You are on page 1of 327

CENTRIFUGE MODEL STUDY OF

TUNNEL-SOIL-PILE INTERACTION
IN SOFT CLAY

ONG CHEE WEE

NATIONAL UNIVERSITY OF SINGAPORE


2009

CENTRIFUGE MODEL STUDY OF


TUNNEL-SOIL-PILE INTERACTION
IN SOFT CLAY

ONG CHEE WEE


(B.Eng.(Hons),UPM, M.Sc.(Civil), NUS, P.Eng.)

A THESIS SUBMITTED
FOR THE DEGREE OF DOCTOR OF PHILOSOPHY
DEPARTMENT OF CIVIL ENGINEERING
NATIONAL UNIVERSITY OF SINGAPORE
2009

DEDICATION

To my dearest parents, my caring wife and my lovely twins

ACKNOWLEDGEMENTS
It was been sweat and tears the past four years to complete my PhD and coupled with
countless trials and errors; the conclusion of this thesis has finally arrived! Foremost, I
would like to extend my heartfelt gratitude and thanks to a special and wonderful
person, none other than my main supervisor, Prof. Leung Chun Fai. To my cosupervisor, Prof. Yong Kwet Yew, who had throughout the research period, gave me
constructive comments which have stimulated the success of this thesis. I would like to
thank them for their willingness to share their vast experience and guidance. In
addition to this, my appreciation goes out to Prof. Chow Yean Khow for his advice in
the constant students group discussion held fortnightly.
Gratitude also goes out to Prof Tan Thiam Soon and A/Prof. Harry Tan Siew Ann for
their advices during my qualifying examination which has propelled me to move
forward towards a clearer and better direction. Prof. Tan always encouraged his
students to think out of the box. One day, when I was exhausted in innovating my new
centrifuge model tunnel (see Figure 3.6), I have suggested the owner of our
engineering canteen, Seton Lin to turn over my model tunnel and use it as cup holder
(as shown in the Figure 1 below). With this turned-over tunnel, you can save your
queuing time next time when you buy a cup of coffee or tea from NUS canteen. The
innovative centrifuge model tunnel is also part of the contributions from Mr. Wong
C.Y, Dr. Okky Purwana, Dr. Shen Rui Fu, Mr. Martin Loh and Ian Kit.

Figure 1 The turned-over model tunnel is now a cup holder in NUS Engineering canteen

I also wish to express my sincere gratitude and thanks to Asst. Prof. Goh Siang Huat,
who has been a source of endless ideas and inspiration, to Dr. Shen Ruifu, who has
been advising me for my centrifuge model test and interpretation of my test results,
and to Prof. K.K. Phoon, who has encouraged and supported me since I was reading
my M.Sc. back in year 2002. I will never forget the wonderful time that I spent with
him in organizing many conferences.
Special thanks are also extended to the technical staffs of NUS Geotechnical
Centrifuge Laboratory,Wong Chew Yuen, Mdm. Jamilah, Lye Heng, Shaja, John Choy,
Loo Leong Huat and Adrian Tan. I would like to extend my sincere gratitude to the
final year students that I had privilege of working with, i.e. Kai Yang (04/05), Eddie
Hu (05/06), Ian Kit and Eliza (06/07) and Teng Hui (07/08)
I also wish to credit the support of the following professionals, associates and friends
for sharing their experiences and knowledge namely Prof. Andrew Palmer, Prof. I.H
Wong, Dr. Dave White, Dr. Johny Cheuk, Mr. Nick Shirlaw, Mr. Ow Chun Nam, Dr.
ii

Jeyatharan Kumarasamy, Dr. Lin Kai Qiu, Mr. Cham Wee Meng, Mr. Lim Tuck Fang,
Mr. Cheang Yew Kee, Mr. Poh Chee Kuan, Mr. Koo Chung Chong, Mr. Lee Hong
Keow, Mr. Edmund Ng, Mr. Jonathan Ang, Mr. Phang Chu Mau, Mr. Chew Eik
Khoon, Mr. Jimmy Chew and Mr. Lau Kim Hwa.
I extend my appreciation to my many colleagues and friends who I have consulted
during the course of the research, particularly Chin Hong, Ran Xia, Xie Yi, Yonggang,
Xiying, San Chuin, Kheng Ghee, Kar Lu, Cheng Ti, Hung Leong, Czhia Yheaw,
Haibo, Teik Lim, Okky, Dominic, William, Wang Lei, Chen Jian, Feijian, Sindhu,
Chng Yi, Poh Hai, Karma, Subhadeep, Khrishna, Jiangtao, Andy, Chong Hun & Ben.
It was a memorable trip to Schofield Centre at Cambridge University in April/May
2007. Thank you for the kind arrangement of Prof. Robert Mair. My stay in Jesus
College was wonderful and fulfilling. I would also like to thank Prof. Malcolm Bolton
who has provided arrays of resource and pertinent pointers in improving my model and
research. To Prof. Andrew Noel Schofield, Prof. Kenichi Soga, Dr. Stuart Haigh, Dr. I
Thusyanthan, Dr. Johnson Chung, Sidney Lam, Fiona Kwok, Gui Chang Shin, Tricia
Lee, Alec Marshall, Richard Laver, Hisham Mohamad, Senthil, Christopher and Dr
Yueyang Zhao, all for giving me valuable advices and encouragement on my research.
Thank you Geotechnical Society of Singapore for the sponsorship to Bangalore, India
to attend the 6th Asian Young Geotechnical Engineers Conference (Dec 2008) in which
my paper co-authored with Prof. C. F. Leung and Prof. K. Y. Yong have won the Best
Contribution Award.
A special acknowledgement is dedicated to Wendy and Angelia for their help and
assistance in the compilation of this thesis.
I would like to extend my gratitude to my parents, my wife, Shin Inn and my twins,
Yee Heng and Yee Huan (born one month after I pursued my PhD), my sisters,
brother-in-laws, grandparents, uncles, aunts and cousins for their never-ending love,
support, tolerance and sacrifice in encouraging me to complete this research. A special
mention to my brother-in-law, Moong Khai Chee for his support, guidance and
technical advice ever since my undergraduate years.
It is also with great pride despite the hectic schedule of my PhD research, I have also
successfully passed the qualifying exams and registered myself as a Professional
Engineer with Board of Engineers Malaysia in 2008; passed the Professional Engineer
Fundamentals Engineering Examination, Singapore in 2007, and being registered as a
Resident Engineer with Building and Construction Authority, Singapore also in 2007.
To the remaining people whom I am unable to list down, please accept my sincere
appreciation and thanks for the feedback, assistance, tolerance and help rendered.
Last but not least, deepest appreciation and thanks to National University of Singapore
for the award of this research scholarship throughout the four years period for without
which this research program would not have been possible.
Ong Chee Wee
28-02-2009

iii

TABLE OF
CONTENTS

Dedication

Acknowledgements

ii

Table of Contents

iv

Summary

List of Tables

xii

List of Figures

xiii

Nomenclature

xxvi

CHAPTER 1

INTRODUCTION

1.1

Background

1.2

Tunnelling-Induced Soil Movements

1.3

Effects of Tunnelling on Piles

1.4

Objective and Scope of Study

1.5

Structure of Thesis

CHAPTER 2

LITERATURE REVIEW

2.1

Introduction

2.2

Techniques for Simulation of Tunnelling in Centrifuge

2.2.1 Simulation Technique 1 - High Density Polystyrene Foam

11

iv

2.3

2.4

2.5

2.2.2 Simulation Technique 2 - Compressed Air

12

2.2.3 Simulation Technique 3 - Liquid Oil / Water

14

2.2.4 Simulation Technique 4 - Mechanical Equipments

16

Tunnelling-Induced Soil Movements

17

2.3.1 Field Studies of Tunnelling-Induced Soil Movement

17

2.3.2 Centrifuge Model Tests of Tunnelling-Induced Soil Movement

22

2.3.3 Predictive Methods of Tunnelling-Induced Soil Movement

24

Tunnelling-Induced Pile Responses

29

2.4.1 Field Studies of Tunnelling-Induced Pile Responses

29

2.4.2 Centrifuge Model Tests of Tunnelling-Induced Pile Responses

33

2.4.3 Predictive Methods of Tunnelling-Induced Pile Responses

36

Summary

39

CHAPTER 3

EXPERIMENTAL SET-UP AND PROCEDURE

67

3.1

Introduction

67

3.2

Geotechnical Centrifuge Modeling

67

3.2.1 Principles of Geotechnical Centrifuge Modelling

67

3.2.2 NUS Geotechnical Centrifuge Facility

69

Experimental Set-Up

70

3.3.1 Model Tunnelling Technique

70

3.3

3.3.1.1 Advantages of Model Tunnel

72

3.3.1.2 Limitations of Model Tunnel

73

3.3.2

Instrumented Model Piles

74

3.3.3

Model Pile Cap

76

3.4

3.5

3.3.4 Strong Box

77

3.3.5

Kaolin Clay

77

3.3.6

Toyoura Sand

78

3.3.7

Potentiometer

78

3.3.8 Pore Pressure Transducers (PPT)

79

3.3.9 Non-Contact Laser Transducers

80

Image Acquisition System

80

3.4.1 High Resolution Camera

80

3.4.2 Lighting System

81

3.4.3 On-Board and Command Computers

81

3.4.4 Post-Processing of Images

82

3.4.5 Assessment of Effectiveness of Image Processing System

83

Experimental Procedure

84

3.5.1 Preparation of The Soil Sample

84

3.5.2 Pre-Consolidation Process

85

3.5.3 Installation of Model Tunnel and PPTs At 1g

86

3.5.4 Preparation Works for PIV Analysis

87

3.5.5 Installation of Model Pile at 1g

87

3.5.6 Test Procedure

88

CHAPTER 4

BASIC TESTS ON VOLUME LOSS

102

4.1

Introduction

102

4.2

Test Program

102

4.3

Tunnelling-Induced Soil Movements (Tests 1, 2)

104

vi

4.4

4.5

4.6

4.3.1 Cumulative Soil Movements

104

4.3.2 Soil Surface Settlement Troughs

105

4.3.3 Subsurface Vertical Soil Movements

107

4.3.4 Subsurface Horizontal Soil Movements

110

4.3.5 Qualitative Assessment On Excess Pore Pressure Response

111

Typical Tunnel-Soil-Piles Interactions (Test 3)

112

4.4.1 Induced Axial Force and Settlement

113

4.4.2 Induced Bending Moment and Deflection

115

Test Series 1- Effects of Volume Loss (Tests 3, 4)

118

4.5.1 Induced Axial Force and Settlement

118

4.5.2 Induced Bending Moment and Deflection

119

Concluding Remarks

122

4.6.1 Tunnelling-Induced Soil Movements

122

4.6.2 Tunnel-Soil-Piles Interaction

123

CHAPTER 5

EFFECTS OF TUNNELLING ON SINGLE PILES

151

5.1

Introduction

151

5.2

Test Series 2- Effects of Pile Tip & Head Conditions (Tests 3, 9, 10, 13)

151

5.2.1 Effects of Pile Tip Conditions

151

5.2.2 Effects of Pile Head Conditions

154

5.3

Test Series 3 - Effects of Pile Length (Tests 3, 7, 8)

157

5.4

Effects Of Distance of Pile From Tunnel

162

5.4.1 Test Series 4 - Free-Head Floating Piles (Tests 3, 5, 16, 6)

162

vii

5.4.2 Test Series 5 - Free-Head End-Bearing Piles (Tests 10, 11, 12)

165

5.4.3 Test Series 6 - Fixed-Head End-Bearing Piles (Tests 13, 14A, 14B)

166

5.4.4 Comparison of Results from Test Series 4, 5 And 6

167

5.5

Effects of Time on Pile Responses in Soft Clay

171

5.6

Comparison of Soil and Single Pile Behaviours due to Inward and


Outward Tunnel Deformations

172

5.6.1 Tunnel-Soil Interaction

172

5.6.1.1 Similarities (Tunnel-Soil Interaction)

173

5.6.1.2 Differences (Tunnel-Soil Interaction)

174

5.6.2 Tunnel-Pile Interaction

5.7

174

5.6.2.1 Similarities (Tunnel-Pile Interaction)

175

5.6.2.2 Differences (Tunnel-Pile Interaction)

175

Concluding Remarks

CHAPTER 6

EFFECTS OF TUNNELLING ON PILE GROUPS

177

208

6.1

Introduction

208

6.2

Floating Pile Group

209

6.2.1 Induced Axial Force and Settlement

210

6.2.2 Induced Bending Moment and Deflection

211

End-Bearing Pile Group

215

6.3.1 Capped-Head

215

6.3.2 Fixed-Head

218

Pile Group Size

220

6.4.1 Capped-Head

220

6.3

6.4

viii

6.5

6.6

6.4.2 Fixed-Head

225

Pile Cap Conditions

227

6.5.1 2-Pile Group

227

6.5.2 6-Pile Group

229

Concluding Remarks

233

CHAPTER 7

7.1

7.2

CONCLUSIONS

281

Concluding Remarks

281

7.1.1 Technique for Simulation of Tunnelling

282

7.1.2 Tunnel-Soil Interaction

283

7.1.3 Tunnel-Single Piles Interaction

283

7.1.4 Tunnel-Pile Groups Interaction

286

Recommendations for Future Studies

288

REFERENCES

290

ix

SUMMARY
Tunnels are often constructed close to existing pile foundation in dense urban areas. It
is challenging to carry out extensive instrumentation and monitoring in the field to
observe the pile responses due to tunneling activities. Hence, centrifuge modelling
emerges as an attractive alternative option to investigate the effects of tunnellinginduced soil movement on adjacent piles. In the present study, a modeling technique
was developed to simulate the inward tunnel deformation due to over-excavation of
tunnel.

Phase 1 of the study was performed to investigate the free-field soil movements due to
tunneling. It is found that the surface settlement trough in clay generally follows the
Gaussian distribution curve in the short-term. The magnitude of maximum ground
surface settlement increases with time and tunnel volume loss. Though the settlement
magnitude is larger in the long-term, the settlement trough is wider and hence the
differential settlement at the ground surface is not as serious as compared to that in the
short-term. The data confirmed that the empirical equation proposed by Mair et al
(1993) is applicable in the prediction of the subsurface settlement troughs in clay in the
short-term. In the short-term, an Immediate Shear Zone with large soil movement
above the tunnel can be identified, while the zone outside the immediate shear zone is
identified as Support Zone. In the long-term, soil settlement is noted to be more
dominant than lateral soil movement.

Phase 2 of the study was conducted to study the tunnelling-induced single pile lateral
and axial responses in both short- and long-term. The effects of factors such as volume
loss, pile tip and head condition, pile length and pile-to-tunnel distance were examined.

It was found that a floating pile is mainly governed by pile settlement and a socketed
pile is likely governed by its material stress when tunneling was carried out adjacent to
it. It was noted that tensile force and relatively large negative bending moments are
induced at the pile head due to total fixity resulting in a reduction in drag load, and
positive bending moment at the mid-pile shaft.

The centrifuge model study was subsequently extended to pile groups to evaluate the
effects of number of piles, pile cap and pile tip condition. It was found that the pile
group is generally beneficial as the average pile responses of a pile group due to
tunneling are smaller than the average of those of single piles at the same locations.
The pile-cap-pile interaction in a capped-head 6-pile group would moderate the
induced pile bending moments among the piles within a pile group. It was noted that
the induced pile bending moments in the middle row is smaller than that of rear row
which is contrary to the induced lateral soil movements at the respective locations. For
the piles in a totally fixed-head 6-pile group, the piles behaved like single piles
standing side by side in terms of axial force and bending moment, except that the
magnitude is affected by the total number of piles in the group.

A common trend was observed for the long-term over short-term ratio of pile
responses for both single pile and pile group. The results reveal that soil and pile
responses increase over time with long-term over short-term pile responses ratio
ranging from 1.32 to 2.4, regardless of pile size, pile head and toe conditions.

Keywords: tunnel; pile; interaction; axial force; settlement; bending moment;


deflection; clay; centrifuge modelling

xi

LIST OF TABLES

Table 3.1 Scaling relation of centrifuge modeling (After Leung et at, 1991)
Table 3.2 Physical properties of Malaysian kaolin clay (After Goh, 2003)
Table 3.3 Physical properties of Toyoura sand (After Teh et. al, 2005)
Table 4.1 Test program and parameters for the basic tests on volume loss
Table 5.1 Test program and prototype parameters in Phase 2 study
Table 6.1 Test program and prototype parameters for pile group tests

xii

LIST OF FIGURES

Figure 1.1

Pile responses induced by tunnel construction: (a) Tunnelling under


pile foundation, (b) Tunnelling adjacent to pile foundation.

Figure 1.2

Pile foundations supported existing buildings normally designed to


resist compression load only.

Figure 2.1

Simulation technique of tunnelling using high density polystyrene


foam (After Sharma et al., 2001 and Feng, 2003)

Figure 2.2

Simplified tunnel lining deformation with time by simulation


technique of tunnelling using high density polystyrene foam (After
Ran, 2004)

Figure 2.3

Simulation technique of tunnelling - applying compressed air (After


Grant & Taylor, 2000)

Figure 2.4

Simulation technique of tunnelling - model tunnel infilled with oil


(After Loganathan et al., 2000)

Figure 2.5

Simulation technique of tunnelling - model tunnel infilled with


water (After Jacobsz, 2002)

Figure 2.6

Sand pouring in process during model preparation with model tunnel


infilled with water (After Jacobsz, 2002)

Figure 2.7

Strong-box with two openings to fix the model tunnel in place (After
Jacobsz, 2002)

Figure 2.8

Simulation technique of tunnelling - mechanical equipment of


miniature shield tunneling machine (After Nomoto et al., 1994)

Figure 2.9

Simulation technique of tunnelling - mechanical equipment of shield


model machine (After Yasuhiro et al., 1998)

Figure 2.10

Simulation technique of tunnelling - mechanical model tunnel used


to simulate the tunnel volume loss by decreasing the diameter of
model tunnel under 1g (After Lee and Yoo, 2006)

Figure 2.11

Gaussian curve approximating transverse surface settlement trough


for MRT project C852, Singapore (After Cham, 2007)

xiii

Figure 2.12

Definition of parameters controlling tunnelling-induced settlement


trough ( After Standing & Burland, 2006)

Figure 2.13

Gaussian curve approximating transverse surface settlement trough


(After Peck, 1969)

Figure 2.14

Variation in surface settlement trough width parameter with tunnel


depth for tunnels in clay (After Lake et al., 1992)

Figure 2.15

Variation of trough width parameter K with depth for subsurface


settlement profiles above tunnels in clay (After Mair et al., 1993)

Figure 2.16

Initial settlement trough at Grimsby increased from 1.5Z to a final


value in excess of 4Z in long-term (After OReilly et al,1991)

Figure 2.17

The maximum surface settlements at Grimsby


significantly over the time (After OReilly et al, 1991)

Figure 2.18

The ratio of the maximum immediate settlement to maximum longterm settlement for Shanghai Metro Tunnel No.2 (After Zhang et al.,
2004)

Figure 2.19

Normalized post-construction surface settlement troughs due to


consolidation of soft clay (After Shirlaw, 1995)

Figure 2.20

Estimated trend of excess pore pressure in normally consolidated


clay surrounding the tunnel (After Schmidt, 1989)

Figure 2.21

Changes in Pore pressure for Shanghai Metro (After Schmidt, 1989)

Figure 2.22

Change in pore pressure measured at Thunder Bay Sewer Tunnel


(Adapted from data in Ng et al, 1986) (After Shirlaw et al, 1994)

Figure 2.23

Transverse movements in Toulouse subway line B were


significantly increased over time, but stabilized after 15 days of
tunnel excavation (After Emeriault et al, 2005)

Figure 2.24

Horizontal soil movement for the Singapores effluent outfall


pipeline in tunnel (After Balasubramanian, 1987)

Figure 2.25

Comparisons of surface settlement troughs in sand (After Feng,


2003) and clay (After Ran, 2004)

Figure 2.26

Ground surface settlement trough over time from a typical test


(After Ran, 2004)

Figure 2.27

Normalized Vertical and horizontal soil movement profile at


different subsurface elevations with best-fit curves: (a) 10mm
below ground level; (b) 30mm below ground level; (c) 70mm below
ground level; (d) 100mm below ground level; (After Grant and

increased

xiv

Figure 2.28

Taylor, 2000)
Comparisons of measured surface settlement and analytical
solutions (After Loganathan et al., 2000)

Figure 2.29

Plastic deformation mechanism for tunnels in clay (After Osman et


al., 2006a)

Figure 2.30

Definition of GAP parameter (After Lee et al., 1992)

Figure 2.31

Oval-shaped soil displacement around tunnel boundary


(After Loganathan and Poulos, 1998)

Figure 2.32

Boundary conditions of prescribed displacement (After Park, 2005)

Figure 2.33

Transverse movements in Toulouse subway line B were


significantly increased over time, but stabilized after 15 days of
tunnel excavation (After Emeriault et al, 2005)

Figure 2.34

A piled bridge pier foundation assessed during the CTRL project


(After Jacobsz et al., 2005)

Figure 2.35

Typical section and instrumentation layout for pile-tunnel


interaction study for MRT North East Line Contract 704 in
Singapore (After Pang et al.,2005)

Figure 2.36

Post-tunneling measurement of the development of axial force in


pile P1 at Pier 20 showing the time-dependent behaviour of dragload MRT North East Line Contract 704 in Singapore (After Pang
et al., 2005)

Figure 2.37

Responses of pile foundation in terms of (a) axial force and


(b)bending moment for MRT North East Line Contract 704 in
Singapore (After Pang et al., 2005)

Figure 2.38

Illustration of positions of existing instrumented piles relative to


tunnels for MRT Circle Line Stage 3 (CCL3) Contract 852 in
Singapore. (After Cham, 2007)

Figure 2.39

Configuration of centrifuge tests (After Loganathan et al., 2000)

Figure 2.40

Tunneling-induced pile bending moments (After Loganathan et al.,


2000)

Figure 2.41

Tunneling-induced pile axial loads (After Loganathan et al., 2000)

Figure 2.42

Variations of (a) induced pile bending moment profiles and (b)


induced pile lateral deflection profiles with time in typical test
(After Ran, 2004)

xv

Figure 2.43
Figure 2.44

(a) Induced pile axial force profile and (b) pile settlement profile at 2
days in typical test (After Ran, 2004)
Zone of influence around tunnel in which potential for large
pile settlements exists (After Jacobsz et al., 2005)

Figure 2.45

Settlement, rotation and load distribution on triple pile group


(After Jacobsz et al., 2005)

Figure 2.46

Layout of basic problem (After Chen et al., 1999)

Figure 2.47

Tunneling-induced pile responses and Greenfield soil movement


(After Chen et al., 1999)

Figure 2.48

Numerical analysis of pile-group responses due to tunnelling (After


Loganathan et al., 2001)

Figure 2.49

Typical 3D finite elements mesh to back-analyse a case history on


the response of pile foundation subjected to shield tunnelling (After
Pang et al., 2005b)

Figure 2.50

Prediction of responses of pile foundation in terms of axial force and


bending moment using 3D finite element analysis (After Pang et al.,
2005b)

Figure 3.1

Schematic diagram of NUS geotechnical centrifuge

Figure 3.2

Photograph of NUS geotechnical centrifuge with the model package


mounted on the platform

Figure 3.3

Sketch of a typical centrifuge model package (All dimensions in


mm)

Figure 3.4

Photograph of a typical centrifuge model package

Figure 3.5

Longitudinal view of model tunnel set up

Figure 3.6

Cross-section of model tunnel

Figure 3.7

Instrumented model pile (All dimensions in mm)

