (Fig. 2.7). In this method, the overall total stresses in soil

can be modelled but the effective soil stresses are difficult to determine as both the

imbalance load simulating the excavation as well as the consolidation of soil affects the

stress changes. This makes the interpretation of the test results very difficult. Another

setback to this approach is that the prototype dimension and stiffness of the model

retaining wall would increase according to the increasing g-level, where it should be

kept constant. Due to these shortcomings, the validity of this method is limited.

12
Chapter 2 Literature Review

Nevertheless, this method marked an innovative use of the technique of centrifuge

modelling. Kimura et al. (1993) carried out this “Increasing-g” method to study the 2D

excavation behaviour in Kaolin clay. They reported that due to the unknown initial

stresses in the experiments, they interpret the undrained shear strength of soil sample

based on unconfined compression tests (for constant strength with depth soil sample) or

multiplying with the preconsolidation pressure at different depths (for increasing

strength with depth soil sample).

2.1.1.2 Excavate and Spin Method

“Excavate and spin” or “Increasing-g” method is another alternative method to simulate

excavation. In this method, the centrifuge would stop at each incremental stage for

excavation to carry out in 1-g condition, before the model is spun again (Zhang and

Zhang, 1994; Frydman et al., 1994; Liu et al. 1994). These procedures are repeated

until the final excavation depth is reached or failure happens. This method is used

mainly for excavation involved complicated construction sequences, such as multi-

strutting and soil nailing in the excavation system. This method is able to deliver broad

trend of the behaviour and suitable for case-specific excavation problem related to the

field. However, due to the lack of knowledge in the model conditions, this method is

not suitable for mechanics study.

2.2.1.3 Heavy Liquid Method

Another method to simulate excavation in centrifuge, which is probably the most

popular method, is to draining a heavy liquid with a unit weight identical to that of the

soil to be excavated (Fig. 2.8). This method is referred to as the “heavy liquid”

method. This method is an improvement of the first method, has been used

13
Chapter 2 Literature Review

successfully by numbers of researchers e.g. Kusakabe (1982), Powrie (1986), Powrie et

al. (1994), Takemura (1992), Toyosawa et al. (1994), Bolton and Stewart (1994),

Richard and Powrie (1998), and Wei (1998).

As compared to the “Increasing-g” method, the stress condition of the model soil is

known and is kept constant throughout the excavation. Moreover, controlling the

draining of the liquid can more realistically simulate the excavation event. In “Heavy

liquid” method, the vertical stress history of the model soil can be correctly modelled as

long as the unit weight of the liquid chosen is identical to that of the original soil to be

excavated. However, the lateral stress history of the model required further

consideration.

It is noted that the total lateral earth pressure coefficient, Ko for the liquid is always

close to unity. Powrie (1986) in his excavation studies on over-consolidated (OC)

London clay explained that the initial lateral earth pressure coefficient of about 2.0 of

the clay would be reduced to a value of between 1.0 and 1.2 during the slurry trench

phase of the retaining wall. Hence, the Ko of unity of the liquid was deemed to be

approximately correct to simulate the soil Ko before the excavation. However, for

normally consolidated clay, which is commonly found in Singapore, which has a Ko

within the range 0.55 to 0.65, this method will not be able to produce the correct initial

horizontal stress condition.

In addition, the coefficient of lateral earth pressure for the soil on the excavation side

will change from Ko to Kp during an excavation with the development of passive soil

pressure. However, as the coefficient of lateral earth pressure of a liquid is a constant

(unity), this means it cannot simulate the development of passive earth pressure against

14
Chapter 2 Literature Review

the wall during an excavation. If this passive resistance is not correctly simulated, the

stress distribution in the model soil will be affected and is likely to be incorrect.

Powrie (1986) explained in his study that since the time taken to drain the liquid

solution is comparatively short (2 to 5 minutes in the centrifuge at 125-G, or 3 to 8

weeks in a prototype time), the error of simulating the lateral earth pressure at this stage

is considered negligible. However, it is in the author’s opinion that the effect of

correct simulation of earth pressure in an excavation needs further consideration. In

the field, excavation usually carried out in stages, and the time required for the complete

excavation to the formation level is usually longer, and the correct lateral earth

pressures are usually considered in the design of retaining wall. This shortcoming is

more severe in a 3D excavation, where the amount of passive earth pressure developed

will vary along the distance from the corner of the wall. This is one of the key effects

in a 3D excavation, and has to be accounted for to obtain a more accurate analysis of

deep excavation.

The researchers at Tokyo Institute of Technology (TIT) have used the above mentioned

alternative methods to model excavation in the centrifuge. In view of the shortcoming

of the above method, especially in modelling excavation in soft clay, they have

developed an in-flight excavator to simulate more accurately the excavation process in

the centrifuge (Kimura et al., 1993, 1994).

2.2.1.4 2D In-Flight Excavation Test In The Centrifuge

The TIT’s in-flight excavator set-up includes a centrifuge container and an excavator

(Fig. 2.9). The excavator consists of a movable table, a cutting blade and a soil-

retaining gate. The cutting blade, powered by stepper motor, is controlled to scrap soil

15
Chapter 2 Literature Review

at an intended speed and depth to model excavation in-flight. Some experiments were

also conducted with “tieback” to the retaining wall to simulate supported retaining wall

excavation. The set-up was designed to operate at 100-g environment, but finally could

only excavate soil at about 50-g (Kimura et al. 1993). The friction of mechanical part

is suspected to be the cause of this drawback. However, due to space limitation, higher

capacity motor cannot be accommodated to improve the performance of the excavator.

From the earth pressure measurements on active and passive sides of model retaining

wall, they showed that the initial stress condition of soil was brought from near-isotropy

to anisotropy or Ko condition before and after centrifugal consolidation. From earth

pressure on retaining wall measurements, they showed that when failure took place, the

horizontal earth pressure in the passive side generally agreed well with the pressure

calculated based on Rankine formula, but at the active side, the pressure measured was

smaller than Rankine pressure. Tanaka (1994) has also reported this phenomenon in a

field retaining wall study and postulates that this is due to the changes in shear strength,

which are dependent on strength anisotropy, swelling of soil, adhesion between ground

and the retaining wall. Kimura et al. (1993) postulates that this phenomenon may be

due to stress anisotropy. While there is no certain explanation on this phenomenon, it

is very clear that only realistic simulation of excavation could deliver such complicated

soil behaviour. This highlighted the importance of realistic modeling of centrifuge

experiment if true mechanisms are to be captured.

Kimura et al. (1994) also modeled the provision of bracing to the retaining wall by

using a tieback system. A load cell measured the load in the tieback wire. Through

16
Chapter 2 Literature Review

this experiment, they highlighted the advantage of preloading in tieback in an

excavation problem.

TIT’s in-flight excavator is a breakthrough in centrifuge modelling of deep excavation.

Realistic excavation can be performed using this excavator on soil sample with well-

controlled boundary and initial conditions.

2.2.1.5 Tunnel Excavation Method

Other than basement excavation, tunnel excavation or tunneling is also a major topic in

geotechnical engineering. Various tunnel excavation methods in centrifuge were also

developed (Onoue et al. 1994; Nomoto et al. 1994; Konig et al. 1994). As tunnel

excavation is a very different construction method compared to basement excavation,

the excavation techniques are also very different. As the focus of the current study is

on vertical cut excavation, the tunnel excavation technique will not be discussed here.

In NUS, other researchers (Wei 1998) as well as the author had used the heavy liquid

method to simulate excavation in the early stage of experiments. However, it was

found in centrifuge consolidation test for soft NC clay with the heavy liquid setup that

the larger pressure from the heavy liquid had pushed the retaining wall towards the

retained soil side. This make the initial stress condition of the sample very complicated

for interpretation. This serious setback reinforced the author’s desire to develop an in-

flight excavator in this study.

17
Chapter 2 Literature Review

2.2.2 The Challenge of 3D In-Flight Excavation Test

In order to perform 3D in-flight excavation test in the centrifuge, an in-flight set-up,

able to excavate soil at an accurate depth and speed under high-g environment, as of

TIT’s excavator, has to be first developed. In addition, the preparation of the soil

sample, model retaining wall and the set-up has to be able to simulate a 3D excavation.

Furthermore, the amount of transducers and control required for a 3D excavation are

significantly more than those for 2D to obtain meaningful results. All of these factors

contribute to the difficulty of conducting 3D in-flight excavation tests. Due to the

above difficulties, it is not surprising that to date, there is no one in the world who can

perform 3D excavation test in the centrifuge, let alone the 3D in-flight excavation test.

In this project, a challenge is taken up to develop and conduct 3D in-flight excavation

tests in the centrifuge to provide reliable physical data. The key challenge is to develop

suitable experimental apparatus to conduct centrifuge test with well-defined initial and

boundary conditions and to carry out model excavation under realistic excavation

sequences. Through these tests, better understanding of 3D behaviour of an excavation

can then be derived.

The development of an in-flight excavation apparatus suitable for 3D excavation is

already a very complicated task. To put in place an in-flight strutting system is more

daunting and thus was not pursued in this project.

In all excavations, the first stage is usually an unbraced excavation. If the excavation

is in soft soil, this first stage of excavation is often a critical stage where significant

movement can take place if it is not handled properly. Thus the results from this study,

18
Chapter 2 Literature Review

besides their intrinsic value to provide reliable data for fundamental study, are also

realistic for actual excavation, especially the first stage. In many of these excavations

especially when contiguous bored piles or diaphragm walls are involved, ground beams

are provided. These are seldom evaluated, mainly because in 2D analyses, such an

effect cannot be accounted for. The ability to understand the optimal depth of

excavation possible in an unbraced excavation is also an important construction

consideration.

2.3 LITERATURE REVIEW ON 3D EXCAVATION BEHAVIOUR

2.3.1 General Appreciation of 3D Excavation Behaviour

The 3D effect in excavation was studied as early as 1970s. Giger and Krizek (1975)

conducted a study on the stability of unsupported excavation with corner angle varying

from 90° to 180° (Figs. 2.10a and b). They reported that the stability of a vertical cut

was significantly influenced by the angle of the corner, where an increase of the corner

angle would reduce the stability of a vertical cut. They showed that when the corner

angle equals to 180°, the stability of the vertical cut was the same of those for 2D

problem and was most critical.

St. John (1975) carried out 2D and 3D analyses of square, unsupported excavation in

London Clay using finite element analyses. He reported that axisymmetric analyses

could provide good agreement with that of 3D analyses, and the 2D analyses over-

predicted horizontal soil movements up to 100%. Burland et al. (1979) also suggested

that axisymmetric analysis is a more appropriate method to estimate ground movement

of a square excavations than using plane strain analysis, as the hoop stress of the

excavation is partially considered. Simpson (1992), in his study of the British Library

19
Chapter 2 Literature Review

excavation, reported that the differences between axisymmetric and plane strain analysis

are negligible and attributed this to the shallow depth of a relatively stiff stratum.

Arul and Kusakabe (1983) and Kusakabe (1984) in their study on axisymmetric

unsupported excavation, reported that the stability number of an axisymmetric

excavation decreases with the function D/ro, where D is the excavation depth and ro the

radius of the excavation cylinder (Fig. 2.11). This implies that for a constant excavation

depth (D), an increase of the radius of excavation would reduce the stability of an

axisymmetric excavation. This highlighted the excavation geometry effect in

excavation behaviour. They argued that their studies are applicable to deep excavation

problem, as the excavation can be approximated as semi-circular in plan and its

behaviour similar to that of a 3D condition. Philips (1986) in his studies on trench

heading problem, found that the stability of an unsupported flat trench heading was

similar to that for the semi-circular heading.

All the above studies (St. John, 1975; Burland et al., 1979; Arul and Kusakabe, 1982;

Philips, 1986; Simpson, 1992) are in agreement that axisymmetric analysis could

approximately explain some of the 3D excavation behaviour. However, as pointed out

by Lee et al. (1998), as excavation are usually irregularly in shaped and supported by

struts, axisymmetric approach is difficult to apply. Further more, the hoop stress

development behaviour does not replicate the lateral flexural action of the waler system.

Hence, while the axisymmetric approach is simple and intuitively plausible, its

extension to explain the 3D corner effect in a supported excavation is limited.

20
Chapter 2 Literature Review

De Moor (1994) in his study on the installation effect on bored pile and diaphragm wall,

investigate the soil stress changes around diaphragm wall trenching. He is aware of the

drawback of 2D analysis which would overestimated the changes around the trench, and

acknowledged that axisymmetric analysis could provide better prediction than the 2D

analysis. However, for a better prediction, he proposed an alternative to consider the

3D behaviour of the trenching work. In his finite element analysis, the 2D plan

horizontal section through a series of wall panels, rather than a vertical cross section

normally used in wall analysis, was used to study the trench behaviour (Fig. 2.12). By

using this approach, he was able to analyze the behaviour of a panel excavation at a

given depth. This method provides an indirect way to account for corner effect as well

as the influence of the overall dimension of an excavation.

The above studies highlighted that the appreciation of 3D effect in excavation and the

traditional approaches to understand and quantify the effect. It is interesting to note

that these early studies are on soil alone, without the presence of retaining wall. This is

probably because the soil behaviour is the primary factor initiated 3D effect and hence

most attempts were devoted to understand it before proceeding into the more

complicated soil-structure interaction in a supported excavation. The lack of

computing resources capable to carry out 3D excavation analysis is also a possible

factor. Since late 1980s, due to the advancement of computer technology, computing

resources able to handle 3D analysis are becoming readily available and affordable.

Hence, many studies on 3D supported excavation were sprouted since then. Some of

these studies were reported by Ou and Chiou (1993), Fernandes et al. (1994), Lee et al.

(1998), Ou et al. (1996) and Lin et al. (2003). Due to the increased awareness on

movement control, the National University of Singapore spearheaded series of research

21
Chapter 2 Literature Review

studies on deep excavation topic. Some of these studies are reported by Yong et al.

(1989), Chew et al. (1998), Tan et al. (1995), Liu et al. (1996) and Lee et al. (1998).

Substantial attentions were devoted to study the 3D behaviour of excavation (Liu et al.,

1996; Yong et al., 1997; Chew et al., 1998; Lee et al., 1998 and Sun, 2003). Chew et

al. (1998) used 2D and 3D analyses to predict the wall deflection of a strutted

excavation underlain by deep deposit of soft clay in Central Business District in

Singapore. They found that the 2D analysis always over predicted the wall deflection,

especially nearer to the corner. They attributed this effect to the soil arching and the

mutual support from the intersecting walls at the corner. Lee et al. (1998) carried out

2D and 3D analyses on a multi-strutted diaphragm wall deep excavation for the

Immigration Building in Singapore. They found that at a section nearer to the corner,

the 2D and 3D analyses over predicted the measured maximum wall deflection by 200%

and 30% respectively. The 2D analysis apparently yield more severe error than that of

3D analysis because it did not account for any 3D effect.

General finding from the above studies was that 3D analysis usually provides more

accurate prediction for the wall movement in an excavation than 2D analysis, especially

for small size excavation (Ou and Chiou, 1993; Chew et al., 1998). Detail discussions

on the findings obtained from the above studies are presented in the following several

sections.

2.3.2 Quantification of 3D Corner Effect

While the 3D corner effects were appreciated and intuitively considered in design, there

was very little information on how to quantify the 3D corner effect. Ou et al. (1996)

defined a plane strain ratio (PSR) to quantify the magnitude of 3D corner effect in

22
Chapter 2 Literature Review

excavation. The PSR was defined as the ratio of the maximum wall displacement (δ) of

a section over the maximum wall displacement (δ2D) of the section. The higher the PSR,

the lesser is the corner effect, and when PSR is unity, the section is in a plane strain

condition. From FEM parametric studies on various widths and lengths of excavation

(3D excavation behaviour), and subsequently compared the results to independent 2D

analyses using the same program and parameters, a PSR chart relates PSR ratios with

excavation width, length and distance from corner was developed (Fig. 2.13). They

argued that in the field, 3D results could be estimated based on 2D analysis, multiply by

the PSR ratios. They defined this approach as “PSR Method”.

Ou et al. (1996) asserted that with this method, the computing resources required to

capture 3D behaviour of excavation could be significantly reduced. They

acknowledged that the PSR chart reported were not accounted for various factor in

excavation, such as the excavation sequence, excavation depth, penetration depth of

support wall, wall stiffness, excavation geometry and soil strength and can only used for

problem with identical parameters. However, the PSR chart can be expanded to include

other factors.

Lin et al. (2003) carried out a study on corner effect in excavation, and reported a case

study on Bangkok Subsoil (Rajavej Hospital Project). Their approach towards

assessing corner effect was based along the PSR Method proposed by Ou et al. (1996).

Also, using independent 2D and 3D analyses, but with the same program and

parameters, PSR charts were developed for this site condition. Subsequently, they used

the PSR Method to estimate the lateral wall movement, and compared it with field

measurement results. Their results shows that the PSR Method only slightly reduces

23
Chapter 2 Literature Review

the over-predicted movement obtained by 2D FEM analysis, and the results were still

significantly different from the measured field results.

The PSR Method proposed by Ou et al. (1996) is innovative and is probably the most

specific method on quantifying the 3D corner effect to date. However, this method has

an apparent problem, that is, the 2D and 3D analyses were both numerical. It is known

that one of the major problems of FEM analysis is the inability to model the full

spectrum of real soil behaviour. Some of the complicated soil behaviour such as the

small strain soil non-linearity (Jardine et al., 1986, 1991; Kavvadas, 1994; Whittle et al.,

1993; Whittle 1997) and anisotropy characteristic (Banerjee et al. 1984, Wroth and

Houlsby 1985; and Banerjee and Yousif , 1986; Jovicic and Coop, 1998) have known to

affect the FEM results. The uncertainty of this soil behaviour under a 2D and 3D

environment alone might cause different set of problem in 2D and 3D analyses. In

addition, the construction details are very difficult to model in FEM analyses. Lee et

al. (1998) reported the difficulty of modelling the out of plane strut forces of diagonal

struts at corner in 2D analysis, and they resorted to approximately model and adjust the

equivalent forces and stiffness of struts to reflect equivalent values per unit width as that

of 3D analysis. Hence, it is clear that the direct comparison between 2D and 3D results

is very difficult. It is in fact likely that the PSR of unity when 3D behaviour ceased at a

section far from the corner may not be achieved at all.

In addition, it is also fundamentally doubtful whether the 2D results obtained in the

FEM analyses represent the actual 2D results in the field, as the actual condition in the

field is always 3D in nature. Hence, is there any 2D excavation condition in the field?

If it is, what is the definition and under what conditions? Lee et al. (1998) in their case

24
Chapter 2 Literature Review

study of a deep excavation in Singapore (the Immigration Building) reported that the 2D

FEM predicted maximum wall displacements were constantly about 30% larger than the

field results, even at the mid-span of the excavation. The mid-span was about 37m

away from the corner. They hence suggested that even at this distance, the 2D

conditions still not fully prevailed. This comparison of 2D FEM results with field

measurements to quantify 3D corner effect was also adopted by others (Nath, 1983).

However, is this approach to quantify 3D corner effect valid? To answer these

provoking questions, it is obvious that an understanding on the 3D corner effect

influence range is necessary.

2.3.3 Influence Range of 3D Corner Effect

Lee et al. (1998) has studied the effects of corners on wall deflection and ground

movement around multi-strutted diaphragm wall deep excavation for the Immigration

Building in Singapore. The excavation plan is about 75m long and 50m wide. The

excavation depth is 17.3m below the ground level with three layers of corner diagonal

struts. Significant corner effects have been observed from inclinometer and ground

settlement data for short and long sides, above and below excavation levels. Lee et

al.(1998) compared the 2D and 3D results from FEM analysis and found that the

maximum wall deflection was well predicted by the 3D analysis, whereas the plane

strain 2D analysis consistently over predicts the deflection by about 30%, which suggest

that some corner constrain effects persist and plane strain conditions do not completed

prevail even at mid span sections of this excavation (37m from corner). Lee et al.

(1998) suggests that the corner effect might be control by the length of excavation (L) to

depth of excavation (h) ratio (L/h), a low ratio would provide more significant corner

effect. In their study, the L/h ratio is about 3.0 to 4.5 and significant corner effect still

25
Chapter 2 Literature Review

persists. They have also provided a summary table on corner effect and relationship

with the L/h ratio. Part of the information provided were used and presented in Table

2.1 for the subsequent discussion.

Nath (1983) in his study on the effect of trench excavation on adjacent buried pipes,

used 3D FEM analysis to study the 3D behaviour of trench excavation. He assessed the

corner effect of a trench by comparing the 3D analysis results with plane strain analysis

results. From a typical analysis on a 10m long x 6m deep x 1m wide trench excavation

(Fig. 2.14), he reported that the when Length (L)/ Depth (h) greater than 2, the

behaviour is plane strain. Lee et al. (1998) also use the factor of L/h to characterise

corner effect. A summary of corner effect influence range and the associated factors

reported by Nath (1983) and others are shown in Table 2.1.

Ou and Chiou (1993) carried out a case study on the Hai-Hua Building in Taipei. The

comparison of 2D and 3D analyses shows significant corner effects at a distance of 1

times depth of excavation from the corner. From the reported results, it is estimated

that the L/h is between 1.0 and 1.5 in their analysis. Wong and Patron (1993) reported

five cases of excavation in Taipei with diaphragm wall system. They have also

suggested that corner effect were present at distance of 1 times excavation depth from

corner or less.

Simpson (1992) reported a case study in stiff London Clay with L/h ratio of 3.3 and no

apparent corner effect was observed. He attributed this to the shallow depth of

excavation and stiff soil stratum.

26
Chapter 2 Literature Review

Ou et al. (1996) in their study on the effect of size of excavation carried out parametric

studies using FEM analyses with various primary and complementary wall lengths (Fig.

2.13a). They found that for a small length of complementary wall (B=20m), the plane

strain condition could be achieved when the primary wall is 40m long (L = 40m).

They also reported that when the complementary wall is short, the distance from the

corner to a section having plane strain condition seems to be independent of the length

of primary wall (Fig. 2.15-B=20m). However, they found that for a given larger length

of complementary wall (e.g., B≥20m), the decrease of primary wall length results in the

decrease of a section’s wall deflection, and for such cases, plane strain condition has not

been achieved even when the primary wall length is 100m long. This implies that the

complementary wall length would also significantly affect the influence range of corner

effect. No explanation is provided for this behaviour in the report. It is in the opinion

that the complementary wall length may influence the behaviour of a small cofferdam

excavation, as the movement at the corner would affect the movement of the primary

wall. However, in a case when the corner has negligible movement, as in the case of

their study, the characteristics of the complementary wall shall have insignificant effect

on the primary wall. In addition, Ou et al. (1996) shows that the 2D movement is

significantly affected by the size of excavation, B. This is quite unreasonable, as the

space in front of a plane strain excavation should not have such major impact on the

overall excavation behaviour. Ou et al. (1996) characterize the corner effect with the

B/L ratio, rather than the L/h ratio used by Nath (1983). However from the reported

data, as the depth of final excavation is 16m, the L/h ratio can be estimated as 2.50 and

6.25 for B = 20m and 100m respectively. These results were also included in the

summary in Table 2.1.

27
Chapter 2 Literature Review

Lin et al. (2003) in their 2D and 3D FEM modelling of a 10m deep excavation project in

Bangkok suggested that the corner effect influence range is 25 to 30m from the corner.

From the above review, the corner effect influence range observed on site is generally

confined to 1 to 2 times depth of excavation or less (Nath, 1983; Simpson, 1992; Ou and

Chiou, 1993; Wong and Patron, 1993). The case study of Immigration Building in

Singapore reported by Lee et al. (1998) registered corner effect range beyond 2 times

depth of excavation, which is larger than those reported. However, as most of the struts

employed in this projects were corner diagonal struts, this stiff corner struts may

amplify the corner effect.

It is noted that the influence range reported by Liu et al. (2003) is 3.0 times excavation

depth. This value was higher than those reported above. It is also reported by Ou et

al. (1996) that up to 3.12 times depth of excavation (for case B = 100m, refer Table

2.1), the corner effect still persists. Ou et al. (1996) further reported that the influence

range would further increase if there were an increase of excavation width (B).

Hence, it is apparent that Liu et al. (2003) and Ou et al. (1996) may be over-estimated

the corner effect influence range. It is interesting to notes that both Ou et al. (1996) and

Lin et al. (2003) influence range were deduced from the comparison of 2D versus 3D

FEM analyses, without verification with field results. Hence, it is possible that the

movement predicted in their 2D analyses might be always higher than those obtain from

3D analyses. It is even more interesting to imagine that when movement registered in

the field are constant with distance from corner, this usually considered as plane strain

condition achieved. This is a common understanding percept by many. Hence, the

corner effect influence range deduced from the comparison of 2D versus 3D analyses

28
Chapter 2 Literature Review

might over-estimate the real influence range. More detail discussions on this issue are

presented in the subsequent Chapters 4 and 5.

2.3.4 Other Parameters Which Might Influence Corner Effect

Simpson (1992) reported a case study of excavation in stiff London Clay and no

apparent corner effect was observed. He attributed this to the shallow depth of

excavation and stiff soil stratum. From the comparison of various case studies, Lee et

al. (1998) suggested that the corner effect might be influenced by the soil profiles,

especially the depth to stiff stratum. They suggested that if soil strata underlie the

excavation, the maximum wall deflection is likely to reach above excavation level, and

corner effect may be suppressed if sufficiently stiff strutting system is used. It is also

observed from a numbers of published case studies (Dysli et al. 1982; Bono et al. 1992;

Wong et al. 1993; Liu 1995; Lee et al. 1998) that the corner effect is usually reported in

excavation within soft ground. This soil strength effect is discussed in greater detail in

Chapter 4.

Fernandes (1986), in his study of the 3D effects of the strutting in an infinitely long

excavation supported by diaphragm wall, concluded that, 3D effects arising from the

strut loads are only significant in the vicinity of the struts and does not play an

important role in the overall behaviour of the wall. This argument is logical for cross

strutting as strut forces/stiffness are distributed evenly along the length of wall and it

would only provide passive support against movement. It is unlikely to contribute to

3D corner effect. This is not true when corner diagonal struts were used, as those

reported by Lee et al. (1998), where significant corner effect was still registered beyond

2 times depth of excavation, which is larger than those generally reported in the field

29
Chapter 2 Literature Review

(Section 2.3.3). This is most likely because corner diagonal struts will provide more

stiffness at the corner and in turns amplified the corner effect. However, Fernandes

(1986) opinion is still mostly correct as the corner strut effect should fall into the

category of excavation configuration, which includes the excavation shape and

geometry. However due to the complexity and case specific nature, the excavation

configuration effect was seldom studied thus far.

2.4 SUMMARY

In the literature review above, the necessity of carrying out 3D analysis was presented.

It was shown that due to the lack of physical data for calibration of numerical analyses,

there is serious lacking of fundamental understanding on the 3D behaviour of

excavation. In particular, how the 3D corner effect is developed in an excavation has

not been studied before. Hence, the significance of carrying out centrifuge model test

is presented. Subsequently, detail reviews on traditional methods used in the centrifuge

to approximately simulate excavation were presented. Then, the current status on the

development of centrifuge technology to simulate excavation realistically in the

centrifuge was presented.

The general understanding and studies on 3D behaviour on excavation was reviewed

and may be categorized into four main parts, which are:

1) General appreciation of corner effect: Previous studies on the general

observation on corner effect were reviewed. Further, the traditional

approximate methods, such as the axisymmetric method to understand the 3D

behaviour of an excavation were presented.

30
Chapter 2 Literature Review

2) The quantification of corner effect: It was clear from the review that although

3D corner effect is recognized and intuitively considered in design, there was

very little, if any, understanding on how to quantify this effect. The first

quantification method, the PSR ratio method to quantify corner effect is

reviewed and examined for its advantages and validity. It was shown that this

method is innovative, but has limitation due to the difficulty to directly compare

2D and 3D FEM results.

3) The corner effect influence range: Numerous site observation and case studies

on corner effects were presented. It was found that the influence range

observed in the field were within 2 times depth of excavation. However, the

range predicted solely from FEM analyses were higher than those observed in

the field.

4) Factors influencing corner effect: Various factors, which may influence the

characteristics of corner effects, were reviewed. These include the excavation

width and length, the soil strength and strata-formation, the strutting and the

excavation configuration effects.

31
Chapter 3 Development of 3D In-Flight Excavator and Centrifuge Model Tests

CHAPTER 3

DEVELOPMENT OF 3D IN-FLIGHT EXCAVATOR AND

CENTRIFUGE MODEL TESTS

3.1 DESIGN AND DEVELOPMENT OF 3D IN-FLIGHT EXCAVATOR

In the present study, an in-flight excavator capable of modelling 3D excavation was

developed. The Tokyo Institute of Technology’s (TIT) (Kimura et al., 1993, 1994) in-

flight excavator has provided much inspiration for the design. The main design criteria

for the 3D in-flight excavator was to allow for proper preparation of soil and its

associated transducers in the laboratory floor, and then the excavator set-up can be

mounted separately to excavate soil in 100-g environment. The proper soil sample

preparation, precision control of excavation mechanism, and careful acquisition of data

are paramount to obtained realistic and accurate data, which is critical for data

interpretation and final in-depth mechanics study. The main objective of the current

study was to gain a better understanding on the fundamental aspect of 3D behaviour of

an excavation. Hence, the approach of the experimental set-up was thus focus on the

establishment of an elementary set-up with most typical and simple excavation

configuration, to ensure accurate interpretation can be carried out subsequently. Thus,

the excavator set-up was based on un-braced excavation, with a right angle corner to

provide simple yet useful data to capture the salient features of 3D effects in an

excavation. However, to ensure that the bracing effect can be included in further study,

the set-up has provision of headroom above the in-flight excavator to cater for bracing

installation. This is an improvement over the TIT’s in-flight excavator where no

headroom is provided and the bracing effect was studied by providing tieback to

retaining wall (Kimura et al., 1993, 1994).

32
Chapter 3 Development of 3D In-Flight Excavator and Centrifuge Model Tests

The author carried out the complete designs, and the fabrication of the excavator set-up

was carried out by a local professional workshop with extensive experience on

centrifuge model set-up.

The excavator set-up consists of two main parts: a detachable lift-shaft and a watertight

centrifuge container, as shown in Fig. 3.1. Figs. 3.2a and b illustrate the various

components of the excavator, such as the detachable lift-shaft, water tight centrifuge

container, Stepper Motor 1, Stepper Motor 2, linear rails, scraper platform, soil retaining

gate. During 1-g and high-g consolidation of soil samples, the detachable lift-shaft was

replaced by a detachable sidewall. The Stepper Motor 1, which is mounted on top of

the centrifuge container, is also detachable and will only be assembled before the final

re-consolidation and excavation period. This provides headroom on the centrifuge

container for surcharge loading to be applied in 1-g consolidation of sample. The

scraper and the soil-retaining gate are mounted on the scraper platform. The vertical

movement of the scraper and soil-retaining gate is synchronised and the top of the gate

is always 1mm lower than the edge of the cutting blade. This is to enable the scraper

to scrape the soil into the lift-shaft for later disposal and ensure that the soil below the

excavation is confined. The vertical movement of the scraper platform is controlled by

Stepper Motor 1 through a bevel gears and precision ball screw system. This stepper

motor has a 1:30 ratio gearbox and can provide a torque of up to 400N.m at 300 RPM.

Stepper Motor 2 powers the horizontal movement of the scraper through a timing belt

and precision ball screw system. This motor which is smaller, has a 1:8 gear-box and

can provide a torque of up to 12N.m at 300 RPM. The vertical and horizontal

movements are guided by two linear rails. Commands to control the motors are sent

33
Chapter 3 Development of 3D In-Flight Excavator and Centrifuge Model Tests

via personal computers located in the remote control room through a pair of

indexers/drivers mounted on-board of the centrifuge. The excavation process is pre-

programmable but can also be controlled interactively using on-line commands. The

3D in-flight excavator scrapes soil at an accurately controlled speed and movement.

Table 3.1 shows the characteristic of the scraping mechanism.

Fig. 3.3 shows the schematic representation of the 3D in-flight excavator set-up on

National University of Singapore (NUS) centrifuge machine. References on NUS

centrifuge machine were presented elsewhere (Lee et al., 1991, 1992; Thanadol, 2002).

During the commissioning stage for the excavator, it was found that by locating the

indexers/drivers in the control room, which also housed the data-acquisition and video

system, the indexers/drivers generate electrical noise which interfered with the low

intensity transducers’ and video’s signals. The interference was so severe that the in-

flight miniature camera failed to transmit viewable image to the control room, which is

paramount for the monitoring of the centrifuge operation of the centrifuge. This

problem also happens in other centrifuge experiments by fellow researchers using

electrical stepper motor. Through some trial and error and troubleshooting, it was

found that by arranging the indexers/drivers on-board the centrifuge arm (Fig. 3.4), the

signal noise is reduced very significantly. To enable the indexer/drivers to be mounted

on board and subjected to high-g force, the internal parts of the indexers/drivers were

strengthened carefully and extensively by silicone sealant (Fig. 3.5). The

indexers/drivers were then mounted on the centrifuge arm, at an available space nearest

to the centrifuge rotating shaft (400mm from the shaft), so as to minimise the

centrifugal force experienced by the indexers/drivers.

34
Chapter 3 Development of 3D In-Flight Excavator and Centrifuge Model Tests

The 3D in-flight excavator developed is able to model approximately excavation for a

corner of a cofferdam (Fig. 3.6). The size of the excavation is 170mm width x 155mm

length in plan, modelling a 17.0m x 15.5m prototype corner excavation in a 100-g field

(refer Section 2.2 for the concept of centrifuge –g field). The centrifuge container has

a width of 435mm and a length of 435mm, simulating an area of research of 43.5m x

43.5m.

Another achievement of the in-flight excavator developed in the current study was its

transformation features. When mounted on a narrow container, a 2D excavation

problem can be simulated with the same excavator in the centrifuge (Fig. 3.7). This is a

very important feature which allow the 2D versus 3D behaviour of an excavation be

captured using the same excavator, minimizing unwanted complication if different

excavators, or set-up were used. This allows the comparative study on the 2D and 3D

behaviour be carried out and the 3D effect can be highlighted easily. This is important

in the current study. Other information about this 3D in-flight excavator can be found

in the earlier publication by the author (Loh et al., 1998).

3.2 EXPERIMENTAL DETAILS

The method and procedures for carrying out 2D and 3D in-flight excavation tests in the

centrifuge are described below.

3.2.1 The Model Retaining Wall

Model retaining walls are prepared prior to any centrifuge test. Two types of materials

were used in this study to make the model retaining walls. In the initial stage of the

experiment, a micro-concrete (concrete mortar) wall (Fig. 3.8) was prepared by casting

35
Chapter 3 Development of 3D In-Flight Excavator and Centrifuge Model Tests

sand and cement paste mixture into a smooth L-shape perspex mould. The

reinforcement of the wall was two layers of 2mm thick square wire mesh. With this

method, it was very difficult to cast a wall thinner than 10mm, in which the stiffness

(EI) is equivalent to a 1m thick concrete wall in the field if the experiment is conducted

at 100-g. On site, diaphragm walls are generally 0.6m to 1.2m thick. For sheet pile or

contiguous bored pile walls, the stiffness is usually less. Hence, the mortar wall can

only model a wall where the stiffness is on the high side of that used in practice. Thus,

in the second stage, aluminium alloy walls were used (Fig. 3.9). This material was

chosen as it is light but relatively strong. This minimised the possibility of its sinking

excessively during consolidation and disturbed the soil condition. The thicknesses of

these walls were 3mm and 1mm, which modelled stiffness equivalent to 430mm and

140mm thick concrete walls respectively, or 500mm x 225mm x 28mm and 400mm x

75mm x 8mm (width x height x thickness) U-type steel sheet pile wall respectively at

prototype scale under 100-g.

The size of the L-shaped model retaining wall was 170mm wide x 155mm long,

modelling a 17.0m panel joined perpendicularly to a 15.5m panel in prototype scale.

The two walls were welded together at right angles. This was intended to simulate a

corner of a rectangular wall in an actual excavation. The edges of the retaining wall

and the edges of the wall and soil in this region were therefore allowed to slide along

the side wall or perspex wall (Fig. 3.10).

Miniature total stress transducers (TSTs; Entran EPL Series) were attached, using

flexible silicone rubber on the model retaining wall. These transducers were used to

capture lateral earth pressures acting on the wall. The sizes of TSTs were 5mm wide x

36
Chapter 3 Development of 3D In-Flight Excavator and Centrifuge Model Tests

10mm long x 1mm thick. TSTs were installed in front and behind the excavation, at

different distances from the corner. The location of the TSTs will be presented in the

individual test report in the next chapter. Twenty-one sets of strain gages (SG) have

also been installed on the model retaining wall to measure vertical and lateral bending

moments at various locations along the wall (Fig. 3.11). The edges of the model walls

were carefully grounded and rubber seals were glued on the edges to ensure that the

edges were watertight during the excavation test.

3.2.2 Soil Sample Preparation

One of the biggest advantages in centrifuge testing is the ability to prepare a well-

controlled soil profile with known properties. For this purpose, remoulded soil sample

was used. The preparation of the clay is carried out by combinations of consolidating

the soil sample through surcharge loading in 1-g environment first and then by self-

weight in high-g centrifugal field. These two-stage consolidation processes were

designed to produce a layer of over-consolidated (OC) soil at the ground surface

overlying normally consolidated (NC) clay. This type of soil profile is more common

in the field, especially in Singapore (Chong et al., 1998).

Kaolin clay was chosen in this study because of its high permeability also because its

physical properties are similar to soft clay. The high permeability will reduce

consolidation time considerably, a major consideration as centrifuge time is at a

premium in NUS. This clay has been used substantially for centrifuge model ground

preparation for excavation related studies, both in NUS (Loh et al., 1998; Thanadol

2002) and other research centres (Philips, 1986; Kimura et al., 1993). The properties of

kaolin clay used are given in Table 3.2. The model ground preparation in this study is

37
Chapter 3 Development of 3D In-Flight Excavator and Centrifuge Model Tests

similar to that adopted by Kimura et al. (1993) and Thanadol (2002). The kaolin clay

was first remoulded to a water content of at least 1.5 times the liquid limit. A 100-litre

capacity clay mixer was used for remoulding the clay to uniform slurry. The

remoulded clay was then transported to a vacuum chamber for de-airing for 2 days to

expel air bubbles from the clay slurry and ensure a saturated soil model.

The clay sample inside the centrifuge model container is normally left for self-weight

consolidation on the laboratory floor for 2 or 3 days prior to surcharging using loading

plate. This is to let the top layer of the clay to gain some strength, which though small,

is important to prevent the clay from flowing out from around the edges of the loading

plate.

After the self-weight consolidation, a layer of geotextile was laid on the soil sample.

The loading plate was carefully lowered onto the soil surface. Linear variable

displacement transducers (LVDTs) were then attached onto the loading plate to monitor

the settlement of the soil sample through a data logger. The sample is consolidated for

14 days until at least 95% degree of consolidation was achieved.

Two types of soil samples were prepared for this study. The first sample consists of

normally consolidated (NC) soil of approximately 240mm thick overlain a thin layer of

over-consolidated (OC) soil achieved by a surcharge loading of 20kPa. The second

sample consists of highly OC soil achieved by applying surcharge of 230kPa. Fig.

3.12 shows the schematic representation of the OCR profiles obtained from the different

soil samples after the centrifuge consolidation described below.

38
Chapter 3 Development of 3D In-Flight Excavator and Centrifuge Model Tests

Druck PDCR81 pore pressure transducers (PPTs) installed in the clay samples were

used to monitor the pore pressure and consolidation behaviour of the clay during

centrifuge tests. For NC clay, where relatively lower surcharge pressure was applied,

PPTs were installed into the clay sample after the 1-g consolidation. This is to ensure

that the clay has obtained enough strength to support the PPTs during the subsequent

centrifugal consolidation. On the other hand, in the preparation of OC clay, PPTs were

installed before applying the final step of pneumatic pressure. The gaps formed during

the PPTs installation will closed up after applying the final surcharge pressure of

230kPa. Before inserting into the soil samples, the PPTs were saturated by submerged

under water and de-aired by boiling (Fig. 3.13). To insert PPTs into the soil sample,

the location for the PPTs were first determined. A miniature auger was then used to

bore a channel at the determined location on the soil sample. The PPT was then

inserted into the channel using a guilder (Fig. 3.14). To replace the soil bored out using

the miniature auger, the channel is then filled up with clay slurry. Typically, more than

ten miniature PPTs were installed behind the model retaining wall at different locations

in each clay model.

When the soil sample had achieved at least 95% average degree of consolidation under

1-g applied pressure, the soil sample was transported to centrifuge for second stage of

consolidation using centrifugal force. During this process, clay sample was always

maintained saturated by providing a thin layer of water on top of the model ground

surface. To prevent excessive evaporation due to long hour of centrifugal spinning, a

thin layer of oil is sprayed on the surface of the water before the centrifuge

consolidation.

39
Chapter 3 Development of 3D In-Flight Excavator and Centrifuge Model Tests

A water standpipe fitted with solenoid valve was mounted at the bottom of the model

container to allow bottom drainage through the sand layer place below the clay sample.

If soil sample modelling one way drainage condition was required, the solenoid valve

will be shut after the desired degree of consolidation was achieved. The soil was then

further consolidated for a few hours to achieve equilibrium under one way drainage

condition.

In the centrifuge consolidation, the degree of consolidation was monitored by measuring

the surface settlement of the clay sample and pore pressure dissipation inside the clay.

The conventional hyperbolic plot method (Tan et al., 1991) was used to predict the

ultimate settlement of the clay. With these final readings known, the degree of

consolidation of the clay during consolidation could then be reasonably predicted.

This centrifuge consolidation was stopped when the clay sample achieved an average

degree of consolidation of at least 95%.

3.2.3 Centrifuge Model Container Preparation

The centrifuge model container was assembled while mixing and vacuuming the clay.

The centrifuge model container was made watertight to prevent water leakage during

consolidation. In order to minimise the wall friction, the container walls were first

greased and then two layers of thin polyethylene sheets with a coating of grease in

between the sheets were then wrapped around the walls. This wall lubrication method

was shown by Khoo et al. (1994) to be effective in reducing friction between concrete-

sand, steel-sand and perspex-sand interfaces in centrifuge tests. As the soil sample

used in this study was kaolin and marine clay, the above lubricating method was also

efficient in reducing wall friction in this case.

40
Chapter 3 Development of 3D In-Flight Excavator and Centrifuge Model Tests

For preparation of a bottom drainage layer, a layer of fine silica sand, about 30mm in

thickness, was carefully laid on the base of the container. The laying of sand was

conducted under a layer of de-aired water to ensure the sand layer was always

maintained saturated. After the sand was laid, a steel plate was used to lightly compact

the sand surface. Finally, a ruler was used to level the sand surface. A water standpipe

was then installed leading from the sand layer to the pre-determined ground water level

inside the container. The de-aired clay sample was then transported from the vacuum

chamber to this centrifuge model container carefully by using a scoop conducted under

water. This soil sample was then ready for consolidation.

3.2.4 Instrumentation and Excavator Set-Up

Before the centrifuge excavation test, the consolidated soil sample was transferred to the

laboratory floor for final stage of instrumentation and excavator set-up. The

instrumented model retaining wall was first installed by slowly pushed into the

consolidated clay sample with the aid of a vertical guide. This process was conducted

very carefully to ensure that the wall was vertical and the edges flush with the side wall

and perspex wall (refer to Fig. 3.2 for the details).

In the field, it is generally known that wall installation would alter the soil stresses

around the wall (De Moor 1994, Powrie et al. 1998). In a diaphragm wall

construction, trenching prior to wall installation usually causes reduction of soil stresses

around the wall. De Moor (1994) conducted studies on diaphragm wall installation

effect using finite element analysis. He suggested that for stiff overconsolidated fine-

grained soils, the lateral earth pressure coefficient, σh’/σv’ (Ki) should be reduced to Ki

= 1.1, instead of Ko = 1.5. For the case of displacement walls, such as soldier and

41
Chapter 3 Development of 3D In-Flight Excavator and Centrifuge Model Tests

sheet pile wall, the wall installation effect is different. Finno and Nearby (1989)

conducted field measurement and finite element analyses on sheet pile installation effect

and suggests that sheet pile installation effects tend to preload the soil before excavation

begins. The preload has a beneficial effect on the active side of the wall where the

shear stresses available to resist movement during an active loading are larger than

would have existed without the preload. On the passive side, the preload reduces the

available shear resistant compared to that which would have existed without the

preload.

For the current centrifuge experiments, the wall was push in. Hence, the wall

installation effect would be that described by Finno and Nearby (1989). However, as

the push-in is carried out in 1g, and the model is further reconsolidated in the centrifuge

to reach new equilibrium before excavation started, the wall installation effect is

expected to be less compared to that in the field.

At this stage, the perspex wall was temporary removed to allow insertion of beads and

drawing of grid lines on the front face of the clay for image processing purposes. Then,

the perspex wall was re-assembled, and the detachable side wall was replaced by the

detachable lift shaft together with Stepper Motor 1. LVDTs to measure lateral

deformations at the top of the model wall and ground surface settlement behind the wall

were installed. Finally, a layer of de-aired water was again laid on the surface of the

clay sample before the sample is transported to the centrifuge for re-consolidation.

The above procedures are conducted quickly and without delay to avoid excessive

swelling or shrinking. A typical completed set-up is shown in Fig. 3.15.

42
Chapter 3 Development of 3D In-Flight Excavator and Centrifuge Model Tests

3.2.5 Excavation Procedures

The completed set-up was re-consolidated under 100-g for about twelve hours for the

soil to re-establish the equilibrium achieved earlier. PPTs and surface settlement

readings were again used to monitor the soil response in the re-consolidation process.

Once the soil had achieved at least 95% of average degree of consolidation, the model

excavation commenced without further stoppage of the centrifuge.

When model excavation began, Stepper motor 1, which controlled the vertical

movement of the scraper, was first activated. The scraper platform, in which the

scraper, stepper motor 2 and soil-retaining gate were mounted, was lowered so that the

cutting blade of the scraper penetrated into the soil (Fig. 3.16b). The interface between

the soil-retaining gate and the soil was properly greased to minimise friction at this

interface during motion. Stepper motor 2 was activated to pull the scraper horizontally

towards the lift-shaft. By doing so, the soil under the edge of the cutting blade was

trimmed and pushed into the base of the lift-shaft (Fig. 3.16c). The scraper platform

was then moved up again to a level slightly higher than the excavated level and then the

scraper was extended out from the lift shaft to a point very close to the front side of the

model retaining wall (Fig. 3.16a). This sequence was repeated until the final

excavation depth was achieved. For the current test, every cycle of excavation would

remove a 5mm layer of soil in 36 seconds. This would simulate a 0.5m thick

excavation in about 4 days in a prototype scale. This is quite realistic speed for a small

excavation. As the excavation stages is continuous, the soil is consolidating when the

readings are taken. To enable the parametric studies be carried out for various tests,

the speed of excavation of all the tests were maintained the same.

43
Chapter 3 Development of 3D In-Flight Excavator and Centrifuge Model Tests

The soil-retaining gate facing the soil is thoroughly lubricated with grease to minimize

friction. However, it is still likely that the movement of the soil-retaining gate would

affect to the soil in front of it. This would likely affect the passive earth pressure there.

However, the model retaining wall is located at 15.5m away from the soil-retaining

gate. For excavation depth within 5m, the distance of the wall from the soil retaining

gate is more than 3 times the depth of excavation. Hence, the soil friction at the soil-

retaining gate effect to the movement of the retaining wall and retained soil behind it

should be insignificant.

3.3 THE CONDITIONS OF IN-FLIGHT EXCAVATION TESTS

The above in-flight excavation test modelled the following conditions:

3.3.1 Area Modelled in the Experiments

The size of the model container for 2D in-flight excavation test at 100g is 155mm width

by 435mm long, simulating an area of research of 15.5m x 43.5m in prototype scale.

For 3D in-flight excavation test at 100g, the size of the model container is 435mm width

by 435mm (Fig. 3.6), simulating an area of research of 43.5m x 43.5m in prototype

scale.

For both 2D and 3D tests, the size of the actual area of excavation in plan view is

155mm width x 170mm length. For 2D excavation, the above excavation area covered

the entire width of the centrifuge container, hence simulating a plane strain excavation.

For 3D excavation, the excavation area occupies one corner of the strong box, thus

simulating a 17.0m x 15.5m prototype corner excavation (Fig. 3.6).

44
Chapter 3 Development of 3D In-Flight Excavator and Centrifuge Model Tests

This size of the container and the excavator set-up is limited by the space available on

the centrifuge platform. From previous studies (Wei 1998), the surface settlement

behind the excavation, which is a measure of the effect due to excavation, generally

diminished after a D/h (distance from the edge of the excavation/excavation depth) ratio

of 2.0. In this study, the minimum distance from the container boundary to the back of

the excavation is 265mm, or 26.5m in prototype scale. This provides a D/h of more

than 2.5 for up to 10m of excavation. Hence, the size of model container used in the

present study is deemed adequate to minimise the boundary effects of the container and

to ensure that enough space is provided to capture the entire excavation behaviour.

3.3.2 Fixities and Water Table Boundary Conditions

The four sides of the centrifuge strong box are flat, smooth and rigid, with proper

greasing (refer to Section 3.2.3) to minimise friction, thus providing effective lateral

restraint, making such boundary conditions easy to simulate in a finite element analysis.

The solid side walls and base of the container also provides impervious boundaries with

no ground water recharge. The soil sample was always maintained saturated and the

ground water table was slightly above the ground surface before and during model

excavation.

For 2D and 3D tests, the edges of retaining wall are sealed with rubber membrane and

silicone grease to prevent water from seeping through the side of the wall. For 2D test,

the retaining wall also straddles across the entire container ensuring that the wall will

deform in a plane strain manner. For 3D test, the edges of the retaining wall were

placed perpendicular to the container wall and allow sliding and rotating slightly against

the container wall. As such, this modelling is not a true modelling of a quadrant of a

45
Chapter 3 Development of 3D In-Flight Excavator and Centrifuge Model Tests

bi-symmetric excavation, where the edges would then have zero slopes at the mid-

points. A more detail discussion on this boundary condition and its limitation is

presented in Chapter 4.

3.4 METHODOLOGY OF TEST

In order to highlight the 3D effect of an excavation, a series of 2D in-flight excavation

tests, using the same set-up and configuration as described in Section 3.1 above were

first conducted. Subsequently, a number of 3D excavations were conducted to highlight

the 3D effect through parametric studies. The parametric studies conducted are

described in the following sub-sections.

3.4.1 Parametric Study on the Effect of Wall Stiffness

The lateral bending stiffness of retaining wall provided by corner restraint acts as an

additional support to restrict wall movement. In order to highlight this effect, 2D in-

flight excavation tests, which are free from such lateral bending stiffening effect, were

conducted and compared to the results of 3D in-flight excavation tests. Subsequently,

model retaining walls with different thickness were used for 2D and 3D in-flight

excavation tests to further evaluate the effect of lateral bending stiffness of walls, and its

influence on the 3D behaviour.

3.4.2 Parametric Study on the Effect of Capping Beam

In the field, capping beam is generally restrained at the corner. The capping beam may

enhance the corner effect by providing extra lateral stiffness to the retaining wall. The

capping beam effect is generally not felt in 2D analysis. This is an important effect,

especially for the first stage of excavation, which is usually without any bracing

46
Chapter 3 Development of 3D In-Flight Excavator and Centrifuge Model Tests

provided. However, as the effect is truly due to a corner, this has not been studied

specifically before. In this study, the capping beam effect in a 3D excavation is

investigated.

3.4.3 Parametric Study on Influence of Geotechnical Effect to 3D Corner Effect

The behaviour of an excavation is very much dependent on the geotechnical aspect of

an excavation, such as the soil strength, type of soil and depth of excavation. To

investigate these geotechnical aspects, comparative test on tests with marine clay and

stiff highly over-consolidated Kaolin clay soil sample are investigated. This

comparison is a basic first step to gain more insight on the geotechnical effect in this

interesting topic.

More detail discussions on the planning of the parametric studies are presented in

Chapter 4.

47
Chapter 4 Experimental Results And Discussions

CHAPTER 4

EXPERIMENTAL RESULTS AND DISCUSSIONS

4.1 INTRODUCTION

For the centrifuge tests carried out in this research, in-flight excavations were carried

out to ensure more realistic simulation of prototype excavation. The results aim to

offer a glimpse of the excavation behaviour around a corner under fairly simple

boundary and initial conditions. The advantage is that this approach will facilitate a

thorough interpretation of the results, which cannot be achieved from a much more

complex problem. One particular point that stands out from the tests carried out in this

research is the fact that no bracing is provided. The main reason is that this is very

complicated to simulate in-flight bracing during a centrifuge test. However, what is

important to recognise is that in any excavation, the very first stage is usually an

excavation without any bracing. In practice, for ease of construction and to expedite the

excavation, many engineers would attempt to carry out this first stage as quickly as

possible. Thus, besides providing a better understanding of the effect of corner, the

tests are also useful to understand the first unbraced stage of an excavation. The effect

of providing a ground capping beam, especially in cases where diaphragm wall is used,

is an important consideration. Attention is also paid to this point in the present study.

4.1.1 Initial Conditions of Tests

As described in Chapter 3 (Sections 3.2.4 and 3.2.5), after the model retaining wall was

installed, various transducers would be installed into the soil model, and then the whole

set-up, including the in-flight excavator would be spun again at 100-g. This process is to

reconsolidate the soil sample to allow it to reach equilibrium again. This

48
Chapter 4 Experimental Results And Discussions

reconsolidation process is monitored via pore pressure transducers (PPTs) and vertical

linear variable displacement transducers (LVDTs) installed on the soil surface.

The LVDTs and PPTs readings are captured real-time. The hyperbolic method (Tan et

al., 1991) is used to analyse the LVDTs readings to provide an accurate estimate of the

expected ultimate settlement. A typical surface settlement versus time plot and its

associated hyperbolic plot for Test 3DK-3 are shown in Fig. 4.1 and 4.2 respectively.

The reconsolidation is considered sufficient when the soil reached at least an average

consolidation ratio of 95%.

4.1.2 Soil Characterization

Soil characterisation tests are important to establish the conditions of the soil sample

prior to the tests and to provide verification of the intended parametric configuration.

In the initial stage of the experiments, characterization tests on soil were conducted

immediately after completion of model testing at 1g. It should be recognised that

characterization tests at 1g may not reflect the actual initial condition of the test due to

soil swelling upon spun down of the centrifuge.

Much of the experimental works described in this thesis were conducted before the in-

flight CPT and T-bar apparatus were introduced in the National University of Singapore

(NUS) centrifuge. Hence, at that time, as an interim solution, in a later stage of the

experiments, a controlled soil sample, with exactly identical preparation procedures as

the actual test is used solely for the purpose of soil characterisation. As no

experimental set-up is attached to this control sample, intensive characterization tests

49
Chapter 4 Experimental Results And Discussions

can be carried on this control model immediately after it is spun down to provide a more

accurate picture of the initial condition. Two soil properties were measured.

The soil characterisation tests carried out in this series of centrifuge tests are:

a) water content test to determine the water content profile with depth and

b) 1-g laboratory vane shear test to estimate undrained shear strength (Cu) profile

with depth.

These tests are not as good as in-flight CPT and T-bar tests as the soil would swell

during spinning down and tests carried out in 1-g would not represent the in-flight soil

parameters.

A minimum of three tests were conducted at the same depth and the average water

content and Cu values were taken. The vane shear method was used to measure the

undisturbed shear strength of very soft, fine grain soil (Bowles, 1996) in the present

study. A typical Cu and density profile (estimated from the water content test) plot of

the soil samples are shown in Figs. 4.3. As expected, the Cu of the model ground

increases with depth, from 7 kPa to 27 kPa from 4m to 24m depth below ground level.

Ladd and Foott (1974) and Trak et al. (1980) proposed that the Cu of a normally

consolidated or lightly over consolidated clay could be estimated by the relationship:

Cu
= 0.22 (4.1)
P' v

where P’v is the effective vertical stress of soil. The Cu versus depth profile estimated

based on Equation (4.1) was also plotted on Fig. 4.3 for comparison, and found to be

generally in agreement with the Cu profile measured by vane shear tests.

50
Chapter 4 Experimental Results And Discussions

4.1.3 Prototype Modelled and Other Consideration

The dimension and set-up of the centrifuge model container was discussed in Section

3.1 (Chapter 3). It is to note that at the edge of the model retaining wall (refer to Fig.

4.4), the edges were allowed to slide and rotate slightly against the boundary wall.

As such, this modelling is not a true modelling of a quadrant of a bi-symmetric

excavation, where the edges would then have zero slope at the mid-points. As the main

focus of this study is to investigate an excavation about a corner, to provide a better

understanding of how the presence of a corner affects the behaviour away from it, this

limitation at the edges is not deemed to be an error. Because of the complexity of this

class of problems, this test in a centrifuge is, to author’s awareness the first in the world.

Thus the understanding arrived at is a crucial first step towards a better understanding of

the behaviour of a three dimensional (3D) excavation.

The 3D model retaining wall is made of two flat aluminium alloy plates butt-welded

perpendicular to each other. One of the concerns of this L-shape wall is that it may

result in a slight rotation during excavation with the pivot at the corner (Fig. 4.4). As

the wall used is thin, this rotation, if any, is unlikely to be significant. To check this, in

one of the tests, Test 3DK-2, a LVDT is installed on another plane of the retaining wall.

The results of LVDTs installed at difference planes are shown in Fig. 4.5. This plot

shows that the lateral wall movement is largely symmetrical, and very little rotation has

taken place. Thus to optimise the layout of the instrumentation and to prevent

cluttering of the space above the ground surface, the displacement transducers in the

experiments were installed at the plane parallel to the perspex window only. This

minimise the numbers of such transducers needed while providing space for installation

of other transducers.

51
Chapter 4 Experimental Results And Discussions

4.2 GENERAL CHARACTERISTICS OF 3D EXCAVATION BEHAVIOUR

Excavation model tests including in-flight excavation in a centrifuge is difficult,

complex and very time consuming. Thus careful planning of the tests to be conducted

is essential so as to minimise the number of tests needed, while ensuring adequate data

were collected on the various aspects associated with deep excavation. Key issues that

need to be considered is this specific research include the difference between plane-

strain (basis of most finite element method analysis) and three-dimensional (3D), and

stiffness of the wall (vary through use of different thickness). In all the 3D cases

except for one, the effect of the presence of a ground capping beam is included through

the provision of a protruded wall.

These tests, their codes and the meaning of the alphabets used in the code are given in

Table 4.1. A chart showing how each test is placed in relation to other conducted is

shown in Fig. 4.6.

In the discussion in this section, the 3D in-flight excavation test with a 2mm thick

model retaining wall, with normally consolidated Kaolin clay as soil model, test label

3DK-2c is used as a basis to provide a general description of the observed behaviour in

a typical excavation around a corner.

The locations of the linear variable displacement transducers (LVDT) to measure the

surface settlements, vertical and horizontal strain gages (SG) to measure the bending

moment of the model retaining wall, and total stress transducers (TST) installed on the

model retaining wall are shown in Figs. 4.7a and b. As can be observed, a

52
Chapter 4 Experimental Results And Discussions

comprehensive instrumentation plan was devised, to optimise the efforts needed for

such a complex test.