Figure 3.8

Model pile caps

Figure 3.9

In-flight undrained shear strength of clay

Figure 3.10

Image acquisition system

Figure 3.11

On board set-up

Figure 3.12

Picture captured by JAI CV-A2 progressive scan camera for PIV


analysis

xvi

Figure 3.13

Image manipulation during PIV analysis. (After White et al., 2003)

Figure 3.14

Evaluation of displacement vector from correlation plane, Rn(s): (a)


correlation of Rn(s); (b) highest correlation peak (integer pixel); (c)
sub-pixel interpolation using cubic fit over 1 pixel of integer
correlation. (After White et al., 2003)

Figure 3.15

Experimental set-up for assessment of effectiveness of image


processing system and comparison of performance of flocks and
beads

Figure 3.16

Results of assessment of effectiveness of image processing system


and comparison of performance of flocks and beads

Figure 3.17

Pore pressure dissipation and settlement during consolidation stage

Figure 3.18

Estimation of ultimate settlement by Asaokas method (1978)

Figure 3.19

Different colours of beads were randomly embedded on the surface


of clay

Figure 3.20

The Perspex window is highly greased to ensure free movement of


soil

Figure 3.21

Set-up of the entire model package in 1g (top view)

Figure 4.1

Schematic of viewing area in tunnel-soil interaction tests


(all dimensions in mm)

Figure 4.2

Example of digital images taken during test for PIV analysis

Figure 4.3 (a)

Vectors and contour plots of soil movements after 2 days (Test 1)

Figure 4.3 (b)

Vectors and contour plots of soil movements after 180 days (Test 1)

Figure 4.3 (c)

Vectors and contour plots of soil movements after 360 days (Test 1)

Figure 4.3 (d)

Vectors and contour plots of soil movements after 720 days (Test 1)

Figure 4.4 (a)

Vectors and contour plots of soil movements after 2 days (Test 2)

Figure 4.4 (b)

Vectors and contour plots of soil movements after 180 days (Test 2)

Figure 4.4 (c)

Vectors and contour plots of soil movements after 360 days (Test 2)

Figure 4.4 (d)

Vectors and contour plots of soil movements after 720 days (Test 2)

Figure 4.5

Surface settlement troughs over time (Test 1)

xvii

Figure 4.6

Surface settlement troughs over time (Test 2)

Figure 4.7

Maximum surface settlements over time (Tests 1 & 2)

Figure 4.8

Settlement troughs at surface, 4.3m and 9.3m depths (Test 1): (a)
comparing with Mair et. al (1993) (b) comparing with Loganathan
and Poulos (1998)

Figure 4.9

Settlement troughs at surface, 5m and 10.9m depths (Test 2): (a)


comparing with Mair et. al (1993) (b) comparing with Loganathan
and Poulos (1998)

Figure 4.10

Distribution of inflection point i with depth in short- and long-term


(Test 2)

Figure 4.11

Comparison of ratio of iLT/iST at different depths (Tests 1 & 2)

Figure 4.12

Horizontal soil movements at different distance from tunnel centerline at 2 and 720 days - Test 1
Horizontal soil movements at different distance from tunnel centerline at 2 and 720 days - Test 2

Figure 4.13

Figure 4.14

Pore pressure changes due to tunnelling (Test 1)

Figure 4.15 (a)

Tunnelling-induced pile axial force (Test 3, 3% free-head floating


long pile)

Figure 4.15 (b)

Tunnelling-induced pile axial force (Pile BP1-G) for MRT Circle


Line Stage 3 (CCL3) Contract 852 in Singapore (After Cham, 2007)

Figure 4.16

Tunnelling-induced pile head settlement (Test 3) and observed freefield soil movement at pile location (Test 1, PIV)

Figure 4.17

Tunnelling-induced (a) maximum pile axial force (b) maximum pile


head settlement and soil surface settlement (Test 1) (c) maximum
pile bending moment (d) maximum pile head deflection and soil
surface lateral movement (Test 1)

Figure 4.18 (a)

Tunnelling-induced pile bending moment (Test 3, 3% free-head


floating long pile)

Figure 4.18 (b)

Tunnelling-induced pile bending moment (Pile BP2-E) for MRT


Circle Line Stage 3 (CCL3) Contract 852 in Singapore (After Cham,
2007)

Figure 4.19

Tunnelling-induced pile deflection (Test 3) and free-field lateral soil


movement at pile location (Test 1)

xviii

Figure 4.20

Variation of pile axial force with volume loss (Tests 3 and 4)

Figure 4.21

Variation of pile head settlement (Tests 3 and 4) and observed freefield soil movement at pile location (Tests 1 and 2) with volume loss

Figure 4.22

Variation of pile bending moment with volume loss (Tests 3 and 4)

Figure 4.23

Variation of pile deflection profiles (Tests 3 and 4) and observed


free-field lateral soil movement at pile location (Tests 1 and 2) with
volume loss

Figure 4.24

Variation of (a) maximum pile axial force (b) pile head settlement
(c) pile bending moment (d) pile head deflection with volume loss
(Tests 3 and 4)

Figure 4.25

Long-term to short-term ratio of pile responses for different volume


losses (Tests 3 and 4)

Figure 5.1

Pile base position investigated in the parametric studies (not to


scale)
Variation of pile axial force with tip condition in (a) Short-term (b)
Long-term (Tests 3, 9, 10 and 13)

Figure 5.2

Figure 5.3

Variation of pile bending moment with tip condition (a) Short-term


(b) Long-term (Tests 3, 9, 10 and 13)

Figure 5.4

Variation of pile deflection with tip condition (a) Short-term (b)


Long-term (Tests 3, 9, 10 and 13)

Figure 5.5

Variation of (a) maximum pile axial force (b) pile head settlement
(c) pile bending moment (d) pile head deflection with tip and head
conditions (Tests 3, 9, 10 and 13)

Figure 5.6

Variation of pile axial force with pile length (Tests 3, 7 and 8)

Figure 5.7

Variation of pile head settlement and soil settlement profile (Test 1)


with pile length (Tests 3, 7 and 8)

Figure 5.8

Variation of pile bending moment with pile length (Tests 3, 7 and 8)

Figure 5.9

Variation of pile head deflection and free- field lateral soil


displacement (Test 1) with pile length (Tests 3, 7 and 8)

Figure 5.10

Variation of (a) maximum pile axial force (b) pile head settlement
(c) pile bending moment (d) pile head deflection with normalized
pile length over tunnel depth (Tests 3, 7 and 8)

Figure 5.11

Short pile to long pile ratio of pile responses for different pile length
over tunnel depth (Tests 3, 7 and 8)

xix

Figure 5.12

Variation of pile axial force with pile-to-tunnel distance for freehead floating piles (Tests 3, 5 and 6)

Figure 5.13

Variation of pile head settlement for free-head floating piles (Tests


3, 5 and 6) and free-field soil settlement (Test 1) with pile-to-tunnel
distance
Variation of pile bending moment for free-head floating piles with
pile-to-tunnel distance (Tests 3, 5 and 6)

Figure 5.14

Figure 5.15

Variation of pile deflection for free-head floating piles (Tests 3, 5


and 6) and free-field lateral soil displacement profile (Test 1) with
pile-to-tunnel distance

Figure 5.16

Variation of pile axial force for free-head end bearing piles with
pile-to-tunnel distance (Tests 10, 11 and 12)

Figure 5.17

Variation of pile bending moment for free-head end bearing piles


with
pile-to-tunnel distance (Tests 10, 11 and 12)

Figure 5.18

Variation of pile deflection for free-head end bearing piles with pileto-tunnel distance (Tests 10, 11 and 12)

Figure 5.19

Variation of pile bending moment for fixed-head end bearing piles


with pile-to-tunnel distance (Tests 13, 14A and 14B)

Figure 5.20

Variation of pile deflection for fixed-head end bearing piles with


pile-to-tunnel distance (Tests 13, 14A and 14B)

Figure 5.21

Variation of (a) maximum pile axial force (b) maximum pile head
settlement and soil surface settlement (Test 1) (c) pile bending
moment (d) maximum pile head deflection for Test Series 4, 5 and 6
with pile-to-tunnel distance

Figure 5.22

Assessment of pile responses for different pile-to-tunnel distance


(Tests 3, 5 and 6)

Figure 5.23

Lateral soil displacement profiles at different pile-to-tunnel distance


(Tests 3, 5 and 6)

Figure 5.24

Long-term to short-term ratio of pile responses for all tests


(Tests 3, 4, 5, 6, 7, 8, 9, 10, 11, 12, 13, 14A, 14B and 16)

Figure 5.25

Comparison of (a) ovalisation of tunnel lining by Ran (2004); and


(b) over-cut of tunnel in the present study

Figure 5.26

Simplified tunnel lining ovalisation with time (not to scale) (Test 1)


(after Ran, 2004)

xx

Figure 5.27

Development of subsurface soil movements at (a) 2 days and (b) 720


days after tunnel excavation (after Ran, 2004)

Figure 5.28

Variation of surface soil settlement troughs with tunnel deformation

Figure 5.29

Variation of maximum surface soil settlement at tunnel central line


with
tunnel deformation

Figure 5.30

Variation of vertical soil settlement at pile location with tunnel


deformation

Figure 5.31

Variation of soil deflection at pile location with tunnel deformation

Figure 5.32

Variation of pile axial force with tunnel deformation

Figure 5.33

Variation of pile head settlement with tunnel deformation

Figure 5.34

Variation of pile bending moment with tunnel deformation

Figure 5.35

Variation of tunnelling-induced maximum pile bending moment


over time for different tunnel deformation

Figure 5.36

Variation of pile deflection with tunnel deformation

Figure 5.37

Long-term to short-term ratio of pile responses over time for


different tunnel deformation

Figure 6.1

Tunnelling-induced pile axial force (Test PG1)

Figure 6.2

Tunnelling-induced front pile (Test PG1) and corresponding single


pile (Test 3) axial force
Tunnelling-induced rear pile (Test PG1) and corresponding single
pile (Test 16) axial force

Figure 6.3

Figure 6.4

Tunnelling-induced pile head settlement (Tests PG1, 3, 16)

Figure 6.5

Tunnelling-induced pile bending moment (Test PG1)

Figure 6.6

Tunnelling-induced front pile (Test PG1) and corresponding single


pile (Test 3) bending moment

Figure 6.7

Tunnelling-induced rear pile (Test PG1) and corresponding single


pile (Test 16) bending moment

Figure 6.8

Tunnelling-induced pile deflection (Tests PG1)

Figure 6.9

Tunnelling-induced front pile (Test PG1) and corresponding single


pile (Test 3) deflection

xxi

Figure 6.10

Tunnelling-induced rear pile (Test PG1) and corresponding single


pile (Test 16) deflection

Figure 6.11

Tunnelling-induced pile head deflection (Test PG1, 3 & 16)

Figure 6.12

Single pile over pile group ratio for front pile (Test 3/ PG1)

Figure 6.13

Single pile over pile group ratio for rear pile (Test 16/ PG1)

Figure 6.14

Tunnelling-induced pile axial force (Test PG2)

Figure 6.15

Tunnelling-induced front pile (Test PG2) and corresponding single


pile (Test 10) axial force

Figure 6.16

Tunnelling-induced rear pile (Test PG2) and corresponding single


pile (Test 11) axial force

Figure 6.17

Tunnelling-induced pile bending moment (Test PG2)

Figure 6.18

Tunnelling-induced front pile (Test PG2) and corresponding single


pile (Test 10) bending moment

Figure 6.19

Tunnelling-induced rear pile (Test PG2) and corresponding single


pile (Test 11) bending moment

Figure 6.20

Tunnelling-induced pile deflection (Test PG2)

Figure 6.21

Tunnelling-induced front pile (Test PG2) and corresponding single


pile (Test 10) deflection

Figure 6.22

Tunnelling-induced rear pile (Test PG2) and corresponding single


pile (Test 11) deflection

Figure 6.23

Tunnelling-induced pile head deflection in the (a) short-term (b)


long-term (Tests PG2, 10 & 11)

Figure 6.24

Single pile over pile group ratio for front pile (Test 10/ PG2)

Figure 6.25

Single pile over pile group ratio for rear pile (Test 11/ PG2)

Figure 6.26

Tunnelling-induced pile axial force (Test PG3)

Figure 6.27

Tunnelling-induced front pile (Test PG3) and corresponding single


pile (Test 13) axial force

Figure 6.28

Tunnelling-induced pile bending moment (Test PG3)

Figure 6.29

Tunnelling-induced front pile (Test PG3) and corresponding single


pile (Test 13) bending moment

xxii

Figure 6.30

Tunnelling-induced rear pile (Test PG3) and corresponding single


pile (Test 14A) bending moment

Figure 6.31

Single pile over pile group ratio for front pile (Test 13/ PG3)

Figure 6.32

Single pile over pile group ratio for rear pile (Test 14A/ PG3)

Figure 6.33

Tunnelling-induced pile axial force (Test PG4)

Figure 6.34

Tunnelling-induced front pile in 2-pile group (Test PG2) and


corresponding front pile in 6-pile group (Test PG4) axial force

Figure 6.35

Tunnelling-induced rear pile in 2-pile group (Test PG2) and


corresponding middle pile in 6-pile group (Test PG4) axial force

Figure 6.36

Tunnelling-induced pile bending moment (a) in the short-term (b) in


the long-term (Test PG4)

Figure 6.37

Tunnelling-induced front pile in 2-pile group (Test PG2) and


corresponding front pile in 6-pile group (Test PG4) bending moment

Figure 6.38

Tunnelling-induced rear pile in 2-pile group (Test PG2) and


corresponding middle pile in 6-pile group (Test PG4) bending
moment

Figure 6.39

Tunnelling-induced pile deflection in the (a) short-term (b) longterm (Test PG4)

Figure 6.40

Tunnelling-induced pile deflection (Tests PG2 and PG4)

Figure 6.41

Tunnelling-induced pile bending moment (Tests PG2 and PG4)

Figure 6.42

2-pile over 6-pile group ratio for front pile (Test PG2/PG4)

Figure 6.43

2-pile over 6-pile group ratio for middle pile (Test PG2/PG4)

Figure 6.44

Tunnelling-induced pile axial force (Test PG5)

Figure 6.45

Tunnelling-induced front pile in 2-pile group (Test PG3) and


corresponding front pile in 6-pile group (Test PG5) axial force

Figure 6.46

Tunnelling-induced rear pile in 2-pile group (Test PG3) and


corresponding middle pile in 6-pile group (Test PG5) axial force

Figure 6.47

Tunnelling-induced pile bending moment in the (a) short-term (b)


long-term (Test PG5)

Figure 6.48

Tunnelling-induced front pile in 2-pile group (Test PG3) and


corresponding front pile in 6-pile group (Test PG5) bending moment

xxiii

Figure 6.49

Figure 6.50

Tunnelling-induced rear pile in 2-pile group (Test PG3) and


corresponding middle pile in 6-pile group (Test PG5) bending
moment
2-pile over 6-pile group ratio for front pile (Test PG3/PG5)

Figure 6.51

2-pile over 6-pile group ratio for middle pile (Test PG3/PG5)

Figure 6.52

Tunnelling-induced front pile in capped-head 2-pile group (Test


PG2) and corresponding front pile in fixed-head 2-pile group (Test
PG3) axial force

Figure 6.53

Tunnelling-induced rear pile in capped-head 2-pile group (Test


PG2) and corresponding rear pile in fixed-head 2-pile group (Test
PG3) axial force

Figure 6.54

Tunnelling-induced front pile in capped-head 2-pile group (Test


PG2) and corresponding front pile in fixed-head 2-pile group (Test
PG3)bending moment

Figure 6.55

Tunnelling-induced rear pile in capped-head 2-pile group (Test


PG2) and corresponding rear pile in fixed-head 2-pile group (Test
PG3) bending moment

Figure 6.56

Capped-head pile over fixed-head pile ratio (front pile, Test


PG2/PG3)

Figure 6.57

Capped-head pile over fixed-head pile ratio (rear pile, Test


PG2/PG3)

Figure 6.58

Tunnelling-induced front pile in capped-head 6-pile group (Test


PG4) and corresponding front pile in fixed-head 6-pile group (Test
PG5) axial force

Figure 6.59

Tunnelling-induced middle pile in capped-head 6-pile group (Test


PG4) and corresponding middle pile in fixed-head 6-pile group (Test
PG5) axial force

Figure 6.60

Tunnelling-induced rear pile in capped-head 6-pile group (Test


PG4) and corresponding rear pile in fixed-head 6-pile group (Test
PG5) axial force

Figure 6.61

Tunnelling-induced front pile in capped-head 6-pile group (Test


PG4) and corresponding front pile in fixed-head 6-pile group (Test
PG5) bending moment

Figure 6.62

Tunnelling-induced middle pile in capped-head 6-pile group (Test


PG4) and corresponding middle pile in fixed-head 6-pile group (Test
PG5) bending moment

xxiv

Figure 6.63

Tunnelling-induced rear pile in capped-head 6-pile group (Test


PG4) and corresponding rear pile in fixed-head 6-pile group (Test
PG5) bending moment

Figure 6.64

Variation of maximum bending moment for front, middle and rear


pile in the short-term (Tests PG4 and PG5)
Variation of maximum bending moment for front, middle and rear
pile in the long-term (Tests PG4 and PG5)

Figure 6.65

Figure 6.66

Capped-head pile over fixed-head pile ratio (front pile, Test


PG4/PG5)

Figure 6.67

Capped-head pile over fixed-head pile ratio (middle pile, Test


PG4/PG5)

Figure 6.68

Capped-head pile over fixed-head pile ratio (rear pile, Test


PG4/PG5)

Figure 6.69

Long-term over short-term ratio (Tests PG1, 2, 3, 4 & 5)

xxv

NOMENCLATURE

Tunnel cover

Cu

Undrained shear strength of clay

Cv

Coefficient of consolidation

Tunnel diameter

EA

Pile axial rigidity

EI

Pile flexural rigidity

Gravitational acceleration (9.8 m/s2)

Gs

Specific gravity

Tunnel depth

Point of inflection

Coefficient of permeability

Trough width parameter

Pile length

Soil settlement induced by tunneling

Smax

Maximum ground surface settlement

U3D

Parameter accounting for three dimensional heading effects (Gap


Parameter Method)

Tunnel Volume loss

Horizontal distance between pile and tunnel centre

Depth below ground surface

Zo

Depth to tunnel axis level

Soil effective friction angle

crit

Critical state friction angle

Effective unit weight of soil

Unit weight of water

min

Minimum dry density

max

Maximum dry density

xxvi

x,z

Ground loss with horizontal and vertical distance

Ground loss ratio

Slope of critical state line in p-q space

Slope of isotropic compression line in p-v space

Slope of swelling line in p-v space

Abbreviations
LL

Liquid Limit

LT

Long-term

NP

Neutral Plane

NSF

Negative Skin Friction

NUS

National University of Singapore

PI

Plasticity Index

PIV

Particle Image Velocimetry

PPT

Pore Pressure Transducer

PL

Plastic Limit

ST

Short-term

VL

Volume loss

xxvii

Chapter 1 Introduction

CHAPTER ONE

INTRODUCTION

1.1

BACKGROUND

In urban areas, tunnels are often constructed close to existing buildings due to space
constraints. As cities develop rapidly, the need for underground transportation and
utility tunnels have greatly increased. Unfortunately, many structures exist long before
the tunnels are planned. As such, it is increasingly complex and challenging to build
tunnels underground as the tunnels almost inevitably run close to or underneath some
piled foundations supporting existing buildings.

Tunnel excavations generally cause ground settlement and deformation nearby,


especially in soft clay. Ground movements due to tunnelling would induce additional
axial (settlement and axial force) and lateral (deflection and bending moment)
responses on adjacent pile foundations (Figure 1.1). As pile foundations supporting
existing buildings are generally designed to resist compression load only (Figure 1.2),
the foundations may not be safe to resist the induced loads due to tunnelling. At
present, there are very few reliable design methods (Chen at al., 1999; Loganathan et
al., 2001; Pang et al., et al. 2005b) to assess induced pile responses due to tunnelling.

Although many studies have been conducted to investigate tunnel-soil-pile


interactions in the short-term, impacts in the long-term are still not well understood. A

Chapter 1 Introduction

published field study covering an 11-year period of post-tunnelling monitoring for the
Haycroft Relief Sewer in Grimsby, UK, (OReilly et al., 1991) in very soft clay
recorded that the soil settlement in the long-term had increased significantly. The
results of centrifuge model tests (Ran, 2004) conducted at the National University of
Singapore (NUS) also revealed that for tunnelling in clay, the ground continues to
deform long after the completion of tunnelling, thus inducing further settlement,
deflection, axial and lateral loads on adjacent piles in the long-term.

In view of the complexity of field instrumentation and monitoring, physical


modelling can be an attractive mean to study the tunnel-soil-pile interaction problem in
both short-term and long-term. One effective way is to conduct centrifuge model tests
employing artificial gravitational field to replicate the prototype stress level as
experienced by the ground in the field. Under a well-controlled environment,
centrifuge tests can provide

flexibility and repeatability to study tunnel-soil-pile

interaction problems that could not be achieved in field tests (Mair et al., 1984;
Loganathan et al., 2000; Ran, 2004; Jacobz et al., 2004).

1.2

TUNNELLING-INDUCED SOIL MOVEMENTS

Many research studies have been carried out to investigate tunnelling-induced soil
movements. Peck (1969), Schmidt (1969), OReilly and New (1982), Lake et al.
(1992), Mair et al. (1993) and others developed empirical formulae from field studies
to predict the soil movements induced by tunnelling. In addition, several centrifuge
model studies including Loganathan (2000) and Ran (2004) were conducted to
examine soil movements due to tunnelling.

Chapter 1 Introduction

Besides empirical formulae and centrifuge model studies, analytical solutions


have been developed by researchers including Sagaseta (1987), Verrujit and Booker
(1996), Loganathan and Poulos (1998), Park (2005) and Osman et al. (2006a) to
predict the ground displacements for various shapes of tunnel deformation. However,
such analytical solutions cannot account for all aspects, in particular the time effects.
Many researchers reported field measurement results for soil movements induced by
tunnelling. However, most of these studies did not report field measurements during
the post-tunnelling period despite some reports on significant long-term settlements
after tunnelling in soft clay; see for example, Haycroft Relief Sewer in Grimsby, UK,
(OReilly et al., 1991) and Shanghai Metro Tunnel No.2, China, (Zhang et al., 2004).

From the above review, it is evident that the mechanism and calculation of
tunnelling-induced soil movements in the short-term are reasonably well studied.
However, the long-term tunnel-soil interaction in soft clay clearly needs further
investigation.

1.3

EFFECTS OF TUNNELLING ON PILES

It is generally not viable to monitor the responses of existing piles due to tunnelling.
This is because the piles usually exist long before the tunnels are constructed and it is
almost impossible to install strain gauges in the existing piles to monitor the pile
responses. However, some valuable field measurements in limited cases have been
made available where the existing piles were instrumented prior to tunnel construction,
see Selemetas et al. (2005), Jacobsz et al. (2005), Pang et al. (2005) and Cham (2007).