4.2.1 Characteristics of Surface Settlement and Wall deflection

The surface settlements at various locations behind the retaining wall during the

progress of excavation are shown in Fig. 4.8. This plot of surface settlement responses

is typical of all the 3D tests conducted in this study. The surface settlement profiles

behind the model retaining wall at 3.5m, 9.1m and 14.5m distances from the corner at

various stages of excavation are shown in Figs. 4.9a to c. The graphs clearly shows

that along a line perpendicular to the wall, the surface settlement is a maximum near the

retaining wall (nearest point measured was 3m behind the wall), and reduces rapidly

away from the wall. To give a broad idea on how this settlement compared to general

ballpark figure and trend, the settlement results at 2m and 4m excavation are plotted

together with 2D tests results carried out in this study (Figs. 4.10a and b). Results

provided by Peck (1969) based on a braced retaining wall system was also plotted on

the same graph as a yardstick. The graphs show that the settlements in 3D tests are

significantly smaller than those from 2D tests. It is also shown that the figure proposed

by Peck (1969) is much smaller than both the 2D and 3D results captured in the

experiments conducted. This is because the experimental results are for unbraced

excavation, while Peck’s results are based on braced excavation.

The surface settlement readings presented above are plotted against distance from the

corner, as shown in Fig. 4.11. Within a distance of 9m from the corner, it is shown that

surface settlements both at 3m and 7m behind wall section decrease as distance from the

corner reduces. Beyond 9m from the corner, the ground settlement becomes almost

53
Chapter 4 Experimental Results And Discussions

constant. It is of general perception that the corner effect would extend to only a certain

distance from corner. Hence, the fact that the surface settlements observed in this test

become constant at 9m from corner raised a question whether the presence of a corner

has ceased to be felt beyond 9m. If not, what does this nearly “constant movement

zone” means and how is it related to 2D condition? This issue will be dealt in greater

depth in Chapter 5.

Visual inspection of the retaining wall after the tests revealed that in this particular test

with a 2mm thick model wall, the wall had yielded and deformed permanently. The

yielding point was observed to be at a section about 7 to 10m distance away from the

corner. Hence, it is likely that the “constant movement zone” observed above is caused

by the yielding of the plate, and it is quite clear that its behaviour is not the same as that

in a plane strain condition, where the latter would not have any lateral interaction at all.

As the boundary walls were greased extensively (Section 3.2.3) as per a proven method

proposed by Khoo et al. (1994), the friction against the container boundary wall is

unlikely to caused the yielding. Hence, it is interesting to investigate how the lateral

yielding occurs in such corner excavation and how it influences the excavation

behaviour. Also, it is crucial to investigate the behaviour under such a condition, as

compared with 3D excavation with an elastic wall and 2D excavation under a plane

strain condition. Further discussion on this important topic is presented in the coming

section on comparative study with two-dimensional (2D) tests.

What is also interesting is the fact that at 3m away from the wall, the settlement profile

changes dramatically away from the corner, but at 7m away from the wall, the changes

is very minimal except for the case of excavation to 5m depth. What is significant about

54
Chapter 4 Experimental Results And Discussions

this observation is that the 3D effect, as represented by the corner is very significant

when near field (relative to the corner) settlement is examined, but not as significant

when far field (relative to the corner) settlement is concerned. In the case of an

unbraced excavation as in this test, up to 5m-excavation, at a distance of 7m away, this

corner effect is already not a major consideration.

For better illustration of the overall surface settlement profile, the above surface

settlement readings are plotted as contours (Fig. 4.12) using the commercial software

Surfer (Golden Software, Inc, 1997). The contour lines are interpolated from the

measured readings. A 6m reference line behind wall is also plotted for easy reference.

It is shown that at 2m-excavation, beyond a distance of 5m from the wall, the surface

settlement is virtually independent of distances from corner, with the settlement contour

nearly parallel to the retaining wall. The maximum settlement is about 0.18m. Even at

this early stage, the presence of a corner in an excavation begins to have different effect

on different regions behind the retaining wall. Close to the wall, the impact is felt, with

the settlement reducing from 0.18m furthest away to about 0.11m at 4m away and

presumably reducing to even smaller value when very near to the corner.

At 4m-excavation, the surface settlement contours become denser, an indication that

changes both along and away from the wall have become more pronounced. Clearly,

the extent of the effect of corner also spreads. Up to a distance of 6m away from the

wall, the settlement increases rapidly as distance from the corner increases. However,

beyond 8m, again the contour is nearly parallel to the wall, indicating that the presence

of the corner is not felt here.

55
Chapter 4 Experimental Results And Discussions

At 7m-excavation, the surface settlement contours clearly indicate that the effect of

corner is now felt everywhere. The massive settlement observed means that the

excavation has failed. Thus what these three figures has shown is that the effect of

corner is also very much related to the depth of excavation, besides its distance away

from the corner. Though the presence of a corner is felt instantly, the extent of this

impact increases with depth of excavation. This is a logical development. With

increasing excavation, the imbalance in load increases while the overall stiffness of the

entire support system decreases. Under such a situation, any stiffening, such as the

presence of a corner, will be called more and more into effect. The relationship of the

extent of corner effect to depth of excavation is investigated in more detail in Chapter 5.

By using the still image capturing and processing technique, the wall deflection profiles

of the model retaining wall abutting the perspex window of the test are obtained (Fig.

4.13). The graph shows that the deflection of the wall looks very similar to that for a

typical cantilever wall in a 2D plane strain excavation. At 18m depth, the deflection is

very small. The soil in this test is soft normally consolidated clay with undrained shear

strength (Cu) ranges from 7 kPa to 27 kPa at the toe of wall (Fig. 4.3). Even though

the wall is “floating” within this soft clay, the wall deflection near the toe is near zero.

This indicates that the wall embedment is sufficient. Hence, the toe of the wall could

be considered approximately as a fixed end, this particular observation will be used later

in Chapter 5 for derivation of an approximate solution to calculate wall deflection under

certain loading and boundary conditions.

56
Chapter 4 Experimental Results And Discussions

Due to limitations of the imaging systems at the time when these tests were conducted,

the accuracy is estimated to be about 1mm (equivalent to 0.1m prototype). This wall

deflection data is not accurate enough for use in the mechanics study.

As discussed in Chapter 2, a key advantage in centrifuge modelling compared to actual

field studies is the control over model parameters, soil profiles and properties, and

excavation sequences. This advantage enables more in-depth parametric studies be

carried out to obtain better insight into the mechanics. One particularly significant

advantage is the ability to study the movement induced during each excavation stage,

from the time the scraper starts scrapping the soil (scraper nearest to retaining wall), to

the time the scraper finish the scrapping (150mm from retaining wall). Fig. 4.14 shows

a schematic sketch of this typical scraping process. In this scrapping stage, only the

soil response to the scrapping is captured. What is very useful about this is that it

allows a more direct cross comparison of various parameters of interest at each and

every stage of excavation. When accumulated values are used and if the construction

process is not neatly sequenced, as is the case with most field studies, it is difficult to

compare the results from different parametric studies as any induced movement is due

to increase in imbalance of loads and resulting decrease in overall system’s stiffness,

both due to removal of soil on the passive side of the retaining wall. The principal

reason is that after the first stage of excavation, it is very difficult to apportion the

subsequent accumulated measurements to these different effects. The use of

incremental results, due solely to the excavation process at each depth, will allow a

comparison for example of the composite overall system’s stiffness at that depth of

excavation as a result of changes to certain parameters in the test (at the same depth of

excavation, there is same amount of load imbalance). Such incremental behaviour is

57
Chapter 4 Experimental Results And Discussions

possible only in centrifuge tests equipped with appropriate instrumentation. This

method was used by fellow researcher (Thanadol 2002) in deep excavation related

studies to establish the mechanism of embedded soil berm in an excavation.

The incremental surface settlement of a single scrapping operation at various stages of

excavation for the series of LVDTs at 7m behind wall and various distance to the corner

(x) are shown in Figs. 4.15a to c. The excavation from 0.5m to 1.0m, 1.5m to 2.0m,

2.5m to 3.0m and 3.5 to 4.0m were denote as 1m exc, 2m exc, 3m exc, 4m exc and 5m

exc respectively in the graphs. The graphs shown that the settlement developed were

larger at the initial stage of scrapping, and the rate of increase in settlement gradually

slowed down when the scraper moves away from the wall. The graphs also show that

the magnitude of incremental settlement increases according to the increase in

excavation depth. For instance, in Fig. 5.15a which is x = 3.5m, it was shown that the

settlement at the end of scrapping (ultimate incremental settlement) gradually increased

from about 0.0015m when excavating from to 1m exc, to about 0.0080m at 5m exc.

LVDTs at x = 9.1m and 14.5m (Figs 5.15b and c) show similar trend, except that the

magnitude of incremental settlement were consistently larger than that at x = 3.5m.

This increase of incremental settlement according to the increase of excavation depth is

expected as excavation proceeds, the imbalance in load increases while the overall

stiffness of the entire support system decreases hence widen the net imbalance. To

highlight the different behaviour at various sections from corner, the incremental

settlement at at x = 3.5m, 9.1m and 14.5m, at the same stage of excavation depth were

plotted (Fig. 4.16). In these graphs, the settlements were normalized by dividing it with

the ultimate incremental settlement at the section x = 14.5m (furthest from corner), for

better comparison. The normalization factors are shown in the graphs. From the graphs

58
Chapter 4 Experimental Results And Discussions

(Figs. 4.16 a to d), which are for 1m to 4m exc, it was shown that the incremental

settlement at the section nearest to corner, that is, at x = 3.5m is always smaller than that

at sections x = 9.1m and 14.5m. The ultimate incremental settlement at excavation 1m

to 4m exc at x = 3.5m are always about 60% of that at x = 9.1m and 14.5m. This is

expected as the corner restraint are more pronounced nearer to corner, hence its

influence to the composite overall stiffness at section nearer to corner are larger. What

is more interesting is the fact that the incremental settlements at sections x = 9.1m and x

= 14.5m were almost identical. This suggests that the resultant “composite overall

system stiffness” at these sections are quite similar, and the corner effects felt at these

sections are about the same. To better illustrate the characteristics of incremental

settlement with the development of excavation depth, the ultimate incremental

settlement are plotted against the excavation stages (Fig. 4.17). The graph shows that

the incremental movement at each stage of scrapping is increasing almost linearly with

depth of excavation, indicating a linear net difference between the increase in imbalance

load and the decrease in system composite stiffness. The above presentation of

incremental movement study is means for general appreciation of this method. More

investigations using this method on parametric studies to carry out more in-depth

investigation on factors influencing 3D corner effect are carried out. This is discussed

in greater detail in the subsequent sections.

4.2.2 Characteristics of Lateral Earth Pressure on Retaining Wall

The lateral earth pressures on the model retaining wall are measured by total stress

transducers (TST) installed on the front (passive) and rear (active) sides of the wall

(refer Section 3.2.1 for detail). The cumulative changes of the lateral earth pressures

corresponding to different depths of excavation are shown in Fig. 4.18. This shows that

59
Chapter 4 Experimental Results And Discussions

the lateral earth pressures on both the active and passive sides decrease with the

progress of excavation. The decrease of lateral earth pressures at the passive side is

largely attributed to the loss of earth in front of the excavation. On the other hand, as

the soil behind the wall is retained, the decrease of lateral earth pressures on the active

side is largely attributed to the mechanical soil movement, reducing the stress state from

a K o condition to K a − active (active state) condition.

To better reflect the lateral earth pressures (P) response due to ground movement, the

P
coefficient of lateral earth pressure, K = with depth of excavation is plotted (Fig.
γy

4.19a). The total stress K term is the total earth pressure measured in the experiment

and includes water pressure. γ is the total unit weight of the soil, which is about 16.0

kN/m3 (Fig. 4.3). At the active side, y is taken as the depth of the sensor from the

original ground level before excavation. At the passive side, y is taken as the depth of

the sensor from the excavated formation level.

It is observed from Fig. 4.19a that the development of passive earth pressures at 8m and

15m from the corner, measured by TST-X13 and X11 respectively, are almost identical.

On the active side, the decrease of active earth pressure at 8m away from the corner

(TST-X12), is smaller than that at 15m away (TST-X10). This suggests that the 3D

effect is not the same on the active and passive side.

To examine approximately whether the earth pressures have reach limit states, the earth

pressures were checked against Rankine’s theory. Rankine’s theory proposed that for

cohesive soil, the active and passive earth pressures are:

60
Chapter 4 Experimental Results And Discussions

Pactive = γ y – 2Cu (4.2)

Ppassive = γ y + 2Cu (4.3)

Based on Equations (4.2) and (4.3) above, the (P - γ y)/2Cu value shall reach ±1.0 in the

limit states. This approach was used by Tanaka (1994) and Kimura et al. (1993) in

their study to characterize earth pressure limit states in excavation.

In the current study, (P - γ y)/2Cu were calculated and plotted against the depth of

excavation (Fig. 4.19b). The terms P, γ and y were explained in the previous paragraphs.

The undrained shear strength, Cu is taken at the TST levels (8m below ground), and

from the soil characterisation tests described in the earlier Section 4.1.2 (Fig. 4.3), the

Cu is 15kPa. From the graph, it is shown that the active state is achieved at 8m (TST-

X12) and 15m (TST – X10) from corner, at excavation depth of 1.5m and 4.5m

respectively. The passive state is achieved at a relatively later stage, where both TSTs

(TST – X13 and TST – X11) at 8m and 15m from corner, achieved passive state at the

end of 5m-excavation. This confirms that the active state is achieved at smaller strain

than the passive state, consistent with the general understanding. The active and passive

earth pressures are generating imbalance load in an excavation. This imbalance load is

restrained by the wall effect. Hence, for a better understanding on excavation

behaviour, the earth pressure and the wall effects must be investigated further for a

better understanding on excavation. This is discussed further in Chapter 5.

The above results present an overview of 3D excavation behaviour and its salient

features that can only be captured from properly conducted centrifuge-modelling test.

In the following sections, parametric studies were carried out to specifically bring out

major characteristics of a corner excavation to allow in depth probing of the mechanics.

61
Chapter 4 Experimental Results And Discussions

4.3 2D VERSUS 3D EXCAVATION BEHAVIOUR

One of the main objectives of the current study is to investigate experimentally the

differences in 2D and 3D-excavation behaviour. To achieve this objective, in addition

to 3D tests, a series of 2D tests were also conducted. These 2D tests were carried out

with the same in-flight excavator, but on a modified set-up which facilitates a 2D

excavation (see Chapter 3). The 2D tests were carried out on NC Kaolin clay soil

model, using 1mm and 3mm thick model retaining wall. The tests were labelled 2DK-1

and 2DK-3 respectively.

The surface settlement profiles behind the wall during excavation for 2D tests, 2DK-1

and 2DK-3 were plotted together with results of 3D test on wall without capping beam

(3DK-2c). If a 3D excavation is sufficiently large, far enough from the corner, the

behaviour should approach a 2D problem. Thus, for the 3D test, readings of LVDTs

furthest away from the corner, that is, 14.5m from the corner were chosen for

comparison (Fig. 4.20). It is shown that even at a distance of 14.5m away from the

corner, the settlement of the 3D test is significantly less than that of 2D tests. Thus,

this means that at that distance away from the corner, and for a relatively thin wall, the

behaviour is still not 2-D. This also suggests that the “constant movement zone”

observed earlier cannot be due to the cessation of corner effect.

The wall deflection profiles at the perspex window (17m from corner) captured using

image processor for tests described above are shown in Fig. 4.21. In the plot, the wall

deflection of the graph is divided with a Normalization Factor (NF), which is the wall

displacement at ground level, to better illustrate the deflection trend. It is noted that the

62
Chapter 4 Experimental Results And Discussions

NF of 2D tests are a few times larger than that of 3D tests, indicating that the effect of a

corner to a 3D test at the distance of 17m is still very significant. It is shown that at the

mid depth, there is significant different in wall deflection profiles between 2D and 3D

Tests. In the graph, it is also shown that the deflection profiles in the 3D tests

extended deep below the formation level, up to about 18m deep from the ground level.

This is very much different from that in 2D tests, where the wall deflections have

mostly attenuated at a depth of about 8m from the ground level. The principal reason

for this is that in the 3D test, the deformation is due to a set of perpendicularly joined

plates. Thus in balancing the removal of soil in the passive side, the whole plate

mechanism is mobilised, unlike the 2D case where only deflection of a beam in a

vertical plane is used. This is crucial to the understanding of the characteristics of

corner excavation, and will be discussed in more detail in Chapter 5.

To allow more in-depth investigation into the mechanics, the incremental settlement

plot (Section 4.2.1) are carried out for the 2D and 3D tests discussed above. The

incremental settlement at various distances from corner, x = 3.5m, 9.1m and 14.5m were

shown in Figs. 4.22a, b and c respectively. In these graph, the incremental settlement

at 3 different stages, namely the excavation from 0.5m to 1.0m (1m-excavation), 1.5m

to 2.0m (2m-excavation), and 2.5m to 3.0m (3m-excavation), are presented. In these

graphs, the incremental settlements were divided by normalized values to highlight the

trend. The normalization factors were denoting as NF2D and NF3D for 2D and 3D test

respectively. For the 2D test, the normalization factors were taken as the ultimate

settlement (at end of scraping) for the corresponding stage of excavation. For 3D test,

the normalization factors were taken as the ultimate settlement at x = 14.5m (furthest

from corner), for the corresponding stage of excavation.

63
Chapter 4 Experimental Results And Discussions

Fig. 4.22a shows that at the section nearest to corner, x = 3.5m, for up to 3m-excavation,

the 3D test (with and without wall protrusion) show movement largely developed in the

initial stage of scraping, and the movement apparently reach equilibrium earlier than

that of 2D test. In this excavation, the immediate scrapping of soil at the passive side

would create an immediate imbalance in the system, which the soil on the passive side

and the retaining wall must respond to. In the current study, as the retaining wall was

“floating” in the soft clay with undrained shear strength ranging from 17 kPa to 27 kPa,

it s expected that the soil behaviour would be more prominent in equilibrate the

imbalance. However, as the soils needs time to adjust to reach new equilibrium, it is

hence expected that there would be time lapsed after the soil immediately in front of the

retaining wall is scraped, for the movement to stabilize. In 3D test, there is another

response which would help to equilibrate the imbalance, that is, the lateral restrained of

the wall developed from the corner. As the structural (wall) response to load is much

faster than that of soil, it is hence logical that the movement at 3D test would reach

equilibrium faster than that of 2D test. It is also observed that for 3D test at sections

further from corner, namely at x = 9.1m and 14.5m, Figs. 4.22b and c show that the

graphs are almost identical to that of 2D test. This indicates that the lateral restrain

effect decreases with increasing distance from the corner, as expected.

After each stage of excavation, soil would undergo consolidation. Hence, except the

very first stage of excavation, all incremental settlement would be the results of

immediate response of soil to the excavation process, plus the consolidation effect. At

7m behind wall, it is expected that the consolidation settlement contribute significantly

to the total settlement caused by immediate scraping of soil. Hence, to capture the

64
Chapter 4 Experimental Results And Discussions

immediate behaviour of the excavation system solely due to the excavation process, the

response of wall provides a better gauge. In this study, all the tests carried out in this

study were equipped with horizontal LVDTs mounted on wall top to monitor its

displacement, except Test 3DK-2c. Test 3DK-2c has no wall protrusion as it was

intended to simulate 3D wall without capping beam. The incremental wall top

displacement of 2D and 3D tests with capping beam, at section x = 9.1m, at 4 different

stages of excavation are plotted in Fig. 4.23a to d. Again, the incremental movements

were divided by normalized factors to highlight the trend. The normalization factors

were denoting as NF2D and NF3D for 2D and 3D test respectively. For the 2D test, the

normalization factors were taken as the ultimate settlement (at end of scraping) for the

corresponding stage of excavation. For 3D test, the normalization factors were taken as

the ultimate settlement for test 3DK-1 (thinnest wall), at x = 14.5m, for the

corresponding stage of excavation. It is shown very clearly in these graphs that at all

excavation stages, the lateral wall top displacement responded to the scraping

immediately, and reaches the ultimate movement almost immediately after the soil in

front of the retaining wall was excavated (scraper at 30m from wall). This is a clear

contrast to that of 2D test where the ultimate movement occur only when the scrapping

was almost completed (scraper at 140mm from the wall). This result confirms the

observation made in the previous paragraph that the wall restraint due to the 3D effect

helps substantially in equilibrate the imbalance, and the response to the imbalance was

much faster.

It is also noted that for all the 3D tests, while the respond time to achieve stabilization

are the same regardless of the wall thickness, the magnitude of ultimate movement of

thicker wall are consistently smaller than that of thinner wall. This shows that in these

65
Chapter 4 Experimental Results And Discussions

series of 3D tests, the wall stiffness highly influences the composite system stiffness of

the excavation system. Correspondingly, the effect of soil parameters was suppressed.

Lee et al. (1995) reported that unstrutted excavation in soft soils could exhibit very

strong 3D effects due to corner restraint. The importance of such effects causes the 2D

analyses to be much more sensitive to soil parameters than the 3D analyses. This is in

agreement with the above experimental findings. More detail study on the wall

thickness effect on 3D behaviour of an excavation will be presented in the subsequent

sections.

Ou et al. (1996) and Lin et al. (2003) have used the concept of plane strain ratio (PSR)

to describe the effect of a corner (Section 2.3.2). In the following discussion, the idea

of PSR is adopted to compare the 3D surface settlement at various distances from corner

with that of 2D results.

As no 2D test is carried out with exactly the same thickness (2mm) as that of the 3D

test, the Test 2DK-3 (3mm wall) is selected for comparison. This test is selected

instead of Test 2DK-1 (1mm wall) because the smaller movement in this test would not

exaggerate the difference between 2D and 3D results. The surface settlements of 3D

test (3DK-2c) at various distances from corner (x) are divided by the surface settlement

of 2D test (2DK-3), at the same stage of excavation to give PSR-sett. The surface

settlements at 7m behind wall are used for this comparison. The PSR-sett against x plot

at every stages of excavation is as shown in Fig. 4.24. It is shown that the PSR-sett of

the 3D test is within 0.12 to 0.27 throughout the wall length, indicating a very

significant corner effect. In addition, the small PSR-sett value clearly indicates that a

constant settlement profile away from the corner, caused by yielding of the wall, does

66
Chapter 4 Experimental Results And Discussions

not imply to the cessation of 3D effect. Also, the above phenomena also indicate that

the excavation in these 3D experiments is not large enough.

An immediate concern thus pops up from these observations: In the field, all excavation

is 3D in nature. Thus how large should an excavation be for 2D condition to occur and

3D (or corner effect) need not be considered? This is also discussed in Chapter 5.

4.4 RETAINING WALL THICKNESS EFFECT IN A CORNER EXCAVATION

The stiffness of retaining wall is an important factor controlling the behaviour of an

excavation. In this study, centrifuge tests were carried out to evaluate the effect of wall

stiffness in a 3D excavation. In this series of tests, except for the different model wall

thickness used, all other procedures for preparation of the models were identical. The

model retaining walls were made of aluminium alloy with thickness of 1mm, 2mm and

3mm. The stiffness of these walls are tabulated and compared with the equivalent

prototype concrete wall thickness in Table 4.1. All the walls are relatively flexible as

compared to those used in practice. The tests identification labels were 3DK-1, 3DK-2

and 3DK-3 respectively. Please refer to the summary table (Table 4.1) presented in

Section 4.3 above for description of the labels.

The broad trends of surface settlement and lateral deflection of wall top in all the three

tests are shown in Figs. 4.25 and 4.26 respectively. The results were fairly consistent,

and more importantly, indicate that the overall behaviours namely the response with

increasing excavation and response away from the corner were highly dependent on the

stiffness of the retaining wall. One useful observation is both the similarity in the

67
Chapter 4 Experimental Results And Discussions

responses and their consistency. This is highly relevant to subsequent effort to explain

the behaviours.

The lateral displacements of wall top at various distances from the corner in these three

tests are shown in Fig. 4.27. In this plot, the maximum lateral displacement at x =

15.2m for the case when a 1mm plate is used, 0.3064m, is used to normalize the other

results. For the case with 2mm retaining wall, the wall top displacement decreases

almost linearly towards the corner. For 3mm retaining wall, only the measurement at

9.8m from corner is captured due to a technical glitch and hence the movement profile

cannot be established. This shortcoming of data is complemented by surface settlement

measurements, which will be discussed subsequently. For 1mm retaining wall, from

the corner to 10m away, the wall top displacement increases quite rapidly. Beyond

10m from the corner, there is a change in the wall displacement profiles, and the

displacement becomes nearly constant with distance from the corner. The surface

settlement results are further examined to better appreciate this observation.

The surface settlement profiles at 7m behind the retaining wall, at various distances

from corner of these tests are shown in Fig. 4.28. In this plot, the maximum surface

settlement at x = 14.5m for the case of 1mm plate, 0.1223m, is used for normalization.

All the surface settlement profiles show a slight reduction as they approach the corner

except for the case of 1mm plate with 4m-excavation. Overall, the results indicate that

at this distance away from the excavation (7m from retaining wall), the effect of the

corner is not distinctly felt. This does not mean there is no effect of the corner, that the

behaviour is 2-D. Instead, what this means is that at that distance away, the effect is

averaged out, an idea similar to St Venant’s Principle in stress concentration. The

68
Chapter 4 Experimental Results And Discussions

characteristics of surface settlement can be seen more clearly in the surface settlement

contours of the tests (Fig. 4.29). As there were some settlements at the corner, the

settlement results were offset with the settlement at the corner to highlight the

differences with distance from the corner. In this exercise, the settlement at the corner

is taken as the intercept of surface settlement at the Y-Axis, as shown in a typical plot in

Fig. 4.30a. The offset-surface settlements versus distance from corner plots for the

tests described above are shown in Figs. 4.30b to d. These graphs show identical trend

as per the wall top displacement plots, where the 2mm and 3mm walls show an almost

linear decreasing of settlement towards the corner, while the 1mm wall show changed in

movement profile and developed a “constant movement zone” at 10m from corner.

Visual inspection of the retaining wall after the tests revealed that the 1mm wall has

yielded and deformed permanently at a section about 8m to 10m distance away from the

corner. There are no sign of yielding observed for the 2mm and 3mm walls. Hence,

the pronounced change in the movement profile above is caused largely by the yielding

of the plate. This phenomenon was also observed in Test 3DK-2c (Section 4.2.1).

Hence, it is clear that the yielding of the plate, which causes the changing shape profile

and “constant movement zone”, are also dependent on the stiffness of the plate and also

the associated flexural capacity. Clearly, this yielding in the wall and the resulting

change in the movement and settlement profiles significantly affect the behaviour of an

excavation around a corner. A more detail study on this issue will be carried out in

Chapter 5.

69
Chapter 4 Experimental Results And Discussions

4.5 PRESENCE OF A CAPPING BEAM IN A 3D EXCAVATION

In the field, a capping beam is usually constructed to bind panels of wall, or contiguous

pile together such that the panel/pile can act together as a contiguous retaining wall.

The capping beam is usually designed as a continuous beam, with stiffness (thickness of

beam) much larger than the wall. In typical design, usually only 2D analyses were

carried out and the effect of this capping beam is not captured as the reinforcement

coming from a capping beam derived mainly from its lateral stiffness and connection to

a nearly stationary corner. An illustration showing the configuration of capping beam

at site and the general assumption made in 2D analysis is shown in Fig. 4.31. In a real

excavation, this could be an important effect, especially during the first stage of

excavation, which is usually carried out without any bracing provided. However, as

the effect is truly due to a corner, to the author’s awareness, this aspect has not received

any attention before. In this study, an in-flight excavation test, labelled 3DK-2, with

identical set-up as for test 3DK-2c, is carried out to study the stiffening effect due to the

presence of a capping beam in 3D excavation. In this study, a monolithic extension to

the wall (model retaining wall protruding from the ground) is used to approximately

simulate the additional stiffening due to a capping beam. For both Tests 3DK-2 (with

capping beam) and 3DK-2c (without capping beam), the retaining walls were made of

2mm thick aluminium alloy, equivalent to a concrete wall of 287mm thick. Hence, a

protrusion of 60mm from ground surface, used for 3DK-2, will approximately model a

0.52m thick x 1m height concrete capping beam. This is within the practical range of

capping beam on site.

Comparison of surface settlement profiles of Tests 3DK-2c and 3DK-2 are presented in

Fig. 4.32a. Clearly in the test with capping beam (3DK-2) the ground settles

70
Chapter 4 Experimental Results And Discussions

significantly less than that for the case without capping beam (3DK-2c) illustrating the

importance of accounting this effect. To highlight the effect of capping beam with the

depth of excavation, settlement at 7m behind wall at various distances from corner are

plotted against various stages excavation (Fig. 4.32b). Further, the ratio in terms of

settlements of Test 3DK-2 (Sett (cap)) divided by settlement of Test 3DK-2c (Sett

(uncap)), versus depth of excavation graph was also plotted (Fig. 4.32c). From the

graph, it is observed that for the first 2m of excavation, the settlements from Tests 3DK-

2 were within 20% to 40% of that for Test 3DK-2c. But beyond 2m-excavation, the

effectiveness of the capping beam as an additional reinforcement start to decrease

significantly. The reduction is almost linear with depth of excavation. At 4.5m-

excavation, the ratio of Sett (cap)/Sett (uncap) was within 50 to 70%. What is also

interesting is that the effect is less important when very near to the corner. It is

important to recognise that this observation shows that with capping beam, the

behaviour is akin to the provision of a bracing at the top of the wall. This “bracing” is

effective to restrain movement in the early stage of an excavation. When excavation

becomes deeper, the vertical spacing between this bracing and formation level widens,

and this bracing effect reduces. This has important practical implications, as almost all

analyses will show that the most significant deflection in a retaining wall is induced

during the initial stage of excavation when conventional bracing cannot be provided.

In this case, the presence of the capping beam, if any, must be included.

The wall deflection profiles at the edge of retaining wall (17m from corner) captured

from image processing for Tests 3DK-2 and 3DK-2c are plotted on Fig. 4.33. The

graphs show that the deflection profiles of the two tests are very different. For the test

3DK-2c, the deflection profiles show a cantilever profile, similar to that of a typical 2D

71
Chapter 4 Experimental Results And Discussions

unbraced cantilever wall profile and the profile expected during the early stages of an

excavation when conventional spacing is not available. For Test 3DK-2, the deflection

profiles of the wall show the top being restrained in a manner akin to that of a wall

provided with a top bracing. This further support the assertion that the capping beam

provides reinforcement effect similar to that provided by a top brace. Further

investigation on the effect of capping beam on 3D corner effect is discussed in more

detail in Chapter 5.

4.6 EFFECT OF SOIL STRENGTH IN 3D EXCAVATION

As reviewed in Section 2.3.4, soil strength of an excavation is reported to influence

corner effect. To investigate the contribution of soil strength, tests with stiff highly

OC (Test 3DK-2o) and relatively soft NC (Test 3DK-2) Kaolin clay soil models are

examined. Though this may not be adequate to establish fully how the soil contributes

in an excavation around a corner, it does provide basic first insight into the soil effect in

this interesting topic. The preparation of these two soil samples was discussed in

Section 3.2.2. The undrained shear strength (Cu) profiles determined using the 1-g vane

shear method (discussed in Section 4.1.2) of these two soil samples are shown in Fig.

4.34.

The lateral displacements at the top of the wall at various distances from corner for the

above tests were shown in Fig. 4.35a. It is shown that the OC clay test generally

having the same trend as that for NC clay test, but with smaller magnitude. To

highlight the influence of corner, the ratio of displacement with distance from corner is

plotted (Fig. 4.35b). The displacements at the measurement point furthest from corner,

that is, at x = 15.2m, are used as normalization factors (NF) to divide the displacements

72
Chapter 4 Experimental Results And Discussions

measured at nearer distances from corner. In this exercise, the lower normalized

displacement ratio reflects larger variation of displacement with distance from corner,

an indication of larger corner effect. It is shown in Fig. 4.35b that except at 2 m-

excavations where the NC tests show large deviation at x = 4.5m to 9.8m, all the plots

are generally showing the same trend and magnitude. This suggests that the soil

strength has no significant effect to the characteristics of corner effect.

To confirm this point, the surface settlement data is examined. The offset-settlement

(settlement relative to settlement at the corner, refer Section 4.4) versus distance from

corner plots was shown in Fig. 4.36a. The graph shows similar trend as that of wall top

displacement plot (Fig. 4.35a). However, it is interesting to note that there are

apparently larger differences in settlement for soft NC clay test than that of stiff OC

clay test. For instance at 6m-excavation, as indicated in the graph (Fig. 4.36a), the

differences of settlement at x = 4.5m and 15.2m were 0.084m and 0.041m for NC clay

and OC clay tests respectively. This shows a clear indication of larger magnitude of

corner effect for the NC clay test. To highlight the influence of corner, the ratio of

settlement with distance from corner is plotted (Fig. 4.36b). The same procedures used

for wall top displacement results discussed earlier were adopted. It is shown in Fig.

4.36b that except at 2m-excavation where the NC clay test show large deviation, all the

plots are generally showing the same trend and magnitude. This shows the

characteristic of corner effect is not sensitive to the soil strength. This finding is

consistent with the wall top displacement results described earlier.

The above results conclude an interesting point; that the more pronounced corner effect

observed in soft ground reported in the field is mainly in terms of magnitude, rather than

73
Chapter 4 Experimental Results And Discussions

relative ratio quoted in the current study. Hence, to estimate the magnitude of corner

effect in an excavation, the soil strength has to be considered.

4.7 SUMMARY

It is to note that thin walls were used in the experiments and large displacements are

captured. This large displacement is not likely to be obtained in the field. This is to

highlight the behaviour of corner effect for theoretical study and not to simulate a field

problem directly.

It is noted that all the experiments were done once due to limited time. It is

acknowledged that it is better to conduct two identical tests at least once to check the

repeatability of the results. However, it is observed that all the experimental data

captured were consistent and some characteristics of corner effects in an excavation

were identified. Firstly, from the 3D tests, it shows that the movement at corner is

smaller than that at a distance away due to the presence of corner restrains effect. This

corner restrain effect is shown to increase with wall stiffness. Secondly, it is also

evident that the presence of a capping beam enhances the wall stiffness, and hence the

corner effect increases accordingly. In Tests 3DK-1 (1mm plate with capping beam)

and 3DK-2c (2mm plate without capping beam), the wall displacement as well as

surface settlement profiles changed shape at 10m and 9m respectively from the corner,

and the movements become constant with distance from the corner. At this “constant

movement” range, the movement is still found to be much lesser than that at 2D

condition. The changing shape is found to be caused by the yielding of the wall.

Hence, it is clear that the “constant movement” range in a 3D excavation does not imply

the cessation of corner effect. The wall yielding and the “changing shape” behaviour

74
Chapter 4 Experimental Results And Discussions

show a fundamental change in the excavation behaviour which has not been studied

before. The influence of wall effect in a corner excavation, incorporating the wall

yielding effect is hence important to be established to gain better understanding on the

behaviour of a corner excavation.

Secondly, from the measurement of lateral earth pressure at different distances from

corner, it was found that the lateral earth pressure varied with distance to the corner.

This is due to the variation of soil strain mobilize in an excavation against the distance

to the corner. As earth pressure causes imbalanced load in an excavation, its

contribution to corner effect is crucial. It was shown that the variation of earth pressure

with distance from corner is increasing with the increase of depth of excavation. Also,

while the soil strength (Cu) has shown to have insignificant effect on the characteristics

of corner effect (in terms of relative ratio), it still would influence the magnitude of

γh
corner effect. Hence, the stability number, S = , where the effect of excavation
Cu

depth (h) and soil strength (Cu) are considered together, is used to correlate with the

variation of earth pressure effect to better represent the “geotechnical effect”

influencing the corner effect of an excavation. This is discussed in Chapter 5.

75
Chapter 5 Fundamental Behaviour of Corner Effect in Excavation

CHAPTER 5

FUNDAMENTAL BEHAVIOUR OF CORNER EFFECT

IN EXCAVATION

5.1 INTRODUCTION

In the earlier chapters, results from experimental studies investigating key issues

concerning corner effect in excavation were presented. Key issues examined include

the plane strain (or 2D) versus 3D behaviour of the retaining system, with emphasis on

the effects of corner, wall thickness and the presence of a capping beam.

Other than the wall stiffness as described above, the experimental studies also show that

the depth of excavation and shear strength of soils will affect the effectiveness of a

corner. These geotechnical issues could be characterized by the stability

γh
number, S = , where the effect of excavation depth (h) and soil strength ( cu ) is
cu

considered together (Peck, 1969).

In this chapter, the above factors are studied in more details to better understand the

mechanics of how a corner in an excavation affects the overall behaviour.

Due to space constraint and the limitation of the 3D in-flight excavator, excavation with

a fixed width and length were conducted in the centrifuge testing stage of this study.

With the current centrifuge at National University of Singapore (NUS), it was not

practical to conduct parametric studies of the size in the centrifuge. Thus, to

complement the understanding on the effect of the size of an excavation and its relation

76
Chapter 5 Fundamental Behaviour of Corner Effect in Excavation

to the presence of a corner, numerical analyses were carried out. As the objective of

the numerical analyses was to carry out parametric studies on excavation dimension to

gain better insight, and not to develop or enhance any numerical program, a reliable and

readily available FE program was used.

5.2 FINITE ELEMENT ANALYSIS

The Finite Element Method (FEM) program, CRISP, was chosen for numerical analyses

in this study. CRISP (CRItical State Program) was first developed by research workers

in the Cambridge University Engineering Department soil mechanics group. This

program has been widely customized and used for the modeling of excavation in 2D

(Powrie 1986; Balasubramanium et al. 1992; Chee 1999; Tan et al. 2001; Sun 2003) and

3D (Liu 1995; Quan 1995; Yong et al. 1997; Lee et al. 1998; Sun 2003) analyses.

Detail information on the overall structure of this program can be found elsewhere (e.g.

Britto and Gunn, 1990, Liu 1995, Lee et al. 1998, Sun 2003).

A typical 3D mesh used for the 3D excavation study is shown in Fig. 5.1. Both the

retaining wall and the surrounding soil are modeled using twenty-node linear strain

brick elements with 8 vertex nodes, with extra pore pressure degrees of freedom for a

consolidation analysis. The mesh has a total number of 10 x 10 x 10 = 1000 elements

with 1331 vertex nodes. In Sun (2003), for a 3D FEM modeling of a similar size

excavation problem, he demonstrated that about 1000 elements are sufficient to reach

convergence when compared with those with much finer FEM mesh.

Two types of material properties were used in this FEM analysis. For the retaining

wall, the property of aluminum alloy used in the experimental studies was modeled

77
Chapter 5 Fundamental Behaviour of Corner Effect in Excavation

using the Mohr-Coulomb elastic perfectly plastic model. The wall is modeled as a

monolithic structure, which is indeed the case in the centrifuge experiments. The soil

was modeled using Modified Cam Clay model where the soil properties were

determined from consolidation tests and the critical state parameter, M, from triaxial

tests. The parameters used in the FEM analyses are as shown in the Table 5.1. These

parameters have been used extensively for a number of similar works, including that by

Thanadol (2002) and Goh (2003) in their 2D FEM analyses.

The boundary conditions imposed for the FEM analyses simulate the condition of the

model test conducted. For 3D analyses, the four vertical side boundaries are constrained

laterally but free to move along the boundary faces. At the base, the mesh is restraint

from movement in all the directions. Drainage was permitted at the top of the soil

model. The drainage conditions for the side and bottom boundaries of the soil

domain are impervious. The initial soil stress condition is the condition after the

centrifuge re-consolidation and before the commencement of excavation. Table 5.2

shows the initial condition simulated in the analyses. In the centrifuge experiments

(Section 3.2.1), the edges of retaining wall abutting the centrifuge container were not

physically connected perpendicularly to the container sidewalls and allowed to slide.

The main reason is technical; it was too complicated and difficult to impose this

condition. As a result of this, the retaining wall might not be always perpendicular to

the container sidewalls during the excavation when it moved. This means that the edge

of the retaining wall may tilt slightly. This is difficult to simulate properly in the FEM

analyses and thus was not modeled. This shortcoming is acknowledged and noted for

the subsequent interpretation of results. In the present study, the main focus is on the

behaviour around the corner and how the effect is extended away from it.

78
Chapter 5 Fundamental Behaviour of Corner Effect in Excavation

The FEM analysis was first checked using results from the centrifuge tests. 3D FEM

analyses were carried out to model tests 3DK-3, 3DK-2, 3DK-1 (all with wall

protrusion) and 3DK-2c (without wall protrusion). The deflection profiles of the

retaining wall in these tests as well as those predicted by the FEM analyses are plotted

in Figs. 5.2 a to d. It is shown that except for Test 3DK-1, the trend and magnitude of

the deflection profiles obtained experimentally for all the tests were captured reasonably

well by the FEM analyses. The large discrepancy between experimental and FEM

results for Test 3DK-1 is likely due to the yielding of the thin wall in the experiment

and the deficiency of FEM analyses to model this yielding, even though the wall was

modeled as an elastic perfectly plastic wall, with an appropriate yield strength specified.

From the above study, it is found that FEM analyses are able to produce the right trend

of behaviour of the excavation. Hence, the FEM program together with the associated

soil parameters, initial and boundary conditions are adopted for the subsequent

parametric study. Where appropriate, results will also be used to provide a deeper

appreciation of the adequacy of the FEM analysis.

It is to note that some preliminary 2D FEM analyses using the same program, soil

parameters, initial and boundary conditions were carried out with an intention to

compare it with the 3D FEM results. However, the 2D FEM model predicted large

displacement, and the model becomes unstable after about 1.5m excavation, due to the

fact that the excavation is in soft clay and is un-braced. Hence, a comparison of 2D and

3D FEM results were not possible.

79
Chapter 5 Fundamental Behaviour of Corner Effect in Excavation

5.2.1 Parametric Studies by Varying Excavation Dimensions of a Corner

Excavation

Ou et al. (1996) in their study of 3D excavations carried out a series of parametric

studies on the effect of varying excavation sizes. Fig. 5.3 shows the configuration of

the excavation cases studied by them. The parametric studies carried out were based

entirely on 3D and 2D FEM analyses. When the movement of the section furthest from

the corner in a 3D analysis coincides with that of the results from the equivalent 2D

analysis, they define that the plane strain condition is achieved in that 3D test.

In the current study, the investigation on the effect of excavation sizes is first carried out

through a series of FEM analyses, with different primary wall length. The definition of

the terms primary (denotes as PL) and complementary wall (denotes as B) length in the

analysis is shown in Fig. 5.4. For this exercise, the Test 3DK-2 is used as a reference

test in the modelling. The primary wall lengths investigated were 17m (the actual

experimental dimension), 25m, 65m and 130m long. The complementary wall length is

15.5m (the actual experimental dimension) for all the analyses.

It is reported by Ou et al. (1996) that the behaviour of the excavation, especially the

deformation of the primary wall, is highly dependent on the length of the

complementary wall. They also show that an increase of complementary wall length

would increase the maximum wall deflection and the corner effect influence range (Fig.

5.5). However, no explanation is provided for this behaviour in the report. As the

complementary wall is connected to the primary wall at the corner, it is plausible that

the length of complementary wall would influence the behaviour of an excavation to a

certain extent. However, in a real case in the field, when the corner of an excavation

80
Chapter 5 Fundamental Behaviour of Corner Effect in Excavation

only moved slightly, as in the reported cases, the characteristics of the complementary

wall shall have insignificant effect to the primary wall behaviour. Hence, in this study,

the length of the complementary wall is remained constant at 15.5m for all the analyses.

Figs. 5.6a-d show the displacement at the top of the wall versus distance from corner for

the above analyses, with experimental data superimposed for comparison. The

experimental data do land support to the ability of the FEM analysis chosen to model

correctly the overall behaviour of an excavation around a corner. It is generally known

that one of the major shortcomings of FEM analysis is the inability to model the full

spectrum of real soil behaviour. Some of the soil behaviours such as the small strain soil

non-linearity (Jardine et al., 1986; Whittle et al., 1993; Whittle 1997) and anisotropy

characteristic (Banerjee et al. 1984, Wroth and Houlsby 1985; and Banerjee and Yousif,

1986; Jovicic and Coop, 1998) have known to affect the FEM results. The uncertainty

of this soil behaviour under a 2D and 3D environment alone might cause different set of

problem in 2D and 3D analyses. Detailed discussion on this issue was presented in

Section 2.3.2. Hence, while it is more convincing to compare the results of 2D and 3D

FEM analyses to quantify the 3D effect, as adopted by Ou et al. (1996) and Lin et al.

(2003), it is difficult to be carried out accurately. It is noted from Ou et al. (1996) and

Lin et al. (2003) studies that when plane strain condition is achieved in their 3D

analyses, the wall displacement graph becomes constant with distance from corner, and

the magnitude of maximum displacement becomes identical for all cases with longer

primary wall (Fig. 5.5). This suggests that the corner effect has ceased to be felt beyond

this distance, and the condition is in plane strain. These characteristics were observed in

the current study for cases with primary wall length (PL) of 65m and 130m (Figs. 5.6c

and d), where the wall displacement becomes virtually constant beyond 60m to 70m

81
Chapter 5 Fundamental Behaviour of Corner Effect in Excavation

from the corner, and the magnitude of maximum displacement is identical for both

cases. Hence, it is concluded that in the above two cases (PL of 65m and 130m), the

condition is in plane strain beyond the distance of 60m to 70m from the corner. In the

same graph (Fig. 5.6), the 2D test results (Test 2DK-3) were also plotted for

comparison. It is shown that the plane strain condition predicted by the 3D FEM

analyses underestimated the movement captured in the 2D test. This suggested the 3D

analyses carried out has its short coming. However, as the main focus in the current

study is to use the 3D FEM analyses to carry out parametric study to establish the

theoretical behaviour, this short coming is deemed not a major issue. The

improvement of 3D FEM program to capture true 2D behaviour is a major task and is

beyond the scope of the current study.

From Figs. 5.6a-b, for shorter PL of 17 and 25m, there is no section in the primary wall

which has not reached plane strain condition.

The distance of 60 to 70m when the corner effect diminished, or plane strain condition

achieved, can be viewed as the extent of the corner effect. It is proposed to denote this

“extent of corner effect” as L2D, as illustrated in Figs. 5.6c and d. Within a distance of

L2D, the corner effect is felt. To investigate the characteristics of L2D, the normalized

wall displacement, in terms of wall displacement divided by the wall displacement at

2D condition, or plane strain ratio (PSR) as proposed by Ou et al. (1996) is plotted

against the distance from corner, x (Fig. 5.7). In this graph, PSR goes to unity for both

analyses at about 60 to 70m from the corner. For an excavation depth of 5m, this

70m
influence range translates to about = 14 times the depth of excavation (L2D/h =
5m

82
Chapter 5 Fundamental Behaviour of Corner Effect in Excavation

14). The comparisons of this influence range with results reported by others would be

discussed in more detail in Section 5.2.3.

5.2.2 Parametric Studies on the Effect of Wall Stiffness and the Presence of

Capping Beam to 3D Corner Effects

To investigate the wall stiffness effect in a corner excavation, parametric studies were

carried out on corner excavation with different wall stiffness, with and without capping

beam. All analyses were modelling the experimental configuration, except a longer

primary wall of 130m is adopted to ensure the corner effect is extended. FEM analyses

were first carried out to model experimental Tests 3DK-1, 3DK-2 and 3DK-3, but with

Lp of 130m. This series of test were carried out with wall protrusion simulating capping

beam at ground level. The analyses were labeled as 3DK-1-130, 3DK-2-130 and 3DK-

3-130 accordingly. The graphs for plane strain ratio (PSR) versus the distance from

corner (x) of these analyses are shown in Figs. 5.8a-c. These graphs clearly show that

the corner effect’s influence range, L2D, is a constant 70m for all three tests. It should

be noted that the wall stiffness (EI) of Tests 3DK-2 (2mm thick wall) and 3DK-3 (3mm

thick wall), are 8 and 27 times of that for Test 3DK-1 (1mm thick wall). Thus, this

finding clearly shows that the corner effect’s influence range is independent of wall

stiffness.

To isolate the effect of capping beam, the above analyses were repeated with wall

without protrusion. These analyses were labeled as 3DK-1c-130, 3DK-2c-130 and

3DK-3c-130 accordingly. Graphs showing PSR versus the distance from corner (x) of

these analyses are shown in Figs. 5.9a-c. These graphs again clearly show that the

corner effect’s influence range, L2D, is virtually constant for all the tests. More

83
Chapter 5 Fundamental Behaviour of Corner Effect in Excavation

interestingly, the results are also the same with those from analyses with capping beam.

This confirms that the corner effect’s influence range is not affected by the wall

stiffness, as the presence of a capping beam is mainly add to the wall stiffness,

especially at the upper part. This result is thus consistent with the previous set.

Another interesting point that emerged from the above analyses is that the L2D is not

affected by the depth of excavation. To probe whether this is also true before the 2D

condition is achieved, the PSR at a section of x = 30m is plotted against the depth of

excavation, for the analyses described above and shown in Fig. 5.10. The graph shows

that the PSR for all the tests (various thicknesses, with and without capping beam) were

within 0.8 to 1.0. This indicates that the effect of depth of excavation is only within

20% variation for the first 5m of excavation. This finding is consistent with that of

experimental results presented in Section 4.3, where the corner effect is shown to be

relatively insensitive to the depth of excavation. In addition, the graph also shows that

the wall stiffness and capping beam have insignificant impact (within 20% variation) on

the way corner affects the soil, in relation to its equivalent 2D behaviour.

5.2.3 Summary of Findings from FEM Studies

From the above FEM analyses, it was found that for a small excavation, with a primary

wall length of 2 x 25 = 50m, the corner affect covered the entire excavation and plane

strain condition cannot be achieved at any section of the wall. However, for a

sufficiently large excavation, plane strain condition is achieved at a distance sufficiently

far away from the corner. In these cases, the corner effect’s influence range is about 60

to 70m away from the corner. In 3D FEM studies by Ou et al. (1996), for a

complementary wall length of 100m, the corner effect’s influence range is more than

84
Chapter 5 Fundamental Behaviour of Corner Effect in Excavation

50m from the corner. Lin et al. (2003) in their 3D FEM study on corner effect

reported that the influence range is about 30m. More detail comparison on corner

effect’s influence zone have been presented in Section 2.3.3. It is to note that the FEM

analyses carried out in the current study is on unbraced excavation, while those carried

out by Ou et al. (1996) and Lin et al. (2003) were on multi-level-braced excavation.

However, all the above analyses reported large influence range. An influence range of

50m implies that even for an excavation with one section up to 100m long, no section

can be considered to be in truly plane strain condition. This is generally larger than the

reported figure from case studies, as presented in Section 2.3.3 and Table 2.1.

The FEM analyses also show that in a situation when an excavation is large enough and

the plane strain condition developed, the corner effect’s influence range is nearly

independent of the excavation depth and wall stiffness. However, within the influence

range, corner effect is indeed more pronounced at deeper excavation level, but the effect

is relatively insignificant of within 20% variation in the first 5m of excavation. The

results also show that the presence of a capping beam would not affect the corner effect

influence range as it is practically providing additional stiffness to the retaining wall.

However, the presence of capping beam did enhance the magnitude of corner effect

within the influence range.

Table 5.3 summaries the findings of FEM analyses. The findings from centrifuge

modelling discussed in Chapter 4 also presented in the same table for comparison.

Discussion on the agreement, disagreement and elaboration between the FEM and

centrifuge findings is presented in Section 5.3.2.4.

85
Chapter 5 Fundamental Behaviour of Corner Effect in Excavation

The FEM analyses above provided useful information pertaining to the effect of

excavation dimension to corner effect to complement the shortcoming of experimental

studies. Equip with this information, further probe into the fundamental behaviour of

corner effect in an excavation is carried out.

5.3 CHARACTERISATION OF CORNER EFFECTS

5.3.1 Major Factors Affecting the Corner Effect

The centrifuge test results and the FEM analysis results have shown that the corner

effect is mainly controlled by the following factors:

1) Structural issues - Stiffness of retaining wall and the restrain condition at the corner.

The restraint of wall at the corner is the source of corner effect and the wall stiffness is

the factor controlling the magnitude of the corner effect. From FEM studies, the wall

stiffness seems to have little effect on the extent of this corner effect, as indicated by

Figs. 5.8 and 5.9 for different wall thickness and for wall with and without a capping

beam. In these cases, the corner effect extends up to a distance of 14 times the depth

of excavation. However, from centrifuge Tests 3DK-1 and 3DK-2c, when the lateral

flexural capacity of the wall is exceeded and the wall yielded permanently, the corner

effect’s influence range apparently shortened and a “constant movement zone”

developed.

2) Geotechnical issues – Responses of lateral earth pressures, depth of excavation and

γh
shear strength of soil. This could be characterized by the stability number, S = ,
cu

where the effect of excavation depth (h) and soil strength ( cu ) is considered together

(Peck, 1969).

86
Chapter 5 Fundamental Behaviour of Corner Effect in Excavation

To understand better the fundamental behaviour of how a corner affects the overall

excavation behaviour, a simple model to explain the salient features and their

controlling factors is desired.

The configuration of a corner excavation carried out in the centrifuge experiment in this

study is illustrated in Fig. 5.11. The retaining wall was made of two aluminum alloy

plates fully welded orthogonally to each other. As the corner of the plate was rigidly

connected there is transfer of moment between the plates. This had been shown in the

experimental results that the two plates were deformed almost symmetrically without

sign of rotation at the corner (Section 4.1.3, Fig. 4.6). As a first approximation, the

boundary condition at the corner of the plate could be approximated as a fixed

boundary, as shown in Fig. 5.11b. As there are negligible movements observed at the

toe of the plate in the experiments, it could also be assumed as fixed for the convenience

of derivation. The schematic representations of the idealized boundary and loading

conditions of the system are shown in Fig. 5.12.

The loadings acting on the plate are the earth pressure in front (passive side) and behind

(active side) of the plate. The conditions of the plate are similar to that of a typical 2D

problem’s cantilever wall subjected to active and passive earth pressures, if the restraint

at the boundary is ignored. Many references are available for this kind of retaining wall

analysis (Bowles 1996, Sevenoaks 1996, CIRIA 1993). The idealized pressure diagram

of the wall is as shown in Fig. 5.13. As described earlier, the corner effect is controlled

by structural and geotechnical factors. The characterizations of these factors are

discussed in the following subsections.

87
Chapter 5 Fundamental Behaviour of Corner Effect in Excavation

5.3.2 Characterization of Structural Factor

The plate problem as shown in Fig. 5.12 is complicated and no standard formula is

available. Hence, in the current study, the wall is first assumed to be in an elastic state

and the plate problem above can be divided into a number of vertical and horizontal

strips to enable some analytical solutions be arrived at (Fig. 5.12). The deformation of

the plate is the combination of displacements due to the horizontal and vertical earth

pressures. This is applicable in a geotechnical soil-structure interaction problem. The

important thing is to ensure that these composite analyses are able to capture the

essential mechanism of an excavation around such a perpendicularly jointed plates

system. The vertical strips are subjected to the imbalance earth pressures that would be

generated in an excavation as shown in Load Cases 1a and 1b in the figure. The

horizontal strips are subjected to the differences between active and passive earth

pressures at the particular strips, or q y = Pp ( y ) − Pa ( y ) , as shown in Load Case 2 in the

figure. Load Cases 1a and 1b are the imbalance loads that cause movement, and hence it

is examined first.

5.3.2.1 Deformation Due to Excavation Induced Imbalance Load

Based on moment equilibrium, the bending moment at any coordinate y on the vertical

strip due to Load Case 1a, which denotes the case with active earth pressure loading, is

1 P y P
M = * a y * y * = a y3 (5.1)
2 b 3 6b

The general equation for bending moment in a beam is (Timoshenko 1958):

d 2δ
M = EI (5.2)
dy 2

Combining Equations (5.1) and (5.2) yield

88
Chapter 5 Fundamental Behaviour of Corner Effect in Excavation

d 2δ Pa 3
EI = y (5.3)
dy 2 6b

where

Pa = Active earth pressure at the toe of the wall

b = Total depth of the wall from ground level to the toe of wall

y = Coordinate along the depth of wall

δ = Deflection at y-axis

E = Elastic modulus of the wall

I = Second moment of area of the wall

Integrating Equation (5.3), give

dδ P
EI = a y 4 + C1 (5.4)
dy 24b

and

Pa 5
EIδ = y + C1 + C 2 (5.5)
120b

Apply the boundary condition of Case 1a to Equation (5.4), where

dδ Pa 3
y = b, =0 , gives C1 = − b (5.6)
dy 24

Pa 4
y = b, δ = 0 , gives C 2 = b (5.7)
30

Substituting the constants C1 and C2 into Equation (5.5) and rearranging yields

Pa P P
δ= y 5 − a b3 y + a b 4
120 EIb 24 EI 30 EI

= Pa ( β1 y 5 + β 2 y + β 3 ) (5.8)

89
Chapter 5 Fundamental Behaviour of Corner Effect in Excavation

where

1 1 1 b3 1 b4
β1 = , β2 = − , β3 =
120 EIb 24 EI 30 EI

The above solution is actually the standard solution for deflection for triangular loading

on simple cantilever beam (Gere and Timoshenko, 1994). The derivation is presented

here for a better illustration of the imbalance loads and the subsequent relationship with

the corner effect. Equation (5.8) is the equation for wall displacement due to the active

earth pressure. This deformation is countered by deformation due to passive earth

pressure, which is Load Case 1b in this exercise.

The solution for wall displacement due to Load Case 1b, below the excavation level, or

when h ≤ y ≤ b , is

Pp Pp Pp
δ =− ( y − h) 5 +
3
h1 ( y − h) − h1
4

120 EIh1 24 EI 30 EI

= Pp γ 1 ( y − h) 5 + Pp γ 2 ( y − h) + Pp γ 3 (5.9)

where,

3 4
1 1 1 h1 1 h1
γ1 = − , γ2 = , γ3 = −
120 EIh1 24 EI 30 EI

h = depth of excavation

h1 = embedment of wall after excavation, measured from the excavated

formation level to the toe of wall

The solution for deformation of wall above the excavation level, which is the focus of

the study, is governed by the equation

90
Chapter 5 Fundamental Behaviour of Corner Effect in Excavation

dδ  Pp Pp 3 
EI =  − ( y − h) 4 + h1 
dy y =h  24h1 24  y =h

Pp 3
= h1 (5.10)
24

Rearrange and integrate Equation (5.10), gives

 P 
δ =  p h13 (h − y )
 24 EI 

= Pp γ 4 (h − y ) (h ≥ y ≥ 0) (5.11)

where

3
h1
γ4 =
24 EI

The combination of displacements due to Load Cases 1a and 1b shall give the total

displacement due to earth pressure. For displacement at ground level, which is the

focus of the current study, combine and solve for Equations (5.8) and (5.11), for

0 ≤ y ≤ h . This gives:

δ = − Pa ( β1 y 5 + β 2 y + β 3 ) + Pp γ 4 (h − y )

{
= − Pa nγ 4 (h − y ) − ( β1 y 5 + β 2 y + β 3 ) } (5.12)

where

Pp
=n
Pa

Equation (5.12) above is the solution for displacement at ground level, for a cantilever

elastic wall subjected to excavation induced imbalance load. This is akin to a typical

plane strain, cantilever wall excavation problem.

91
Chapter 5 Fundamental Behaviour of Corner Effect in Excavation

5.3.2.2 The Effect of Structural Restraint at Corner

For Load Case 2, the load qy is assumed as a uniformly distributed load (UDL) along the

x-axis (distance from wall). It is changing along the y-axis (the depth from ground).

The UDL assumption is only approximately accurate as the earth pressures may vary

with distance from corner, especially during excavation, but this simple assumption is

necessary so that the derivation can proceed. A more detailed discussion on the effect

of earth pressure is given in the following few sections.