Chapter 1 Introduction

Although a good number of centrifuge model studies (Loganathan et al., 2000; Feng,
2003; Ran, 2004 and Jacobsz et al., 2005) and numerical analysis and analytical
solutions (Chen et al., 1999; Loganathan et al., 2001 and Pang et al.; 2005) have been
attempted to investigate the effects of tunnelling on piles, the results of literature
review presented in Chapter 2 reveal that there are still major gaps for a better
understanding of pile responses due to tunnelling. Two examples of major gap include
limited field studies on long-term tunnel-soil-pile interaction, and lack of predictive
methods available to evaluate the effects of pile group or soil-structure interaction. The
development of a qualitative and quantitative framework to assess potential impact of
tunnel construction on existing piled foundations is needed to improve the current
knowledge on the prediction of pile responses due to tunnelling.

1.4

OBJECTIVE AND SCOPE OF STUDY

The main aim of the present study is to investigate tunnel-soil-pile interaction in soft
clay. More specifically, the objectives of this research are:-

1.

To develop a simulation technique of tunnel excavation associated with


inward tunnel deformation pattern using centrifuge modelling technique. This
pattern is chosen because it is often observed in practice. However, the wishin-place model tunnel is a simplification and idealization of a cavity
contraction, but with an oval-shaped GAP and well controlled volume loss to
simulate the soil movements induced by tunnelling when the tunnel
excavation has passed a particular section. The minimum volume loss that

Chapter 1 Introduction

can be simulated by the present model tunnel is only 3% in order to maintain


the accuracy of the experiment results.

2.

To study tunnelling-induced soil movement in free-field experimentally with


subsequent Particle Image Velocimetry (PIV) analysis. The purpose of this
investigation is to examine the mechanism of tunnel-soil interaction. Longterm post-tunnelling soil movement was also studied.

3.

To study the induced pile responses due to tunnelling experimentally. The


vertical (axial force and settlement) and lateral (bending moment and
deflection) pile responses were examined. In addition, parametric studies
were performed to evaluate the effects of various tunnel volume losses, pile
tip conditions, pile lengths, pile distances to tunnel and pile groups. Longterm post-tunnelling pile performance was also studied.

1.5

STRUCTURE OF THESIS

This thesis consists of seven chapters and the contents of subsequent chapters are
briefly described as follows:
a) Chapter 2 reviews the existing literature on to tunnel-soil-pile interaction. The
review is divided into three parts. Firstly, various methods of simulating tunnel
excavation in centrifuge tests are reviewed. Subsequently, ground movements
caused by different methods of tunnel excavation are examined. Finally, existing
research studies concerning pile responses due to tunnelling are highlighted.
b) Chapter 3 describes the present centrifuge model set-up and experimental

Chapter 1 Introduction

procedures. A detailed description of the novel technique for simulation of


tunnelling during centrifuge flight is also presented.
c) Chapter 4 first presents the experimental results on tunnelling induced soil
movements analysed by PIV. The test results on the basic tests of volume loss for
tunnelling on single piles are then presented in detail.
d) Chapter 5 presents the results of parametric studies on the effects of tunnelling on
single piles
e) Chapter 6 presents the results on effects of tunnelling on pile groups
f)

Chapter 7 summarizes the main findings of the present study and proposes future
works.

Chapter 1 Introduction

Existing
Building
Depth (m)
5

Induced Pile
Settlement

10

Settlement
Trough

15
20
25

Induced
BM

Tunnel
Contraction

30
35

Induced Axial Loads

(NTS)

Figure 1.1 Pile responses induced by tunnel construction: (a) Tunnelling under pile
foundation, (b) Tunnelling adjacent to pile foundation.

Chapter 1 Introduction

Existing
Building
Depth (m)
5

Steel reinforcement

10
15

Existing Pile

20
25

New tunnel

30
(NTS)

Figure 1.2 Pile foundations supported existing buildings normally designed to resist
compression load only.

Chapter 2 Literature Review

CHAPTER TWO

LITERATURE REVIEW

2.1

INTRODUCTION

Tunnelling would cause soil movements that would in turn induce axial and lateral
loads on adjacent pile foundation. Hence, it is of great interest to understand the effects
of tunnelling induced soil movements on existing piles. Many studies on free-field soil
movements due to tunnelling have been reported. However, few studies concerned the
pile responses subject to soil movements caused by tunnelling, especially for long-term
pile behaviours.

The methods, effects and difficulties faced in investigating the

behaviour of piles and soil due to tunnelling will be reviewed in detail in this chapter.
The topics investigated include: (a) model simulation technique of tunnelling; (b)
tunnelling-induced soil movements and (c) tunnelling-induced pile responses.

2.2

TECHNIQUES FOR SIMULATION OF TUNNELLING IN


CENTRIFUGE

Prior to the study of pile behaviours due to tunnelling, the correct simulation of tunnel
excavation plays an important role. The soil movement pattern is directly caused by
the model tunnel and affects the pile responses.

Chapter 2 Literature Review

Tunnel construction is three-dimension in nature and time dependent. In


general, the settlement due to shield tunnelling can be derived from the face loss
during tunnel advancement, shield loss due to over-cut, tail void closure and
consolidation (Shirlaw et al., 2003). With the increasing availability of large-scale
computing resources, 3D finite elements simulations of tunnels are reported in the
literature in recent years (Komiya et al., 1999; Augarde and Burd, 2001; Lim, 2003;
Lin et al., 2002; Melis et al., 2002; Ng et al., 2004; Lee and Ng, 2005; Ng and Lee,
2005; Lee et al., 2006; Phoon et al., 2006). However, realistic three-dimensional finite
element analysis has still not reached a point of development where it can be routinely
used in engineering design. Two common problems associated with three-dimensional
analyses mean very large memory requirement and long computing times (Lee et al.,
2006; Phoon et al. 2006).

In order to compare the performance of 2D and 3D analyses, Moller (2006)


performed a back analysis for the case of Second Heinenoord Shield Tunnel. It is
reported that the surface settlements of 2D analysis compare well with the 3D analysis,
as the face pressure does not much contribute to the development of the surface
settlement. Moreover, from the field monitoring of Mass Rapid Transit North East
Line Contract C704 studied by Pang (2006), it is apparent that the maximum axial
force developed in the pile when the Earth Pressure Balance machine (EPBM) was
directly adjacent to the pile. Besides, the induced pile bending moment in the
longitudinal direction is either equal or smaller than the induced pile bending moment
in transverse direction. This study clearly demonstrated that the transverse section is
more critical, and hence the simulation of plane strain model of a long tunnel is
comparable to 3D modelling in this case.

10

Chapter 2 Literature Review

With the availability of centrifuge modelling technique, a prototype problem


can be simulated in a scaled laboratory model hence overcoming the limitation of 1-g
model. Most of the centrifuge tests on tunnel-soil interaction that have been carried out
(Bezuijen and Schrier, 1994; Hergarden et al., 1996; Loganathan, 1999; Ran et al.,
2003; Jacobsz et al., 2002; Feng et al., 2002; Lee and Chiang, 2004) thus far
considered only a wish-in-place plane strain tunnel in the simulation. This is probably
because the development of 3D model tunnel in centrifuge is very difficult and has its
limitation of boundary conditions and constraints. Therefore, If thein a situation
whenre the tunnel excavation has passed a particular section is considered, the vectors
of the ground movement developed will be more or less in the plane perpendicular to
the tunnel axis. Consequently it is reasonable to assume that a plane strain model of
long tunnel section would be a good representation of tunnelling-induced soil
movements; this is usually referred to as a two-dimensional simulation (Taylor, 1998).
Nevertheless, if the tunnel face passes directly under a foundation, the 3D soil response
around tunnel face would be important. However, this scenario has not been examined
in this research.

The review of various methods of simulating tunnel excavation in the


centrifuge is briefly discussed here in order to gather useful information and guidance
for the development of model tunnelling technique in the present study.

2.2.1 Simulation Technique 1 - High Density Polystyrene Foam


Sharma et al. (2001) presented a method to simulate tunnel excavation in the
centrifuge by dissolving polystyrene foam quickly using an organic solvent. Figure 2.1
shows the arrangement for the inflow of solvent into the tunnel core. The polystyrene

11

Chapter 2 Literature Review

foam core was placed tightly inside the model tunnel lining, which was made by
wrapping a half hard brass foil around the foam core and soldering the lap joint with
the help of tin solder and an electronic gun. The flow of this liquid into the polystyrene
foam (model tunnel) is controlled by using solenoid manifold and solvent reservoir.
The stiffness of the filled tunnel can approximately be made to be equivalent to that of
the parent soil. The lining is left in place when the foam core has been dissolved. This
technique is an improvement over other methods of modelling tunnel excavation, such
as reduction of air pressure supporting the tunnel lining or gradually draining heavy
liquid from within the lining. A limitation of this modelling technique is that this
approach may not correctly simulate the actual tunnel excavation situations in the field
as it can only simulate excavation cases with tunnel spring line moving outwards.

The technique proposed by Sharma et. al. (2001) for the simulation of tunnel
excavation in the centrifuge has been adopted by Feng (2003) and Ran (2004) at
National University of Singapore (NUS). Feng (2003) carried out a centrifuge model
study to investigate tunnel-soil-pile interaction in sand. Ran (2004) extended Fengs
study on tunnel-soil-pile interaction to clay.

Figure 2.2 shows the shape of the deformed tunnel lining over time obtained
from a typical test. It can be seen that the lining deforms into an oval shape with tunnel
spring lining protruding slightly outwards. In most cases in practice, the overexcavation of tunnel and the gap between the shield tunnel machine and lining would
cause the tunnel spring line to move towards the tunnel resulting in inward soil
movements at the tunnel spring line.

12

Chapter 2 Literature Review

2.2.2 Simulation Technique 2 - Compressed Air


One of the earlier model tunnelling technique has been developed at the University of
Cambridge (Potts, 1976; Mair et al., 1984) in which air pressure is applied to control
tunnel support conditions. This technique has subsequently been used by Chambon et
al. (1991), Knig et al. (1991), Lee et al. (1999), Grant and Taylor (2000) and Bilotta
and Taylor (2005). Figure 2.3 shows the schematic diagram of a typical plane strain
centrifuge model tunnel using compressed air (Grant and Taylor, 2000). In this
modelling technique, the soil mass, which has to be removed during the excavation, is
represented in the model by a rubber membrane pressurized with air pressure. The
shape of the rubber membrane is identical with the contour of excavation. The
equilibrium is kept between the internal air pressure and earth pressure during model
preparation and increasing g-level. The excavation is simulated by decreasing the
internal air pressure.

To keep equilibrium between internal air pressure and the earth pressure, it
would be necessary to apply internal support according to the theoretical earth pressure
at rest. This support cannot be achieved by the air pressure technique, because air
pressure remains constant along the contour of excavation whereas earth pressure
varies with the change of orientation from vertical to horizontal and due to increase in
vertical stress with depth.

13

Chapter 2 Literature Review

2.2.3 Simulation Technique 3 - Liquid Oil / Water


The limitation of simulation of tunnelling by air pressure can be eliminated by
applying fluid pressure inside the rubber membrane. This was realized in several
studies related to excavation of shafts and trenches (Lade et al., 1981; Mair et al.,
1984). A zinc chloride solution has been used to achieve a fluid with a density similar
to that of the surrounding soil. This technique is well suited to study the relationship
between the displacements of soil into the excavation and surface settlements or to
evaluate the failure conditions. However, the technique needs to be evaluated carefully
if the stress strain conditions (depending on the tunnelling technique, soil stiffness,
stiffness of lining as well as load concentration at the end of the tunnel lining) in the
soil close to the excavation are most relevant for the investigated effect. (Knig, 1998).

Figure 2.4 shows a cross-section of the model tunnel infilled with oil developed
by Loganathan et al. (2000). The stability of tunnelling procedure was represented by
the equivalent surface ground loss values. A technique of decreasing tunnel diameter
during centrifuge flight was adopted to model the required ground loss. The inner core
of the tunnel was made of a long aluminium tube, and a very thin rubber membrane
was placed on top of the inner core cylinder. The rubber membrane was attached at
both ends of the inner core cylinder by end caps which prevented any leakage of oil
from between the inner core and the membrane. A small hole was drilled through the
end cap and inner core to allow the passage of oil. The cylindrical face of the assembly
was then covered by a 0.5-mm thick smooth-surfaced overlapping PVC spring to
enhance the tunnel lining stiffness, and to ensure a uniform change in the tunnel
diameter during the test. A syringe pump was fabricated to control the volume of oil in

14

Chapter 2 Literature Review

the tunnel assembly. The advantage of this model is that various volume losses can be
simulated in one test by extracting oil slowly.

The recent model tunnel reported by Jacobsz (2002) was infilled with water to
study the tunnelling effects on single piles in sand. The model tunnel consisted of a
brass mandrel surrounded by a 1-mm thick latex membrane as shown in Figures 2.5
and 2.6. During centrifuge tests, the approximately 4-mm thick annulus between the
mandrel and the membrane was filled with water that could be extracted to accurately
impose volume losses from 0% to about 20% on the surrounding soil. The outer
diameter of the model tunnel was 60 mm, representing a 4.5-m diameter tunnel at
prototype scale. A pressure transducer was incorporated into the tunnel control system
to monitor the pressure in the annulus of water between the latex membrane and the
brass mandrel.

The tunnel was connected via a solenoid valve to a standpipe in which a


constant water level was maintained to automatically balance the tunnel pressure with
the overburden pressure during the acceleration of the centrifuge. After the desired
acceleration had been achieved and the model piles installed, the solenoid valve was
closed and the volume loss was imposed by extracting water slowly from the annulus
around the tunnel.

Nevertheless, according to Knig (1998), there are indeed short-comings of the


above model tunnels which control the volume loss by changing the volume of gas or
fluid to simulate the volume loss. This is because the density of gas or fluid used is
different from that of soil. Thus, the initial stress condition of the model tunnel in this

15

Chapter 2 Literature Review

simulation technique is one of the major concerns.

Jacobsz (2002) reported that due to the self-weight of the water in the tunnel, a
hydrostatic pressure gradient would have existed in the tunnel, with the pressure near
the crown 22 kPa lower and the pressure at the invert 22 kPa higher than measured.
Besides, a small amount of water trapped near the tunnel ends, which would have been
difficult to extract completely, would also have increased the registered tunnel pressure.
In addition, the stiffness of the tunnel membrane and the fact that the contraction of the
model tunnel occurred non-uniformly, would have affected the measured pressures.
Most factors above would have resulted in the pressure being over-registered.

Moreover, the model tunnel is fixed at both ends of the strong box, as shown in
Figure 2.7, which would induce boundary condition and is not realistic as the middle
part of the tunnel will suffer much higher stress and displacement as compared to the
both ends of the tunnel when subject to high-g acceleration.

2.2.4 Simulation Technique 4 - Mechanical Equipments


Relatively complicated mechanical equipments have been developed to simulate the
process of shield tunnelling in centrifuge. Figure 2.8 shows an in-flight miniature
shield tunnelling machine developed by Nomoto et al. (1994). This machine allowed
the in-flight excavation of soil and also an in-flight installation of the lining. This time
the whole shield tunnelling process was successfully reproduced in a centrifuge force
field. Owing to the small size of the centrifuge and the restrictions in terms of
dimensions and weight of the model, the attempt to evaluate earth pressure acting on

16

Chapter 2 Literature Review

the tunnel lining quantitatively was not yet fulfilled. The tail void was also too large
compared with the prototype machine (Nomoto et al., 1994).

A plane-strain shield model machine was developed by Yasuhiro et al. (1998)


to perform centrifuge tests in both sand and clay to study the tunnelling-induced
ground movement. The shield model machine consists of steel rings and a wedge
shaped shaft, which are able to simulate the tail void and backfill grouting in flight by
contraction and expansion of shield model rings controlled by a motor, see Figure 2.9.

Recently, Ghahremannejad at el. (2006) simulated volume loss by varying the


diameter of the model tunnel along the tunnel. The desired volume loss was achieved
when the smaller aluminum tubes were pushed through the sand box. Lee and Yoo
(2006) adopted another technique simulating the volume loss by decreasing the
diameter of model tunnel as shown in Figure 2.10. However, the recent model tunnels
developed by Ghahremannejad at el. (2006) and Lee and Yoo (2006) can only simulate
tunnel volume loss under 1g condition.

2.3

TUNNELLING-INDUCED SOIL MOVEMENTS

2.3.1 Field Studies of Tunnelling-Induced Soil Movement


Peck (1969), Schmidt (1969) and subsequently many other researchers have shown
that the transverse settlement trough after tunnel excavation in short-term (Cham, 2007)
can be well-described by a Gaussian distribution curve as shown in Figure 2.11. The
method needs an estimate of volume loss (V) and the trough width parameter (K) to

17

Chapter 2 Literature Review

obtain the maximum ground surface settlement (Smax) and subsequently the surface
settlement profile, as defined in Figure 2.12. The ground settlements are generally
negligible beyond an offset of 3i from the tunnel centre line for Pecks (1969)
proposed curve (Figure 2.13). The surface settlement trough,S, and volume loss,V, are
approximated as follows:

x2
S = S max exp 2
2i

V = 2 iS max

(2.1)

(2.2)

where x is the offset to the tunnel vertical line and i is point of inflection.

The trough width parameter K is relatively easy to quantify as it is largely


independent of construction method and operation experience (Fujita, 1981; OReilly
and New, 1982). Numerous field data have been collected and hence many estimates
of trough width parameters have been proposed. However, a comprehensive summary
done by Lake et al. (1992) on tunnelling data from many countries on clay shown in
Figure 2.14 has established the general variations of i as follows:
Simple approximate relationship i = Kz o
Where K = 0.4 to 0.6 for clays (soft and stiff), and
K = 0.25 to 0.45 for sands and gravels.
zo= depth to tunnel axis

Lake et al.s (1992) field studies supported the various proposals that K can be
assumed solely as 0.5 for tunnelling in clay (OReilly and New, 1982; Mair et al.
1993).

18

Chapter 2 Literature Review

The subsurface settlement profiles can also be reasonably approximated by a


Gaussian distribution curve in the same way as surface settlements. However, based on
the field data collected as shown in Figure 2.15, Mair et al. (1993) proposed that at a
depth z below the ground surface and above a tunnel depth Z0, the trough width
parameter for tunnels constructed in clay, i, can be expressed as:

i = K ( zo z)

(2.3)

0.175 + 0.3251 z
zo

K=
1 z

z o

(2.4)

Trough width parameter is shown to increase with depth and significantly


underestimated if a constant value is assumed. Similar patterns of increase in K was
observed in studies by Moh et al. (1996) and Dyer et al. (1996) irrespective of the soil
conditions encountered. Centrifuge studies by Grant and Taylor (2000) confirmed that
the proposed variation of K with depth for clay by Mair et al. (1993) provided a good
fit to the data obtained from tests within a certain range between the ground surface
and tunnel axis level.

Ran et al. (2003) observed that the Gaussian Curve can only describe the shortterm surface settlement well. A comprehensive review of field data of postconstruction settlements above tunnels in soft clay has been carried out by Shirlaw
(1993), Shirlaw et al. (1994) and Shirlaw (1995). Shirlaw (1993) recorded examples of
long-term settlements where the long-term component had the effect of increasing the

19

Chapter 2 Literature Review

short-term settlement by up to a factor of 10. The more typical settlement increase in


the long-term is in the order of 30% to 100%.

A case history study covering an 11-year period on the Haycroft Relief Sewer
at Grimsby in very soft clay soil had recorded that the initial settlement trough
increased from 1.5Z to a final value in excess of 4Z, as shown in Figure 2.16. The
maximum surface settlements over time is shown in Figure 2.17 reveals that the
consolidation of the disturbed settlements around the tunnel in long-term cannot be
neglected. Similar time dependent responses can also be found in Shanghai Metro
Tunnel No.2 (Zhang et al., 2004). The ratio of the maximum immediate settlement to
maximum long-term settlement shown in Figure 2.18 clearly demonstrates that the
long-term settlements were significant, and hence cannot be neglected.

Thus, generally it has been clearly shown that post-construction settlements can
be significant, particularly for tunnels in soft and compressible clay. Some case
histories shown in Figure 2.19 have much wider settlement troughs in the long-term.
Similar widening of settlement trough has been reported by a number of authors (e.g.
Glossop, 1978; Howland, 1980; OReilly et al., 1991 and Bowers et al., 1996).

Mair and Taylor (1997) concluded that the long-term settlement troughs are
similar to classical Gaussian curve associated with short-term settlement when positive
excess pore pressure are generated during tunnelling; whereas wider long-term
settlements are related to tunnel lining acting as drain and the development of steady
state seepage towards the tunnel. Hence, the major factors influencing the development
of long-term settlements are (i) initial magnitude and distrbution of pore water pressure;

20

Chapter 2 Literature Review

(ii) compressibility and permeability of the soil; and (iii) pore pressure boundary,
particularly the permeability of the tunnel lining relative to the permeability of the soil.

For normally consolidated clays, significant zones of positive excess pore


pressure can be generated even for a tunnel where unloading occurs, as shown by
Schmidt (1989). Negative excess pore pressures were noted included near the tunnel,
while positive excess pore pressures due to shearing may result a short distance away.
The pattern of pore pressure is shown in Figure 2.20, whereby two similar case
histories were observed for Shanghai Metro (Schmidt, 1989) in Figure 2.21 and
Thunder Bay Sewer Tunnel (Shirlaw et al., 1994) in Figure 2.22. The extent of the
positive excess pore pressure in Thunder Bay Sewer Tunnel spanned over six tunnel
diameters from the tunnel centre line.

For lateral soil movement due to tunnelling in long-term, very limited field data
are available. Emeriault et al, 2005 showed that in Toulouse subway line B, the
transverse movements were significantly increased over time, but stabilized after 15
days of tunnel excavation, as shown in Figure 2.23. Besides, Balasubramanian (1987)
reported both vertical and horizontal soil movement over almost a year time for the
Singapores largest effluent outfall pipeline in tunnel through various types of soils,
mostly marine clay. The data shown in Figure 2.24 demonstrated the time dependent
behavior of horizontal soil movements.

21

Chapter 2 Literature Review

2.3.2 Centrifuge Model Tests of Tunnelling-Induced Soil Movement


With the advances of new image processing technology, centrifuge modelling emerges
to be an attractive technique to investigate the effects of tunnelling-induced soil
movement on adjacent piles. The results enable an in-depth understanding into the
mechanism of ground responses associated with tunnel construction in terms of surface
and subsurface soil movements, as well as soil stresses.

Feng (2003) and Ran (2004) adopted the technique proposed by Sharma et al.
(2001) for the simulation of ovalisation of tunnel using high density polystyrene foam.
Figure 2.25 shows the surface settlement troughs under approximate similar volume
loss in clay and sand, as reported by Ran (2004). The two measured settlement troughs
follow Gaussian curve well. However, it is evident that the settlement troughs are
markedly different as the sand settlement trough is much narrower than that of clay.
This indicates different settlement propagation mechanisms in clay and sand. In clay,
the deformation of the soil propagates gradually upwards and outwards from the
tunnel cavity to the ground surface. However, the deformation in sand propagates
sharply and almost vertically from the tunnel to the ground surface. Therefore, the
different mechanisms suggest that the ground deformation in sand may cause more
severe damages to the ground surface or structures above and nearby the tunnel. For
clay, the soil movements cause differential settlement spreading a wider range. This
may explain why sinkholes on the ground surface associated with tunnelling are
mainly spotted in competent soils like sand in the field; while in clay, such drastic
settlements are less common.