The loading and boundary conditions are actually the standard equation of a cantilever

beam subjected to uniformly distributed load. The general expression for displacement

for this problem is (Gere and Timoshenko, 1994):

qy
δ= ( x 4 − 4 Lx 3 + 6 L2 x 2 ) (5.13)
24 EI

where

L = length of the cantilever beam, and

x = distance from the fixed end (in the current study, the corner)

The displacement for this Load Case 2 shall be combined with that for Load Case 1a

and 1b to give an approximate solution for a corner excavation. It is noted that the

displacement of the system is actually driven by the imbalance of earth pressure, which

is captured in Load Case 1a and 1b. The structural beam effect in Load Case 2 is

actually the wall retraining effect, which is the source of the corner effect. The

expression for the beam displacement is controlling the development of this corner

restrain.

92
Chapter 5 Fundamental Behaviour of Corner Effect in Excavation

It is noted that qy, the difference between passive and active earth pressure at a given

depth is a variable dependent on depth of excavation, distance from corner, type of soil,

status of consolidation and so on. Hence, to simplify the problem, lets the restrain of

Load Case 2 at any section of the beam be described in terms of normalization ratio,

which is the displacement at any distance divided by the maximum displacement at

distance furthest from corner:

δ
λs = (5.14)
δ max

Where

δ = displacement at any distance from corner, x

δmax = displacement at distance furthest from corner, or at x = L

Fig. 5.14 illustrates this assumption.

Substituting Equation (5.13) into Equation (5.14) and simplified, gives the normalized

ratio as

4 3 2
1 x 4 x x
λ s =   −   + 2  (5.15)
3 L 3L L

Combining Equations (5.12) and (5.15), the wall displacement of a corner excavation

satisfying all the boundary conditions, at ground level (y = 0), is

δ = Pa {nγ 4 h + β 3 }λs (5.16)

where

3
h4 h
γ4 = − + 1
24 EIh1 24 EI

1 b4
β3 =
30 EI

93
Chapter 5 Fundamental Behaviour of Corner Effect in Excavation

5.3.2.3 Corner Effect Due to Structural Restrain at Corner

As the first expression Pa {nγ 4 h + β 3 } in Equation (5.16) is the displacement of wall top

under a plane strain condition, this means that the 3D displacement of the wall can be

estimated based on the plane strain results, multiply by the factor λs.. In that sense, the

factor λs serves a similar role to that of the plane strain ratio (PSR) proposed by Ou et al.

(1996) (Section 5.2.1). The only difference between the PSR and λs in this study is that

the λs is at the furthest distance from the corner, and only when the distance is far

enough, and 2D condition is developed, that λs becomes PSR.

Following the philosophy first proposed by Ou et al. (1996), in this study, it is proposed

to consider λs as a “corner effect factor”, which is an “active” contributor providing

corner effect in a 3D excavation. The displacement of a retaining wall subjected to

corner effect can then be estimated as

δ max
δ= (5.17)
λs

Curve 1 of Fig. 5.15 shows a typical plot of normalized wall top displacement versus

distance from corner calculated from FEM analysis modeling Test 3DK-2-130. It is

shown that the wall reached 2D condition (PSR = 1) at x = 70m. At this distance, the

corner effect has ceased completely, and the x value is L2D, which is the corner effect’s

influence range. More importantly, the corner effect is felt most acutely in the initial

range of about 30m from the corner. At x = 25m, the PSR achieved is already 85%,

and the remaining 15% of corner effect is obtained from distance between 30m to 70m.

94
Chapter 5 Fundamental Behaviour of Corner Effect in Excavation

Hence, for easier illustration, it is not unreasonable to assume the corner effect influence

range as 25m, and this distance is termed Ll, as shown in the graph.

The normalized beam deflection curve based on the structural beam Fixed-end

condition, as represented by Equation (5.15) above, is also reflected in the same graph

with a prefix PSR-Fixed (Fig. 5.15). The L is taken as Ll, or 25m in the calculation. It

is shown that the PSR-Fixed curve is significantly smaller than that from the FEM

analysis, especially at distance near to the corner. This suggests that using a fixed-end

condition over-estimated the effect of the corner restrain. This is reasonable as the

fixed-end condition actually assumed a perfect lateral bending moment transfer from a

stiff, fixed boundary. The connection of the wall at the corner is unlikely to achieve this

perfect fixed-end condition.

The other extreme condition at the corner can be a pin-end with zero moment transfer.

The illustration and derivation of this end condition is shown in Fig. 5.16. For this

end condition, the expression of the corner restrain normalization ratio is:

x  8 4  x 
2 3
δ  1 x 
λs= =  −   +    (5.18)
δ max L l  5 5  Ll  5  Ll  

Also, the end condition may be pin, but with certain moment transfer at the corner. The

relationship of the development of restrain under this condition is:

δ x  8 x 1  x  
2

λs= =  − 2 +    (5.19)
δ max Ll  3 Ll 3  Ll  

The complete derivation and illustration of this end condition is shown in Fig. 5.17.

95
Chapter 5 Fundamental Behaviour of Corner Effect in Excavation

The normalized displacement curves based on the above end restrains are shown in the

graph (Fig. 5.15) with prefix Pin and Pin-Moment accordingly. It is shown in the

graph that both the pin and pin-moment end conditions underestimated the effect of

corner restrain where the displacement curves are always higher than the actual FEM

results.

Hence, a logical situation of the above end-condition should be in between fixed and pin

end condition. However, it is impossible to assess the end condition accurately. Hence,

linear approximation between that for the fixed and pin conditions is proposed to

illustrate the corner effect due to the corner restrain easier. The linear relationship is

x x
, which is shown as PSR- curve in the graph. The Ll is assumed as 25m for all
Ll Ll

the corner effect ratios presented above. Though not perfect, the results of this case

agree much better with that from the FEM analyses.

The above expressions show to some extent of how the effect of a corner is propagated.

This is true not only at the wall top, but also for the entire depth of the wall. However,

the above discussions are confined only to the general evaluation of the development of

corner effect. To quantity the corner effect, it is clear that the corner effect influence

range, L2D or Ll have to be understood better.

5.3.2.4 The Corner Effect’s Influence Range

Literature review on previous studies and observations on the corner effect influence

range was presented in Section 2.3.3. From the above review, the corner effect

influence range observed on site is generally confined to 1 to 2 times depth of

excavation or less (Nath, 1983; Simpson, 1992; Ou and Chiou, 1993; Wong and Patron,

96
Chapter 5 Fundamental Behaviour of Corner Effect in Excavation

1993). Also from the above review (Section 2.3.3), the corner effect influence range

obtained from FEM analyses (Ou et al., 1996; Lin et al. 2003) is more than 3 times

depth of excavation, which is significantly larger than that observed in the field.

In the current study, the FEM analyses simulating the centrifuge tests show that the

corner effect influence range is up to 70m away from the corner. The influence range

is almost constant regardless of excavation depth, wall thickness, with or without

capping beam. Table 5.3 compared the findings from FEM analyses and centrifuge

modelling. In contrast, centrifuge modelling described in Chapter 4 shows that for thin

wall (1mm and 2mm), a “constant displacement zone” developed at about 10m from

corner. This “constant displacement zone” is usually considered as an indication of the

cessation of corner effect. Hence, the FEM analyses carried out seems overestimated

the corner effect influence range. This is consistent with the literature review findings

discussed in the previous paragraph.

From the 3D centrifuge tests (Chapter 4), it is found that the “constant displacement

zone” is actually caused by the yielding of the retaining wall in the lateral direction.

When the lateral flexural capacity was exceeded, the wall cannot extend the corner

effect far enough to approach a plane strain condition but instead developed into a

region with nearly constant displacement. In the FEM analyses, the retaining walls

were assumed to be isotropic homogenous monolithic structures and the lateral yielding

was not captured.

As a matter of fact, in the actual retaining wall construction, such as sheet pile, soldier

pile, contiguous bored pile or diaphragm walls, there are minimal lateral moment

97
Chapter 5 Fundamental Behaviour of Corner Effect in Excavation

transfers between the piles/panels. The lateral flexural capacities of the walls are

usually much smaller than the vertical capacity. The relatively lower lateral flexural

stiffness of the retaining wall reduces the corner effect which was not able to be

modeled by the 3D FEM due to the deficiency of 3D FEM in modeling the anisotropy

of bending moment capacity of the retaining wall (Sun 2003).

In view of the above findings, it is concluded that the extent of corner effect is also

limited by the lateral flexural capacity of wall. In the condition that the wall is strong

enough, the effect of the corner would propagate far enough to achieve a plain strain

condition. Hence, the corner effect influence range can also be written as:

La= min .{Lm , L2 D } (5.20)

Where Lm = Length from corner when lateral flexural capacity is exceeded.

L2D = Length from corner when plane strain condition is achieved.

Fig. 5.24 illustrates the scenario of Equation (5.20) as Curve 3 in the graph. Curves 1

and 2 are presented earlier (Fig. 5.15) for the situation if lateral flexural capacity of the

wall is not exceeded, and 2D condition is achieved. In Curve 2, the L2D is simplified

as Ll as most corner effect is contained within Ll, as discussed in Section 5.3.2.3.

In the centrifuge experiments, due to the relatively short length of the wall, 2D

condition cannot be achieved. In Tests 3DK-1 and 3DK-2c, where yielding occurred,

the Lm condition is achieved. This scenario is illustrated in Fig. 5.24. In the field, due

to the small lateral flexural capacity of wall compared to the vertical direction, it is also

likely that the Lm condition is achieved rather than the L2D condition. From this

98
Chapter 5 Fundamental Behaviour of Corner Effect in Excavation

deduction, it is clear that the corner effect within the Lm zone also deserved in-depth

investigation.

It is noted that in the experiments, one end of the wall is welded to the corner while the

other end is a free-end. Hence, the structural corner restrain effect is controlled only

by the corner end condition, regardless of the wall length. As such, it is not counter

intuitive to assume that the displacement versus x curve of a short length wall, is the

same as that of a longer wall, which is represented by Curve 3 in the graph (Fig. 5.24).

From Curve 3 in the graph (Fig. 5.24),

δ x
= = λ 2 D if lateral flexural capacity is not exceeded (5.21)
δ 2 D Ll

δ x
= = λ y if lateral flexural capacity is exceeded, as in the experiments. (5.22)
δ y Lm

Though approximate, this is an useful expression as it suggests that if Lm is known, the

corner effect at any distance from the corner, within the Lm zone can be estimated.

Similarly, if the Ll or L2D is known, the corner effect at any distance can also be

quantified relative to that of the plane strain condition. It also means that the lateral

flexural capacity has to be evaluated; this is an important extension to Ou et al. (1996)’s

concept of a PSR.

From the experimental studies of Test 3DK-2, it was found that negative lateral bending

moment developed near to the corner due to the welding at the corner (Fig. 5.25a).

Nearer to the other end, the bending moment is decreasing due to the free-end condition.

At other span of the wall, the lateral bending moment is increasing with distance from

corner. The capture bending moment profiles agreed well with the theoretical lateral

99
Chapter 5 Fundamental Behaviour of Corner Effect in Excavation

bending moment profile shown in Fig. 5.25b. When the bending moment (BM)

developed exceeded the yield stress of the wall at certain distance from the corner, the

wall will yield. However, due to the limited information that could be obtained from the

measurements, a more detailed bending moment profile to allow for in-depth analysis of

yielding behaviour cannot be carried out. Nevertheless, from the results, it clearly

suggests that at a certain distance from the corner, the lateral bending movement

reached the lateral flexural capacity of the wall and the wall yielded.

In the present series of tests, the case of a 1mm thick wall with capping beam (Test

3DK-1) showed yielding, at about 10m from corner. Theoretically, when the wall

stiffness is zero, Lm should be zero, reflecting no wall retrain effect at all. In addition,

from the test results, tests with 2mm and 3mm thick wall show no sign of yielding. The

accurate estimation of Lm according to the wall stiffness is difficult as the problem

involves soil-structure interaction. Hence, based on the above observations, as a first

approximation, it is assumed that Lm is linearly proportional to the wall thickness (Fig.

5.26a), with intercept at zero. The wall thickness can be represented by the first

moment of wall section area (EA), to more reasonably define the moment capacity.

Fig. 5.26b illustrates this the Lm versus EA relationship. Based on this hypothesis,

values of Lm for 2mm and 3mm thick wall are deduced as 20 and 30m respectively.

If the above hypothesis is true, by using the deduced Lm and applying Equation (5.22)

above, the movement at various distances from corner, of various stiffness walls, would

be normalized to the value of δy, as illustrated in Fig. 5.27. These hypotheses are

examined with the experimental test results in the coming sections.

100
Chapter 5 Fundamental Behaviour of Corner Effect in Excavation

5.3.3 Characterization of Geotechnical Factors

For Test 3DK-2c, the lateral earth pressures at the front and back faces of the wall were

measured at two locations from corner, which are, at 8m and 15m from the corner (Fig.

5.28 and also described in Section 4.2.2, Figs. 4.18 and 4.19). While the data are

limited, the variation of earth pressure values captured provides useful information for

the characterization of corner effect on the earth pressure due to the excavation depth

and soil strength.

As discussed in Section 4.2.2, the changes of total lateral earth pressure coefficient, K ,

during excavation show the following interesting trend (Fig. 5.28) :

a) At the active side, the changes of K a ( K at active side) at a further distance

from the corner (15m) drop significantly faster than that nearer (8m) from the

corner.

b) At the passive side, the changes of K p ( K at passive side)at 8m and 15m from

corner during the excavation were almost identical.

To understand the first observation, it is important to reiterate that the soil behind the

wall should be at K o condition before excavation and the lateral earth pressure would

reduce as the shear strength was mobilized and the wall deflection increased. The

lateral earth pressure coefficient should reduce from K o to K a − active ( K at active state),

the exact value is dependent on the magnitude of soil strain mobilized. Nearer to

corner, the wall displacements are smaller hence the drop of K a is accordingly smaller.

101
Chapter 5 Fundamental Behaviour of Corner Effect in Excavation

It was shown in Section 4.2.2 that for TSTs at 15m from corner, the active state was

achieved by the time excavation reached 1.5m depth. At the same location on the

passive side, the passive limit state has just reached at the end of 5m-excavation.

Hence, the above phenomenon is consistent with the general understanding that much

more significant development of strain is needed to reach passive state. Nevertheless,

the magnitudes of change in the passive stress are larger than that for the active stress,

as evident from the experimental results shown in Fig. 5.28. K p changes from 0.85 to

1.60, while K a changes from 0.85 to 0.60. Intuitively, the larger changes on the passive

side should mean more variation in the lateral earth pressure with distance from the

corner. However, the experimental results clearly show otherwise. Hence, from the

above observation, it is not unreasonable to consider that, for the above centrifuge tests,

the variation of earth pressure due to corner effect would mainly influence the active

side only and to a very insignificant extent on the passive side. A schematic diagram

illustrating the above scenario is shown in Fig. 5.29.

The active earth pressures of the 3D test described above (Test 3DK-2c) are plotted

together with active earth pressure of 2D test (Test 2DK-3) and showed in Fig. 5.30. It

is shown that the active earth pressure for the 3D test, measured by total stress

transducer (TST) at 15m from corner ( Pa ( 3 D ) −15 m ), shows almost identical response

during the excavation as that for the 2D test ( Pa ( 2 D ) ). For TST at 8m from corner

( Pa (3 D ) −8 m ) of the 3D test, the active earth pressure decreases significantly less during

the excavation as compared to that of the TST at 15m from corner or that for the 2D

test.

102
Chapter 5 Fundamental Behaviour of Corner Effect in Excavation

To better illustrate the relationship between the active earth pressure at various distances

away from the corner for the 3D test, as well as to distinguish the differences of 3D and

Pa ( 3 D ) γh
2D earth pressure behaviour, relationship with the stability number is plotted
Pa ( 2 D ) Cu

in Fig. 5.31. The stability number combines the effect of excavation depth (h) and the

soil strength and is widely used to characterize excavation (Peck, 1969; Clough and

Denby, 1977). The graph shows two interesting points:

Pa ( 3 D ) −15 m
1) The ratio is always near to unity throughout the excavation. This shows
Pa ( 2 D )

that the earth pressure at 15m away from the corner is already close to the plane strain

condition. It is important to note that here it is the soil condition that is approaching 2D

condition, not the overall system including the structural stiffness of the wall. This is

also consistent with general understanding of earth pressure theory that only a small

mobilized strain is needed to reach an active state.

Pa ( 3 D ) −8 m
2) the ratio is unity before excavation, but increases almost linearly with the
Pa ( 2 D )

γh
increase of stability number, and reaches 1.20 when is 5. This trend suggests that
Cu

the corner effect on the development of active earth pressure would increase according

to excavation depth. In other word, the deeper the excavation, the more significant is the

effect of corner on the active earth pressure.

Pa ( 3 D )−8 m
For , the best-fit linear curve of the graph is in a form of
Pa ( 2 D )

103
Chapter 5 Fundamental Behaviour of Corner Effect in Excavation

Pa ( 3 D ) γh
= 1+ R (5.23)
Pa ( 2 D ) Cu

Where

R = slope of the graph

Pa ( 3 D )
The variation of magnitude of at 8m and 15m distances from the corner described
Pa ( 2 D )

above is due to the differences of wall movement at the two points. Hence, it indicates

Pa ( 3 D )
that is not only dependent on the excavation depth, but also influenced by the
Pa ( 2 D )

distance from corner. Due to the small size of the scale model in a centrifuge test, only

four TSTs were installed. To complement the limited information, FEM analyses

Pa ( 3 D ) γh
modelling the same test, Test 3DK-2c, were carried out. The versus
Pa ( 2 D ) Cu

relationship obtained is shown in Fig. 5.32. The Pa(2D) in the plot is taken from the

FEM analysis for 3DK-2c-130, where earlier, it was shown that with a 130m long wall,

plane strain condition could be developed (Section 5.2.1). The experimental results

were also plotted in the same graph for comparison.

Fig. 5.32 shows the followings:

Pa ( 3 D )−15.8 m
1) ratio is close to unity throughout the excavation. This shows that
Pa ( 2 D )

the earth pressure at 15m away from the corner is already close to or at plane

strain condition. This finding is consistent with the experimental results.

Pa ( 3 D )
2) The ratios at sections 1.7m, 6.3m and 11.8m from the corner increase
Pa ( 2 D )

almost linearly with an increase in the stability number. This trend is again

104
Chapter 5 Fundamental Behaviour of Corner Effect in Excavation

consistent with the experimental results, and supports the observation that the

Pa ( 3 D )
ratio indeed increases proportionally to an increase in excavation depth.
Pa ( 2 D )

Pa ( 3 D )
However, Fig. 5.32 also shows that the slope (R) governing the increase in
Pa ( 2 D )

increases as section is closer to the corner. To see this relation better, the slope (R) is

plotted against the distance from corner, as shown in Fig. 5.33. The graph shows a

relationship of the form:

 x
R = K 1 −  (5.24)
 a1 

Where a1 is the distance when the earth pressure first reached that of the plane strain

condition, or when the corner effect is fully diminished. In the current exercise, a1 is the

intersect of the graph at X-axis, which is 13m. This value is slightly shorter than the

Pa ( 3 D )
15m observed above where at 15m is close to unity during excavation. As shown
Pa ( 2 D )

in Fig. 5.33, K in Equation (5.24) above is a constant, which is the intercept of the graph

against the R-axis.

Substituting Equation (5.24) into Equation (5.23), yield

Pa (3 D ) x γh
= 1 + K (1 − ) (5.25)
Pa ( 2 D ) a1 C u

Since the displacement of a corner excavation is actually directly related to Pa at plane

Pa ( 3 D )
strain condition, the relationship of as shown in Equation (5.25) above is actually
Pa ( 2 D )

a “corner effect factor”, as shown below:

105
Chapter 5 Fundamental Behaviour of Corner Effect in Excavation

Pa ( 2 D ) 1
=λ a = (5.26)
Pa ( 3 D ) x γh
1 + K (1 − )
a1 Cu

At corner when x =a1, λa = 1, which is at the plane strain condition.

It is noted that while the passive earth pressure is shown to be not influenced by the

corner effect, it may be case specific for the current experimental study only. With

variation in soil type and soil strength, the corner effect may also influence the passive

earth pressure. However, with limited information, this cannot be assessed in the

current work.

5.4 COMBINED CHARACTERIZATION OF CORNER EFFECT

From the above study, it is clear that the corner effect results mainly from two factors,

the structural restrain and the geotechnical factors. Hence, the general equation for

wall top displacement of a corner in an excavation can be approximately expressed as:

δ 3 D = δ y * λs * λa (5.27)

Where

δy = Maximum wall top displacement when the wall lateral flexural capacity is

exceeded

λs = Corner effect factor due to corner structural restrain

λa = Corner effect factor due to geotechnical factors

In the condition that the wall is strong enough, the effect of the corner would propagate

far enough to achieve a plain strain condition, and δy shall becomes δ2D, and the above

equation shall becomes:

106
Chapter 5 Fundamental Behaviour of Corner Effect in Excavation

δ 3 D = δ 2 D * λs * λa (5.28)

It is to note that these corner effect factors are isolated and can be quantified separately.

These two factors have been shown to explain major aspects of 3D corner effects in an

excavation. And if these two factors are used to normalize the displacements at various

distances from corner to take into account how the corner affects the overall behaviour,

the displacement after normalization should approach the value of δy, as illustrated in

Fig. 5.27.

However, the discussions on the characteristics of corner effect in this section have been

mostly theoretical to this point. It is necessary to apply these principles to evaluate the

experimental results.

5.4.1 Evaluation Of Corner Effect Hypotheses

The wall top displacement of Test 3DK-1, 3DK-2, 3DK-3 is as shown in Figs. 5.34a,

5.35a and 5.36a. As described in Section 5.3.2.4, Test 3DK-3 has only one

measurement point due to technical difficulty in this test. To isolate the corner effect

due to the structural corner restrain effect, the displacements in the graph above are

divided by the respective λs accordingly (Fig. 5.34b, 5.35b and 5.36b). The simplified

λs = x/Lm relationship, with the restrain condition of between the fixed and pin-end

restraint condition, described in Section 5.3.2.3, is used for this exercise. The Lm

deduced in Section 5.3.2.4 are 10, 20 and 30m respectively for 1mm, 2mm and 3mm

thick wall (with wall protrusion), and the x is the distance from corner. The graphs

show that the normalized displacements versus x approaching a flat line against the

distance form corner, indicating the removal of contribution from the lateral structural

107
Chapter 5 Fundamental Behaviour of Corner Effect in Excavation

stiffness of the wall. This also suggests that the observed effects of a corner in these

tests are derived from the structural restraint at the corner.

The above results are further normalized to remove the corner effect due to the

geotechnical factors as a result of the structural effect. For this, the displacements

divided by λs were further divided by the respective λa accordingly (Fig. 5.34c, 5.35c

and 5.36c). Due to the limited earth pressure measurement results, λa is assumed to be

the same for all the tests. λa is estimated based on Equation (5.26), by using K =

0.0857, and a1 = 15m (Fig. 5.33). The graphs show that the normalized displacements

versus x approaching a constant value, highlighting that these are the two main factors

controlling how a corner affects an excavation.

The above normalization shows that the ideas developed thus far in this chapter are able

to capture the mechanics of the corner on an excavation. But thus far, only the

displacement at the wall top is examined. Next, the surface settlement is examined.

The surface settlement behind an excavation and wall displacement can generally be

related with a ratio. For example, Mana and Clough (1981) reported that the

maximum ground settlement is within 0.5 to 1.0 times the maximum wall movement.

Hence, it is possible to use the corner effect factors, developed thus far to interpret the

surface settlement behind the wall with distance from corner directly.

The surface settlement picked up by displacement transducers installed nearest to the

wall, which is at 7m behind the retaining wall are used for this exercise because the

settlement here is the maximum. Figs. 5.37a, 5.38a and 5.39a show the offset-surface

108
Chapter 5 Fundamental Behaviour of Corner Effect in Excavation

settlement of tests 3DK-1, 3DK-2 and 3DK-3. The settlements were offset against the

settlement at the corner to highlight the corner effect (Section 4.4). To account for the

corner effect due to the structural restrain from the corner, the displacements in the

graphs above are divided by the respective λs accordingly (Fig. 5.37b, 5.38b and 5.39b).

The simplified λs used was the same as that used earlier to interpret the displacement at

the top of the wall. The results obtained are then divided by the respective λa

accordingly (Fig. 5.37c, 5.38c and 5.39c) to remove the corner effect due to the

geotechnical factors. From the graph, it is observed that the surface settlement

divided by λs versus x for all the tests approaching a constant value, suggesting that the

structural restrain from the corner play a very significant role. When the results

obtained is further divided by λa, a further improvement is obtained.

To better compare the results of all the tests, the wall top displacement versus depth of

excavation of the above tests are plotted on a same graph. In these graphs, the results

of Tests 2DK-1 and 2DK-3 were also plotted to provide an overall comparison. Fig.

5.40a shows the wall top displacements of all the tests before normalization. Fig. 5.40b

shows the wall top displacements divided by λs. The graph shows that the data band

narrowed significantly, illustrating the impact of the lateral structural restraint arising

from a corner. When the above results in Fig. 5.40b are further divided by λa, as

shown in Fig. 5.40c, the data band narrowed further, and at each stage of excavation, the

data are converging to a constant value. This shows that the two factors are able to

account for a large extent the corner effect in an excavation.

109
Chapter 5 Fundamental Behaviour of Corner Effect in Excavation

A similar exercise is carried out for the offset-surface settlements (at 7m behind wall)

and the results are shown in Figs. 5.41a to c. The results are consistent with that

observed for the displacement at the wall top.

It is obvious from the graphs (Figs. 5.40c and 5.41c) that the normalized wall top

displacement and surface settlement has not reached the plane strain condition (Tests

2DK-1 and 2DK-3). This is because the normalization above is based on the

movement at a situation where the lateral flexural capacity is exceeded, where “constant

corner effect” developed, rather than that at plane strain condition. This “constant

corner effect” zone is the focus of the current study, and it provides important elemental

findings on how the corner effects are developed in an excavation. In the field, due to

the fact that the walls lateral flexural capacity is much smaller than that of vertical

panels, it is also more likely that the corner effect at “constant corner effect zone” are

developed, rather than the ideal plane strain condition. More importantly is the fact that

these findings are fundamentally valid and applicable even for a large excavation where

plane strain problem may develop at sections far from corner.

However, it is noted that there are still some scatter in the data after all the

normalizations (Figs. 5.40c and 5.41c). This spread of the data is likely due to the fact

that the corner effect factors described above are only based on generalized parameters

and also simplified mechanics. Further fine-tuning of the above factors based on more

accurate estimate of wall and soil characteristics, such as the wall yielding criteria,

bending moment transfer at the corner, soil strength and inclusion of passive earth

pressure effect, can further improve the normalization.

110
Chapter 5 Fundamental Behaviour of Corner Effect in Excavation

Lee et al. (1998) carried out a study on the 3D corner effect in excavation with some

reported case studies. From the assessment of corner effect at various sites, with

different wall length and final excavation level, they suggested that the length of wall

over the depth of excavation ratio might be a possible contributory factor for corner

effects. The lower the ratio, the more significant is the corner effect. In the discussion

earlier (Sections 5.3.2.3 and 5.3.2.4), the proposed corner effect ratios, λs and λa, are a

function of distance from corner (x). The further from the corner, the smaller is the

corner effect. Hence, for the same excavation depth, the shorter is the wall; the lesser

is the corner effect. This is consistent with the suggestion made by Lee et al. (1998).

However, for a same excavation where the length of the wall is a constant, Lee et al’s

suggestion implied that the excavation depth would increase the corner effect. This is

only partially true as the corner effect shown in the current study is not as sensitive to

excavation depth. It is valid in terms of magnitude, rather than relative ratio quoted in

the current study.

Another major factor of excavation parameters in the field that was not investigated in

the current study is the effect of bracing on the influence of a corner. It is to note that

cross bracing only provides passive resistance, even for the case of braces with pre-

loading. Liu (1995) and Lee et al. (1998) in their 3D-excavation study reported that the

presence of steel strutting might have actually suppressed the corner effect above

excavation level. In their excavation analysis where thick soft clay was found below

the final formation level, they observed that the corner effect was more pronounced

below the excavation level where there were a thick soft clay, likely because there were

no suppressing effect by the stiff strutting system. However, it is to note that the

bracings at corner often are diagonal bracings, which may be stiffer than cross braces at

111
Chapter 5 Fundamental Behaviour of Corner Effect in Excavation

other areas far from corner. This may have increased the structural stiffness at the

corner wall, contributing to corner effect. The inclusion of these factors is well beyond

the scope of the present study, but would be needed to ensure a comprehensive

understanding of how a corner affects an excavation.

Giger and Krizek (1975) conducted a study on the stability of vertical cut with variable

corner angle. They postulated the 3D-failure mechanism of vertical cuts and showed

that the stability of a vertical cut is significantly dependent on the corner angle. In the

current study, the corner of the excavations, at the initial stage is always at 90°. In the

field, many excavations are not having 90° corner. The variation of corner angle is

expected to have a significant effect on the end-restrain condition, which would then

affect the development of the structural restrain. In addition, complicated passive and

active earth pressure effect would also develop due to the variation of corner angle.

However, the purpose of the current study is to develop a basic understanding on the

behaviour of a corner effect in an excavation, and the approach was based on the

understanding of a simple excavation with a set of perpendicular wall. Consideration of

other factors mentioned above is beyond the scope of the present work. In particular,

new techniques need to be developed to carry out such tests in a centrifuge, this is not

an easy task.

112
Chapter 6 Conclusions

CHAPTER 6

CONCLUSIONS

6.1 DEVELOPMENT OF 3D IN-FLIGHT EXCAVATOR

1) A 3D in-flight excavator was developed in the current study to simulate excavation

around a corner in the centrifuge. The results obtained from the centrifuge tests are

used for mechanistic study on three-dimensional (3D) behaviour of an excavation

affected by the presence of a corner.

2) The in-flight excavator developed is able to model both 2D (plane strain) and 3D

(corner) excavations in the centrifuge. This has made the comparison between 2D

and 3D tests more straightforward without complication caused by using different

set-up for different tests.

3) The results of the centrifuge tests present a glimpse of the excavation behaviour

around a corner under controlled boundary and initial conditions. This facilitates

the interpretation of the results, something a much more complex problem in the

field cannot offer. In addition, the in-flight excavator has the advantage over the

traditional “increasing-g” and “heavy liquid” methods to simulate excavation as the

in-situ soil stresses and changes in earth pressures due to excavation could be

modeled more correctly, especially in normally consolidated soft clay. The effect

of variation of lateral earth pressure with distance from corner, which proven to be

an important parameter influencing corner effect, was captured in the centrifuge

test.

113
Chapter 6 Conclusions

6.2 SUMMARY OF FINDINGS

The following findings may be concluded from the current study:

1) The presence of capping beam, which usually ignored by practicing engineers in

analysis, would enhance the corner effect. The effect of capping beam is more

pronounced in the early stage of excavation. This is expected as the capping beam

is akin to a bracing at the top of wall which effectively restrains wall movement

when the excavation is near to the top of wall level. The capping beam effect is

significance as the first stage of any excavation is often an unbraced open

excavation, and is also the stage where significant movement occurs.

2) Comparative study on excavation tests with stiff highly over consolidated and soft

normally consolidated samples found that the presence of a corner would affect the

wall deflection and surface settlement more pronounced in soft soil than that in stiff

soil. However, this is in terms of absolute magnitude. In terms of relative ratio

of movement as mechanistic study is concerned, there is no apparent correlation

between the behaviour of corner effect with the soil strength.

3) From centrifuge tests and finite element analyses, it was found that the behaviour of

corner effect is insensitive to the depth of excavation.

4) If the retaining wall flexural capacity is strong enough, the influence range of the

corner is given by the point where the 2D plane strain condition starts. In this case,

the influence range is independent of the wall’s flexural capacity or whether there is

a presence of ground capping beam. If the wall is not strong enough, the influence

range is decided by the point when the flexural capacity of the wall was exceeded,

which is at a shorter distance from the corner.

5) The retaining wall in the current study is essentially a structural plate problem

subjected to varying earth pressures. The wall movement subjected to the earth

114
Chapter 6 Conclusions

pressures is actually similar to that of a typical 2D problem, if there is no restraint

at the boundary. Thus, the boundary condition at the edge is actually the source of

the 3D corner effect.

6) The development of corner restrain from the corner of a “plate” was further

simplified to that of a beam problem assuming that the wall is in an elastic state.

The corner effect factor due to this structural restrain was established and denotes

as λs. λs is dependent on the wall fixities at the corner, and it shows to some

extent how the effect of a corner is propagated. This is true not only at the wall

top, but also for the entire depth of the wall.

7) The variation of wall movement along the distance from corner inevitably results in

the variation of earth pressures along this axis. The variation of earth pressure

γh
effect can be correlated with the stability number, , to better represent the
Cu

geotechnical factors influencing how a corner affects an excavation. This

contribution on the corner effect was quantified separately from that of structural

restrain, using λa, which is the corner effect factor due to geotechnical factors. A

semi-empirical relationship to calculate λa was given by Equation 5.26.

6.3 EFFECT OF A CORNER IN A THREE-DIMENSIONAL EXCAVATION

From the above findings, the 3D corner effect measured based on the wall displacement

can be expressed as:

δ 3 D = δ 2 D * λ s * λa (6.1)

Where δ3D = Wall top displacement within the 3D corner effect influence zone

δ2D = Wall top displacement in 2D condition

λs = Corner effect factor due to corner structural restrain

115
Chapter 6 Conclusions

λa = Corner effect factor due to geotechnical effect

Equation (6.1) is valid if the retaining wall is strong enough and the influence range of

the corner is given by the point where the plane strain condition starts. The general

expression considered the situation if the flexural capacity of the wall was exceeded, is:

La = min .{Lm , L2 D } (6.2)

Where Lm = Length from corner when lateral flexural capacity is exceeded.

L2D = Length from corner when plane strain condition is achieved.

For the present series of experimental results, the length of influence was confined to

the Lm zone, rather than the L2D. In the field, due to the small lateral flexural capacity

of wall, it is also likely that the Lm condition would occur. Based on Equation (6.2), the

expression for 3D corner effect of an excavation as shown in Equation (6.1), can be

revised to:

δ 3D = δ y * λ s * λ a (6.3)

Where δy = Wall displacement when the wall lateral flexural capacity is exceeded

The suitability of Equation (6.3) was examined using centrifuge experimental data.

The results show that the equation is able to estimate the corner effect influence to the

displacement at the top of wall, as well as surface settlement at 7m behind the wall.

This shows that the λs and λa are able to explain major aspects of 3D corner effects in

the excavations.

The findings and hypotheses obtained in the current study provide a better

understanding on how a corner affects an excavation, in particular the range of influence

116
Chapter 6 Conclusions

of a corner and the factors controlling how the corner propagates its presence. The

corner effect factors proposed, λs and λa, provided insight into how to quantify the

corner effect. Nevertheless, more in depth study is required to estimate λs and λa more

accurately, so that it can be applied into actual field condition.

6.4 RECOMMENDATION FOR FUTURE STUDIES

It is apparent from the above study that the influence range of a corner is dependent on

the lateral flexural capacity of wall which deserves more detailed study. In particular,

the characteristics of yielding related to the bending stresses need to be investigated to

allow for a more specific evaluation of the length of influence, Lm. In the actual

retaining wall construction, such as sheet pile, soldier pile, contiguous bored pile or

diaphragm walls, there are minimal lateral moment transfers between the piles/panels.

The lateral flexural capacities of the walls are usually much smaller than the flexural

capacity in the vertical direction. Hence, it is recommend that centrifuge modeling and

3D FEM analysis of retaining wall constructed by panels, instead of monolithic wall

used in the current study, be carried out in the future studies.

The current study was carried out for excavation with a right-angled (orthogonal)

corner, which is a typical case. Nevertheless, there are many excavations with non-

orthogonal walls. The variation of corner angle would affect the corner restrain

conditions as well as the earth pressure distribution around the corner. This in turn, will

affect the way the corner affects an excavation.

Another major factor not investigated in the current study is the effect of bracing on the

corner effect. While it is important, the task of conducting such studies in a centrifuge

117
Chapter 6 Conclusions

is non-trivial and would need additional development of robotics to carry out the

bracing. In spite of the difficulty, this is a very important aspect that needs more in-

depth look.

From the study, it was found that the corner effect in terms of relative ratio, is quite

insensitive to the excavation depth. However, in terms of absolute displacement

magnitude, the differences of displacement at various distances from corner actually

change according to the excavation depth. This is the reason why it is generally

perceived in the field that the corner effect is more pronounced at a greater excavation

depth. Also, the different excavation rate might influence the corner effect as the

consolidation of soil would change the magnitude of movement, hence influence the

absolute movement magnitude. But again, this should have insignificant effect to the

corner effect in terms of relative ratio. However, it is noted that the above findings

are derived from homogeneous soil. In cases where the soil have significant different

undrained shear strengths, the corner effect behaviour might change according to the

excavation depth. This might be worthwhile for future centrifuge studies. Also,

centrifuge experiments with sandy soil might be carried out to investigate the influence

of soil friction to corner effect.

It is noted that the 3D FEM analyses carried out in the current study have some

limitations. Firstly, the excavation dimension is anisotropy, and this might influence

the results. Secondly, the boundary the retaining wall is modeled as having a zero

slope, which is different from the centrifuge experiment where a rotation is allowed.

Thirdly, the retaining wall is assumed in-placed, rather than simulating the actual 1g

push-in wall. And finally, the drainage condition of the FEM model needs to further

118
Chapter 6 Conclusions

refined to better reflect the actual drainage condition of the centrifuge test model. The

above limitation is deemed acceptable in the current study as the main focus is to bring

out the salient mechanism of corner effect. But if the FEM results are to be applied

directly to an actual problem, the above limitations shall be fixed.

119
Table 2.1 Summary on Corner Effect Influence Range Reported/Proposed

a) Influence range deduced from case study and comparison with FEM analyses
Reference Corner effect Influence Range (m) Length (L)/ Remarks
Depth (h)
ratio

Nath (1983) Reported that beyond 2 x depth away 1.7 Trench


(h = 6m) excavation.
from corner, already plane strain No retaining
wall.

Ou and Chiou Reported that significant corner effect 1.0to 1.5 Top down
(1993) within 1x depth away from corner (Estimated diaphragm
from report) wall
construction.

Wong and Authors suggested corner effect 2.1 to 6.0 Multi-strutted


Patron (1993) presence at 1 x depth away from corner diaphragm
wall
Lee et al. Corner effect extended beyond 35m 3.0 to 4.5 Multi-strutted
(1998) from corner. Hence, corner effect (h = 17.3m) diaphragm
apparent beyond 35/17.3 = 2.0 x depth wall. Most
away from corner struts are
corner
diagonal struts

b) Influence range deduced solely based on FEM analyses


Ou et al. FEM analysis results: 2.5 Multi-strutted
(1996) For excavation width (B) = 20m, corner (h = 16m) diaphragm
effect up to 20m away from corner wall

For B = 100m, corner effect extended


beyond 50m away from corner. Hence, 6.3
corner effect apparent beyond 50/16 = (h = 16m)
3.12 x depth away from corner

Lin et al. Authors suggested corner effect 6.0 Multi-strutted


(2003) confined within 30m from corner. (h = 10m) diaphragm
Hence, corner effect within 30/10 = 3.0 wall
x depth away from corner

120
Table 3.1 Speed and Stroke of NUS’s 3D In-flight Excavator

Horizontal Direction Vertical Direction


Min. Speed 0.06 mm/s 0.40 mm/s
Max. Speed 31.0 mm/s 2.5 mm/s
Effective Stroke 170mm 170mm
Accuracy of Displacement ± 1.0mm ± 0.5mm

Operational environment: 100-g in the centrifuge

121
Table 3.2 Properties of Malaysian Kaolin clay

Properties Value
Specific Gravity (Gs) 2.60

Liquid Limit (WL) 80


Plastic Limit (WP) 35
Compression Index (Cc) 0.55
Swelling Index (Cs) 0.14
Permeability at 100 kPa on NC Clay 3.2 x 10-8 m/s

122
Table 4.1 Summary of In-flight Excavation Tests Carried out

No Test Dimension Type of Clay Wall Prototype Equivalent Wall

Code of Test Thickness Stiffness, EI Prototype Embedment

(kN/m2) Concrete

Wall

Thickness

(mm)

1 2DK-1 2D NC Kaolin Clay 1mm 6417 150 240mm

2 2DK-3 2D NC Kaolin Clay 3mm 173250 450 240mm

3 3DK-2c 3D NC Kaolin Clay 2mm 51333 300 240mm

4 3DK-1 3D NC Kaolin Clay 1mm 6417 150 240mm

5 3DK-2 3D NC Kaolin Clay 2mm 51333 300 240mm

6 3DK-3 3D NC Kaolin Clay 3mm 173250 450 240mm

7 3DK-2o 3D OC Kaolin Clay 2mm 51333 300 240mm

NOTES:
1) Embedment is the depth of model retaining wall penetration into the soil, before
excavation.
2) NC is clay with 20kPa 1-g surcharge loading, followed by 100-g self-weight
consolidation without surcharge on the surface
3) OC is clay with 230kPa 1-g surcharge loading, followed by 100-g self-weight
consolidation without surcharge loading on the surface
Additional Information

Symbol Legend c : without capping beam

K for Kaolin Clay o : Highly OC clay (1-g


consolidation to 230 kPa
before centrifuge
consolidation)
3D K – 2 c

3D for 3D Test Thickness of Model


2D for 2D Test Wall In mm

123
Table 5.1 Summary of soil profile and soil parameters used for FEM analyses

Depth Material Soil Model and properties used


(mm)
0 –10 Highly OC Clay Modified Cam Clay
Ko = 1.505, κ = 0.079, λ=0.244, ecs = 2.221, M=0.9, ν=0.3,
γbulk=16.7kN/m3, kx = ky = 2.0x10-8 m/sec

10 - 30 OC Clay Modified Cam Clay


Ko = 0.792, κ = 0.079, λ=0.244, ecs = 2.221, M=0.9, ν=0.3,
γbulk=16.7kN/m3, kx = ky = 2.5x10-8 m/sec

30 - 270 NC Clay Modified Cam Clay


Ko = 0.6012, κ = 0.079, λ=0.244, ecs = 2.221, M=0.9,
ν=0.3, γbulk=16.7kN/m3, kx = ky = 3.0x10-8 m/sec

Structure Material Soil Model and properties used

Model Aluminum Mohr-Coulomb


retaining
E=7.7 x 107 kPa, v =0.33, γbulk=16.7kN/m3,
wall
kx = ky = 1.0x10-15 m/sec

124
Table 5.2 Initial Stress Conditions of FEM Analyses

Strata Soil Vertical σx ’ σy’ σz’ Total Water Pc’ Remarks


Reference Type Coordinates Pressure, Uss
(m) kPa kPa kPa kPa kPa

1 Soil 0.27 0 0 0 0 20 At ground surface


clay
2 Soil 0.26 9.76 16.0 9.76 0 20 0.01m below ground. Assumed KO =
clay 0.61
3 Soil 0.00 107.93 176.94 107.93 255.06 176
clay

Note: The FEM analysis is simulating centrifuge test, with 100-G gravitational force σv ‘ = 0.01 * 100* 16
= 16 kPa
Soil Ground Level
Ground Water Table Level

0.27m 0.26m

σv ‘ = 16.0 + 0.26 *100


*(16.0 – 9.81)
= 176 94 kPa
σv = 0.27 * 100* 16.0 U = 0.26*100*9.81 125
= 432.0 kPa = 255.06 kPa
Table 5.3 Summary of Findings from Centrifuge Modelling and FEM analyses

S/N Description Centrifuge Modelling FEM Analyses Agreement/disagreement and elaboration


1 Wall thickness Corner effect increases with Corner effect increases with the Agreement.
effect the increase of wall increase of wall thickness.
thickness.
2 Capping beam Capping beam increases the Capping beam increases the Agreement.
effect wall stiffness hence enhance wall stiffness hence enhance
the corner effect. the corner effect.
3 Wall stiffness on a) For thin wall (1mm and a) Not able to model the lateral Disagreement.
corner effect 2mm), “constant movement” yielding of wall.
influence range range developed beyond If the retaining wall flexural capacity is strong enough, the
about 10m from the corner b) The dimension of FEM influence range of the corner is given by the point where
due to lateral yielding of mesh able to enlarge until the the 2D plane strain condition starts, as shown by the 3D
wall. plane strain condition is FEM analyses. In this case, the influence range is
achieved. The influence range independent of the wall’s flexural capacity or whether
b) For thicker wall (3mm), is 70m, which is much larger there is a presence of ground capping beam.
the corner effect still felt at than the “constant movement”
the edge of the wall furthest range registered in the If the wall is not strong enough, the influence range is
from the corner. This is centrifuge modelling. decided by the point when the flexural capacity of the wall
limited by the centrifuge was exceeded, which is at a shorter distance from the
container size. c) The influence range is the corner. This is shown in the centrifuge modelling.
same for all wall thickness or However, FEM unable to show this behaviour. A general
whether there is a presence of equation explaining the corner effect influence range
ground capping beam. considered the situation if the flexural capacity of wall is
exceeded is then established (Eq. 5.20).

The above finding highlighted the advantage of centrifuge


modelling of able to show the actual mechanism.
4 Geotechnical Geotechnical effect Geotechnical effect Agreement.
effect characterized by the stability characterized by the stability
γh γh
number S = influencing number S = influencing
Cu Cu
the corner effect. the corner effect.

126
Fig. 1.1 A schematic diagram showing a typical excavation carried out in the field

127
Notes: (1) Zone I - Well-braced excavations with slurry wall or substantial berms left permanently in place.
(2) Zone II - Excavation with temporary berms and raking strut support.
(3) Zone III – Excavations with ground loss from caission construction or insufficient wall support.
(4) In this context, the term ‘berm’ was used for a passive buttress.

Fig. 2.1 Observed settlements behind strutted excavation in Chicago


(after O’ Rourke et al., 1976)

Notes: (1) Zone I – Sand and soft to hard clay, average workmanship.
(2) Zone II – (a) Very soft to soft clay.
(i) Limited depth of clay below bottom of excavation.
(ii) Significant depth of clay below bottom of excavation but Nb < 5.14.
(b) Settlement affected by construction activities.
(3) Zone III – Very soft to soft clay to a significant depth below bottom of excavation and with Nb > 5.14.
where Nb = γH/Cub and Cub is as defined in Figure 31.
(4) The data used to derive the three zones shown in this figure are taken from excavation supported by soldier
piles or sheet piles with cross-lot struts or tie-backs.

Fig 2.2 Observed settlements behind excavation (after Peck, 1969)

128
Fig. 2.3 Relationship between factor of safety against basal heave and maximum lateral wall
movement from case histories (after Clough et al., 1979)

Fig. 2.4 Relationship between maximum ground settlements and maximum lateral wall
movement from case histories (after Mana & Clough, 1981)

129
Fig. 2.5 Apparent pressure diagrams for computing strut loads in braced cuts (after Terzaghi et al. 1996)

130
Fig. 2.6 Distress caused to a buried service by a shallow trenching operation
(after Needham and Howe, 1984)

131
h = fixed

Ng = increasing
σ = increasing
σ’ = unknown

Fig. 2.7 Schematic representation of the “Increasing-g” method to


simulate excavation

γliquid = γsoil
Ng = constant
σ = constant
σ’ = known

Fig. 2.8 Schematic representation of the “Heavy Liquid” method


to simulate excavation

132
Fig. 2.9 TIT’s in-flight excavator setup (after Kimura et al. 1994)

133
12
φ = 40°
10

Stability Factor, Ns = Hc . γ/c


30°
∝ 8
20°
10°
6

H 4

2 90° Hc

0
45 90 135 180
Corner Angle, ∝ (degrees)

Fig. 2.10b Stability factor Ns as a function of the


Fig. 2.10a Vertical cut with corner angle ∝ (after
corner angle ∝ (after Giger and Krizek, 1975)
Giger and Krizek, 1975)

134
Fig. 2.11 Stability Number versus Depth of Excavation Divided by Radius
(after Britto and Kusakabe, 1983)

135
Fig. 2.12 2D Section Used in Trench Excavation Analysis
(after De Moor, 1994)

136
a) Configuration of excavation case studied

b) PSR Chart: Relationship between B/L and distance from corner for various PSR

Fig. 2.13 PSR Chart (After Ou et al. 1996)

137
Fig. 2.14 Typical Analysis on a 10m long x 6m deep x 1m wide trench
excavation. Ground movement along line e – e parallel to trench and located
2m below surface: Comparison between results from three-dimensional and
plane strain analyses (after Nath, 1983)

138
Fig. 2.15 Variation of maximum wall displacement with the distance from for
constant sizes of complementary wall and various sizes of primary wall, L =
Length of primary wall; B = length of complementary wall. After Ou et al.
(1996)

139
Detachable
lift-shaft
Watertight centrifuge
container

Scrape

Fig. 3.1 3D in-flight excavator set-up on centrifuge

Stepper Motor

140
(1) Precision ball screw
(2) Bevel gear
(3) Trust bearing Scale
(4) Stepper motor 2
0 200 mm
(5) Detachable lift shaft
(6) Linear rail
(7) Scraper platform
(8) Scraper
(9) Soil-retaining gate
(10) Stepper motor 1
(11) Soil sample
(12) Excavated soil
(13) Retaining wall
(14) Perspex window

a) Front view b) Cut-section A - A

Fig. 3.2 Schematic diagrams of the 3D in-flight excavator

141
Data Acquisition
3D in-flight System
Indexers/
excavator set-up
drivers
Sliprings
Motor 1 Control
Terminal
Counter
Centrifuge balance
platform Motor 2 Control
Terminal

Video Camera
Monitor

Centrifuge Chamber Control Room

Fig. 3.3 Schematic representation of 3D in-flight excavator set-up on the centrifuge during testing

142
Indexer/Drivers for
stepper motors

Fig. 3.4 Indexers/drivers mounted on-board the centrifuge

silicone sealant

Fig. 3.5 Internal parts of indexers/drivers strengthened by silicone sealant.


The indexers/drivers was mounted near to the centrifuge rotating shaft to
minimized centrifugal force during spinning

143
43.5m

43.5m
17.0m
15.5m

Fig. 3.6 Schematic drawing of a corner of an excavation simulated

144
Detachable liftshaft (excavator)
from 3D in-flight excavator

150mm

Fig. 3.7 2D in-flight excavation test set-up on a narrow centrifuge


container

145
Fig. 3.8 Micro-concrete wall used in the study

Fig. 3.9 One of the aluminum alloy wall used in the study

146
Fig. 3.10 Plan view of the excavation and LVDT set-up for Test 3DK-2

Fig. 3.11 Locations of SGs and TSTs in the experiment: Test 3DK-2c
147
OCR Ratio

0 1 ∝

About
30mm 1)
2)
Depth from Ground

1) 20kPa laboratory floor consolidation to


achieve NC soil overlain by a thin layer of
OC soil sample after centrifuge spinning
without surcharge

2) 230kPa laboratory floor consolidation to


achieve OC soil sample after centrifuge
spinning without surcharge

Fig. 3.12 Schematic representation of OCR profiles intended

148
Fig. 3.13 De-airing of pore pressure transducers by boiling

Fig. 3.14 Installation of pore pressure transducers into soil sample

149
Stepper motor 1 controlling
vertical movement of cutting
blade

Cutting Vertical
blade LVDTs

Horizontal
LVDTs
Detachable
liftshaft

3.15a) Completed 3D in-flight excavation test set-up before transporting to centrifuge room

Miniature
camera on
camera frame

3.15b) 3D in-flight excavation test set-up after excavation test on centrifuge platform

150
a) Scraper
extended close to
retaining wall

b) Scraper lowered
to penetrate into soil
and then scrapping
soil towards the
liftshaft

c) Finished
scrapping. Scraper
in liftshaft

Fig. 3.16 Excavation in Progress (captured by miniature camera mounted in


front of the test sample)

151
0
Surface Settlement (mm) Model Scale

20

29.104mm

40

0 5000 10000 15000 20000


Elapsed Time (s)

Fig. Fig.
4.1 4.1 Surface
Surface settlement
settlement versusversus elapsed
elapsed reconsolidation
reconsolidation time3DK-3)
time (Test (Test 3DK-3)

800
Estimated final settlement = 1/0.0331022 = 30.209 mm

Average % consolidation = 29.104/30.209 = 96.3%


t/sett (s/volts)

400
Fit Equation:
Y = 0.0331022 * X + 22.427

0 5000 10000 15000 20000


Elapsed Time (s)
Fig. 4.2 Hyperbolic plot, Elapsed time/Settlement versus Elapsed Time (Test 3DK-3)
Fig. 4.2 Hyperbolic plot, Elapsed time/Settlement versus Elapsed Time (Test 3DK-3)

152
Unit Weight (kN/m3) Undrained Shear Strength, Cu (kPa)
16 16.5 17 17.5 18 0 10 20 30 40
0 0

5 5
Depth from Ground Level (m) Prototype Scale

Depth from Ground Level (m) Prototype Scale


1-g Vane Shear Test
10 10

15 15

Cu = 0.22 P’v
20 20

25 25

30 30

a) Density profile of NC soil used in the experiments b) Undrained shear strength profile of NC soil used in
the experiments
Fig. 4.3 Density and undrained shear strength profiles of NC soil used in the experiments

153
Fig. 4.4 Schematic plan view of model retaining wall edge at container wall
face

154
8
Horizontal LVDTs Layout (Plan)
Lateral Displacement (mm)

98

98
4
Primary wall investigated

0 10 20 30
Elapsed Time (mins)

Fig. 4.5 Responses of LVDTs installed at difference planes (Test 3DK-2)


Fig. 4.6 Responses of LVDTs installed at difference planes (Test 3DK-2D)

155
Fig. 4.6 Flow chart showing centrifuge tests conducted
156
165
70
35

Fig. 4.7a) Plan view of the excavation and LVDT set-out: Test 3DK-2c

Fig. 4.7b) Locations of SGs and TSTs in the experiment: Test 3DK-2c 157
Depth of Excavation (m) Prototype
0.00 1.00 2.00 3.00 4.00 5.00 6.00
0.00

0.10
Surface Settlement (m) Prototype

0.20

D=3, x=3.5
0.30 D=3, x=9.1
D=3, x=14.5
0.40 D=7, x=3.5
D=7, x=9.1
D=7, x=14.5
0.50 D = distance behind wall (m)
D=16.5, x=3.5
D=16.5, x=9.1 x = distance from corner (m)
0.60 D=16.5, x=14.5

0.70

Fig. 4.8 Test: 3DK-2c: Surface settlement at various location behind the retaining wall

158
Distance behind Wall (m) Prototype Scale
0 5 10 15 20
Surface Settlement (m) Prototype
0.00
0.10
0.20 1m exc

0.30 2m exc
0.40 3m exc
0.50 4m exc
0.60 5m exc
0.70

a) 3.5m from corner

Distance behind Wall (m) Prototype Scale


0 5 10 15 20
Surface Settlement (m) Prototype

0.00
0.10
0.20 1m exc
0.30 2m exc
0.40 3m exc
4m exc
0.50
5m exc
0.60
0.70
b) 9.1m from corner

Distance behind Wall (m) Prototype Scale


0 5 10 15 20
0.00
Surface Settlement (m) Prototype

0.10
0.20 1m exc
0.30 2m exc
3m exc
0.40
4m exc
0.50 5m exc
0.60
0.70
c) 14.5m from corner

Fig. 4.9 Surface settlement profiles behind wall, at various section from corner

159
D/h
0 2 4 6 8 10 12
0%

2%

4%

6%

8% 2DK-1
S/h

10% 2DK-3
3DK-2c: x =3.5
12%
3DK-2c: x = 9.1
14% 3DK-2c: x = 14.5

16% Upper bound of Zone III by Peck (1969) for very soft clay (for soldier
pile or sheet pile wall with cross-bracing or tiebacks)
18%

a) at 2m excavation

D /h
0 1 2 3 4 5 6
0%

5%

10%

15% 2D K -1
S/h

2D K -3
20%
3D K -2c: x =3.5
25% 3D K -2c: x = 9.1
3D K -2c: x = 14.5
30%
Upper bound of Zone III by Peck (1969) for very soft clay (for soldier
35% pile or sheet pile wall with cross-bracing or tiebacks)

b) at 4m excavation

Fig. 4.10 Surface settlement behind wall: Test 3DK-2c compare 2D tests and
published data (after Peck 1969)
S = Surface settlement
h = Depth of excavation
D = Distance behind wall
160
0.70

1m Exc
0.60 2m Exc
Surface Settlement (m) Prototype

3m Exc
0.50 4m Exc
5m Exc
0.40
3m behind wall

0.30

0.20
7m behind
wall
0.10

0.00
0.00 2.00 4.00 6.00 8.00 10.00 12.00 14.00 16.00
Distance from Corner (m) Prototype

Fig. 4.11 Surface settlement profiles at various distances


from corner: at 3.0m and 7.0m behind wall (Test 3DK-2c)

161
14.00

12.00
Distance From Corner (m)

10.00

8.00

6.00

4.00
4.00 6.00 8.00 10.00 12.00 14.00 16.00
Distance Behind Wall (m)

a) after 2m excavation

14.00

12.00
Distance from Corner (m)

10.00

8.00

6.00

4.00
4.00 6.00 8.00 10.00 12.00 14.00 16.00
Distance Behind Wall (m)

b) after 4m excavation

14.00

12.00
Distance from Corner (m)

10.00

8.00

6.00

4.00
4.00 6.00 8.00 10.00 12.00 14.00 16.00
Distance Behind Wall (m)
c) after 7m excavation
Fig. 4.12 Surface settlement contour behind retaining wall (Test 3DK-2c)
162
0.00

5.00
Depth from Ground (m) Prototype

10.00

15.00

20.00
3DK-2c, 1m exc

3DK-2c, 2m exc

25.00 3DK-2c, 3m exc

3DK-2c, 4m exc

30.00
0.00 0.10 0.20 0.30 0.40 0.50 0.60
Lateral Displacement of Wall (m) Prototype

Fig. 4.13 Wall deflection profiles at Perspex window (Test 3DK-2c)

150mm

5mm

Fig. 4.14 A typical scraping of 1 layer of soil in 3D in-flight excavation test

163
0.000
0.002
Settlement (m) Prototype
Incremental Surface
0.004
0.006 1.0m exc
0.008 2.0m exc
3.0m exc
0.010
4.0m exc
0.012
5.0m exc
0.014
0 20 40 60 80 100 120 140 160
Distance of Scraper from Wall (mm) Model Scale
a) at x = 3.5m from corner

0.000
0.002
Settlement (m) Prototype

0.004
Incremental Surface

0.006 1.0m exc


0.008 2.0m exc
0.010 3.0m exc
0.012 4.0m exc

0.014 5.0m exc

0 20 40 60 80 100 120 140 160


Distance of Scraper from Wall (mm) Model Scale
b) at x = 9.1m from corner

0.000
Settlement (m) Prototype

0.002
Incremental Surface

0.004
1.0m exc
0.006
2.0m exc
0.008 3.0m exc
0.010 4.0m exc
0.012
0.014
0 20 40 60 80 100 120 140 160
Distance of Scraper from Wall (mm) Model Scale
c) at x = 14.5m from corner

Fig. 4.15 Incremental surface settlement of a single scraping operation.