22

Chapter 2 Literature Review

Figure 2.26 shows the surface settlement trough over time in clay (Ran, 2004).
It is noted that the measured short-term surface settlement trough follows the Gaussian
distribution curve fairly well. After the completion of tunnel excavation, the soil
continues to settle with time and the rate of increase in settlement decreases with time.
The incremental soil settlements become negligible after 720 days of tunnel excavation.
However, Gaussian curve is found to be inappropriate to depict the measured longterm surface settlement troughs. The measured final trough has a somewhat wider
parabolic shape than that of Gaussian curve. Furthermore, Gaussian distribution curve
largely underestimates the measured settlement at the far end of the ground surface,
showing that the spread of the surface settlement trough increases over time. Grant and
Taylor (2000) carried out a series of centrifuge tests to investigate tunnelling-induced
ground movements in clay. As discussed in Section 2.2.2, tunnelling was simulated by
reducing the compressed air pressure in a model tunnel lined with a latex membrane
and this caused a uniform radial contraction of tunnel. Figure 2.27 shows the profile of
normalized vertical and horizontal ground movements at different subsurface
elevations. In the near surface region, the horizontal movements are not well described
by assuming an average vector focus (Based on Grant and Taylor (2000), the average
position of the vector focus can be used to give average distribution of horizontal
movement) but the agreement is very good at all other elevations. The studies also
suggested that a constant trough width regardless of volume loss is common to all of
the tests and the centrifuge studies confirmed that the proposed variation of K with
depth for clays by Mair et al. (1993) provides a good fit to their centrifuge test data.
However, the long-term behaviour of the soil movements is not studied, probably due
to the soil condition of moderately stiff clay with less significant

in long-term

settlement.

23

Chapter 2 Literature Review

Simulation of tunnelling using liquid pressure was proposed by Loganathan


et al. (2000). The tunnel deformation pattern is uniform radial contraction. Figure 2.28
shows that the ground settlement troughs measured in the centrifuge tests match well
with his analytical prediction (Loganathan et al., 1998) which is based on oval-shaped
deformation. However, the long-term effect has not been studied by Loganathan et al.
(2000).

2.3.3 Predictive Methods of Tunnelling-Induced Soil Movement


Although attractive as a predictive tool, analytical methods are mathematically limited
in the efforts required to derive solutions accounting for material non-linearity,
complex geometries and time effects. Thus, relatively few analytical solutions are
available.

Sagaseta (1987) presented an analytical solution to predict tunnelling induced


ground movements for a weightless incompressible soil by simulating ground loss
around a tunnel in the form of a point sink. The tunnel is first assumed to be located
within an elastic infinite medium where it collapses uniformly. The soil is modelled as
a linear-elastic material and the solution is based on fluid mechanics concepts.

Solutions derived by Sagaseta (1987) are subsequently extended by Verrujit


and Booker (1996) to account for compressible materials and the effects of ovalisation
of the excavated tunnel boundary. However, the method gives a wider surface
settlement profile and larger horizontal movements than observed in practice.
Loganathan and Poulos (1998) modified Verruijt and Bookers solution to give

24

Chapter 2 Literature Review

narrower settlement troughs and to account the construction effects empirically.

Osman et al. (2006a) developed a kinematic plastic solution for ground


movements around a shallow, unlined, tunnel embedded within an undrained clay layer.
In this solution, the pattern of deformation around the tunnel is idealised by a simple
plastic deformation mechanism (see Figure 2.29). Within the boundaries of the
deformation zone, the soil deforms compatibly following a Gaussian distribution.
Outside this zone, the soil is assumed to be rigid. This mechanism does not incorporate
slip surfaces. Osman et al. (2006a) demonstrated that the upper bound theorem applied
to distributed shearing mechanism offered a reasonable assessment for collapse, and
also matches the displacement field observed in centrifuge tests of tunnel failure in
clay (Mair, 1979). It was also demonstrated that the shape of the surface settlement
profile remained the same as the magnitude of settlement increased towards failure
(Osman et al. , 2006a).

In addition, Osman et al. (2006b) also demonstrated that an upper bound style
of calculation is also capable of predicting ground displacements at any stage prior to
failure, representing the soil as a strain-hardening plastic material. A simplified closedform solution is provided for the prediction of maximum surface ground settlement for
the particular case of deep tunnelling. This solution is obtained by integrating the
equilibrium equations along the tunnel centre-line from the tunnel circumference up to
the ground surface and by invoking radial symmetry. A simple power curve was used
to model the stressstrain relations. These analytical solutions for maximum surface
settlements have also been validated against the centrifuge test data, and gave close
correspondence for deep tunnels but under predicted tunnel support pressure by about

25

Chapter 2 Literature Review

20% for shallow tunnels (Depth to tunnel crown/ depth of tunnel axis, C/D <3).
Analytical solutions providing the most convenient way in predicting
tunnelling induced ground movements. However, arguably accurate empirical method
has certain limitations in accounting for the effect of ground conditions, construction
methods, post-construction ground responses and precise tunnelling model.

In view of the above, a well-calibrated GAP method was proposed by Rowe


et al. (1983). The GAP parameter is defined as the magnitude of the equivalent twodimensional (2D) void formed around the tunnel due to the combined effects of the
three-dimensional (3D) elastoplastic ground deformation at the tunnel face, overexcavation of soil around the periphery of the tunnel shield, and the physical gap that
is related to the tunelling machine, sheild, and linming geometry.

The GAP parameter can be estimated using a theorectical method developed by


Lee et al. (1992), once the details of the machine support system and the soil
parameters are given. Comprehensive guidelines have been provided to calculate the
gap parameter (Lee et al. 1992), as

GAP = Gp + U3D + w

(2.5)

where Gp represents the difference between the cutter head and outer lining diameter
while U3D

and w account for 3D heading effects and workmanship quality,

respectively.

26

Chapter 2 Literature Review

In practice, as pointed out by Rowe et al. (1983), the radial ground deformation
is not uniform since the equivalent 2D gap (tail void) around the tunnel is non-circular
(e.g. typically oval-shaped) as shown in Figure 2.30. The possible reasons for the
formation of an oval-shaped gap around the tunnel are: (1) tunnel operators advance
the shield at a slightly upward pitch relative to the actual design grade to avoid the
diving tendency of the shield; (2) the tunnel lining settles on the ground when the tail
piece is removed; and (3) 3D elasto-plastic movement of the soil occurs at the tunnel
face.

Loganathan and Poulos (1998) extended the study from Lee et al. (1992) by
incorporating the shape of tunnel deformation to predict tunnelling-induced undrained
ground movements around a tunnel in soft ground. The traditional definition of the
ground loss parameter is redefined as equivalent ground loss parameter with respect
to the gap g parameters and incorporated in the analytical solutions. The nonlinear
ground movement due to the formation of an oval-shaped gap is then modelled by
adopting an exponential function to the equivalent undrained ground loss with
appropriate boundary conditions. Hence, the analytical solution models the effect of
non-uniform soil convergence around a deforming tunnel as shown in Figure 2.31.

The analytical solution of ground loss with horizontal and vertical distance x,z
from the tunnel centre is given as:

1.38 x 2
0.69 z 2
+

2
H 2
( H + R )

x , z = 0 exp

(2.6)

27

Chapter 2 Literature Review

where 0 is the ground loss ratio, H is the tunnel depth, z is the depth below ground
surface and x is the lateral distance from tunnel centre-line.

Although the method has been successfully used to back analyze some case
histories in clay, calculated results have to be treated with caution as certain important
conditions necessary in the derivation of analytical solutions are violated and the
volume loss is not conserved for undrained cases when empirical assumptions are
introduced (Cheng, 2003).

The accuracy of the oval-shaped radial displacement of tunnelling has been


further verified by the analytical solutions presented by Park (2005) through the case
studies. The oval-shaped ground deformation pattern (Loganathan and Poulos, 1998;
Park, 2004) is imposed as the boundary condition of the displacement at the opening to
consider real non-uniform ground deformation pattern. The gap parameter (Lee et al.,
1992) is used to describe the displacement at the opening.

Four simple boundary conditions, one for uniform radial deformation pattern
(B.C.-1) and three for oval-shaped deformation pattern (B.C.s-24), are considered at
the tunnel opening as shown in Figure 2.32. The boundary condition B.C.-2 (ovalshaped) is chosen for further study to give a conservative estimation for lateral
displacement. Five case studies have been used to check the applicability of the
proposed analytical solutions. The surface and maximum subsurface settlements and
the lateral displacement predicted using the proposed analytical solutions are
comparable to those from the methods of Verruijt and Booker (1996) and Loganathan
and Poulos (1998), and in reasonable agreement with field observations for tunnels in

28

Chapter 2 Literature Review

uniform clay. Hence, it can be concluded that the oval-shaped radial displacement is a
better tunnelling model that provides more realistic predictions of ground movement in
uniform clay.

2.4

TUNNELLING-INDUCED PILE RESPONSES

2.4.1 Field Studies of Tunnelling-Induced Pile Responses


Selemetas et al. (2005) presented the results of a full-scale trial investigating the
effects of tunnelling on piles. The field study took place in Channel Tunnel Rail Link
(CTRL) Contract 250, in Dagenham, Essex, UK. The Contract involved the
construction of 5.2 km of twin 8-m diameter bored tunnels using two Lovat Earth
Pressure Balance (EPB) machines with tail-skin grouting. The study involved the
installation, loading and monitoring of four instrumented piles along the route of the
twin tunnels . A comparison was made between the resulting pile head settlements due
to tunnelling with the surrounding ground surface movements. The results identified
three zones of influence in which the pile settlements were correlated to the ground
surface settlements (see Figure 2.33):
1. Piles in Zone A settled 2-4mm more than the ground surface,
2. Piles in Zone B settled by the same amount as the ground surface,
3. Piles in Zone C settled less than the ground surface.
4. The well-established Gaussian curve describing the magnitude of ground surface
settlement due to tunnelling could be used as a reference frame for the assessment
of pile settlement due to tunnelling.

29

Chapter 2 Literature Review

The study confirmed that pile head and ground surface settlement can be
correlated as presented in other studies by Kaalberg et al. (2005) and Jacobsz et al.
(2001).

Jacobsz et al. (2005) reported the case studies for the construction of the
tunnels for the CTRL project in London on the effects of tunnelling on piled
foundation. Three piled bridge pier foundation are described, one with end-bearing
piles and the other two friction piles. In the case of end-bearing piles, the settlement of
the superstructure was judged to be the same as the soil (Terrace Gravels) at pile toe
level. These were estimated and the bridge structure deemed safe for the level of
movement anticipated. Figure 2.34 shows a section of the friction pile case studies
where the pile toes were very close to the tunnels. The Terrace Gravels were grouted
as a mitigation measure to increase shaft capacity at that elevation and to create a
pseudo-slab beneath the pile caps. Total surface settlements of 8 mm to 10 mm were
observed (volume loss of 0.6%) with no detrimental effects on the bridge. In the third
case, the strains along the length of the pile, both vertically and laterally (to obtain
bending strain) were estimated from the ground movements with depth assuming full
friction at the soil-pile interface. The results indicated that the piles would not be overstressed and that assuming that the pile movement is the same as that for the Free-field
surface settlement is conservative. No mitigation measure was implemented and no
damage was sustained for the bridge. It is recommended that the pile capacities should
be re-evaluated as there is potential redistribution of loads in the piles.

Pang et al. (2005a) presented data from part of the MRT North East Line
Contract 704 in Singapore, where forward-thinking enabled instrumentation to be

30

Chapter 2 Literature Review

installed in bridge pier piles so that the influence of future planned tunnels, running
parallel to the bridge could be assessed. The data from 2 piles of a four-pile group
supporting a bridge pier are presented. The piles are 62 m long and 1.2 m in diameter
with four sets of strain gauges installed orthogonally, in pairs, to enable average axial
loads and bending moments in transverse and longitudinal directions to be determined,
as shown in Figure 2.35. The data presented related to a 6.3-m diameter EPBM tunnel,
constructed in residual soils, at 1.6m from the nearest piles at a depth of 21 m (to its
axis). The surface settlement profile due to tunnelling follows a Gaussian form with a
maximum value of about 18 mm. Correlating the developing settlements with TBM
position has enabled the volume losses related to the different phases of the tunnel
process to be identified. It is reported that the range of volume loss was between 0.32
and 1.45% only.

Information from the strain gauges within the piles reveals that the piles
experience down-drag, registered as increasing axial force, with greater force
developing in the pile nearer as might be expected. Calculations indicate that downdrag loads were between 9 and 43% of the structural capacity of piles with peak value
occurring when the face of the TBM was in line with the piles. Post-tunneling
measurement of the development of axial force in pile P1 at Pier 20 showing the timedependent behavior of drag-load as indicated in Figure 2.36. Clear trends in bending
moment distributions along the length of pile are also shown, with maximum values,
although small, occurring in the close vicinity of the tunnel (see Figure 2.37). Also
evident is shielding of the outer pile by the inner pile between it and the tunnel (see
Figure 2.37). Some interesting relation between volume loss and axial force and
bending moment are also presented, showing increase in both quantities with volume

31

Chapter 2 Literature Review

loss. It is also concluded that a volume loss up to 1.5% does not seem to have a
significant effect on the piles.

Cham (2007) analysed the responses of instrumented piles and Free-field


ground instrumentations to tunnelling processes such as tail void grouting, application
of face pressure, tunnel advancement and thrust force in a field study involving twin
tunnelling in MRT Circle Line Stage 3 (CCL3) Contract 852 in Singapore. The tunnel
diameters are 6m and tunnelling was carried out with 2 Earth Pressure Balance
Machines (EPBM) at a depth of 23m below the ground surface, through residual soil of
Bukit Timah Granite, Kallang Formation and Old Alluvium. Tunnelling of the second
tunnel commenced 4 months after the first and at the site of field study, the twin
tunnels have a lateral clearance of 2.5 m. Figure 2.38 illustrates the positions of
existing instrumented piles relative to tunnels.

The instrumented piles are reinforced-concrete bored piles with diameter of


either 600mm or 800mm, and of penetration depth ranging from 27m to 33m. At the
closest points, the piles lie within 1m of the tunnel lining. It was observed that downdrag forces developed when the TBMs were directly adjacent to the piles. The
maximum induced axial force due to twin tunnel advancement ranged from 7% to 72%
of the pile structural capacity. In addition, piles experienced additional bending
moments due to tunnelling and the induced bending moment was observed to increase
with increasing volume loss.

The field studies of tunnel-soil-pile interaction help to improve the current


knowledge on for the prediction of pile responses due to tunnelling. Prior planning and

32

Chapter 2 Literature Review

arrangements are made to instrument the pile as reported by Pang et al. (2005) are
strongly recommended to study the influence of future planned tunnels to existing piles.

2.4.2 Centrifuge Model Tests of Tunnelling-Induced Pile Responses


Nearly all existing field cases are limited in various aspects of measurements, which
may be due to deficiencies in instrumentation planning or difficulties in collecting field
data. Thus, centrifuge modelling is an alternative method to study the tunnel-soil-pile
interaction problem.

Loganathan et al. (2000) presented the model tunnel in-filled with oil to
simulate the uniform radial contraction in centrifuge to the pile responses due to tunnel
excavation in overconsolidated clay. The scope of the study focused on friction piles
(single pile and a 2x2 pile group). The effects of pile tip elevation relative to tunnel
axis level and volume loss on the displacements and performance of piles were
investigated to study the interaction problem. The relative position of the piles in
various tests is shown in Figure 2.39. Three tests were performed with the tunnel
located above, at and below the pie tip level. The induced bending moment and axial
force profiles at a volume loss of 1% are presented in Figures 2.40 and 2.41,
respectively. It is observed that both the induced maximum bending moment and axial
force occurred approximately at the tunnel spring line level in long pile cases where
the pile tips were below the tunnel spring line. The maximum bending moments
occurred just above the pile tips and the pile axial force increased from the pile head to
the pile tip in a short pile case with pile tip at or above the tunnel spring line level. The
comparison of the three tests showed that for single piles, the maximum bending

33

Chapter 2 Literature Review

moment was the largest when the pile tip was located at the tunnel spring elevation,
whilst the maximum axial force was the largest when the pile tip was above the tunnel
spring line. It was concluded that the maximum measured bending moments vary
almost linearly with ground loss values below 5%. As such, it was postulated that an
elastic analysis may be performed to predict tunnelling-induced pile behaviour if the
ground loss value was less than 5%.

Feng (2003) performed a series of centrifuge tests to investigate the pile


responses associated with a lined tunnel in dry sand at NUS and Ran (2004) extended
the studies by Feng (2003) to clay. The simulation method of tunnel excavation
proposed by Sharma et al. (2001) using polystyrene foam was adopted and the tunnel
deformation is oval shape. Ran (2004) reported two major series of tests involving the
effects of pile-to-tunnel distance and pile length, as well as on the effects of volume
loss. Figures 2.42 and 2.43 present the typical pile axial and lateral responses over time.
From the studies, it is demonstrated that the pile responses in clay are time-dependent.
Moreover, it is evident that the induced pile responses in clay are smaller than those in
sand (Feng, 2003). The limitation in the studies is that the model tunnel simulating the
ovalisation of tunnel, and hence the induced pile bending moment demonstrates an
opposite behaviour as most practical cases.

A recent study on the influence of tunnelling on piled foundations was


completed at the University of Cambridge by Jacobsz et al. (2005). The centrifuge
model study (Jacobsz, 2002; Jacobsz et al., 2004) focused on tunnelling near driven
piles in dense sand. Both single piles and pile groups were considered and the model
was constructed at a scale of 1:75. The model tunnel comprised a 50-mm brass pipe

34

Chapter 2 Literature Review

surrounded by a 1-mm thick latex rubber membrane. The 4-mm thick annulus between
the pipe and membrane was filled with water which could be discharged accurately to
impose volume losses from 0% to approximately 20% on the surrounding sand. It was
intended to use the model tunnel to impose relatively realistic plane-strain tunnellingrelated ground movements on the surrounding ground, rather than model the
progressive advance of a tunnel face and uniform radial displacement around tunnel is
simulated. A number of instrumented model piles were located at various offsets and
depths during the tests.

A zone of influence around a tunnel was established (see Figure 2.44) in which
significant base load reduction, accompanied by large pile settlements, were noted
should a certain volume loss be exceeded. Piles with their bases outside the zone of
influence did not suffer large settlements even at volume loss up to 10%. The stresses
exerted by piles on the surrounding ground result in a subsurface settlement profile
different from the Free-field situation. Pile settlements can however be approximated
by the Free-field surface settlement should the pile shaft capacity not be exceeded due
to volume loss. The piles in the centrifuge study possessed significant reserve
(immobilized) shaft capacity. Should piles not have the reserve shaft capacity, e.g.
where piles are end-bearing in sand with the shafts surrounded by soft clay, volume
loss may cause more rapid settlement. Pile groups behave in a similar fashion to
volume loss than individual piles (see Figure 2.45). Load transfer from one pile to
another within a group only occurs once the shaft capacity of a given pile has been
mobilized causing its settlement to become significant. For pile groups, these usually
occurred at large volumes which are undesirable in practice.

35

Chapter 2 Literature Review

In practice, the over-excavation of the tunnel and the gap between the shield
tunnel machine and lining would cause the tunnel spring line to move inwards into the
tunnel resulting in inward soil movements towards tunnel at the tunnel spring line. Lee
et al. (1992), Loganathn and Poulos (1998) and Park (2005) had successfully proposed
analytical solutions based on inward tunnel deformation or oval-shaped deformation.
Hence, it is of great interest to improve the model tunnelling technique to study the
pile responses due to inward tunnel deformation. It should be noted that based on St
Venants Principle, the exact deformation shape should be immaterial if the piles are
located sufficiently far away. However, in the present study, the deformation shape of
the tunnel is of great importance as the piles are close to the tunnel and the soil
movements would significantly affect the pile responses.

2.4.3 Predictive Methods of Tunnelling-Induced Pile Responses


Chen et al. (1999) presented a simple approach to assess tunnelling induced pile
responses where a two-stage uncoupled method was introduced. In the method, freefield tunnelling induced ground movements at the pile location is first approximated
based on the quasi-analytical method proposed by Loganathn and Poulos (1998) . The
movements were then applied to the soil elements surrounding the pile using separate
numerical programs (PALLAS and PIES) to assess the lateral and vertical pile
responses. The approach started with a basic problem as illustrated in Figure 2.46
where an existing single pile is situated adjacent to a tunnel under construction. The
induced pile responses together with the free-field soil movement of the basic problem
are shown in Figure 2.47. Subsequent parametric studies provided valuable insight into
the various factors affecting pile performance, in particular the variation of maximum

36

Chapter 2 Literature Review

induced bending moment and axial force with pile-to-tunnel distance and relative
position of pile tip to tunnel axis level. In general, the maximum bending moment and
axial force values decrease to insignificant magnitudes (less than 10% of value at
X=1D) beyond a respective distance of 2D and 5D from the tunnel centre line. At a
given horizontal offset from the tunnel centre line, the pile bending moment is
generally the greatest when its tip is below the tunnel axis level, decreasing as the pile
tip moves upwards. However, the pile lateral deflection profiles are almost identical in
shape and magnitude to imposed free-field soil displacements. This is probably due to
the low flexible stiffness of the pile and the homogeneous clay profile with constant Cu
and Youngs modulus with depth used in the analysis.

Free-field displacements are movements of the soil that occur at a distance


from the pile such that the displacements are not affected by the presence of the pile. A
free-field soil displacement method, in which a pile was represented by beam elements
and the soil was idealized using the modulus of subgrade reaction, was proposed by
Chow and Yong (1996). The magnitude of soil movement profile serves as input to the
method. With this idealization, non-homogeneous soil can be easily represented. This
approach requires the knowledge of pile bending stiffness, distribution of lateral soil
stiffness and the correct limiting soil pressure acting on the pile with depth.
Comparisons with available well-documented case histories suggest that the method
gives reasonable prediction of behaviour of pile subject to lateral soil movements.

Loganathan et al. (2001) incorporated the analytical solutions for tunnellinginduced ground movement (Loganathan and Poulos, 1998) into the computer
programme GEPAN (Xu and Poulos, 2000) to compute the response of a 2 x 2 pile

37

Chapter 2 Literature Review

group as shown in Figure 2.48. Generally, the front pile has slightly higher responses
than the rear pile. The lateral deformation and bending moment profiles for single
piles and piles in group are almost similar, except for a small difference in bending
moment at the pile cap location due to the fixity condition. However, for single
isolated piles, the settlement is slightly higher that the piles in the group. In addition,
the axial down-drag force estimated for a single pile is about 20% higher than the
down-drag force induced in a pile within the pile group.

As the pile responses due to tunnelling is essentially a 3D problem, Pang et al.


(2005b) carried out three-dimensional finite element analysis to back-analyse a case
history on the response of pile foundation subjected to shield tunnelling as described in
Section 2.4.1. The analysis back-analysed the behaviour of an instrumented pile group
where the full shield tunnelling process, including the application of face pressure,
shield tunnel advancement, over-cutting, tail void closure and installation of lining,
was simulated (see Figure 2.49). The trend of pile group deformation was noted to
deflect towards the tunnel transverse direction. In the longitudinal direction, prior to
the tunnel arriving at the pile group, the piles deflected towards the tunnel with
maximum movement at the pile head level and not at the tunnel spring line.
Subsequently, as the tunnel advanced past the pile group, the piles were gradually
pushed forward in the same direction of tunnel advancement. In addition, the front pile
was found to be subjected to higher response of up to 2.5 times compared to the rear
pile. Good agreement between the results from analysis and measured data was
obtained which validated the FE analysis shown in Figure 2.50.

38

Chapter 2 Literature Review

2.5

SUMMARY

A review of the literature to-date on effects of tunnelling on adjacent pile is reported in


this chapter. An overview of the published field studies, centrifuge model test and
predictive methods for tunnelling induced soil movements and pile responses are
presented. In addition, the short-comings from existing literature are discussed and
summarised as follows.