LVDTs at 7m behind wall, at various distances from corner (x)
164
0.0 0.0
Normalized Factor = 0.002197m prototype Normalized Factor = 0.005112m prototype
0.2 0.2
Normalized Incremental

Normalized Incremental
Surface Settlement

Surface Settlement
0.4 0.4

0.6 0.6

0.8 0.8 x = 3.5m


x = 3.5m
x = 9.1m x = 9.1m
1.0 1.0
x = 14.5m x = 14.5m
1.2 1.2
0 50 100 150 200 0 50 100 150 200
Distance of Scraper from Wall (mm) Model Scale Distance of Scraper from Wall (mm) Model Scale
a) 0.5m to 1.0m-excavation b) 1.5m to 2.0m-excavation
0.0 0.0
Normalized Factor = 0.007939m prototype Normalized Factor = 0.011265m prototype
0.2 0.2
Normalized Incremental

Normalized Incremental
Surface Settlement

Surface Settlement
0.4 0.4

0.6 0.6

0.8 0.8
x = 3.5m x = 3.5m
x = 9.1m x = 9.1m
1.0 1.0
x = 14.5m x = 14.5m
1.2 1.2
0 50 100 150 200 0 50 100 150 200
Distance of Scraper from Wall (mm) Model Scale Distance of Scraper from Wall (mm) Model Scale
c) 2.5m to 3.0m-excavation d) 3.5m to 4.0m-excavation
Fig. 4.16 Normalized incremental surface settlement at 7m behind wall of a single scraping operation: At
165
various distances from corner at different excavation stages (Test 3DK-2c)
0.000
Incremental Surface Settlement at the end
of each scraping stages (m) Prototype

0.002

0.004

0.006

0.008

0.010 x= 3.5m

x= 9.1m
0.012
x= 14.5m
0.014
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5
Depth of Excavation (m) Prototype

Fig. 4.17 The final settlement at each stage of scrapping.


x = distance from corner

166
120
P = measured lateral earth (kPa)

100

X13 Passive
X12 Active 8m
80
24m

X11 Passive
X10 Active TST-X10 Active Side
60 TST-X11 Passive Side
8m TST-X12 Active Side
*Prototype Scale
15m
TST-X13 Passive Side

40
0 1 2 3 4 5 6
Depth of Excavation (m) Prototype

Fig. 4.18 Lateral earth pressures versus depth of excavation. Total stress
transducers (TSTs ) at passive and active sides, at 8m and 15m from
corner

167
X13 Passive
X12 Active 8m

2.0 24m

X11 Passive
X10 Active
1.6

8m
*Prototype Scale
15m
1.2

P
K=
γy 0.8
TST-X10 Active Side

0.4
TST-X11 Passive Side
TST-X12 Active Side
TST-X13 Passive Side
0.0
0 1 2 3 4 5 6

Depth of Excavation (m) Prototype


P
Fig. 4.19a Coefficient of total stress lateral earth pressure coefficient, K = ,
γy
versus depth of excavation. P = measured lateral earth pressure, γ =
density of soil (16 kN/m3), y = depth of TST from soil level

1.5
Rankine’s Passive State
1.0

0.5

0.0
P − γy
2C u -0.5
Rankine’s Active State
-1.0

-1.5

-2.0
0 1 2 3 4 5 6
Depth of Excavation (m) Prototype

P − γy
Fig. 4.19b versus depth of excavation. P = measured lateral earth
2C u
pressure, γ = density of soil (16 kN/m3), y = depth of TST from soil
level, Cu = undrained shear strength at the TST level

168
Distance behind Wall (m) Prototype Scale
0 5 10 15 20 25
0.00

0.20

0.40

0.60
Surface Settlement (m) Prototype

0.80

2DK-1, 2m exc
1.00

2DK-1, 4m exc
1.20
2DK-3,2m exc
1.40
2DK-3,4m exc
1.60
3DK-2c, at 14.5m from corner, 2m exc
1.80
3DK-2c, at 14.5m from corner, 4m exc
2.00

Fig. 4.20 Surface settlement profiles behind wall. Test 3DK-2c compare 2DK-1 and 2DK-3

169
0.00
2m excavation

4m excavation

5.00
Depth from Ground (m) Prototyp

10.00

2DK-1, 2m exc: NF=0.480m


15.00 2DK-1, 4m exc: NF=1.600m

2DK-3, 2m exc: NF=0.148m

2DK-3, 4m exc: NF=0.884m

20.00 3DK-2c, 2m exc: NF=0.050m

3DK-2c, 4m exc: NF=0.299m

0.00 0.50 1.00 1.50 2.00 2.50 3.00


Normalized Lateral Displacementof Wall
(Normalized against Wall Top Displacement)

Fig. 4.21 Normalized wall deflection profiles. 3D test compare 2D tests

170
Excavate from 0.5 to 1.0m Excavate from 0.5 to 1.0m Excavate from 0.5 to 1.0m
0.0 0.0 0.0
NF2D = 0.029500m
0.4 0.4 0.4 NF3D = 0.002197m
Normalized Incremental Surface Settlement

Normalized Incremental Surface Settlement

Normalized Incremental Surface Settlement


3DK-2c
3DK-2c
3DK-2c
0.8 0.8 0.8
NF2D = 0.029500m 2D NF2D = 0.029500m
NF3D = 0.002197m NF3D = 0.002197m 2D
1.2 1.2 1.2 2D
0.0 Excavate from 1.5 to 2.0m 0.0 Excavate from 1.5 to 2.0m 0.0 Excavate from 1.5 to 2.0m
0 50 100 150Stabilization
200 0 50 100 150 200 0 50 100 150 200
0.4 of 2D 0.4 0.4
movement
0.8 Stabilization 0.8 0.8
NF2D = 0.046000m of 3D NF2D = 0.046000m NF2D = 0.046000m
1.2 NF3D = 0.005112m movement NF3D = 0.005112m 1.2 NF3D = 0.005112m
1.2
0.0 Excavate from 2.5 to 3.0m 0.0 Excavate from 2.5 to 3.0m 0.0 Excavate from 2.5 to 3.0m
0 50 100 150 200 0 50 100 150 200 0 50 100 150 200
NF2D = 0.046200m NF2D = 0.046200m
0.4 0.4 0.4
NF3D = 0.007940m NF3D = 0.007940m

0.8 NF2D = 0.046200m 0.8 0.8


NF3D = 0.007940m
1.2 1.2 1.2
0 50 100 150 200 0 50 100 150 200 0 50 100 150 200
Distance of scraper from wall (mm) Distance of scraper from wall (mm) Distance of scraper from wall (mm)

a) x = 3.5m b) x = 9.1m c) x = 14.5m

Fig. 4.22 Incremental surface settlement at 7m behind wall: 2D (Test 2DK-3) versus 3D (Tests 3DK-2 and 3DK-2c)
NF2D = Normalization factor for 2D test
NF3D = Normalization factor for 3D test

171
0.0
0.2 Test 3DK-3
0.4
a) 0.5 to 1.0m
excavation 0.6
Test 3DK-2
0.8
NF2D=0.07920m Test 2DK-3
1.0
NF3D=0.00684m
1.2
Test 3DK-1
1.4
1.6
0.0 0 20 40 60 80 100 120 140 160
0.2
Normalized Incremental Wall Top Displacement

0.4
b) 1.5 to 2.0m 0.6
excavation
0.8
NF2D=0.12089m 1.0
NF3D=0.02110m 1.2
1.4
0.0
1.6
0.2 0 20 40 60 80 100 120 140 160
c) 2.5 to 3.0m
excavation 0.4

NF2D=0.14888m 0.6
NF3D=0.05058m
0.8

1.0

1.2
0.0
0 20 40 60 80 100 120 140 160
d) 3.5 to 4.0m 0.2
excavation
0.4
NF2D=0.15566m
NF3D=0.06358m 0.6

0.8

1.0

1.2
0 20 40 60 80 100 120 140 160
Distance of scraper from wall (mm)

Fig. 4.23 Incremental wall top displacement: 2D Test (Test 2DK-3) versus
3D Test (Tests 3DK-1, 2 and 3), x = 9.1m
NF2D = Normalization factor for 2D test
NF3D = Normalization factor for 3D test
172
0.30

0.25

0.20
PSR-sett

0.15

0.10 1m Exc
2m Exc

0.05 3m Exc
4m Exc
0.00
0 2 4 6 8 10 12 14 16
x , Distance from Corner (m)

Fig. 4.24 PSR-sett at 7m behind wall versus distance from corner (x) Test 3DK-2c

173
h, Depth of Excavation (m) Prototype
0 1 2 3 4 5 6 7 8 9 10
0.00

0.05

0.10
Settlement (m) Prototype

3DK-1: x = 3.5m
0.15
3DK-1: x = 9.1m
0.20 3DK-1: x = 14.5m
0.25 3DK-2: x = 3.5m
3DK-2: x = 9.1m
0.30 Long null
3DK-2: x = 14.5m period after
0.35 5m-
3DK-3: x = 3.5m
excavation
3DK-3: x = 9.1m in this
0.40 experiment
3DK-3: x = 14.5m
0.45

Fig. 4.25 Surface settlement at 7m behind wall versus depth of excavation: 3D


tests with various wall thickness

0.7
3DK-1: x = 4.2m
3DK-1: x = 9.8m
0.6
δ, Wall top displacement (m) Prototype

3DK-1: x = 15.2m
3DK-2: x = 4.2m
0.5
3DK-2: x = 9.8m
3DK-2: x = 15.2m
0.4
3DK-3: x = 9.8m

0.3

0.2

0.1

0.0
0 1 2 3 4 5 6 7 8 9 10
h , Depth of Excavation (m)Prototype

Fig. 4.26 Lateral wall top displacement (δ) versus depth of excavation: 3D
tests with various wall thickness

174
1.2
3DK-1, 2m Exc
Wall top displacement ratio (1=0.3064m)

3DK-1, 4m Exc
1
3DK-2, 2m Exc
3DK-2, 4m Exc
0.8
3DK-3, 2m Exc
3DK-3, 4m Exc
0.6

0.4

0.2

-0.2
0 2 4 6 8 10 12 14 16
x, Distance from Corner (m)
Fig. 4.27 Lateral displacement at wall top at various distances from
corner: 3D tests with various thickness

1.2

1
Surface Settlement Ratio (1 = 0.1223m)

3DK-1, 2m Exc
0.8 3DK-1, 4m Exc
3DK-2, 2m Exc
3DK-2, 4m Exc
0.6 3DK-3, 2m Exc
3DK-3, 4m Exc

0.4

0.2

0
0 2 4 6 8 10 12 14 16
x, Distance from Corner (m)

Fig. 4.28 Surface settlement profiles at 7m behind wall at various


distances from corner: 3D tests with various thickness

175
14.00
x, Distance from corner (m) Prototype
12.00

10.00

8.00

6.00

4.00
8.00 10.00 12.00 14.00 16.00 18.00 20.00 22.00
Distance behind wall (m) Prototype
a) Test 3DK-1
x, Distance from corner (m) Prototype

14.00

12.00

10.00

8.00

6.00

4.00
8.00 10.00 12.00 14.00 16.00 18.00 20.00 22.00
Distance behind wall (m) Prototype
b) Test 3DK-2
14.00
x, Distance from corner (m) Prototype

12.00

10.00

8.00

6.00

4.00
8.00 10.00 12.00 14.00 16.00 18.00 20.00 22.00
Distance behind wall (m) Prototype
c) Test 3DK-3

Fig. 4.29 Surface settlement contour behind retaining wall after 7m-
excavation: 3D tests with various wall thickness
176
0.14

Surface Settlement (m) Prototype


0.12
1m Exc 0.20
0.10

Offset-Settlement (m) Prototype


2m Exc
0.08 3m Exc
0.16
4m Exc
0.06 5m Exc
6m Exc 0.12
0.04
7m Exc
0.02 0.08
0.00
0.04
0 5 10 15 20
x, distance from corner (m) Prototype
0.00
a) Typical surface settlement versus distance from
0 5 10 15 20
corner plots (Test 3DK-3). The settlement profiles x, distance from corner (m) Prototype
are projected to Y-Axis, and the intercepts are taken
as the settlement at corner for offset b) Offset-surface settlement versus x graph (Test 3DK-1)

0.20 0.06
Offset-Settlement (m) Prototype

Offset-Settlement (m) Prototype


0.16 0.05
0.04
0.12
0.03
0.08
0.02
0.04 0.01
0.00 0.00
0 5 10 15 20 0 5 10 15 20
x, distance from corner (m) Prototype x, distance from corner (m) Prototype
c) Offset-surface settlement versus x graph (Test 3DK-2) d) Offset-surface settlement versus x graph (Test 3DK-3)

Fig. 4.30 Illustration of offsetting surface settlements with settlement at the corner 177
A E
B
B’

F’

a) 3D view of a corner of an open excavation

Capping
Beam

Retaining wall

b) Section B-B’ - Typical 2D modelling configuration

Fig. 4.31 Sketch illustration of capping beam restrained at corner

178
Distance behind wall (m) prototype
0 5 10 15 20 25
Surface settlement (m) Prototype

0.00

0.02

0.04
3DK-2c, 2m exc
0.06 3DK-2c, 4m exc
3DK-2, 2m exc
0.08
3DK-2, 4m exc
0.10
a) Surface settlement behind wall: Test 3DK-2c compare Test 3DK-2
Cumulative surface settlement (m)

0.00

0.04 3DK-2: x = 3.5


3DK-2: x = 9.1
0.08 3DK-2: x = 14.5
3DK-2c: x = 3.5
0.12 3DK-2c: x = 9.1
3DK-2c: x = 14.5
0.16
0 1 2 3 4 5 6
Depth of excavation (m)
b) Surface settlement at 7m behind wall versus depth of excavation: Test
3DK-2c compare Test 3DK-2

80%
Sett (cap)/Sett (uncap) %

60%

40%
x = 3.5
20% x = 9.1
x = 14.5
0%
0 1 2 3 4 5 6
Depth of excavation (m)
c) Differences of settlement at 7m behind wall versus depth of excavation:
Test 3DK-2c compare Test 3DK-2

Fig. 4.32 Surface settlement: Test 3DK-2c compare Test 3DK-2 179
-10.00

-5.00

Ground surface
0.00
Depth from Ground (m) Prototype

5.00

10.00

15.00

3DK-2c, 1m exc
3DK-2c, 2m exc
20.00 3DK-2c, 3m exc
3DK-2c, 4m exc
3DK-2, 1m exc
3DK-2, 2m exc
25.00 3DK-2, 3m exc
3DK-2, 4m exc

30.00
0.00 0.05 0.10 0.15 0.20 0.25 0.30 0.35
Lateral Displacement of Wall (m) Prototype

Fig. 4.33 Wall deflection profiles at the edge of retaining wall (17m
from corner): Test 3DK-2c compare 3DK-2

180
Undrained Shear Strength, Cu (kPa)

0 10 20 30 40 50
0
Depth from Ground Level (m) Prototype Scale

10

15

20
NC (Test 3DK-2)

25 OC (Test 3DK-2o)

30

Fig. 4.34 Undrained shear strength profiles of NC and


OC soils determined by 1-g vane shear test

181
0.45
2m exc
Lateral Wall Top Displacement (m) Prototype

0.40
4m exc
NC (3DK-2)
0.35
6m exc
0.30
OC (3DK-2o)
0.25

0.20

0.15

0.10

0.05

0.00
0 5 10 15 20

x, Distance from corner (m) Prototype


Fig. 4.35a Lateral wall top displacement versus distance from corner. Soft
NC compares Stiff OC soils

1.20
Normalized Lateral Wall Top Displacement

1.00

0.80

0.60

0.40 3DK-2o: 2m exc. NF = 0.0537m


3DK-2o: 4m exc. NF = 0.1594m
3DK-2o: 6m exc. NF = 0.5041m
0.20 3DK-2: 2m exc. NF = 0.0115m
3DK-2: 4m exc. NF = 0.1927m
0.00 3DK-2: 6m exc. NF = 0.4120m

0 5 10 15 20
x, Distance from corner (m) Prototype

Fig. 4.35b Lateral wall top displacement normalized by dividing with


displacement at x = 15.2m (furthest from corner), at the particular
stage of excavation. Soft NC compares Stiff OC soils

182
0.08
NC (3DK-2)
2m exc
Offset-Settlement (m) Prototype

0.06
4m exc
OC (3DK-2o)
6m exc 0.084m
0.04
0.041m

0.02

0.00
0 5 10 15 20
x, Distance from corner (m) Prototype
Fig. 4.36a Offset-settlement versus distance from corner. Soft NC compares
Stiff OC soils

1.20

1.00
Normalized Offset-Settlement

0.80

0.60

3DK-2o: 2m exc. NF = 0.0033m


0.40
3DK-2o: 4m exc. NF = 0.0208m
3DK-2o: 6m exc. NF = 0.0527m
0.20 3DK-2: 2m exc. NF = 0.0048m
3DK-2: 4m exc. NF = 0.0250m
3DK-2: 6m exc. NF = 0.0763m
0.00
0 5 10 15 20
x, Distance from corner (m) Prototype
Fig. 4.36b Offset-settlement normalized by dividing with displacement at x =
14.5m (furthest from corner), at the particular stage of excavation.
Soft NC compares Stiff OC soils

183
Additional 27 wall elements
to model wall with protrusion

Wall elements

Fig. 5.1 Typical 1000 elements FEM mesh used in the study

184
Deflection (m) Prototype Deflection (m) Prototype
0.00 0.20 0.40 0.60
0.00 0.10 0.20 0.30 0.40 0.50 0.60
-5
-5

These 2 points not


so clear in the 0
0 experiments

5
5

Prototype Depth (m)


Prototype Depth (m)

10 10

15 15 3DK-2D: 2m Exc (Exp)


3DK-1D: 2m exc (Exp)
3DK-2D: 4m Exc (Exp)
3DK-1D: 4m exc (Exp)
20 20
3DK-1D: 2m exc (FEM) 3DK-2D: 2m exc (FEM)

3DK-1D: 4m exc (FEM) 3DK-2D: 4m exc (FEM)


25 25
a) Test 3DK-1 b) Test 3DK-2

Deflection (m) Prototype Deflection (m) Prototype


0.00 0.10 0.20 0.30 0.40 0.50 0.60 0.00 0.20 0.40 0.60
-5 -5

0 0

5 5
Prototype Depth (m)

Prototype Depth (m)

10 10

3DK-3D: 2m Exc too small


3DK-3D:
to capture 2m Exc
in experiment
15 15 3DK-2Dc: 2m exc (Exp)
3DK-3D: 4m Exc (Exp)
3DK-2Dc: 4m exc (Exp)

20 3DK-3D: 2m exc (FEM) 20 3DK-2Dc: 2m exc (FEM)


3DK-3D: 4m exc (FEM)
3DK-2Dc: 4m exc (FEM)
25 Fig. 5.3 FEM predicted wall deflection s 25
c) Test 3DK-3 d) Test 3DK-2c

Fig. 5.2 Wall deflection profiles at the edge (on Perspex window in the experiment)

185
Fig. 5.3 Configuration of excavation case studied by Ou et al. (1996)

Distance from corner (x)


Length of primary wall (PL)
Length of
complementary
wall (B)

Fig. 5.4 Schematic illustration of size of excavation studied

186
Fig. 5.5 Variation of maximum wall displacement with the distance from for
constant sizes of complementary wall and various sizes of primary wall, L =
Length of primary wall; B = length of complementary wall. After Ou et al.
(1996)

187
0.80 0.80
5m exc. at plane strain condition 5m exc. at plane strain condition

Wall Top Displacement (m) Prototype


Wall Top Displacement (m) Prototype
0.70 0.70

0.60 0.60
1m Exc (Exp)
(3D Exp)
0.50 0.50
4m exc. 4m exc. 2m Exc (Exp)
(3D Exp)
0.40 0.40

0.30 3m exc. 0.30 3m exc. (3D Exp)


3m Exc (Exp)
0.20 0.20
2m exc. 2m exc.
4m Exc (Exp)
(3D Exp)
0.10 1m exc. 0.10 1m exc.

0.00 0.00 5m Exc (Exp)


(3D Exp)
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 0 5 10 15 20 25
Distance from Corner (m) Distance from Corner (m)
1m exc (FEM)
a) PL = 17m (actual experimental length) b) PL = 25m
0.80 0.80
2m exc (FEM)
5m exc. at plane strain condition 5m exc. at plane strain condition
Wall Top Displacement (m) Prototype

Wall Top Displacement (m) Prototype


0.70 0.70

Reached plane strain condition


3m exc (FEM)
0.60 0.60
Reached plane strain condition

0.50 0.50 4m exc (FEM)


4m exc. 4m exc.
0.40 0.40
5m exc (FEM)
0.30 3m exc. 0.30 3m exc.
0.20 2m exc (2D Exp)
2m exc. 0.20 2m exc.
0.10 1m exc. 0.10 1m exc. 3m exc (2D Exp)
0.00
L2D
L2D 0.00
0 5 10 15 20 25 30 35 40 45 50 55 60 65 0 10 20 30 40 50 60 70 80 90 100 110 120 130
Distance from Corner (m) Distance from Corner (m)
c) PL = 65m d) PL = 130m
188
Fig. 5.6 Wall top displacement at various distances from corner, FEM
analyses with primary wall length (PL) = 17, 25, 65 and 130m
1.20 Corner effect influence range, L2D

1.00

0.80
PL = 65m
PL = 130m
PSR

0.60

1m Exc (FEM)
0.40
2m Exc (FEM)
3m Exc (FEM)
0.20
4m Exc (FEM)
5m Exc (FEM)
0.00
0 10 20 30 40 50 60 70 80 90 100 110 120 130
xd = Distance from Corner (m)

Fig. 5.7 PSR versus distance from corner plot: FEM analyses results. PL = 65
and 130m

189
1.2
L2D = 70m
1.0

0.8
PSR

0.6 1m exc: Max. Disp = 0.0680m


PSR 2m exc: Max. Disp = 0.1290m
0.4 = 0.92 PSR 3m exc: Max. Disp = 0.2593m
= 0.78
4m exc: Max. Disp = 0.6138m
0.2
5m exc: Max. Disp = 1.5450m
0.0
0 20 40 60 80 100 120 140
xd(m)
(m)
a) 3DK-1-130 (1mm thick wall in model scale)

1.2
L2D = 70m
1.0

0.8
PSR

0.6
1m exc: Max. Disp = 0.0670m
0.4 2m exc: Max. Disp = 0.1386m
3m exc: Max. Disp = 0.2441m
0.2 4m exc: Max. Disp = 0.4144m
5m exc: Max. Disp = 0.6853m

0.0
0 20 40 60 80 100 120 140
x (m)
d (m)
b) 3DK-2-130 (2mm thick wall in model scale)
1.2
L2D = 70m
1.0

0.8
PSR

0.6 1m exc: Max. Disp. = 0.0680m


2m exc: Max. Disp. = 0.1416m
0.4 3m exc: Max. Disp. = 0.2366m
4m exc: Max. Disp. = 0.3674m
0.2
5m exc: Max. Disp. = 0.5523m
0.0
0 20 40 60 80 100 120 140
x (m)
d (m)
c) 3DK-3-130 (3mm thick wall in model scale)
Fig. 5.8 Variation of PSR against distance from corner, x
Wall with capping beam, with different wall stiffness
190
1.2
L2D = 70m
1.0

0.8
1m exc
PSR

0.6
2m exc
0.4 3m exc
4m exc
0.2
5m exc
0.0
0 20 40 60 80 100 120 140
x d , distance from Corner (m)
a) 3DK-1c-130 (1mm thick wall in model scale)

1.2
L2D = 70m
1.0

0.8
PSR

0.6
1m exc
0.4 2m exc
3m exc
0.2 4m exc
5m exc
0.0
0 20 40 60 80 100 120 140
d , distance from Corner (m)
x,
b) 3DK-2c-130 (2mm thick wall in model scale)

1.2
L2D = 70m
1.0
0.8
PSR

0.6
1m exc
0.4 2m exc
3m exc
0.2 4m exc
5m exc
0.0
0 20 40 60 80 100 120 140
xd , distance from Corner (m)
c) 3DK-3c-130 (3mm thick wall in model scale)

Fig. 5.9 Variation of PSR against distance from corner, x


Wall without capping beam, with different wall stiffness 191
1.00
0.90
0.80
0.70
0.60
PSR

0.50 3D-1-130
0.40 3D-2-130
3D-3-130
0.30
3DK-1c-130
0.20 3DK-2c-130
0.10 3DK-3c-130

0.00
0 1 2 3 4 5 6
Depth of Excavation (m)

Fig. 5.10 PSR at a section of x = 30m against the depth of excavation.


For tests with and without capping beam

192
Primary wall
being studied Fixed end Primary wall
at corner being studied
x

Pa

Pp

a) corner excavation carried out in the b) assume primary wall is fixed at


experiments the corner

Fig. 5.11 Illustration of a simplified corner excavation

193
x
L Free edges
0 δ

h1

Pp Pa

y
Front View Cross Sectional View

Case 1b

Pp
Case 1a

Case 1 y 0

Pa
δ
b

UDL = q = Pp ( y ) − Pa ( y )

Case 2 x

Fig. 5.12 Schematic representation of the boundary and loadings conditions.

194
Retaining Wall
Retaining Wall

Load Case 1b: Load Case 1a:


Passive earth Active earth
Pp1 pressure pressure
Pa1
Pa
Pp
Pa2 Pp2

a) Idealized pressure diagram b) Simplified pressure diagram

Fig. 5.13 Active and passive earth pressure distribution on cantilever retaining wall

195
L

δ δ max

Fig. 5.14 Illustration of corner restrains effect

196
PSR or δ/δ2D

1.20

1.00
δ2D
Curve 1: FEM analysis modelling Test 3DK-2-130
0.80

3DK-2: 1m exc.
0.60
δy PSR-x/a

PSR-Fixed
0.40
PSR-Pin

0.20 PSR-Pin
Moment
L2D
Lm Ll
0.00
0 10 20 30 40 50 60 70 80
x(m)

Fig. 5.15 Normalized wall top displacement versus x profiles: FEM 3D excavation
modelling compare beam theories

197
S
L L

x
qx
δ= ( S 3 − 2Sx 2 + x 3 )
24 EI

Let S = 2L

qx
The maximum δ at L = δ max = (8L3 − 4 Lx 2 + x 3 )
24 EI

x  8 4  x  1  x  
2 3
δ
λs = =  −   +   
δ max L  5 5  L  5  L  

Fig. 5.16 Illustration and derivation of end restrain effect with a Pin-end condition

S
L L

Mox
δ= (2 S 2 − 3Sx + x 2 )
6 LEI

Let S = 2L

Mox
The maximum δ at L = δ max = (8 L2 − 6 Lx + x 2 )
12aEI

x  8 
2
δ x 1 x 
λs = =  −2 +   
δ max L  3 L 3 L 

Fig. 5.17 Illustration and derivation of end restrain effect with a Pin-Moment end
condition
198
0.35
1m exc
0.30 2m exc
3m exc
δ, wall top displacement (m) Prototype

0.25 4m exc

0.20

0.15

0.10

0.05

0.00
0 5 10 15 20

x, distance from corner (m) Prototype

Fig. 5.18 Wall top displacement versus distance from corner (x): Test 3DK-1
199
0.25
1m exc
2m exc
3m exc
δ, wall top displacement (m) Prototype

0.20
4m exc

0.15

0.10

0.05

0.00
0 5 10 15 20

x, distance from corner (m) Prototype

Fig. 5.19 Wall top displacement versus distance from corner (x): Test 3DK-2

200
0.09
1m exc
0.08
δ, wall top displacement (m) Prototype 2m exc
0.07 3m exc
4m exc
0.06

0.05

0.04

0.03

0.02

0.01

0.00
0 5 10 15 20

x, distance from corner (m) Prototype

Fig. 5.20 Wall top displacement versus distance from corner (x): Test 3DK-3

201
0.14
1m exc

0.12 2m exc
3m exc
Offset-Surface Settlement (m) Prototype

0.10 4m exc
5m exc
0.08 6m exc
7m exc
0.06

0.04

0.02

0.00
0 5 10 15 20
-0.02

x, distance from corner (m) Prototype

Fig. 5.21 Offset-surface settlement versus distance from corner (x): Test 3DK-1

202
0.12
1m exc
2m exc
Offset-Surface Settlement (m) Prototype 0.10 3m exc
4m exc

0.08 5m exc
6m exc
7m exc
0.06

0.04

0.02

0.00
0 5 10 15 20

x, distance from corner (m) Prototype

Fig. 5.22 Offset-surface settlement versus distance from corner (x): Test 3DK-2

203
0.06
1m exc
2m exc
0.05
3m exc
Offset-Surface Settlement (m) Prototype

4m exc
0.04 5m exc
6m exc
7m exc
0.03

0.02

0.01

0.00
0 5 10 15 20

X, distance from corner (m) Prototype

Fig. 5.23 Offset-surface settlement versus distance from corner (x): Test 3DK-3

204
PSR or δ/δ2D
1.20
Curve 2: Simplified curve if wall is
elastic (lateral flexural capacity not
1.00 exceeded)
δ2D
Curve 1: Actual curve if wall is elastic
0.80 (lateral flexural capacity not exceeded)
3DK-2: 1m exc. FEM

0.60 PSR-x/a
δy
PSR-Fixed
0.40
Cuve 3: Simplified PSR-Pin
curve if wall’s
0.20 lateral flexural PSR-Pin Moment
capacity exceeded

Lm Ll L2D
0.00
0 10 20 30 40 50 60 70 80
x(m)
Region studied in the
centrifuge tests

Fig. 5.24 Normalized wall top displacement versus x profiles: FEM 3D excavation
modelling compare beam theories

205
150 Theoretical moment capacity for 2mm wall = 420 kNm
Lateral Bending Moment (kNm) Prototype

100

50

-50 4.5m from GL, 1m exc


4.5m from GL, 2m exc
-100
4.5m from GL, 3m exc
-150 4.5m from GL, 4m exc
4.5m from GL, 5m exc
-200
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17
Distance from corner (m) Prototype