Considerable research studies have been carried out to simulate the process of
tunnels excavation in centrifuge. It should be noted that all these model
tunnels could not exactly replicate the prototype tunnelling process in the
field. The techniques for simulation of tunnelling in centrifuge developed
thus far have certain limitations such as maintaining initial stress-strain
behaviour of soil prior to tunnelling, precise tunnel deformation pattern,
boundary effects of model tunnel, and the complication in implementing the
excavation process during centrifuge flight.

Very few centrifuge studies had been carried out regarding the soil
movements and pile behaviours associated with inward tunnel deformation,
which is common in practice.

Opposite induced-pile bending moment was observed in centrifuge tests


carried out by Ran (2004) who had simulated the ovalisation of tunnel lining.
Hence, an improved simulation of tunnelling technique should be developed.

Limited field studies on the long-term tunnelling-induced ground movements


and pile responses. Centrifuge model tests emerge to be an alternative to
study such long-term behaviour, particularly in soft clay.

39

Chapter 2 Literature Review

Predictive methods available do not take into account the effects of pile group
or soil-structure interaction. Moreover, design charts proposed by Chen et al.
(1999) are confined to free head piles and linear elastic soil model.

Certain important conditions necessary in the derivation of analytical


solutions may be violated (e.g. volume loss is not conserved for undrained
cases developed by Loganathan and Poulos (1998)) when empirical
assumptions are introduced. Hence, extra care has to be exercised when
employing the analytical solution.

Although analytical solutions provide the most convenient way in predicting


tunnelling induced ground movements, these methods have certain
limitations in accounting for the effect of different ground conditions,
construction methods, post-construction ground responses and precise
tunnelling model.

The following chapters aim to present the detailed results of the present
centrifuge model study to investigate the observations and mechanisms of tunnel-soilpile interaction addressing some key issues raised in this chapter.

40

Chapter 2 Literature Review

Brass
lining

Polystyrene foam

Figure 2.1 Simulation technique of tunnelling using high density polystyrene foam
(After Sharma et al., 2001 and Feng, 2003)

Figure 2.2 Simplified tunnel lining deformation with time by simulation technique of
tunnelling using high density polystyrene foam (After Ran, 2004)

41

Chapter 2 Literature Review

Figure 2.3 Simulation technique of tunnelling - applying compressed air (After Grant
& Taylor, 2000)

Figure 2.4 Simulation technique of tunnelling - model tunnel infilled with oil (After
Loganathan et al., 2000)

42

Chapter 2 Literature Review

Figure 2.5 Simulation technique of tunnelling - model tunnel infilled with water
(After Jacobsz, 2002)

Figure 2.6 Sand pouring in process during model preparation with model tunnel
infilled with water (After Jacobsz, 2002)

43

Chapter 2 Literature Review

Opening to fix
the model tunnel

Figure 2.7 Strong-box with two openings to fix the model tunnel in place (After
Jacobsz, 2002)

Figure 2.8 Simulation technique of tunnelling - mechanical equipment of miniature


shield tunneling machine (After Nomoto et al., 1994)

44

Chapter 2 Literature Review

Figure 2.9 Simulation technique of tunnelling - mechanical equipment of shield model


machine (After Yasuhiro et al., 1998)

Figure 2.10 Simulation technique of tunnelling - mechanical model tunnel used to


simulate the tunnel volume loss by decreasing the diameter of model tunnel under 1g
(After Lee and Yoo, 2006)

45

Chapter 2 Literature Review

Figure 2.11 Gaussian curve approximating transverse surface settlement trough for
MRT project C852, Singapore (After Cham, 2007)

Figure 2.12 Definition of parameters controlling tunnelling-induced settlement trough


(After Standing & Burland, 2006)

46

Chapter 2 Literature Review

Figure 2.13 Gaussian curve approximating transverse surface settlement trough (After
Peck, 1969)

Figure 2.14 Variation in surface settlement trough width parameter with tunnel
depth for tunnels in clay (After Lake et al., 1992)

47

Chapter 2 Literature Review

Figure 2.15 Variation of trough width parameter K with depth for subsurface
settlement profiles above tunnels in clay (After Mair et al., 1993)

Figure 2.16 Initial settlement trough at Grimsby increased from 1.5Z to a final value in
excess of 4Z in long-term (After OReilly et al, 1991)
48

Chapter 2 Literature Review

Figure 2.17 The maximum surface settlements at Grimsby increased significantly over
the time (After OReilly et al, 1991)

Figure 2.18 The ratio of the maximum immediate settlement to maximum long-term
settlement for Shanghai Metro Tunnel No.2 (After Zhang et al., 2004)

49

Chapter 2 Literature Review

Figure 2.19 Normalized post-construction surface settlement troughs due to


consolidation of soft clay (After Shirlaw, 1995)

Figure 2.20 Estimated trend of excess pore pressure in normally consolidated clay
surrounding the tunnel (After Schmidt, 1989)

50

Chapter 2 Literature Review

Figure 2.21 Changes in Pore pressure for Shanghai Metro (After Schmidt, 1989)

Figure 2.22 Change in pore pressure measured at Thunder Bay Sewer Tunnel (Adapted
from data in Ng et al, 1986) (After Shirlaw et al, 1994)

51

Chapter 2 Literature Review

Figure 2.23 Transverse movements in Toulouse subway line B were significantly


increased over time, but stabilized after 15 days of tunnel excavation (After Emeriault
et al, 2005)

Figure 2.24 Horizontal soil movement for the Singapores effluent outfall pipeline in
tunnel (After Balasubramanian, 1987)

52

Chapter 2 Literature Review

Distance From Tunnel Centre-Line (m)


-25

-20

-15

-10

-5

10

15

20

25

Settlement (m)

-0.3

-0.6

-0.9

-1.2
Surface settlement trough in clay (measured)
Surface Settlement Trough in Sand (measured)

-1.5

Gaussian curve (clay)


Guassian curve (sand)
-1.8

Figure 2.25 Comparisons of surface settlement troughs in sand (After Feng, 2003)
and clay (After Ran, 2004)

Distance from tunnel centre-line (m)


-25

-20

-15

-10

-5

10

15

20

25

Surface Settlement (mm)

-40

-80
2 days
30 days
180 days

-120

360 days
720 days
1080 days
Gaussian curve (2 days)

-160

Gaussian curve (1080 days)

Figure 2.26 Ground surface settlement trough over time from a typical test (After Ran,
2004)

53

Chapter 2 Literature Review

Figure 2.27 Normalised Vertical and horizontal soil movement profile at different
subsurface elevations with best-fit curves: (a) 10mm below ground level; (b) 30mm
below ground level; (c) 70mm below ground level; (d) 100mm below ground level;
(After Grant and Taylor, 2000)

Figure 2.28 Comparisons of measured surface settlement and analytical solutions


(After Loganathan et al., 2000)

54

Chapter 2 Literature Review

Figure 2.29 Plastic deformation mechanism for tunnels in clay (After Osman et al.,
2006a)

Figure 2.30 Definition of GAP parameter (After Lee et al., 1992)

55

Chapter 2 Literature Review

Figure 2.31 Oval-shaped soil displacement around tunnel boundary


(After Loganathan and Poulos, 1998)

Figure 2.32 Boundary conditions of prescribed displacement (After Park, 2005)

56

Chapter 2 Literature Review

Figure 2.33 Zone of influence due to Earth Pressure Balance (EPB) shield tunneling in
London clay for Channel Tunnel Rail Link (CTRL) Contract 250, in Dagenham, Essex,
UK (After Selemetas et al., 2005)

Figure 2.34 A piled bridge pier foundation assessed during the CTRL project
(After Jacobsz et al., 2005)

57

Chapter 2 Literature Review

Figure 2.35 Typical section and instrumentation layout for pile-tunnel interaction study
for MRT North East Line Contract 704 in Singapore (After Pang et al., 2005a)

Figure 2.36 Post-tunneling measurement of the development of axial force in pile P1


at Pier 20 showing the time-dependent behaviour of drag-load MRT North East Line
Contract 704 in Singapore (After Pang et al., 2005a)

58

Chapter 2 Literature Review

Figure 2.37 Responses of pile foundation in terms of (a) axial force and (b) bending
moment for MRT North East Line Contract 704 in Singapore (After Pang et al., 2005a)

Figure 2.38 Illustration of positions of existing instrumented piles relative to tunnels


for MRT Circle Line Stage 3 (CCL3) Contract 852 in Singapore. (After Cham, 2007)

59

Chapter 2 Literature Review

Figure 2.39 Configuration of centrifuge tests (After Loganathan et al., 2000)

Figure 2.40 Tunneling-induced pile bending moments (After Loganathan et al., 2000)

60

Chapter 2 Literature Review

-5

-5

Depth below GL (m)

Depth below GL (m)

Figure 2.41 Tunneling-induced pile axial loads (After Loganathan et al., 2000)

-10

-15

-20

-10

-15

-20

-25

-25
-250

-200

-150

-100

-50

Pile bending moment (kNm)

50

-2

Pile lateral deflection (mm)

Figure 2.42 Variations of (a) induced pile bending moment profiles and (b) induced
pile lateral deflection profiles with time in typical test (After Ran, 2004)

61

Chapter 2 Literature Review

-5

Depth below GL (m)

Depth below GL (m)

-5

-10

-15

-20

-10

-15

-20

-25

-25

50

100

150

Pile axial force (kN)

200

250

5.85

5.9

5.95

6.05

6.1

Pile settlement (mm)

Figure 2.43 (a) Induced pile axial force profile and (b) pile settlement profile at 2 days
in typical test (After Ran, 2004)

Figure 2.44 Zone of influence around tunnel in which potential for large
pile settlements exists (After Jacobsz et al., 2005)

62

Chapter 2 Literature Review

Figure 2.45 Settlement, rotation and load distribution on triple pile group
(After Jacobsz et al., 2005)

63

Chapter 2 Literature Review

Figure 2.46 Layout of basic problem (After Chen et al., 1999)

Figure 2.47 Tunneling-induced pile responses and Greenfield soil movement


(After Chen et al., 1999)
64

Chapter 2 Literature Review

Figure 2.48 Numerical analysis of pile-group responses due to tunnelling (After


Loganathan et al., 2001)

Figure 2.49 Typical 3D finite elements mesh to back-analyze a case history on the
response of pile foundation subjected to shield tunnelling (After Pang et al., 2005b)

65

Chapter 2 Literature Review

Figure 2.50 Prediction of responses of pile foundation in terms of axial force and
bending moment using 3D finite element analysis (After Pang et al., 2005b)

66

Chapter 3 Experimental Set-up and Procedures

CHAPTER THREE

EXPERIMENTAL SET-UP AND


PROCEDURES

3.1

INTRODUCTION

This chapter presents the details of centrifuge modelling technique, experimental setup and procedures adopted in the present study. A technique of simulating tunnelling
in the centrifuge is developed and described in detail in this chapter. The preparation of
the strong box, soil specimens, fabrications and configurations of the instrumented
model piles as well as associated equipments are then elaborated. Finally, the reducedscale model set-up and test procedure are described in detail.

3.2

GEOTECHNICAL CENTRIFUGE MODELING

3.2.1 Principles of Geotechnical Centrifuge Modelling


Recently, there has been rapid development in geotechnical centrifuge modelling
technology world-wide, and centrifuge testing is now commonly used to study
geotechnical and geo-environmental problems. Geotechnical centrifuge modelling has
also been employed to complement conventional numerical analysis and field
monitoring (Schofield, 1998; Ng et al., 1998; Kimura, 1998). Each approach has its
own advantages in terms of quality of result, time and cost. Particularly in cases where
there are uncertainties in the applicability of a proposed design methodology, use of
67

Chapter 3 Experimental Set-up and Procedures

more than one approach permits calibration of results against each other and
verification of conclusions drawn.

Conventional scaled physical models, in spite of their advantages of wellcontrolled soil condition and extensive data monitoring, have significant limitations in
their usefulness as the fundamental mechanical behaviour of soil is highly non-linear
and stress-level dependent. However, by subjecting the 1/Nth scaled model in a
geotechnical centrifuge to an enhanced gravitational field N times the earth gravity, the
prototype stress can be reproduced in the reduced model, and the model test results can
be used to interpret the prototype behaviour in a rational manner.

An important principle of centrifuge modelling is to simulate the prototype


stress conditions in the model. This is done by subjecting the model components to an
enhanced body force, which is provided by a centrifugal acceleration of magnitude Ng,
where g is the acceleration due to the Earth gravity (i.e. 9.81 m/s2). Stress replication in
an Nth scale model is achieved when the imposed "gravitational" acceleration is equal
to Ng. Thus, a centrifuge is suitable for modelling stress dependent problems.
Moreover, significant reduction of test time for model tests such as consolidation time
can be achieved by using a reduced size model.

For centrifuge model tests, model scaling laws are generally derived through
dimensional analysis from the governing equations for a phenomenon, or from the
principles of mechanical similarity between a model and a prototype (Schofield, 1980,
Tan & Scott, 1985, Taylor, 1995a). Various commonly used scaling relations between
model and prototype can be deduced as summarized by Leung et al. (1991) in Table.
3.1. It can be readily deduced from Table 3.1 that the stress level of a 30-m deep clay

68

Chapter 3 Experimental Set-up and Procedures

can be correctly modelled by subjecting a 30-cm deep clay model to an elevated


"gravitational" acceleration of 100g (i.e., N=100). Also, a 4-hour centrifuge modelling
at 100g can correctly simulate a prototype soil settlement problem consolidated for
more than 4.5 years (i.e., 4xN2 or 4x1002 hours). Substantial time reduction and hence
cost savings can be achieved by adopting the centrifuge modelling technique.

3.2.2 NUS Geotechnical Centrifuge Facility


Figures 3.1 and 3.2 show the NUS centrifuge facilities. The centrifuge primarily
consists of a conical case, a driven shaft, and rotating arm, and two swinging platforms.
It has a capacity of 40,000 g-kg and operates up to a maximum g-level of 200g,
implying that the allowable payloads at 200g and 100g are 200 kg and 400 kg,
respectively. The structure of the centrifuge is based on the conventional dual swing
platform design.

The model package is normally loaded onto one of the swing platforms with
the opposing platform counter balanced by either counterweights or the other model
package with identical weights. When fully spun up during test operation, the distance
from the axis of rotation to the base of the platform is 1.871 m. The centrifuge is
driven by a hydraulic motor delivering up to about 37 kW power. The swing platform
has a working area that measures 750 mm x 700 mm and headroom of 1180 mm. A
stack of electrical slip rings is mounted at the top of the rotor shaft for signals and
power transmission between the centrifuge and the control room.

DC voltage is transmitted through the slip rings to the transducers mounted on


the centrifuge or the model package from the control room. Similarly, registered
69

Chapter 3 Experimental Set-up and Procedures

signals from the transducers are then transmitted via the slip rings. The signals are first
filtered by an amplifier system at 100 Hz cut-off frequency to reduce interference or
signal noise pick-up through the slip rings. The amplified signals are then collected by
a data acquisition system at a regular interval in the control room. Software called
Dasylab is used to process the signals whereby the signals are smoothened using a
block average. Two closed circuit cameras, which are mounted on the centrifuge,
enable the entire in-flight test process to be monitored in the control room. The NUS
centrifuge is described in detail by Lee at al. (1991) and Lee (1992).

3.3

EXPERIMENTAL SET-UP

Figures 3.3 and 3.4 show the sketch and photograph of the model package for the
present study, respectively. The main features of the centrifuge model are described as
follows.

3.3.1 Model Tunnelling Technique


There are many modes of ground movement associated with tunnel construction. In a
situation where the tunnel excavation has passed a particular section is considered, the
vectors of the ground movement developed will be more or less in the plane
perpendicular to the tunnel axis. Consequently it is reasonable to assume that a plane
strain model of long tunnel section would be a good representation of tunnellinginduced soil movements; this is usually referred to as a two-dimensional simulation
(Taylor, 1998).

70

Chapter 3 Experimental Set-up and Procedures

In the present study, an innovative model tunnelling technique has been


developed to simulate the inward tunnel deformation due to over-excavation. An ovalshape ground deformation pattern is imposed as the boundary condition and the gap
parameter (GAP) proposed by Lee et al. (1992) is used to quantify the amount of
tunnel over-cut. Loganathan & Poulos (1998) and Park (2005) evaluated that an ovalshape deformation pattern is in reasonable agreement with tunnel deformations
observed in the field, as reported in Chapter 2.

The longitudinal and cross section of the innovative model are shown in
Figures 3.5 and 3.6, respectively. The model tunnel is made of a circular rigid outer
plate and a hollow metallic circular tube of 60 mm diameter, simulating a 6-m
diameter prototype tunnel at 100g. The rigid plate helps to maintain a uniform GAP for
the entire model tunnel. The two radial bearings inside the model tunnel help to
facilitate a smooth movement of the sliding rod and provide support to the solid
aluminium sliding rods. There are nine small rods which are inserted into the
respective holes of the model tunnel. A rigid circular plate is then used to encircle the
model tunnel and an oval-shape GAP is created between the rigid circular plate and the
point of contact of nine small rods. The whole mechanism works as such when there is
a force pushing the aluminium sliding rod, the small rods will fall onto the three
thinner parts of sliding rod of smaller cross-sectional area. As such, the GAP in crosssectional view will close up and this simulates the inward tunnel deformation of the
oval-shape GAP.

71

Chapter 3 Experimental Set-up and Procedures

3.3.1.1 Advantages of Model Tunnel


The present model tunnel is able to simulate the precise volume loss during the process
of tunnelling. The percentage of volume loss has been calibrated by calculating the
area of surface settlement against the GAP (see section 2.3.3 )created in the model
tunnel at the undrained stage.

Moreover, the circular rigid outer plate shown in Figure 3.6 can provide a very
uniform oval-shaped of the GAP throughout the entire length of the model tunnel. This
ensures that the volume loss is constant along the model tunnel.

For the innovative mechanism created with the control of hydraulics system,
the closure of the GAP between the tunnel linings can be more effectively controlled.
With a switch of hydraulics pump, the hydraulic force can immediately push the
sliding rod inside the model tunnel forward and the small rods will then fall onto the
thinner part of the sliding rod. This causes an immediate closure of the GAP between
the tunnel linings and simulating the volume loss.

The accuracy and repeatability of volume loss control are good, as the model
tunnel is mechanically controlled. The test has been repeated and consistent test results
are obtained as in each test, the settlement trough is measured and the volume loss is
validated.

72

Chapter 3 Experimental Set-up and Procedures

3.3.1.2 Limitations of Model Tunnel


The model tunnel is a two-dimensional plane strain model. The three-dimensional
effects of tunnelling before tunnel approaching and after tunnel passing by cannot be
modelled. Nevertheless, the results of the present study are still a good representation
of tunnelling-induced soil movements as elaborated by Taylor (1998).

Owing to the constraints and difficulties faced in the model set-up in the
centrifuge, the model tunnel is pre-installed at 1g instead of in-flight tunnel excavation.
This does not simulate the real tunnel excavation process whereby the process of
tunnelling shall include excavation, installation of tunnel lining, jet grout etc. Although
the rigid outer aluminum lining may exert some stress around lining during
centrifuge spinning up, it is believed that this effect is unlikely to be significant as the
test is properly conducted after the forced acceleration field is stabilized. The wish-inplace model tunnel is a simplification and idealization of a cavity contraction, but with
an oval-shaped GAP and well controlled volume loss to simulate the soil movements
induced by tunnelling when the tunnel excavation has passed a particular section.

The minimum volume loss that can be simulated by the present model tunnel is
only 3% in order to maintain the accuracy of the experiment results. The accuracy of
volume control will be demonstrated in Section 4.3.2. To achieve the volume loss of
3% in 100g, a GAP parameter of 100-mm in prototype scale is needed. This means
that the GAP is as small as 1-mm in model scale. Few attempts have been carried out
to model lower volume losses in centrifuge but the results so far are not satisfactory.
Nevertheless, the magnitude of volume loss depends primarily on the method of
tunnelling and soil conditions. Although improvements in tunnelling technology have
73

Chapter 3 Experimental Set-up and Procedures

significantly reduced the volume loss due to tunnel excavation, The Civil Design
Criteria for Road and Rail (LTA, 2009) recommends the contractor to demonstrate the
suitability of the selected volume loss values in relation to the values of volume loss
that would occur during tunnelling. The typical values for tunnels up to 6.6-m diameter
in marine clay are in the range of 2% to 3.5%, depending on the tunnelling method
(15% volume loss should be considered if using TBM with compressed air) (LTA,
2009). In view of the above, a volume loss of 3% is simulated in the present study. To
evaluate the detrimental effect of higher volume loss, 6.5% is also simulated.

3.3.2 Instrumented Model Piles


Two different types of instrumented model piles are used in the present study to
examine the responses piles due to tunnelling. They were fabricated using square
aluminium tubes of 9.53 mm external width and 6.35 mm internal width. Ten pairs of
strain gauges were attached along the pile shafts to measure the bending moments
along one type of pile (termed bending pile) and axial forces along the other type of
pile (termed axial pile), see Figure 3.7. The strain gauges were protected by a thin
layer of epoxy resin for waterproofing. The final external width of each pile shaft is
12.6 mm corresponding to 1.26 m in prototype scale.

The strain gauges were wired and then connected with a TDS-300 strain meter
mounted on the centrifuge to form a full Wheatstone bridge circuit utilizing the
dummy strain gauges provided in the strain meter to produce a temperature
compensation system. The bending and axial piles were connected to the strain
meter with half-bridge mode and full-bridge mode, respectively. The detailed
connection principles and load-output relations were elaborated in Feng (2003). A very

74

Chapter 3 Experimental Set-up and Procedures

thin and light PVC plate with smooth and dark surface was attached to the bending
pile to facilitate reflection of laser-rays, for the purpose of measuring the pile
deflection. Conical tip was chosen to minimise the deviation of the piles from the
vertical during installation.

The flexural rigidity, EI, of the model pile, is 3.97x106 kNm2 at 100g, which is
equivalent to that of a 1300-mm diameter Grade 40 concrete bored pile. It should be
noted that it is not possible to correctly simulate the pile axial rigidity, EA, and
flexural rigidity, EI of a prototype pile simultaneously. The flexural rigidity is more
crucial as the pile bending moments and lateral deflection are more sensitive. A pile
with a higher flexural rigidity tends to attract larger bending moments but a lower pile
deflection. On the other hand, the pile axial force and settlement is less sensitive to
pile axial rigidity. For example, the settlement of a loaded pile will be mainly due to
the compression of soil while the elastic shortening of the pile shaft is normally
negligible as long as the pile axial rigidity is relatively high as in the present case.

The calibration of the pile bending moment and axial force was conducted
separately prior to the tests. Bending pile was calibrated by fastening the pile head
with a G-clamp and hanging mass centrally at the pile tip, the strain gauge outputs
were then related to the calculated bending moments. The axial pile was calibrated
by applying incremental loads on the top of the pile resting on a digital balance. The
corresponding strain gauge outputs were then related to the axial force.
The model piles are installed in 1g and positive excess pore water pressures are
generated during installation. Hence, the reconsolidation of the soil before simulating
the tunnel excavation is deemed to be necessary in order to recover the initial stress

75

Chapter 3 Experimental Set-up and Procedures

level of the soil and to allow the full dissipation of excess pore water pressure. Pore
pressure transducers are installed to monitor and ensure that the equilibrium state is
achieved before tunnel excavation.

3.3.3 Model Pile Cap


The model pile cap is made of aluminium with a thickness of 25 mm or 2.5 m thick at
100g. Two types of pile caps were fabricated for the 2- and 6-pile group configurations.

This is an improved design compared to previous pile caps fabricated at NUS


(Lim, 2001; Ong, 2004) as the pile caps were specifically designed to enable each pile
in the group to be tightened individually by tow rows of bolts in both directions. Thus,
the rotation or movement of the pile-pile cap connection can be minimized. Figure 3.8
shows the pile caps used in this study.
The prototype pile cap bending rigidity for the 2 groups is 3.24 x 108 kNm2 .
For the 6-pile group, the pile cap bending rigidity depends on the configuration of the
piles facing the tunnel excavation. If the 6-pile group consists of 3 rows of 2 piles per
row (2 piles x 3 rows) facing the excavation, the prototype EI of the pile cap is similar
to the 2-pile group case. However, in a 6-pile group of 3x2 configurations, the EI of
the cap is 1x 109 kNm2.

3.3.4 Strong Box


The strong box is made of stainless steel alloy to contain the soil specimens. It has
internal dimensions of 525 mm 200 mm 490 mm (length width height). One

76

Chapter 3 Experimental Set-up and Procedures

sidewall of the strong box is made of a 75-mm thick transparent Perspex plate, which
allows image acquisition by a video camera mounted to the centrifuge platform. A
measuring tape is attached to the Perspex wall to provide reference co-ordinates in
order to check the depth of the clay. Both the front (Perspex plate) and back walls of
the strong box can be removed to facilitate the installation of model tunnel and
transducers during the model set-up. To minimize the soil/strong box friction, all the
inner walls of the strong box are heavily greased. This would help to ensure the
deformation of the model ground is under plane strain condition.

3.3.5 Kaolin Clay


The physical properties of Malaysian kaolin clay used in the present study are
summarized in Table 3.2 (Goh, 2003). It has a liquid limit (LL) of 80%, plastic limit
(PL) of 40 % and hence a plasticity index (PI) of 40%, and a specific gravity, Gs, of
2.65. The coefficient of consolidation Cv and permeability at pressure of 100 kPa, are
40 m2/year and 210-8 m/s, respectively. The effective internal friction angle, , is
23o . Kaolin clay has critical state parameters of 0.244, average of 0.053, N of 3.35
and M of 0.9.

Figure 3.9 shows the measured in-flight undrained shear strength profile of the
Kaolin clay used at NUS, using miniature T-bar developed by Stewart and Randolph
(1991). The undrained shear strength profiles from the five tests are consistent and
repeatable The profile indicates an over consolidated layer down to 40 mm, below
which the shear strength increases nearly linearly with depth, and is consistent with
that for normally consolidated clay.

77

Chapter 3 Experimental Set-up and Procedures

3.3.6 Toyoura Sand


The sand that underlies the clay serves as drainage channel and socket for the pile. The
physical properties of Toyoura sand are listed in Table 3.3 (Teh et al., 2005). It has an
average particle size of 0.2 mm and specific gravity, Gs, of 2.65. The minimum and
maximum density of the sand is 1335 kg/m3 and 1645 kg/m3, respectively. The critical
state friction angle is 32o.

3.3.7 Potentiometers
Potentiometers (model LP-50F-61) were used to measure the surface settlements and
pile head settlements during the tests. This model has a measuring range of 50 mm and
an independent linearity of 0.2 %. The working part of the instrument consists of a
resistant and a rod whose stretch can alter the resistance of the resistor and hence the
output voltages. The output voltages are then linearly translated to the measured
distance. A round plastic plate is attached to the tail end of the rod to stop it
penetrating into the clay.

3.3.8 Pore Pressure Transducers (PPT)


Druck PDCR81 miniature pore pressure transducers (PPT) were used to monitor the
variations in pore water pressures during the centrifuge tests. Before the test was
carried out, the PPTs were de-aired using an electronic vacuum pump to release
trapped air bubbles in the PPTs to prevent acquisition of inaccurate readings. Each

78

Chapter 3 Experimental Set-up and Procedures

PPT comes with its own manufacturers calibration factor and this is incorporated to
determine the magnitude of pore water pressure. To confirm the manufacturers factors,
a digital air pump and a multimeter were used to calibrate the PPTs. The calibration
check was conducted by pumping air into the PPTs and recording simultaneously the
air pressure as well as the PPTs output voltage readings measured by the multimeter.

3.3.9 Non-Contact Laser Transducers


NAIS micro laser sensor LM10 (model ANR1250) were used to measure the lateral
pile head deflections during and after tunnel excavation. This model of sensors has a
centre point distance (distance between sensor and target) of 50 mm and a measurable
range of 10 mm within the centre point distance. The light source comes from a laser
diode and has a wavelength 685 nm and beam dimension of 0.6 mm x 1.1 mm at the
centre point distance. It has a linear resolution of 0.5 m, which translates to a linear
error of 0.5 mm at prototype scale.

The laser sensor has three main components, namely, the sensing body, the
relay cable and the controller/display unit. The sensing body houses the laser diode and
its function is to emit laser beam upon connected to a power supply of 24V DC. The
relay cable connects the sensing body to the DC power supply. The controller/display
unit is used to control and set the measuring limit of the sensor.

Calibration was carried out by securely attaching a 100-mm travel


potentiometer to the sensing body of the laser sensor. The transducer was connected to
a multimeter so that the digital display of the voltage could be displayed. The
transducer could take up to a maximum of 10 V. Hence, a direct relationship between
79

Chapter 3 Experimental Set-up and Procedures

displacement and voltage could be established, i.e. 1 V per 10 mm movement of the


transducer. The laser sensor has a specified optimum range of measurement to ensure
accuracy of the reading. However, readings outside this optimum range can still be
measured by the laser sensor but to a lesser degree of accuracy. Therefore, calibration
is ensured to lie only within this optimum range. As such, the transducer serves as an
indication or a ruler for the calibration of the laser sensor. The output voltage reading
on the laser sensor display unit varies with the displacement. Each set of readings of
the transducer and the laser sensor were recorded at every specified displacement
intervals so that correlation between displacement and voltage could be established.

3.4

IMAGE ACQUISITION SYSTEM

An advanced technique of image analysis has been developed at NUS as a method of


acquiring soil movement profiles from high-solution images captured in the centrifuge
model tests. Subsequently, Particle Image Velocimetry (PIV) is used to process the
resulting images (White et al., 2003; Zhang et al., 2005).

3.4.1 High Resolution Camera


JAI CV-A2 progressive scan camera coupled with a Tamron lens was mounted in
front of the Perspex window of the model container, as shown in Figure 3.10. The
cameras maximum grabbing speed is 15 frames per second (fps). However, the
capturing rate was set at 0.5 fps during tunnel excavation and at slower rate during post
tunnelling.

80

Chapter 3 Experimental Set-up and Procedures

3.4.2 Lighting System


Lighting system is important in producing high quality images for the purpose of post
processing the data. As shown in Figure 3.10, two spot lights, each with a 50 W
halogen bulb, were each mounted at the frame arm to provide uniformly distributed
lighting across the soil sample. The florescent lights inside the centrifuge enclosure
were turned off during the capturing of images while the centrifuge was in operation.

3.4.3 On-Board and Command Computers


The progressive scan camera is connected to a computer installed on-board of the
centrifuge, as shown in Figure 3.11. The on-board computer is capable of sustaining
high gravitational force without being damaged. All captured images during an inflight centrifuge test were stored in this on-board computer, in which the hard disk of
the on-board computer was specially designed to provide greater resilience to physical
vibration, shock and extreme temperature fluctuations.
The on-board computer was remotely controlled by another commands
computer in the control through a wireless connection to activate the capturing of live
images of the soil movement during an in-flight centrifuge test. All captured images
could be retrieved from the on-board computer after tests. Figure 3.12 displays the
picture captured by the JAI CV-A2 progressive scan camera, which will be
processed subsequently.

81

Chapter 3 Experimental Set-up and Procedures

3.4.4 Post-Processing of Images


New technique of image analysis, particularly Particle Image Velocimetry (PIV)
(White et al., 2003; Zhang et al., 2005) has been recently applied to geotechnical
centrifuge modelling. The PIV technique increases the details and precision of
deformation measurements. Take & Bolton (2004) and Zhang et al. (2005) confirmed
that this high precision can be achieved in centrifuge conditions. The GeoPIV software
(White and Take, 2002) has been used to implement the PIV technique for post
processing of the images captured during the centrifuge test.

The principles of PIV analysis are summarized in Figures 3.13 and 3.14. PIV
operates by tracking the texture (i.e. the spatial variation of brightness) within an
image of soil through a series of images. The initial image is divided into a mesh of
PIV test patches. Consider one of these test patches, located at coordinates (u1,v1) in
image 1. To find the displaced location of this patch in a subsequent image, the
following operation is carried out. The correlation between the patch extracted from
image 1 (time = t1) and a larger patch from the same part of image 2 (time = t2) is
evaluated. The location at which the highest correlation is found indicates the
displaced position of the patch (u2,v2). The location of the correlation peak is
established to sub-pixel precision by fitting a bicubic interpolation around the highest
integer peak. This operation is repeated for the entire mesh of patches within the image,
and then repeated for each image within the series, to produce complete trajectories of
each test patch. Details are presented in White and Take (2002), White et al. (2003)
and Zhang et al. (2005).

82

Chapter 3 Experimental Set-up and Procedures

3.4.5 Assessment of Effectiveness of Image Processing System


The image processing technique, PIV is used to track the soil markers so that the soil
movement due to tunnelling can be quantified. For clay such as kaolin, there is not
enough natural texture (e.g. sand) for the application of PIV technique (Zhang et al.,
2005). Hence, it is necessary to provide an artifical texture on the clay surface as PIV
technique requires random and unique textures of soil patches within the images for an
accurate analysis.

A typical reconsolidation stage of a centrifuge test has been conducted to assess


the effectiveness of the image processing system, as well as the performance of flocks
and beads as material in creating artificial textures for PIV analysis purpose.
Two methods of measuring the surface settlement are adopted. The first
method is the direct measurement of the surface settlement using potentiometer and the
second method involves measuring the soil settlement by tracking the movement of
flocks and beads using the image processing method.

On the left handside of the model tunnel, the exposed clay surface was
sprinkled with black and gray flocks, while on the right handsied of the model tunnel,
1-mm diameter black/blue/red beads were randomly embedded on the surface of plain
white clay. The set-up of the test is shown in Figure 3.15. Figure 3.16 shows the
settlement measured by these methods. It is observed that the measured settlement
analysed by image processing method agrees well with the direct measurement of the
surface settlement using potentiometer, with an error less than 5%. Amongst these two
different materials, beads demonstrate a higher accuracy compared with the flocks. It

83

Chapter 3 Experimental Set-up and Procedures

is probably the beads are embedded in the soil and move freely together with the soil,
but the flocks are only spread on the surface of clay and hence influenced by the
greased applied on the Perspex windows.

In view of this, the image processing analysis can be considered a competent


and reliable method to measure the soil movements in the present study, and different
colours of beads are used to create unique textures.

3.5

EXPERIMENTAL PROCEDURE

3.5.1 Preparation of the Soil Sample


The model ground was remoulded from Malaysian kaolin clay powder and water at a
weight ratio of 1 to 1.2 in a de-airing mixer. After 4 hours of mixing, the clay sample
was carefully poured into the strong box, in which a 30-mm thick sand layer is placed
at the bottom. To prevent air trap in the clay, water is poured into the strong box before
the clay is poured. Three rubber tubes are used to act as drainage for pore water of the
soil under consolidation. The lower part of the tubes is covered by the sand. A thin
geotextile was placed on top of the sand to separate the sand and clay. The clay slurry
was then carefully scooped into the strong box by immersing into water body to
prevent air trap until it reaches a predetermined height. The underlying sand and
rubber tubes together act as drainage function in this case. After that, the soil container
was shifted to a pneumatic loading frame. The clay was then pre-loaded under a
pressure of 20 kPa. During the loading process, some of the pore water may come out
from the rubber tubes, which further helps to act as drainage.

84

Chapter 3 Experimental Set-up and Procedures

3.5.2 Pre-Consolidation Process


Pore pressure and soil settlement throughout the in-flight consolidation and
reconsolidation were monitored by PPTs and potentiometers. Since the position of the
PPT was fixed and the elevation of the free water was known from the potentiometer
attached with floating ball, the corresponding hydrostatic pressure at PPT can be
calculated from the difference between the two measurements.

The pore pressure and soil surface settlement of a consolidation test over 13
hours are recorded so as to optimise the duration of spinning and for the ease of the
analysis. The results are shown in Figure 3.17. The pore pressure and soil surface
settlement appear to be stabilized after 8 hours of consolidation, whereby further
analysis using Asaokas method (1978) depicts that the final settlement after 8 hours
approaching 100% degree of consolidation (see Figure 3.18). Hence, this can eliminate
the uncertainty in term of long term self-weight soil consolidation in the analysis of the
test data after tunnelling.
The same monitoring was carried out at the reconsolidation stage for every test
to ensure that complete consolidation was restored, as swelling of the soil sample was
noted due to stress release during the set-up installation at 1g. Based on the
observation, a total of 6 hours are required to restore the final soil elevation during the
earlier preconsolidation stage.

85

Chapter 3 Experimental Set-up and Procedures

3.5.3 Installation of Model Tunnel and PPTs in 1g


Following the completion of self-weight consolidation, the back wall of the strong box
was opened. Another layer of grease is again applied to the back wall to make sure that
the clays boundary will act as free-roller support. The wooden tunnel installation
guide was designed in such a way that it can be used to make sure that the model
tunnel is inserted into the clay perpendicularly so as to minimize any soil disturbance.
This will enhance the accuracy of the whole experiment. The tunnel installation guide
will be fixed in position by 2 G-clamps. Along the tunnel installation guide, the
stainless steel tube (60 mm in diameter, 0.8 mm wall-thickness) was inserted into the
clay and excavates a cylindrical cavity through the two openings. The model tunnel
was then subsequently inserted into the cylindrical cavity. The model tunnel end cap is
connected to the hydraulic hose through an aluminium tube. The aluminium tube,
which was secured by a small frame, was then be connected to the hydraulic hose,
whereby the hydraulic force will be used to push the sliding rod inside the model
tunnel so as to simulate the closure of the oval-shaped tunnel gap.

Similarly, PPTs installation guide was used to ensure that the PPTs can be
carefully inserted into the clay perpendicular to the soil surface. PPTs will then be used
to monitor the change in pore water pressure throughout the entire experiment.

3.5.4 Preparation Works for PIV Analysis


After installed the model tunnel and PPTs, the back wall was then fixed back and the
front wall was then removed. As shown in Figure 3.19, different colours of 1-mm

86

Chapter 3 Experimental Set-up and Procedures

diameter beads were randomly embedded on the surface to produce an artificial texture
for the subsequent analysis of PIV. These beads are made of light PVC so that they
could move with the soil freely. The beads were pushed into the soil by the highly
greased Perspex window (see Figure 3.20) of the strong box to ensure a full perfect
contact and the beads can move together with the soil. Permanent control markers dots
with known centre to centre distance were marked on the Perspex window in order to
provide reference points to the subsequent image analysis by PIV.

3.5.5 Installation of Model Pile at 1g


Using the pile installation guide, the model piles were then carefully installed at the
predetermined distance at 1g. The pile installation guide was designed in such a way to
ensure that the model pile could be inserted vertically into the clay and this also
minimizes the disturbance to the clay. The pile guide is made of Perspex.

Two transducers were placed on the pile head to measure the model pile
settlement. Two non-contact laser transducers were used to measure the lateral
deflection of the pile head. The distance between the laser transducers and the bending
pile is about 50-mm. The transducers were attached to a stainless steel holder mounted
tightly onto the top of the container.

3.5.6 Test Procedures


The completed set-up of the entire model package is shown in Figure 3.21. The model
package was then spun up to 100g in 10 steps at 5 minute intervals for reconsolidation
87

Chapter 3 Experimental Set-up and Procedures

of the clay. After about 6 hours, the total pore pressure would be restored to the same
state as that in the consolidation stage and the test then began when the hydraulic valve
was switched on and the hydraulic force would push the sliding rob inside the
mechanical tunnel forward. When the sliding rod was pushed forward, the small rods
lying on the sliding rod would drop onto the thinner part of the sliding rod with smaller
cross-sectional area. This caused the gap between the rigid aluminium plate and the
model tunnel to close and the inward tunnel deformation at the tunnel spring line was
thus simulated in this way. The model tunnel was left in place to simulate the tunnel
lining to study the post-excavation ground deformation and pile responses. The
centrifuge would be kept at 100g for 3 hours after the completion of tunnel excavation.
All instruments were monitored regularly throughout the test.

88

Chapter 3 Experimental Set-up and Procedures

Table 3.1 Scaling relation of centrifuge modeling (After Leung et at, 1991)
Parameter

Prototype

Centrifuge Model at Ng

Linear dimension

1/N

Area dimension

1/N2

Volume dimension

1/N3

Density

Mass

1/N3

Acceleration

1/N

Velocity

Displacement

1/N

Stress

Strain

Force

1/N2

Time (viscous flow)

Time (seepage)

1/N2

Flexural rigidity

1/N4

Axial rigidity

1/N2

Bending moment

1/N3

89

Chapter 3 Experimental Set-up and Procedures

Table 3.2 Physical properties of Malaysian kaolin clay (After Goh, 2003)
Property

Value

Liquid limit, LL

80%

Plastic limit, PL

40%

Specific gravity, Gs

2.65

Coefficient of consolidation at 100 kPa, cv

40 m2/year

Coefficient of permeability at 100 kPa, k

210-8 m/s

Angle of internal friction, '

23o

Particle size*

3.0~5.5 m

Modified Cam-clay parameters:


M

0.9

0.244

0.053

3.35

* Manufacturer data
Table 3.3 Physical properties of Toyoura sand (After Teh et. al, 2005)
Property
Specific gravity, Gs
Average particle size, d50

Value
2.65
0.2 mm

Particle size, d10

0.163 mm

Minimum dry density, min

1335 kg/m3

Maximum dry density, max

1645 kg/m3

Critical state (constant volume)

32o

Friction angle, crit

90

Chapter 3 Experimental Set-up and Procedures

Rotating
arm
(In-flight position)

Drive
shaft

Swing platform
Bearings
(Static position)
Conical base

Figure 3.1 Schematic diagram of NUS geotechnical centrifuge

Slip Rings
On-board
camera

Strain
Meter

Balance
Arm

Counter
Weight
Payload
Conical
Base

Figure 3.2 Photograph of NUS geotechnical centrifuge with the model package
mounted on the platform

91

Chapter 3 Experimental Set-up and Procedures

Figure 3.3 Sketch of a typical centrifuge model package (All dimensions in mm)

Potentiometers
Lasers
Beads
Tunnel

Camera

Figure 3.4 Photograph of a typical centrifuge model package

92

Chapter 3 Experimental Set-up and Procedures

Figure 3.5 Longitudinal view of model tunnel set up

GAP
GAP

Figure 3.6 Cross-section of model tunnel

93

Chapter 3 Experimental Set-up and Procedures

9.53

25 25 25 25 25 25 25 25 25 25 15

50

70

40
12.6

Strain gauge
Plate to measure
deflections by lasers
Epoxy
coating

Aluminum
tube

End cap

Figure 3.7 Instrumented model pile (All dimensions in mm)

Figure 3.8 Model pile caps

94

Chapter 3 Experimental Set-up and Procedures

Undrained Shear Strength (kPa)


0

10

15

20

25

30

35

40

0
2
4
6

Depth (m)

8
10
12
14
16
18
20
22

TEST 1
TEST 2
TEST 3
TEST 4
TEST 5

24

Figure 3.9 In-flight undrained shear strength of clay

Lighting
system
Camera

CCTV

Figure 3.10 Image acquisition system

95

Chapter 3 Experimental Set-up and Procedures

Wireless
system

On board
Imaging PC

On-board
Camera

Strain meter

Hydraulic hose for


connection to tunnel

Figure 3.11 On board set-up

Control
marker

Model tunnel

Texture clay

Figure 3.12 Picture captured by JAI CV-A2 progressive scan camera for PIV
analysis

96

Chapter 3 Experimental Set-up and Procedures

Figure 3.13 Image manipulation during PIV analysis. (After White et al., 2003)

Figure 3.14 Evaluation of displacement vector from correlation plane, Rn(s): (a)
correlation of Rn(s); (b) highest correlation peak (integer pixel); (c) sub-pixel
interpolation using cubic fit over 1 pixel of integer correlation. (After White et al.,
2003)

97

Chapter 3 Experimental Set-up and Procedures

LVDT

LVDT

Control
marker

Flocks

Beads

Figure 3.15 Experimental set-up for assessment of effectiveness of image processing


system and comparison of performance of flocks and beads

Potentiometer 1
-694mm

Potentiometer 2
-702mm

Flocks
PIV -662mm

Beads
PIV -690mm

Error=-4.5%

Error=-1.5%

Figure 3.16 Results of assessment of effectiveness of image processing system


and comparison of performance of flocks and beads

98

Chapter 3 Experimental Set-up and Procedures

400

Total pore pressure (kPa)

300

200

100

PPT 11403 T
Hydrostatic

0
0:00

1:00

2:00

3:00

4:00

5:00

6:00

7:00

8:00

9:00 10:00 11:00 12:00 13:00 14:00

Model time- consolidation (hours)

Model time- consolidation (hours)


0:00

1:00

2:00

3:00

4:00

5:00

6:00

7:00

8:00

9:00 10:00 11:00 12:00 13:00 14:00

Model settlement (cm)

0
1
2
3
4
5

Figure 3.17 Pore pressure dissipation and settlement during consolidation stage
5

Settlement, Si (cm)

Sult= 0 / (1-1)
4

0=2.4615038
3

1=0.99925896

Sult= 3.325 cm

U=Sf /Sult= 3.325/3.325=100%

0
0

Settlement Si-1 (cm)

Figure 3.18 Estimation of ultimate settlement by Asaokas method (1978)

99

Chapter 3 Experimental Set-up and Procedures

Model tunnel

Texture clay

Figure 3.19 Different colours of beads were randomly embedded on the surface of clay

Control
marker

Figure 3.20 The Perspex window is highly greased to ensure free movement of soil

100

Chapter 3 Experimental Set-up and Procedures

Model pile

Potentiometer

Lasers

Model tunnel

Figure 3.21 Set-up of the entire model package in 1g (top view)

101

Chapter 4 Basic Test on Volume Loss

CHAPTER FOUR

BASIC TESTS ON VOLUME LOSS

4.1 INTRODUCTION
In order to interpret the pile behaviour due to tunnelling-induced soil movement, it is
important to examine the mechanism of tunnel-soil interaction. Particle Image
Velocimetry (PIV) technique (White et al., 2003; Zhang et al., 2005) has been used in
centrifuge model tests to obtain more accurate and detailed information on soil
displacements, as described in Chapter 3. With a better understanding and results
obtained in free-field experimentally in the present study, further evaluations can then
be made on tunnelling-induced pile responses.

4.2 TEST PROGRAM


The first half of this chapter presents the results of study on free-field soil movements
due to tunnelling. The test results on the effects of tunnelling with the same volume
loss on single free head piles are then presented in the second half of this chapter. The
complete test program presented in this chapter is shown in Table 4.1. The centrifuge
tests were performed under 100g. Unless otherwise stated, the test configurations and
results are presented in prototype scale hereinafter. For all tests, the thickness of kaolin
clay layer and the underlying Toyoura sand layer is 24 m and 3.5 m, respectively. The
tunnel cover C (distance from ground surface to tunnel crown) and tunnel diameter D

102

Chapter 4 Basic Test on Volume Loss

is 12 m and 6 m, respectively. The schematic and digital image of a typical test are
shown in Figures 4.1 and 4.2, respectively.