Fig. 5.25a Lateral bending moment versus distance from corner (x): Test 3DK-2

Bending moment increases with


distance from corner, and approach
to zero near to free end edge

BM
Fixed

Fig. 5.25b Theoretical lateral bending moment versus distance profile

206
Lm , Length from corner when

Additional length of wall


lateral yielding (m) occur
30

from corner for lateral


Measured

yielding to occur
20

10

1.7
1mm 2mm 3mm
Wall thickness
Test 3DK-1 Test 3DK-2 Test 3DK-3 (Model Scale)
Fig. 5.26a Measured Lm for Test 3DK-1 and hypothesis of Lm for tests with no
yielding (Tests 3DK-2 and 3DK-3)

lateral yielding to occur


wall from corner for
Additional length of
3
Lm/Lm1

1
1.7

E1 A1 E 2 A2 E 3 A3
E1 A1 E1 A1 E1 A1
Test 3DK-1 Test 3DK-2 Test 3DK-3
Ratio of EA
Fig. 5.26b Hypothesis on the relationship between Lm and wall area:
Normalized Lm versus Normalized EA, where
E = Elastic modulus of wall, same for all the test (E1 = E2 = E3)
A = Cross sectional area of the wall, which is total depth of wall * thickness
of wall. Hence:
A3 = 3 * A1 (A of Test 3DK-3 = 3 times of A of Test 3DK-1)
A2 = 2 * A1 (A of Test 3DK-2 = 2 times of A of Test 3DK-1)
Lm1 = Distance from corner when wall of Test 3DK-1 yielded = 10m
207
Curve 2: Simplified curve if wall is elastic (lateral
PSR or δ/δ2D flexural capacity not exceeded)
1.20

δ2D
1.00

0.80 Curve 1: Actual curve if wall is elastic


(lateral flexural capacity not exceeded)

δy
0.60

0.40 Cuve 3: Simplified


Normalized curve if wall’s
corner effects lateral flexural
towards δy capacity exceeded
0.20

0.00
0 5 10 15 20 25 30 35 40
x(m)

Fig. 5.27 Illustration showing the hypothesis of normalization of corner effects to achieved δy

208
X13 Passive
X12 Active
8m

24m

X11 Passive
X10 Active

8m *Prototype Scale
15m
= Total Lateral Earth Pressure Coefficient

2.0

Rankine’s Passive State


1.6

1.2

0.8 Rankine’s Active State


TST-X10 Active Side

0.4
TST-X11 Passive Side
TST-X12 Active Side
TST-X13 Passive Side
0.0
P
K= 0 1 2 3 4 5 6
γ y
Depth of Excavation (m) Prototype Scale

Fig. 5.28 Total lateral earth pressure coefficient at 8m and 15m from corner, at
active and passive sides, versus depth of excavation

209
Primary wall
being studied Fixed end Primary wall
at corner being studied
x

h1

h1
h2

h2

K oγh1

K aγh1

K pγh2

Fig. 5.29 Schematic diagram illustrating the variation of earth pressure effect at active side

210
120
Lateral Earth Pressure at Active Side,

100

80
Pa (kPa)

60
Pa ( 2 D )
2DK-3S
40 Pa (3 D )−8 m
3DK-2Dc: 8m from corner

20 3DK-2Dc: 15m from corner Pa ( 3 D ) −15 m

0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Depth of Excavation (m)

Fig. 5.30 Lateral earth pressure at the retained soil side versus depth of excavation

211
1.35

1.30
8m from corner (Exp) Pa ( 3 D ) − 8 m
1.25

1.20 15m from corner (Exp) P a ( 3 D ) − 15 m

Pa ( 3 D ) 1.15 R = 0.04
Pa ( 2 D ) Best fit curve for TST at 8m
from corner
1.10

1.05

1.00

0.95

0.90
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
γh
Cu
Pa ( 3 D ) γh
Fig. 5.31 versus for TSTs at 8m and 15m from corner for Test 3DK-2c 212
Pa ( 2 D ) Cu
1.35 8m from corner (Exp)
15m from corner (Exp)
1.30
1.7m from corner (FEM)
1.25 6.3m from corner (FEM)
Pa ( 3 D ) 11.8m from corner (FEM)
1.20
Pa ( 2 D ) 15.8m from corner (FEM)
1.15

1.10 R

1.05

1.00

0.95

0.90
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
γh
Cu
Pa ( 3 D ) γh
Fig. 5.32 versus for Test 3DK-2c (Experiment and FEM)
Pa ( 2 D ) Cu

0.1
0.09 FEM
0.08
Experiment
0.07
R = -0.0057x + 0.0857
0.06
R
0.05
0.04
R = -0.0031x + 0.04
0.03
0.02
0.01
0
0 2 4 6 8 10 12 14 16
x, distance from corner (m)

Fig. 5.33 Relationship between R and distance from corner

213
0.40
1m exc
2m exc
0.30
3m exc
4m exc

a) 0.20
δ (m)

0.10

0.00
0 5 10 15 20
x, distance from corner (m)

0.40

0.30
δ/λs (m)

b)
0.20

0.10

0.00
0 5 10 15 20
x, distance from corner (m)

0.40

0.30
δ/(λs *λa) (m)

c)
0.20

0.10

0.00
0 5 10 15 20
x, distance from corner (m)

Fig. 5.34 Wall top displacement versus distance from corner: Test 3DK-1
a) δ before normalization
b) δ normalized with λs 214
c) δ normalized with λa and λs
0.20
1m exc
2m exc
3m exc
4m exc
0.10
a)
δ (m)

0.00
0 5 10 15 20
x, distance from corner (m)
0.40

0.30

0.20
δ/λs (m)

b)

0.10

0.00
0 5 10 15 20

x, distance from corner (m)


0.40

0.30
δ/(λs *λa) (m)

c) 0.20

0.10

0.00
0 5 10 15 20
x, distance from corner (m)

Fig. 5.35 Wall top displacement versus distance from corner: Test 3DK-2
a) δ before normalization
b) δ normalized with λs 215
c) δ normalized with λa and λs
0.40
1m exc
2m exc
0.30 3m exc
4m exc
0.20
a)
δ (m)

0.10

0.00
0 5 10 15 20
x, distance from corner (m)

0.40

0.30
δ/λs (m)

b)
0.20

0.10

0.00
0 5 10 15 20
x, distance from corner (m)

0.40

0.30
δ/(λs *λa) (m)

c)
0.20

0.10

0.00
0 5 10 15 20
x, distance from corner (m)

Fig. 5.36 Wall top displacement versus distance from corner: Test 3DK-3
a) δ before normalization
b) δ normalized with λs 216
c) δ normalized with λa and λs
0.14
1m exc
0.12 2m exc
3m exc
Offset-Sett (m) 0.10 4m exc
5m exc
0.08 6m exc
7m exc
a) 0.06

0.04

0.02

0.00
0 5 10 15 20
x, distance from corner (m)

0.20

0.16
Offset-Sett /λs (m)

0.12
b)
0.08

0.04

0.00
0 5 10 15 20
x, distance from corner (m)
0.20

0.16
Offset-Sett /(λs *λa) (m)

0.12
c)

0.08

0.04

0.00
0 5 10 15 20
x, distance from corner (m)

Fig. 5.37 Offset-surface settlement at 7m behind wall versus x: Test 3DK-1


a) Offset-Sett before normalization
b) Offset-Sett normalized with λs 217
c) Offset-Sett normalized with λa and λs
0.12
1m exc
2m exc
3m exc
Offset-Sett (m)
0.08 4m exc
5m exc
6m exc
7m exc
a)
0.04

0.00
0 5 10 15 20
x, distance from corner (m)

0.20

0.16
Offset-Sett /λs (m)

0.12
b)
0.08

0.04

0.00
0 5 10 15 20
x, distance from corner (m)
0.24

0.20
Offset-Sett /(λs *λa) (m)

0.16
c)
0.12

0.08

0.04

0.00
0 5 10 15 20
x, distance from corner (m)

Fig. 5.38 Offset-surface settlement at 7m behind wall versus x: Test 3DK-2


a) Offset-Sett before normalization
b) Offset-Sett normalized with λs 218
c) Offset-Sett normalized with λa and λs
0.06
1m exc
2m exc
3m exc
Offset-Sett (m)
0.04 4m exc
5m exc
6m exc
7m exc
a)
0.02

0.00
0 5 10 15 20
x, distance from corner (m)

0.20

0.16
Offset-Sett /λs (m)

0.12
b)
0.08

0.04

0.00
0 5 10 15 20
x, distance from corner (m)
0.20

0.16
Offset-Sett /(λs *λa) (m)

0.12
c)

0.08

0.04

0.00
0 5 10 15 20
x, distance from corner (m)

Fig. 5.39 Offset-surface settlement at 7m behind wall versus x: Test 3DK-3


a) Offset-Sett before normalization
b) Offset-Sett normalized with λs 219
c) Offset-Sett normalized with λa and λs
1.0 2DK-1
2DK-3
3DK-1: x=4.2
0.8 3DK-1: x=9.8
3DK-1: x=15.2
3DK-2: x=4.2
0.6 3DK-2: x=9.8
a)
δ (m)
3DK-2: x=15.2
3DK-3: x=9.8
0.4

0.2

0.0
0 1 2 3 4 5 6
h, depth of excavation (m)
1.0

0.8

0.6
δ/λs (m)

b)

0.4

0.2

0.0
0 1 2 3 4 5 6
h, depth of excavation (m)
1.0

0.8
δ/(λs *λa) (m)

c)
0.6

0.4

0.2

0.0
0 1 2 3 4 5 6
h, depth of excavation (m)

Fig. 5.40 Wall top displacement versus depth of excavation: Tests 3DK-1, 3DK-2 and 3DK-3
a) δ before normalization
b) δ normalized with λs 220
c) δ normalized with λa and λs
0.60 2DK-3
3DK-1: x=3.5
3DK-1: x=9.1

Offset-sett (m)
0.50
3DK-1: x=14.5
0.40 3DK-2: x=3.5
3DK-2: x=9.1
a)
3DK-2: x=14.5
0.30
3DK-3: x=3.5
3DK-3: x=9.1
0.20 3DK-3: x=14.5

0.10

0.00
0 1 2 3 4 5 6 7
h, depth of excavation (m)

0.60

0.50
Offset-sett /λs (m)

0.40
b)
0.30

0.20

0.10

0.00
0 1 2 3 4 5 6 7
h, depth of excavation (m)
0.60
Offset-sett /(λs *λa) (m)

0.50

0.40
c)
0.30

0.20

0.10

0.00
0 1 2 3 4 5 6 7

h, depth of excavation (m)

Fig. 5.41 Offset-settlement versus depth of excavation: Tests 3DK-1, 3DK-2 and 3DK-3
a) Offset-Sett before normalization
b) Offset-Sett normalized with λs 221
c) Offset-Sett normalized with λa and λs
References

REFERENCES

Arul, M.B. and Kusakabe, O. Stability of axisymmetric excavations in clays. Journal

of Geotechnical Engineering. (109), No. 5: 666 – 681. 1983.

Atkinson, J.H. An Introduction to The Mechanics of Soils and Foundations: Through

Critical State Soil Mechanics. McGraw-Hill, 1993.

Balasubramanium. A.S., Loganathan. N., Fernando. G.S.K., Indraratna. B., Phien-wej.

N., Bergado. D.T., and Honjo. Y. Advanced Geotechnical Analysis. Report of

Asian Institute of Technology, Thailand. 1992.

Banerjee, P.K. and Yousif , N.B. A plasticity model for the anisotropically

consolidated clays. Int. J. for Numerical and Analytical Methods in

Geomechanics. (10), 521 – 541. 1986.

Banerjee, P.K., Kumbhojkar, A.S. and Yousif, N.B. Finite element analysis of the

stability of a vertical cut using an anisotropic soil model. Can. Geotech. J. (25),

119 – 127. 1988.

Bica, A.V.C. and Clayton, C.R.I. An experimental study of the behaviour of

embedded lengths of cantilever walls. Geotechnique (48), No. 6: 731 – 745. 1998.

Bjerrum, L. and Eide, O. Stability of excavation in clay. Geotechnique (6), 32 - 47.

1956.

Bjerrum, L. Embankments on soft ground. 5th PSC, ASCE, (2): 1 - 54.

Bolton, M.D. and Powrie, W. Behaviour of diaphragm walls in clay prior to

collapsed. Geotechnique, (38), No. 2: 167-189. 1988.

Bolton, M.D. and Powrie, W. The collapsed of diaphragm walls retaining clay.

Geotechnique, (37), No. 3: 335-353. 1987.

Bolton, M.D. and Stewart, D.T. The effect on propped diaphragm walls of rising

groundwater in stiff clay. Geotechnique, (44), No. 1: 111-127. 1994.

222
References

Bosscher, P.J. and Gray, D.H. Soil arching in sandy slopes. J. of Geotechnical

Engineering (112), No. 6 : 626 – 645. 1986.

Bowles, J.E. Foundation Analysis and Design. McGraw-Hill. 5th Edition, 1996.

Brassinga, H.E. and Van Tol, A.F. Deformation of a high-rise building adjacent to a

strutted diaphragm wall. Earth Retaining Structures and Deep Excavation. Proc.

XECSMFE, (2) : 787 – 790. 1991.

Britto, A.M. and Kusakabe, O. Stability of axisymmetric excavation in clays. J.

Geotechnical Engineering, (109), No. 5: 666 – 681. 1983.

Britto, A.M. and Kusakabe, O. Stability of unsupported axisymmetric excavation in

soft clay. Geotechnique (32), No. 3: 261-270. 1982.

Burland, J.B., Simpson, G.B., St John, H.D. Movement around excavation in London

Clay. National Press. 1979.

Chew, S.H., Yong, K.Y. and Lim, Y.K.A. Three-dimensional finite element analysis

of a strutted underlain by deep deposits of soft clay. 1998.

Chong, P.T., Tan, T.S., Lee, F.H., Yong, K.Y. and Tanaka, H. Characterisation of

Singapore lower clay by in-situ and laboratory tests. Proc. Of the Int. Symp. On

Problematic Soils, IS Tohoku, Sendai, Japan: 641-644. 1998.

CIRIA, Special publication 95. The design and construction of sheet piled

cofferdams. Thomas Telford. London, 1993.

Clough G.W. and Denby G.M. Stabilizing berm design for temporary walls in clay.

ASCE J. Geotech. Engng., (1), 13 – 29. 1977.

Clough, G.W. et al. Prediction of support excavation movements under marginal

stability conditions in clay. Proc. 3rd Int. Conf. Numerical Methods, Aachen, (4),

1485-1502. 1979.

Craig, R.F. Soil Mechanics. Chapman & Hall. 5th Edition. 1992. (Reprinted 1995).

223
References

Craig, W. H. Proc. Symp. On the application of centrifuge modelling to geotechnical

design. University of Manchester, UK. 1984 .

Craig, W.H. and Rowe, P.W. Operation of a geotechnical centrifuge from 1970 –

1979. Geotechnical Testing Journal, (4), No. 1, March 1981: 19-25.

Das, B.M. Principles of geotechnical engineering. 2nd Edition, PWS-Kent Publishing

Company. 1990.

De Moor, E.K. An analysis of bored pile/diaphragm wall installation effects.

Technical Note, Geotechnique, (44), No. 2: 341 –347. 1994.

Fernandes, M.M.A. Three-dimensional Analysis of Flexible Earth Retaining

Structures. Proc. Second NUMOG (Eds. G.N. Pande & W.F. van Impe), M.

Jackson & Sons, UK. 1986.

Fernandes, M.M.A., Cardoso, A.J.S, Trigo, J.F.C. and Marques, J.M.M.C. Finite

element modelling of supported excavations. Soil-Structure Interaction: Numerical

Analysis and Modelling (Ed. J.W.Bull), E & FN Spon. 1994.

Finno, R.J. and Nerby, S.M. Saturated clay response during braced cut construction.

J. Geotech. Engrg. ASCE, (115), No. 8: 1065 – 1084. 1989.

Finno, R.J., Lawrece, S.A., Allawh, N.F. and Harahap, I.S. Analysis of performance

of pile groups adjacent to deep excavation. Journal of Geotechnical Engineering.

(117), No. 6: 934-955. 1991

Fourie, A.B. and Potts, D.M. A numerical and experimental study of London Clay

subjected to passive stress relief. Geotechnique. (41), No. 1: 1 – 15. 1991.

Frydman, S., Baker, R. and Levy, A. Modelling the soil nailing – excavation process.

Proc. Centrifuge 94, Singapore: 669 – 674. 1994

Gere. J.M. and Timoshenko. S.P. Mechanics of materials. 3rd SI Edition. Chapman &

Hall. 1994.

224
References

Giger, M.W. and Krizek, R.J. Stability analysis of vertical cut with variable corner

angle. Soils and Foundations, (15), No. 2: 63 – 71. 1975.

Goh, T.L. Stabilization of an excavation by an embedded improved soil layer. PhD

Thesis, National University of Singapore, 2002.

Goldberg, D.T. Jaworski, W.E., Gordon, M.D. Lateral support systems and

underpining. Report FHWA-RD-75-128, (1), Federal Highway Administration,

Washington, D.C. Apr 1976.

Golden Software, Inc. Surfer (Win 32) Version 6.04, 1997.

Griffiths, D.V. & Koutsabeloulis, N. Finite element analysis of vertical excavations.

Computers and Geotechnics, (1) : 221-235. 1985.

Gunn, M.J. Satkunananthan, A. and Clayton, C.R.I. Finite element modelling of

installation effect. In Retaining Structures, 46-55. London: Thomas Telford.

(1993).

Hashash, Y.M.A. and Whittle, A.J. Ground movement prediction for deep excavation

in soft clay. J. of Geotechnical Engneering (122), No. 6: 474 – 486. 1996.

Hsi, J. P. & Small, J. C. Ground settlements and drawdown of the water table around

an excavation. Canadian Geot. J. (29): 740 – 756. 1992.

Jardine, R.J., Potts, D.M., Fourie, A.B. and Burland, J.B. Studies of the influence of

non-linear stress-strain characteristics in soil-structure interaction. Geotechnique

(36), No. 3: 377 – 396. 1986.

Jovicic, V. and Coop, M.R. The measurement of stiffness anisotropy in clays with

bender element tests in the triaxial apparatus. Geotech. Testing J. (21), No. 1: 3 –

10. 1998.

Khoo, E., Okumura, T. & Lee F.H. Side friction effects in plane strain models. Proc.

Centrifuge 94, Singapore: 115-120. 1994.

225
References

Kimura, T. Centrifuge research activities in Japan. Centrifuge in Soil Mechanics.

Graig, James & Schofield (eds), Balkema, Rotterdam. 1988.

Kimura, T., Takemura, J., Hiro-oka, A., Suemasa, N. & Kouda, N. Stability of

unsupported and supported vertical cuts in soft clay. Proc. 11th Southeast Asian

Geo. Conf. Singapore: 61-70. 1993.

Kimura, T., Takemura, J., Hiro-oka, A., Okamura, M. & Park. J. 1994. Excavation in

soft clay using an in-flight excavator. Proc. Centrifuge 94, Singapore: 649 – 654.

1994.

Konig, D., Jessberger, H.L., Chambon, P. and Dangla, P. Behaiour of a tunnel lining

embedded in a bentonite quartz flour water mixture in granular soil. Proc.

Centrifuge 94, Singapore: 705 – 711. 1994

Kusakabe, O. Stability of excavation in soft clay. Ph.D Thesis, Cambridge

University. 1982.

Kusakabe, O. Centrifuge model tests on the influence of an axisymmetric excavation

on buried pipes. Geotechnical Centrifuge Model Testing. Proc. Int. Symposium

on Geotechnical Centrifuge Model Testing. Tokyo: 87 – 93. 1984.

Ladd, C.C. and Foott, R. New design procedure for stability of soft clays. ASCE, J. of

GED, (100) (GT7): 763 – 783. 1974.

Lambe, T.W. and Whitman, R.V. Soil Mechanics, SI Version. John Wiley & Sons.

1979.

Lee, F.H. The National University of Singapore Geotechnical Centrifuge User

Manual. Research Report No. CE001. July 1992.

Lee, F.H., Tan, T.S., Yong, K.Y., Karunaratne, G.P. and Lee. S.L. Development of

geotechnical centrifuge facility at the National University of Singapore. Proc. Int.

Conf. Centrifuge 1991: 11-17. 1991.

226
References

Lee, F.H., Yong, K.Y. and Liu, K.X. Three-dimensional analyses of excavation in soft

clay. Proc. 11th African Regional Conf. Cairo’95, Egyption Geotechnical Society,

Cairo, Egypt. 1995.

Lee, F.H., Yong, K.Y., Quan, K.C.N. and Chee, K.T. Effect of corners in strutted

excavations: Field monitoring and case histories. Journal of Geotechnical and

Geoenvironmental Engineering, (124), No. 4 : 339-349. 1998.

Lee, K.M. and Rowe, R.K. An analysis of three-dimensional ground movements: the

Thunder Bay tunnel. Can. Geotech. J. (28): 25 – 41. 1991.

Leung, C.F., Lee, F.H. and Tan, T.S. Principles and applications of geotechnical

centrifuge model testing. Journal of the Institution of Engineers, Singapore, (31),

No. 4: 39-44. 1991.

Lin. D.G., Chung. T.C. and Phien-wej. N. Quantitative evaluation of corner effect on

deformation behavior of multi-strutted deep excavation in Bangkok subsoil. Journal

of the Southest Asian Geotechnical Society. 41 –57. 2003.

Liu, K. X. Three-dimensional analyses of deep excavation in Soft Clay, M.Eng.

Thesis, National University of Singapore. 1995.

Liu, K.X., Yong, K.Y. and Lee, F.H. A numerical study on 3-D behaviour of

excavation-support system. Proc. 2nd. Int. Conf. On soft soil engineering, Nanjing,

China. 1996.

Liu, W. Lee, K.M. and Zhang, S.D. Modelling of a large underground excavation in

China. Proc. Centrifuge 94, Singapore: 675 – 680. 1994

Loh, C.K., Tan, T.S. and Lee, F.H. Three dimensional excavation tests in the

centrifuge. Proc. Of the Int. Conf. Centrifuge98, Tokyo, Japan. 1998.

Lyndon, A. and Schofield, A.N. Centrifuge model test of a short term failure in

London clay. Geotechnique, (21): 440-442. 1970.

227
References

Mana A.I. and Clough G.W. Prediction of movements for braced cuts in clay. J.

Geotech. Engng, ASCE, (107), June: 759 – 777. 1981.

Nath, P. Trench excavation effects on adjacent buried pipes: Finite element study. J.

Geotechnical Engineering, (109), No. 11: 1399 – 1415. 1983.

Needham, D. and Howe, M. Why Pipes Fail III. Report E. 419, Engineering

Research Station, British Gas Corporation. 1984.

Ng, N. N. C. and Lings, .L. M. Effects of modelling soil nonlinearlity and wall

installation on back analysis of deep excavation in stiff clay. J. of Geotechnical

Engineering, (121), No. 10: 687 – 695. 1995.

Nomoto, T., Mito, K., Imamura, S. Ueno, K. and Kusakabe, O. A miniature shield

tunnelling machine for a centrifuge. Proc. Centrifuge 94, Singapore: 699 – 704.

1994.

O’Rourke, T.D., Cording, E.J. & Boscardin, M. The ground movements related to

braced excavation and their influence on adjacent buildings. U.S. Department of

Transportation, Report no. D0T-TST 76, T-23, 1976.

Onoue, A., Kazama, H., Hotta, H., Kimura, T. and Takemura, J. Behaviour of stacked-

drift-type tunnels. Proc. Centrifuge 94, Singapore: 687 – 692. 1994.

Ou, C. Y. and Chiou, D. C. Three-dimensional finite element analysis of deep

excavation. Proc. 11th Southeast Asian Geotechnical Conf. Singapore: 769-774.

1993.

Ou, C. Y., Chiou, D. C. & Wu, T. S. Three-dimensional finite element analysis of

deep excavations. J. of Geotechnical Engineering, (122) No. 5 : 337 – 345. 1996.

Ou, C.H. and Lai, C.H. Finite-element analysis of deep excavation in layered sandy

and clayey soil deposit. Can. Geotech. J. (31): 204 - 214. 1994

Peck. R.B. Deep excavations and tunnelling in soft ground. Prof. 7th Int. Conf.

228
References

S.M.F.E., Mexico, State-of-the-art volume: 225 –290. 1969.

Philips, R. Ground Deformation in The Vicinity of A Trench Heading. PhD Thesis,

Cambridge University. 1986.

Poulos and Chen. Pile response due to excavation-induced lateral soil movement.

Journal of Geotechnical and Geoenvironmental Engineering, (123), No. 2: 94 – 99.

1997.

Powrie. W. The behaviour of diaphragm walls in clay. PhD Thesis, Cambridge

University. 1986.

Powrie, W. Richards, D.J., and Kantartzi, C. Modelling diaphragm wall installation

and excavation processes. Proc. Centrifuge 94, Singapore: 655 – 661. 1994.

Powrie, W., Pantelidou, H. and Stallebrass, S.E. Soil stiffness in stress paths relevant

to diaphragm walls in clay. Geotechnique, (48), No. 4: 483 – 494. 1998.

Richards, D.J. and Powrie, W. Centrifuge model tests on doubly propped embedded

retaining walls in overconsolidated kaolin clay. Geotechnique, (48), No. 6: 833-

846. 1998.

Schofield, A.N. Cambridge geotechnical centrifuge operation. Geotechnique 30, UK.

1980.

Schofield, A.N. An introduction to centrifuge modelling. Centrifuge in soil

mechanics, Craig, James & Schofield (eds), Balkema, Rotterdam. 1988.

Scotts, R.S. Physical and numerical models. Centrifuge in soil mechanics, Craig,

James & Schofield (eds), Balkema, Rotterdam. 1988.

Sevenoaks, Kent. Deep excavations: a practical manual. 1996. Thomas Telford,

London.

Simpson, B. Retaining structure: displacement and design. Geotechnique. (42), No

7: 541 –576. 1992.

229
References

Smith, I. M. & Ho, D. K. H. Influence of construction technique on the performance

of a braced excavation in marine clay. Int. J. for Numerical and Analytical

Methods in Geomechanics, (16): 845-867. 1992.

St. John, H.D. Field and theoretical studies of the behaviour of ground around deep

excavation in London clay. Ph.D Thesis. Cambridge University. 1975.

Sun. J. H. Development of plate element for FEM modeling of deep excavation.

MEng Thesis, National University of Singapore. 2003.

Tan, T.S., Inoue, T. and Lee, S.L. Hyperbolic method for consolidation analysis.

Journal of Geotechnical Engineering, ASCE, (117), No. 11, November: 1723 -

1737. 1991.

Tan, T.S., Yong, K.Y., Lee, F.H. and Leung, C.F. Deep excavation problem in

Singapore. Proc. Of the Regional Symposium on Infrastructure Development in

Civil Engineering, Bankkok, Thailand. 1995.

Tan. T.S., Ng. T.G., French, D., Wong, F-H. and Takeda, T. Use of an improved soil

berm for stabilization in deep excavation. Proc. of Underground Singapore 2001,

Singapore. 2001.

Tanaka. H. Behaviour of a braced excavation in soft clay and the undrained shear

strength for passive earth pressure. Soils and Foundations. (34), No. 1: 53 – 64.

1994.

Taylor, R.N. (editor). Geotechnical Centrifuge Technology, Blackie Academic &

Professional. An imprint of Chapman & Hall. 1994.

Terzaghi, K., Peck, R.B. and Mesri, G. Soil Mechanics in Engineering Practice. John

Wiley & Sons. 3rd Edition. 1996.

Thanadol, K. Behaviour of an embedded improved soil berm in an excavation. PhD

Thesis, National University of Singapore, 2002.

230
References

Timoshenko. S.P. Strength of materials-part 2: advanced theory and problems. 3rd

Edition. D. Van Nostrand Company. 1958.

Tominaga, M., Hashimoto, T. and Fukuwaka, M. Application of stress path method to

a large excavation. Proc. 11th Int. Conf. On Soil Mechanics and Foundation

Engineering. 1985.

Toyoyawa, Y., Norii, N., Tamate, S., Hanayasu, S. and Ampadu, S.K. Deformation

and failure characteristics of vertical cuts and excavations. Proc. Centrifuge 94,

Singapore: 663 – 668. 1994

Trak, B., La Rochelle, P., Tavenas, F., Leroueil, S. and Roy, M. A new approach to

the stability analysis of embankments on sensitive clays. Canadian Geotechnical J.

(17), No. 4: 526-544. 1980.

Ugai, K. Three-dimensional stability analysis of vertical cohesive slopes. Soils and

Foundations, (25), No. 3: 41 – 48. 1985.

Wei, J. Centrifuge modelling of deep excavation, M.Eng Thesis, National University

of Singapore. 1998.

Whittle, A.J. Evaluation of a constitutive model for overconsolidated clays.

Geotechnique, (43), No. 2: 289 – 315. 1993.

Whittle, A.J. Prediction of Excavation performance in clays. In Civil Engineering

Practice: 65 – 88. 1997.

Whittle, A.J. and Hashash, Y.M.A. and Whitman, R.V. Analysis of a deep excavation

in Boston. J.Geotech. Engrg. (119), No. 1: 69 – 91. 1993.

Whittle, A.J., Degroot, D.J. Ladd, C.C. and Seah, T.H. Model prediction of

anisotropic behaviour of Boston blue clay. J. Geotech. Engrg. (120), No. 1: 199 –

224. 1994.

Whittle, A.J. and Kavvadas, M.J. Formulation of MIT-E3 constitutive model for

231
References

overconsolidated clay. J. Geotech. Engrg. (120), No. 1: 173 – 198. 1994.

Wong, L.W. and Patron, B.C. Settlements induced by deep excavation in Taipei.

Proc., 11th Southeast Asian Geotech. Conf., The Institution of Engineers, Malaysia,

KL., M’sia: 787 – 791. 1993.

Wroth, C.P. and Houlsby, G.T. Soil mechanics – Property characterization and

analysis procedures. Proc., 11th Int. Conf. On Soil Mechanics and Foundation

Engrg. San Francisco, (1), 1 – 55. 1985.

Yong, K. Y., Lee, F.H and Liu, K.X. Three dimensional finite analysis of deep

excavation in marine clay, Proc. 12th SEAG Conference and the 4th Int. Conf. On

Tropical Soil, Kuala Lumpur, Malaysia. (1), 435-440. 1997.

Yong, K.Y., Lee, F.H., Parnploy, U., and Lee, S.L. Elasto-Plastic Consolidation

Analysis for Strutted Excavation in Clay, Comput. Geotech., 8 (4), 311-328. 1989.

Zhang, S.D. and Zhang, H.D. Stability of deep excavation in soft clay. Proc.

Centrifuge 94, Singapore: 643 – 647. 1994.

232

You might also like