In the present study, the terminology short-term (ST) refers to the stage in
which tunnel excavation has just been completed, under undrained condition. On the
other hand, Long-term (LT) refers to the stage when the soil has completed
consolidation due to tunnelling. After 720 days, it was observed that changes in the
ground movement and pile responses are negligible. Hence the time of 720 days after
tunnel excavation is taken as reaching the long-term stage.

As discussed in Section 3.1.3.1.2, the magnitude of volume loss depends


primarily on the method of tunnelling and soil conditions. The typical volume loss
design values for tunnels of up to 6.6m diameter in marine clay are in the range of 2%
to 3.5%, depending on the method of tunnelling (15% volume loss should be
considered if using TBM with compressed air) (LTA, 2009). In view of the above, a
volume loss of 3% (Test 1) is simulated in the present study. To evaluate the
detrimental effects of higher volume loss, 6.5% (Test 2) is also simulated.

Test 1 has been repeated to evaluate the repeatability and consistency of the test
with and without the presence of pile. During both tests (with and without presence of
pile), the beads were randomly embedded on the surface to produce an artificial texture
for the subsequent analysis of PIV. Images captured in days 2, 180, 360, 540 and 720
were analyzed. It is observed that the results are consistent for both tests as long as the
pile is installed far enough behind the Perspex window of the strong box. Both Tests 1
and 2 presented in this chapter refer to centrifuge model tests conducted without pile.

103

Chapter 4 Basic Test on Volume Loss

In addition, Tests 3 and 4 were carried out in series 1 to investigate the


behaviours of long pile (pile tip below tunnel invert) under two different tunnel volume
losses. The pile-to-tunnel distance and pile embedment length are kept constant as 6 m
and 22 m, respectively, in these tests.

4.3 TUNNELLING-INDUCED SOIL MOVEMENTS (TESTS 1 & 2)


4.3.1 Cumulative Soil Movements
The cumulative soil displacement vectors and contours at different times after tunnel
excavation can be obtained using the PIV technique. Figures 4.3 and 4.4 show the
cumulative soil movement contours and vectors over time for both Tests 1 and 2,
respectively. For ease of comparison, the contour of cumulative soil movement of 10
mm is highlighted as bold dash lines in the plots. The 10 mm movement is selected as
the bench mark because the maximum allowable settlement for shallow foundation as
specified by the Civil Design Criteria for Road and Rail (LTA, 2006) is 20 mm and 10
mm is often set as the alert level. In the short-term, principal soil movements are
concentrated within a zone indicated in Figure 4.3(a). This zone may be identified as
the Immediate Shear Zone as the soil within this zone has likely been unloaded due
to tunnel excavation. For clay, the soil does not settle as a rigid body but gradually
deforms by arching, causing the radial stress in the immediate shear zone to be reduced
due to stress relief. This leads to the observed soil movement pattern and the settlement
trough at the ground surface. On the other hand, the zone outside the immediate shear
zone may be identified as the Support Zone, as the circumferential soil stresses
increase within this zone to support the arches formed in the immediate shear zone.

104

Chapter 4 Basic Test on Volume Loss

Qualitatively, it is expected that volumetric soil strain in the long-term would


increase due to soil consolidation. This might cause the soil movements to increase in
both the horizontal and vertical directions, as observed in Figures 4.3 to 4.4. It can be
observed from the figures that the shear zone propagates with time and becomes wider
over time due to post-tunnelling soil reconsolidation. This observation demonstrates
that with relatively large volume loss (>3%), the long-term effects of soil movement
induced by tunnelling could be substantial.

4.3.2 Soil Surface Settlement Troughs


The surface settlement trough along a plane transverse to the tunnel can be described
by the Gaussian curve (Peck, 1969). The popularity of the Gaussian curve as a
prediction tool for the magnitude of surface settlement due to tunnelling lies in its
simplicity and efficiency. The surface settlement curve, S, is given in Equation 2.1.

Figures 4.5 and 4.6 show the measured surface settlement troughs over time
obtained from PIV and potentiometers with volume loss of 3% and 6.5%, respectively.
It is evident that in the short-term (2 days), the surface settlement troughs follow a
Gaussian distribution curve with a maximum ground surface settlement of 41 mm for a
volume loss of 3% and 92 mm for a volume loss of 6.5%. This corresponds to the
respective imposed volume loss with simulated tunnel opening of approximate GAP =
100 and 200 mm (refer to definition of GAP in Section 2.3.3). Based on the above
findings, there is further evidence that the accuracy of volume control of the model
tunnel is good and reliable. As expected, the volume loss at the ground surface is close
to the tunnel volume loss under such undrained condition. The point of inflection, i, is

105

Chapter 4 Basic Test on Volume Loss

determined from the settlement trough at the point when the change of gradient is zero.
The point of inflection, i, of the settlement trough is determined to be approximately
7.5 m for both tests. This value is identical to the prediction of 7.5 m by Peck (1969),
using a trough width parameter k of 0.5 suggested by Mair et al. (1993) for tunnels in
clay. Thus it can be established that the observed settlement trough in the short-term
can be reasonably predicted using existing methods.

In the long-term, the ground settlement continues to increase with time, as


shown in Figures 4.5 & 4.6. It should be noted that the soil has practically completed
its self-weight consolidation before tunnel excavation, as illustrated in Figure 3.19.
The remaining self-weight consolidation settlement of the soil should be very small.
Hence the long-term soil movement is mainly due to stress relief of clay due to tunnel
excavation and the settlement trough S are noted to become wider over time. Shirlaw
(1993) presented case studies of long-term settlements and reported that the ground
settlement due to tunnelling and the extent of settlement trough can increase
significantly in the long-term in some cases. In contrast, Gaussian distribution curve is
found to be inappropriate for representing the long-term surface settlement trough with
a wider parabolic shape. Nevertheless, although the magnitude of maximum long-term
ground settlement is larger, the differential settlement for a wider settlement trough is
not as significant as that in the short-term.

Figure 4.7 clearly shows that that the maximum surface settlements measured
from potentiometers and derived from PIV match well, confirming the accuracy and
effectiveness of the PIV image processing technique. Similar good comparisons

106

Chapter 4 Basic Test on Volume Loss

between the settlement measurements obtained by potentiometers and PIV are


demonstrated in Figures 4.5 & 4.6 as well.

The maximum surface settlement often occurs at the tunnel crown for a single
tunnel case. Figure 4.7 demonstrates that the maximum surface settlements for both
Tests 1 & 2 increase over time. The rate of increase in settlement is significantly
reduced after a period of 360 days and becomes very small after approximately 720
days. This finding implies that for tunnel in soft clay with relatively large volume loss,
the surface settlements in the field should be monitored for the first 1 to 2 years after
tunnelling.

4.3.3 Subsurface Vertical Soil Movements


The vertical soil movements can provide clues on the mechanisms associated with
tunnel-soil-pile interaction, particularly on the induced pile axial forces and settlements.

Figures 4.3(a) and 4.4(a) indicate that in the short-term (ST, 2 days), the largest
vertical soil movements are spotted in the immediate shear zone above the tunnel.
However, this zone becomes wider in the LT as shown in the contour plots over time
in Figures 4.3(b) to (e) and 4.4(b) to (e). The propagation of vertical soil movement
trough seems to be an inverted half-ripple. This large vertical deformation zone is
critical and must be taken into consideration.

Figures 4.8 and 4.9 show the ST surface and subsurface settlement troughs for
Tests 1 and 2, respectively, in comparison with existing predictive methods proposed

107

Chapter 4 Basic Test on Volume Loss

by Mair et al. (1993) and Loganathan and Poulos (1998). Mair et al. (1993) proposed
that at a depth z below the ground surface, and above a tunnel depth of zo, the trough
width parameter for tunnels constructed in clays are given by Equations 2.3 and 2.4.
The solution of vertical displacement around a tunnel excavation proposed by
Loganathan and Poulos (1998) is given in Equation 2.7. It is noted that the method
proposed by Mair et al. (1993) yields a better prediction as compared to the method
proposed by Loganathan and Poulos (1998). In addition, the influence zone predicted
by Loganathan and Poulos (1998) is much greater than the measured data and
prediction by Mair et al. (1993). Hence, care should be exercised when employing
quasi-analytical methods to predict soil displacements due to tunnelling as certain
conditions in the derivation of analytical solutions may not be valid (e.g. volume loss
may not be conserved (Loganathan and Poulos, 1998)).

For the subsurface settlement troughs at various depths, the subsurface


settlement profiles generally follow the prediction by Mair et al. (1993). It should be
noted that the maximum subsurface settlements measured in the experiments,
especially when close to the tunnel, may not be accurate. This is mainly due to the
over-sizing of the tunnel end cap to prevent water seepage. The over-sized tunnel end
cap greatly influenced the tracking of soil displacements which were subsequently
analysed by PIV. Despite the above shortcoming, the back analysis generally validates
the use of Mair et al.s (1993) method to predict the subsurface settlements in the
short-term.

Figure 4.10 compares the measured short-term and long-term inflection point, i
at different depths with the empirical method proposed by OReilly and New (1982)

108

Chapter 4 Basic Test on Volume Loss

and Mair et al (1993). Both methods are based on Equation 2.3, but OReilly and New
(1982) assumed K=0.5 while Mair et al (1993) assumed that K varies with depth as
given by Equation 2.4.

For the present tests in clay, the measured distribution of inflection point (i)
with depth can be described by a straight line and the results are consistent with the
prediction of Mair et al (1993). The distribution of inflection point (i) with depth in
clay can be simplified as Equations 4.1 and 4.2 for the method proposed by OReilly
and New (1982) and Mair et al (1993), respectively.

Zo-Z = 2i

(4.1)

Zo-Z = 3i-8

(4.2)

In the present study, a finding on the long-term behaviour is established


revealing that the distribution of inflection point (i) with depth in clay can be
reasonably approximated by a straight line parallel to Equation 4.2 that yields
convergence points lower than those in the short-term. Hence the proposed distribution
of the inflection point (i) with depth in clay in the long-term can be presented as

Zo-Z = 3i-12

(4.3)

It is worth examining more closely the relationship between inflection point (i)
with depth in the short-term and long-term. Figure 4.11 shows the ratio of iLT / iST at

109

Chapter 4 Basic Test on Volume Loss

different depths. It can hence be deduced that iLT is approximately between 1.21 to
1.29 times iST.

4.3.4 Subsurface Horizontal Soil Movements


The short-term and long-term lateral soil movements at various distances from tunnel
centre-line are plotted in Figures 4.12 & 4.13. The proportion of horizontal to vertical
movements at the surface is considerably greater than that at greater depths, especially
when the distance from the tunnel centre-line increases. This observation is similar to
the finding obtained from the centrifuge model tests conducted by Grant and Taylor
(2000).

As expected, the horizontal soil movement caused by tunnelling diminishes


with increasing distance away from the tunnel. It is noted that the lateral soil
movements form a bulb shape at the tunnel spring line. However, the soil movements
diminish rather rapidly in the horizontal direction and become negligible at distance of
approximately 1.5D from the tunnel circumference, i.e 12 m from tunnel centre-line.

The results from the analytical solution proposed by Loganathan and Poulos
(1998) are also presented in the figures. However, the predictions by Loganathan and
Poulos (1998) do not agree well with the measured data. This may be attributed to the
condition that volume loss has not been conserved for undrained cases in their
formulation and other factors.

110

Chapter 4 Basic Test on Volume Loss

4.3.5 Qualitative Assessment on Excess Pore Pressure Responses

Pore water pressure changes in the ground are monitored using pore pressure
transducers (PPTs) during Test 1, as described in Chapter 3. To minimize the effect of
reinforcement that the PPTs have on the ground, only 2 PPTs were used, of which one
PPT is located within the immediate shear zone and the other one is located outside the
zone. Figure 4.14 shows the schematic location of the PPTs placed in the clay near the
tunnel lining and the trend of the pore water pressure changes obtained from the PPTs
throughout the test.

For the first 50 minutes of the test, the pore water pressure increases in 10 steps.
This is because the acceleration of the centrifuge from 0g to 100g is divided into 10
steps with an interval of 5 minutes per step. Subsequently, the pore water pressure
starts to drop and stabilize. This is because the excess pore water pressure induced by
the increased acceleration field dissipates. This process continues until the effective
stress in the ground is equivalent to the preconsolidation pressure. At this state, the soil
sample is normally consolidated. As PPT1 is at a higher elevation, the initial pore
pressure at PPT1 is lower than that at PPT2.

Tunnel excavation causes stress relief on the clay surrounding the tunnel lining
and thus a sharp drop in the pore water pressure is observed immediately after the
tunnelling process for PPT1 which is located inside the immediate shear zone.
Subsequently, the pore water pressure gradually increases over time due to dissipation
of pore water pressure. In contrast, an opposite trend is observed for PPT 2 located
outside the immediate shear zone. It is observed that additional excess pore pressure is

111

Chapter 4 Basic Test on Volume Loss

being induced in the clay, as indicated by a sharp increase in pore water pressure
immediately after tunnel excavation, caused by the shearing process of the affected soil
due to soil arching.

Further observation suggests that the pore water pressure stabilizes about after
two and a half hours (720 days in prototype scale) after the tunnel excavation. This
observation shows that the excess pore water pressure due to tunnelling has practically
fully dissipated and approaches the steady state pore pressure.

The above changes in the pore water pressure regime once again confirm that the
behaviour of clay can be time-dependent due to low permeability of the clay sample.
The soil will continue to deform with time as a result of dissipation of excess pore
pressures. This observation reiterates the importance of studying the long-term
behaviour of tunnelling-induced soil movement and pile responses for tunnels with
relatively large volume loss.

4.4 TYPICAL TUNNELLING-INDUCED PILE RESPONSES

The detailed results of Test 3 are reported here as an illustrative example of test results
from the beginning to the end of a typical test. For the sign convention used in the
present study, positive lateral soil movements refer to soil movement towards the
tunnel. Likewise, the deflection of pile towards tunnel is taken as positive. Bending
moment inducing pile shaft curvature towards the tunnel is considered as positive.
Lastly, downward vertical movement is regarded as positive.

112

Chapter 4 Basic Test on Volume Loss

In order to study the post-excavation ground deformation and pile responses,


the centrifuge would be kept at 100g for 3 hours after the completion of tunnel
excavation. All instruments were monitored regularly throughout the entire test. In
Test 3, both the axial and bending piles used are long floating piles (pile tip lower than
tunnel invert) with free heads and tips. The ultimate axial capacity of the 22-m long
model pile is estimated to be 2300 kN in soft clay having an almost linearly increasing
undrained shear strength profile, as shown in Figure 3.9. The ultimate bending moment
capacity of the pile is determined to be 3000 kNm.

4.4.1 Induced Axial Force and Settlement


The induced pile axial forces are directly measured from the readings of semiconductor strain gauges installed along the piles. The four semi-conductor strain
gauges are bonded on the external surface of the aluminium rod at an appropriate
elevation to form a full-bridge configuration to measure the induced axial force on the
pile. A strain meter is used to record the strain gauge signals along the model pile.

Figure 4.15(a) shows the induced pile axial force profile at 2 days, 180 days,
360 days and 720 days after tunnel excavation. It is noted that the induced pile axial
force increases with depth and reaches a maximum value approximately at the tunnel
spring elevation, after which the induced axial force gradually decreases till the pile tip.
The observed trend is consistent with the field data reported by Pang et al. (2005a) for
the MRT North East Line Contract 704 and 3D finite element analysis by Cheng
(2003). It is evident that the settling soil drags the pile down and induces negative skin
friction on the pile. This is consistent with the observed downward vertical soil

113

Chapter 4 Basic Test on Volume Loss

movement above the tunnel spring line due to soil over-cut in the process of tunnelling,
as observed in Figure 4.16. The neutral plane elevation becomes deeper over time. The
plot of maximum pile axial force with time shown in Figure 4.17 (a) reveals that the
drag load along the upper pile shaft increases with time and reaches a maximum
magnitude after about 720 days. This observation is consistent with that for Test 1
where the soil settlement does not increase further after 720 days (Fig. 4.7). The
readings reveal that there is a noticeable increase in maximum axial force in the longterm, from 198 kN after 2 days to 370 kN after 720 days. The total increment is about
90%.
Figure 4.15(b) shows field data reported by Cham (2007) for MRT Circle Line
Stage 3 Contract 852 in Singapore and the results are compared with those from Test 3
(Fig. 4.15(a). Although the general configuration of tunnel-pile and soil condition for
both cases is not identical, the general induced axial force profiles are consistent for
both cases. The soil settlement due to tunnelling would induce negative skin friction on
the pile shaft and maximum negative skin friction occurs at the tunnel axis. It can also
be observed that a smaller increase in down-drag force acts on the upper 20m of the
pile shaft. This is an indication of the effectiveness of the de-bonding system in
Contract 852. As non-zero value is observed near to the head of each pile, the
tunnelling-induced pile settlements could have resulted in some re-distribution of
structural loads on the piles after tunnel advancement. This is possible as the piles were
connected by transfer beams and slab. However, , the down-drag forces measured near
to the pile head are negligible in Test 3.

Figure 4.16 shows the observed free-field vertical soil movement profile over
time at the pile location obtained from Test 1 and the corresponding pile head

114

Chapter 4 Basic Test on Volume Loss

settlement shows the relationship between soil movement and pile settlement. From
the subsurface soil settlement observed through the marker beads movement analyzed
by PIV in Test 1 and the measured pile settlement in Test 3, it can be deduced that in
the short-term, the neutral plane is at a depth of about 14.2 m with subsurface soil
settlement very close to the pile settlement.

However, in the long term, the neutral

plane shifts to a lower depth of 16.1 m. The results also illustrate that the measured
pile head settlement for Test 3 increases substantially from 6 mm in the short-term to
17 mm in the long term. These data prove that the pile responses are time-dependent,
as shown in Figure 4.17(b). The significant increase in pile settlement is likely due to
the pile tip floating in the soft clay. Nevertheless, the pile undergoes much smaller
settlement than the soil. It is observed that the pile continues to settle after the
completion of tunnel excavation until long-term ground movement has been stabilized
at about 720 days after tunnel excavation.

4.4.2 Induced Bending Moment and Deflection


The induced pile bending moments are determined from the readings of strain gauges
attached along the pile shaft with the pile calibration conducted at 1g. The strain
gauges were connected to the strain meter with half-bridge mode and were properly
calibrated so that the relationship between microstrain, (registered by the strain
gauges) and the induced pile bending moment (result of hanging dead weights at the
tip of the cantilevered pile) could be established. Two appropriate faces of the
instrumented square pile were calibrated. As such, during centrifuge experiments, the
microstrain readings acquired from the data acquisition system can be readily related

115

Chapter 4 Basic Test on Volume Loss

to its corresponding bending moment from the appropriate scaling law of centrifuge
modelling.
Figure 4.18(a) shows the induced pile bending moment profiles at 2 days, 180
days, 360 days and 720 days after tunnelling. It is shown that the pile bending moment
increases with depth and the maximum induced bending moment toward the tunnel
occurs approximately at the tunnel central axis for long piles with tips well beneath the
tunnel. As expected, the bending moment at the pile head and tip are both zero as they
are not restrained.

The experimental results show that in the short-term, tunnel excavation induces
a maximum bending moment of 47 kNm. The induced bending moment increases
significantly by 98% to 93 kNm after 720 days. It is also noted that the shape of the
bending moment profile remains fairly constant over time and the maximum bending
moment remains practically unchanged beyond 720 days after tunnel excavation
Figure 4.17(c), revealing that the induced bending moment has stabilized. The findings
are consistent with the trend of soil movements induced by tunnelling as discussed in
Section 4.3.
The field measured bending moment profiles reported by Cham (2007) plotted
in Fig. 4.18(b) are compared to those obtained from Test 3. In addition, the measured
bending moment responses of a pile in pier 20 for C704 NEL (Pang, 2005) are also
included in Figure 4.18(b) for comparison. It is noted that the bending moment profiles
reported by Cham (2007) and Pang (2005) are similar to those observed in Test 3. In
general, the maximum transverse bending moment is noted to be at the tunnel axis
level.

116

Chapter 4 Basic Test on Volume Loss

The pile head deflection at the ground surface is obtained by geometry from the
two displacement readings obtained at 2 different pile elevations above the ground.
The free-field lateral soil displacements can be obtained from Test 1 whereas the pile
lateral displacement profile can be obtained by integrating the bending moment
profiles twice with two specified boundary conditions using the measured pile head
displacements at 2 elevations for the present study. The lateral pile displacement is
related to the free-field lateral soil movements, pile bending stiffness and pile-soil
interaction. A comparison of pile and free-field horizontal soil displacements for the
tests is shown in Figure 4.19. As expected, both the pile and soil move towards the
tunnel.

It is evident that the lateral soil movement profile has a roughly similar trend as
the pile deflection profile, showing that the pile basically deforms with the soil in a
similar fashion. As the pile can be considered a rigid body, the magnitude of pile
deflection is much smaller than that of the soil. The pile deflection profile is also
smoother than the soil movement profile due to the large pile bending rigidity. For
Test 3, the magnitude of pile deflection increases with time and the maximum lateral
pile deflection occurs at the pile head having a magnitude of 5 mm in the short-term
and 12.1 mm in the long-term.

117

Chapter 4 Basic Test on Volume Loss

4.5 TEST SERIES 1 - EFFECTS OF VOLUME LOSS (TESTS 3, 4)

In Test 4, the simulated tunnel opening for this model has a GAP of approximately 200
mm in prototype scale with an equivalent imposed volume loss of 6.5%, as compared
to that of 3% for Test 3. The pile-to-tunnel distance and pile length are kept constant in
these tests.

4.5.1 Induced Axial Force and Settlement


Figure 4.20 compares the induced pile axial force due to tunnel excavation in Tests 3
and 4. Generally, the maximum induced axial force on the pile increases with tunnel
volume loss and time. It is observed that the pile axial load transfer profiles along the
piles in Tests 3 and 4 are similar in trend, with a larger magnitude of negative skin
friction for the test with a larger volume loss. For Test 4 with a volume loss of 6.5%,
the maximum negative skin friction is 265 kN in the short-term (2 days after
excavation) and 422 kN in the long term (720 days after excavation). When compared
with Test 3, the increment in negative skin fraction is only 34% in the short-term and
14% in the long term despite an increase in volume loss of over 100% (i.e. from 3% to
6.5%).

Figure 4.21 shows the observed short- and long-terms free-field vertical soil
movement profile at the pile location obtained from PIV analysis from Tests 1 & 2,
and the measured pile head settlements from Tests 3 & 4. The results illustrate a
significant increase in pile head settlement when the volume loss increases from 3% to
6.5%. In the short-term, the pile head settlement increases significantly from 6 mm to

118

Chapter 4 Basic Test on Volume Loss

17 mm (increment of 183%) and drastically from 14.7 mm to 51.3 mm (increment of


250%) in the long-term when the volume loss increases from 3% to 6.5%.

The moderate increment of negative skin friction as compared to drastic


increment in settlement may be due to the fact that 3% volume loss has already
induced substantial relative pile-soil movement for the full mobilisation of negative
skin friction. As such, further ground settlement does not induce further negative skin
friction significantly when volume loss increases. For the induced base resistance in
the short- and long-terms, the pile base load obtained from Test 4 is higher than that of
Test 3. This is because the moderate increase in down drag has to be resisted by the
base resistance. In additional, when vertical soil movement increases with a larger
volume loss, additional pile settlement is necessary to mobilise sufficient positive shaft
resistance and base resistance to maintain pile equilibrium.

4.5.2 Induced Bending Moment and Deflection


Figure 4.22 shows the short- and long-terms induced pile bending moment profiles
obtained from Tests 3 and 4. The results demonstrate that the induced pile bending
moment profile has a double curvature with the moment magnitude increasing over
time. The bending moments at the pile head and tip are zero as they are not restrained.
The maximum induced bending moment occurs approximately at the tunnel springline.
The trend of the bending moment profile remains fairly constant over time and the
maximum bending moment remains practically unchanged after 720 days of tunnel
excavation. Owing to tunnel over-cut, the soil moves towards the tunnel spring
elevation, resulting in the pile bending towards the tunnel.

119

Chapter 4 Basic Test on Volume Loss

In the short-term, the maximum induced bending moment on the pile is 47


kNm, and increases to 93 kNm after 720 days for Test 3. The pile bending moment for
Test 4 is larger than that of Test 3 at all times, especially in the long term due to a
higher volume loss and hence larger soil movements. The maximum induced shortterm pile bending moment of 136 kNm from Test 4 is almost triple of that observed in
Test 3. The long term pile bending moment of 316 kNm is more than triple of that
observed in Test 3. It is evident that the maximum induced pile bending moments
increase significantly when the volume loss increases from 3% to 6.5%.

Figure 4.23 shows the observed short- and long-terms free-field lateral soil
displacement profiles at the pile location. The displacement profiles are obtained from
the PIV analysis of Tests 1 and 2, while the corresponding measured pile deflection
profiles are from Tests 3 and 4. The results reveal that the pile deflection increases
with volume loss, due to increase in lateral soil movement with larger volume loss and
time. As expected, the pile deflection is much smaller than that of the soil. The largest
lateral soil movement occurs at the pile head location, hence induces the largest pile
deflection at this elevation. The magnitude of lateral soil movement decreases with
depth and so did the pile deflection. The pile deflection profiles are similar for both
tests and the pile moves toward the tunnel, with the largest pile deflection observed at
the pile head. This deflection profile is due to increasing soil stiffness with depth.
Besides that, both the lateral soil movement and lateral pile deflection increase with
time. Comparing the pile deflection with horizontal soil movement, the pile deflection
has a smaller magnitude due to the large bending stiffness of the pile. Hence, the
measured pile head deflection in Test 3 is only 5 mm in the short-term and 12.1 mm in
the long-term, as compared to a much larger pile head deflection of 10 mm in the shor-

120

Chapter 4 Basic Test on Volume Loss

term and 28 mm in the long-term for Test 4. The increment of pile deflection is
significant when the volume loss increases from 3% to 6.5%.

Figures 4.24(a) to (d) show a summary of maximum pile axial force, pile head
settlement, pile bending moment and pile head deflection with volume loss for Tests 3
and 4. A consistent trend is observed that all pile responses increase with volume loss
and time. Nevertheless, it is observed that in the present floating pile condition, the
excessive pile movement (settlement and deflection) are the critical pile responses. On
the other hand, the significant increment of bending moment under a large volume loss
is detrimental to the structural integrity of the pile, especially if the pile foundation
supported the existing building only designed to resist the compression load as
illustrated in Figure 1.2. Thus, it is important to keep the tunnel volume loss as small
as possible in order that the long term induced pile responses are not a concern.
Figure 4.25 shows the long-term to short-term ratio of pile responses (pile axial
force, pile bending moment, pile head settlement, pile head deflection) for the two
volume losses. The results reveal that the pile responses increase over time and the
ratio of all long-term/short-term (LT/ST) pile responses except axial force also
increases with volume loss.

121

Chapter 4 Basic Test on Volume Loss

4.6 CONCLUDING REMARKS


4.6.1 Tunnelling-Induced Soil Movement
The centrifuge model tests with the application of PIV have provided useful data to
examine the patterns of soil movements induced by tunnelling in soft clay. The main
aim is to investigate the induced soil movement patterns over time.

The surface settlement trough in clay generally follows the Gaussian


distribution curve in the short-term. The magnitude of maximum ground surface
settlement increases with time and tunnel volume loss. The settlement magnitude is
larger in the long-term and the settlement trough is wider as compared to that in the
short-term. The data confirmed that the empirical equation proposed by Mair et al.
(1993) is applicable in the prediction of the subsurface settlement troughs in clay in the
short-term. Empirical equations in the short-term and long-term were proposed for the
distribution of inflection point in soft clay. On the other hand, an immediate shear zone
with large soil movement above the tunnel can be identified in the short-term. In the
long term, the significant soil movement zone extends much wider. In addition, soil
settlement is noted to be more dominant than lateral soil movement in the long term.
Qualitative assessment on the excess pore pressure responses has provided an
understanding on the development of negative excess pore pressure in the immediate
shear zone and positive excess pore pressure in the support zone.

122

Chapter 4 Basic Test on Volume Loss

4.6.2 Tunnel-Soil-Piles Interaction


Two centrifuge model tests with different volume loss have been performed to
investigate the effects of tunnelling on single free head piles. Owing to downward soil
movement, negative skin friction and bending moment are induced on the pile and the
magnitude of negative friction and moment increases with time and tunnel volume loss.
Both the neutral plane and maximum induced pile moment take place about the tunnel
centre-line elevation. The lateral soil movement profile has a similar trend as the pile
deflection profile, but the magnitude of the pile deflection is much smaller than that of
the soil. The pile deflection profile is smoother than the soil movement profile due to
the large pile rigidity.

Test Series 1 studies the effects of volume loss on pile performances. The test
results shed light on the actual performance of single floating pile due to tunnelling
with volume loss of 3% to 6.5%. It is found that the induced pile bending moment
triples and the pile settlement and deflection increase by almost 2.5 times when
volume loss increases from 3% to 6.5%, in this particular case.

123

Chapter 4 Basic Tests on Volume Loss

Table 4.1 Test program and parameters for the basic tests on volume loss

Phase 1 (Free-field soil movement) -Effects of volume loss


Test No.

Configuration

Common parameters

Individual
parameters

Volume loss =
1
3%
Kaolin
clay

C = 12 m
24m

D=6m
Volume loss =

Toyoura Sand

3.5m

6.5%

Phase 2 -Effects of tunnelling on single piles


Test series 1 Effects of volume loss
Test
No.

Configuration
Bending Pile

C = 12 m

Volume loss = 3%

D=6m

L
D

Individual
parameters

Axial Pile

3
Typical

Common
parameters

24m
X
2m

L = 22 m
X= 6 m

3.5m

Volume loss =
6.5%

124

Chapter 4 Basic Tests on Volume Loss

Water

30

Kaolin Clay

120

Control marker

60

Model Tunnel
60

30

Toyoura Sand
520

Figure 4.1 Schematic of viewing area in tunnel-soil interaction tests


(all dimensions in mm)

200

Texture clay

Control
marker

Y-coordinate (pixel)

400

600

Model tunnel
800

1000

1200

200

400

600

800
X-coordinate (pixel)

1000

1200

1400

1600

Figure 4.2 Example of digital images taken during test for PIV analysis

125

Chapter 4 Basic Tests on Volume Loss

Vectors of soil movements after 2 days


0

Depth below ground level (m)

-5

-10

-15

-20

50mm

-25

10

15

20

25

Distance from tunnel centre-line (m)

Contour plots for total soil movements after 2 days


0
30
40

20

10

-5

10

-10

30

Depth below ground level (m)

20
40

10
20
-15
10

-20

(unit mm)

-25
0

10
15
Distance from tunnel centre-line (m)

20

25

Figure 4.3 (a) Vectors and contour plots of soil movements after 2 days (Test 1)

126

Chapter 4 Basic Tests on Volume Loss

Vectors of soil movements after 180 days


0

Depth below ground level (m)

-5

-10

-15

-20

50mm

-25

10

15

20

25

Distance from tunnel centre-line (m)

Contour plots for total soil movements after 180 days


0
20

70

90
50

80

30

40

60

30

-5

10

40

70

-10

40

60

10

40

50

20

30

-15
10

20

10

Depth below ground level (m)

30

50

20

-20

(unit mm)

-25

10
15
Distance from tunnel centre-line (m)

20

25

Figure 4.3 (b) Vectors and contour plots of soil movements after 180 days (Test 1)

127

Chapter 4 Basic Tests on Volume Loss

Vectors of soil movements after 360 days


0

Depth below ground level (m)

-5

-10

-15

-20

50mm

-25

10

15

20

25

Distance from tunnel centre-line (m)

Contour plots for total soil movements after 360 days


0
90

60

50

80

0
10

30

40

-5

40

20

3
50 0
60

70

90

10

40

80

-10
70

20

50

60

30

10

40

20

10

30

-15

50

Depth below ground level (m)

30
50

20
10

10

-20

(unit mm)

-25

10
15
Distance from tunnel centre-line (m)

20

25

Figure 4.3 (c) Vectors and contour plots of soil movements after 360 days (Test 1)

128

Chapter 4 Basic Tests on Volume Loss

Vectors of soil movements after 720 days


0

Depth below ground level (m)

-5

-10

-15

-20

50mm

-25

10

15

20

25

Distance from tunnel centre-line (m)

Contour plots for total soil movements after 720 days


0
10
0

80

70

60

11
0

90

40

30

-5
20

50

80

30

70

10 0

40

50

60

60

40

30

20

-10

50

70

30

20

40

-15

30

40

60
50 0
6 70

Depth below ground level (m)

90

20
10

10

20

10

10

-20

(unit mm)

-25

10
15
Distance from tunnel centre-line (m)

20

25

Figure 4.3 (d) Vectors and contour plots of soil movements after 720 days (Test 1)

129

Chapter 4 Basic Tests on Volume Loss

Vectors of soil movements after 2 days


0

Depth below ground level (m)

-5

-10

-15

-20

100mm

-25

10

15

20

25

Distance from tunnel centre-line (m)

Contour plots for total soil movements after 2 days

40

70

50

90

80

10

20

60

30

10

-5

60

20

10 0

50
40

90
80
70

Depth below ground level (m)

11 0

-10

30

10

20

60
50

-15

40 30

10
20

10

-20

(unit mm)

-25

10
15
Distance from tunnel centre-line (m)

20

25

Figure 4.4 (a) Vectors and contour plots of soil movements after 2 days (Test 2)

130

Chapter 4 Basic Tests on Volume Loss

Vectors of soil movements after 180 days


0

Depth below ground level (m)

-5

-10

-15

-20

100mm

-25

10

15

20

25

Distance from tunnel centre-line (m)

Contour plots for total soil movements after 180 days

0
10

40
50

90

60

14
0

70

11
0

0
12

80

15 0

60

13
0

-5

11
0
10
0

12
0

80

90

70

Depth below ground level (m)

60

50

-10
0
15
140
13 0
120
110
100

80

70

50

40

60

90

40
50

80
70

-15

40

60

30

50
40
30
20

30

20

20
20

-20 10
10

10

10

(unit mm)

-25

10
15
Distance from tunnel centre-line (m)

20

25

Figure 4.4 (b) Vectors and contour plots of soil movements after 180 days (Test 2)

131

Chapter 4 Basic Tests on Volume Loss

Vectors of soil movements after 360 days


0

Depth below ground level (m)

-5

-10

-15

-20

100mm

-25

10

15

20

25

Distance from tunnel centre-line (m)

Contour plots for total soil movements after 360 days


0
16

0
18

0
19

-5

0
13

0
16
17
0

0
18

0
15

0
14

0
14

0
15

13 0

60
870
0
90

0
10 110

17 0

12
0

10 0

0
12

11 0

90

0
10

90

18 0

Depth below ground level (m)

80

-10

0
13
0
14

160

12 0 11 0
0
10
90

15 0

70
80

70

60

80
12 0

-15

11 0

70

10 0
90

60

80
70
60
50
40

50

60

50
50
40

40

30

30

30

-20
20

20

20

10

-25

(unit mm)

10
15
Distance from tunnel centre-line (m)

20

25

Figure 4.4 (c) Vectors and contour plots of soil movements after 360 days (Test 2)

132

Chapter 4 Basic Tests on Volume Loss

Vectors of soil movements after 720 days


0

Depth below ground level (m)

-5

-10

-15

-20

100mm

-25

10

15

20

25

Distance from tunnel centre-line (m)

Contour plots for total soil movements after 720 days


0
16

17
0

0
19

0
20

21 0

18
0

0
0
15

0
14

13 0

12 0

-5
0
18

0
19

200

0
17

0
16

0
15

14 0
13 0

11 0

12 0

0
14 13 0
180

17 0

16
0

Depth below ground level (m)

11 0

-10
0
15

90

10 0

12 0

11 0

80

90

10 0

80

14 0 0
13
12 0
11 0
10 0

-15

70

90

80

70

70
60

90
80
70
60
50

60

50

50
40

40

40

40

30

-20

30

30

30
20

20

-25

10
15
Distance from tunnel centre-line (m)

(unit mm)

20

25

Figure 4.4 (d) Vectors and contour plots of soil movements after 720 days (Test 2)

133

Chapter 4 Basic Tests on Volume Loss

Distance from tunnel centre-line (m)

Distance from tunnel centre-line (m)


0

10

12 14

16

18

20 22

24

26

inflection
point, 'i"

-40

-40

-60

-60

-80
-100
-120
-140
-160
-180
-200
-220
-240

10

12

14

16

18 20

22

24

26

-20

Settlement (mm)

Settlement (mm)

-20

Volume loss = 3%

2 days (from PIV)


90 days (from PIV)
180 days (from PIV)
360 days (from PIV)
720 days (from PIV)
2 days (from Potentiometer)
90 days (from Potentiometer)
180 days (from Potentiometer)
360 days (from Potentiometer)
720 days (from Potentiometer)
2 days (Gaussian curve)

Figure 4.5 Surface settlement troughs over time (Test 1)

inflection
point, 'i"

-80
-100
-120
-140
-160
-180
-200
-220
-240

Volume loss = 6.5%

2 days (from PIV)


90 days (from PIV)
180 days (from PIV)
360 days (from PIV)
720 days (from PIV)
2 days (from Potentiometer)
90 days (from Potentiometer)
180 days (from Potentiometer)
360 days (from Potentiometer)
720 days (from Potentiometer)
2 days (Gaussian curve)

Figure 4.6 Surface settlement troughs over time (Test 2)

134

Chapter 4 Basic Tests on Volume Loss

Time (days)
0

180

360

540

720

900

1080 1260 1440

0
-20
-40

settlement (mm)

-60
-80
-100
-120
-140
-160
-180
-200
-220
-240
-260
-280
-300

Test 1 (from PIV)


Test 1 (from Potentiometer)
Test 2 (from PIV)
Test 2 (from Potentiometer)

Figure 4.7 Maximum surface settlements over time (Tests 1 & 2)

135

Chapter 4 Basic Tests on Volume Loss

Distance from tunnel centre-line (m)

10

12

14

16

18

Distance from tunnel centre-line (m)

20

22

24

26

10

12

14

16

-40

Surface settlement, Test 1 (PIV)


Surface settlement, Mair et al (1993)

-60

4.3m, Test 1 (PIV)

-20
Settlement (mm)

Short-term
(VL=3%)

-20
Settlement (mm)

18

20

22

24

26

Short-term
(VL=3%)

-40
Surface settlement, Test 1 (PIV)

-60

Surface settlement, Loganathan & Poulos (1998)


4.3m, Test 1 (PIV)

4.3m, Mair et al (1993)

-80

9.3m, Test 1(PIV)

-80

4.3m, Loganathan & Poulos (1998)


9.3m, Test 1(PIV)

9.3m, Mail et al (1993)

-100

-100

(a)

9.3m, Loganathan & Poulos (1998)

(b)

Figure 4.8 Settlement troughs at surface, 4.3m and 9.3m depths (Test 1): (a) comparing with Mair et. al (1993) (b) comparing with Loganathan
and Poulos (1998)

136

Chapter 4 Basic Tests on Volume Loss

Distance from tunnel centre-line (m)

Distance from tunnel centre-line (m)

10

12

14

16

18

20

22

24

26

-20

-20

-40
-80
-100
-120

Surface settlement, Test 2 (PIV)

-140

Surface settlement, Mair et al (1993)

-160
-180
-200
-220
-240

10

12

14

16

18

20

22

24

26

-40

Short-term
(VL=6.5%)

-60
Settlement (mm)

Settlement (mm)

-60

-100
-120
-140

5m, Test 2 (PIV)

-160

5m, Mair et al (1993)

-180

10.9m, Test 2(PIV)


10.9m, Mair et al (1993)

Short-term
(VL=6.5%)

-80

-200
-220
-240

Surface settlement, Test 2 (PIV)


Surface settlement, Loganathan & Poulos (1998)
5m, Test 2 (PIV)
5m, Loganathn & Poulos (1998)
10.9m, Test 2(PIV)
10.9m, Loganathan & Poulos (1998)

Figure 4.9 Settlement troughs at surface, 5m and 10.9m depths (Test 2): (a) comparing with Mair et. al (1993) (b) comparing with Loganathan
and Poulos (1998)

137

Chapter 4 Basic Tests on Volume Loss

20
16
12

Proposed long-term
equiation, Zo-Z = 3i-12

Zo-Z (m)

4
0
-4

-8

10

11 12

(m)

-12
O'Reilly and New (1982)
Mair et al.(1993)
Test 1, 2 days (ST)
Test 2, 2 days (ST)
Test 1, 720 days (LT)
Test 2, 720 days (LT)
Proposed long-term equation

-16
-20

Figure 4.10 Distribution of inflection point i with depth in short- and long-term
(Tests 1 & 2)
1.6
1.5
1.4

Ratio of iLT / iST

1.3
1.2
1.1
1
0.9
0.8
0.7
0.6
0

10

11

12

Depth, Z(m)
Test 1 (PIV)

Test 2 (PIV)

Figure 4.11 Comparison of ratio of iLT/iST at different depths (Tests 1 & 2)

138

Chapter 4 Basic Tests on Volume Loss

-40

-30

-20

-10

-40

-30

-20

-10

12m from tunnel


Soil movement (mm)
0

-40

-30

-20

-10

15m from tunnel


Soil movement (mm)
0

-40

-30

-20

-10

-5

-5

-5

-5

-5

-10

-10

-10

-10

-10

-15

Tunnel

-15

-20

-20

-25

-25

-30

-30

-15

-15

Depth below GL (m)

Depth below GL (m)

Depth below GL (m)

Depth below GL (m)

-40 -30 -20 -10

9m from tunnel
Soil movement (mm)

-15

-20

-20

-25

-25

-30

-30

-20

Depth below GL (m)

6m from tunnel
Soil movement (mm)

4m from tunnel
Soil movement (mm)

-25

2 days
Loganathan
et al 1998

-30

720 days

Figure 4.12 Horizontal soil movements at different distance from tunnel center-line at 2 and 720 days - Test 1
139

Chapter 4 Basic Tests on Volume Loss

Soil movement (mm)


-60

-40

-20

-80

-80

-60

-40

-20

-80

-60

-40

-20

-5

-5

-5

-5

-10

-10

-10

-10

-5

-10

-20

-20

Tunnel

-40

-15

-60

15m from tunnel


Soil movement (mm)

-15

-20

-25

-25

-30

-30

-15

-20

-15

Depth below GL (m)

-80

12m from tunnel


Soil movement (mm)

Depth below GL (m)

Depth below GL (m)

-40 -20

Depth below GL (m)

Soil movement (mm)


-80 -60

9m from tunnel
Soil movement (mm)

-15

-20

-20

-25

-25

-30

-30

-25

Depth below GL (m)

6m from tunnel

4m from tunnel

2 days
Loganathan
et al 1998

-30

720 days

Figure 4.13 Horizontal soil movements at different distance from tunnel center-line at 2 and 720 days - Test 2
140

Chapter 4 Basic Tests on Volume Loss

Excess pore pressure (kPa)

1)Spinning up
from 1g to 100g

240
220
200
180
160
140
120
100
80
60
40
20
0
-20
-40
-60

2) Consolidation at 100g

3)Tunnelling
4)Posttunnelling

5)Spinning
down

E
Point 2
D
(support zone)

Point 1
(shear zone)

500

1000

1500

2000

2500

F
C
B

3000

3500

4000

4500

5000

5500

6000

6500

Time (day)
Point1

Point2
PPT Point 1
PPT Point 2

Figure 4.14 Pore pressure changes due to tunnelling (Test 1)

141

Chapter 4 Basic Tests on Volume Loss

Axial Force (kN)


0

100

200

300

400

500

600

-5

Depth (m)

-10

-15

T unnel

Approximate
Neutral Plane

-20

-25
2 days

180 days

360 days

720 days

-30

Figure 4.15(a) Tunnelling-induced pile axial force (Test 3, 3% free-head floating long
pile)

Volume loss = 0.7%


Tunnel ID=5.8m
Pile dia. =600mm
Pile Length =31m

Figure 4.15(b) Tunnelling-induced pile axial force (Pile BP1-G) for MRT Circle Line
Stage 3 (CCL3) Contract 852 in Singapore (After Cham, 2007)

142

Chapter 4 Basic Tests on Volume Loss

Settlement (mm)
0

20

40

60

80

100

120

-5

Depth (m)

-10

-15

T unnel
Approximate
Neutral Plane

-20
Pile head settlement (Test 3)
-25

Free-field soil settlement (Test 1, PIV)


2 days

180 days

360 days

720 days

-30

Figure 4.16 Tunnelling-induced pile head settlement (Test 3) and observed free-field
soil movement at pile location (Test 1, PIV)

143

Max. pile axial force (kN)

Chapter 4 Basic Tests on Volume Loss

400

(a)

350
300
250
200
150
0

180

100

540
720
T ime (days)

900

1080

1260

1440

(b)

80

Settlement (mm)

360

60

Free-field soil surface settlement (Test 1)

40

Pile head settlement

20
0
0

180

360

540
720
T ime (days)

900

1080

1260

1440

360

540
720
T ime (days)

900

1080

1260

1440

150

Max. pile bending


moment (kN)

(c)
100

50

Lateral movement (mm)

30

180

Free-field lateral soil movement (Test 1)

(d)

25

Pile head deflection

20
15
10
5
0
0

180

360

540
720
T ime (days)

900

1080

1260

1440

Figure 4.17 Tunnelling-induced (a) maximum pile axial force (b) maximum pile head
settlement and soil surface settlement (Test 1) (c) maximum pile bending moment (d)
maximum pile head deflection and soil surface lateral movement (Test 1)

144

Chapter 4 Basic Tests on Volume Loss

Bending Moment (kNm)


-100

-50

50

100

150

200

250

300

-5

Depth (m)

-10

-15

T unnel

-20

-25

2 days

180 days

360 days

720 days

-30

Figure 4.18 (a) Tunnelling-induced pile bending moment (Test 3, 3% free-head


floating long pile)

Volume loss = 0.7%


Tunnel ID=5.8m
Pile dia. =800mm
Pile Length =33.4m

Figure 4.18 (b) Tunnelling-induced pile bending moment (Pile BP2-E) for MRT Circle
Line Stage 3 (CCL3) Contract 852 in Singapore (After Cham, 2007)

145

Chapter 4 Basic Tests on Volume Loss

Lateral deflection (mm)


0

10

20

30

40

-5

Depth (m)

-10

-15

T unnel

-20

Free-field lateral soil movement (Test 1, PIV)


2 days
180 days

-25

360 days
Pile deflection (Test 3)
2 days
360 days

720 days
180 days
720 days

-30

Figure 4.19 Tunnelling-induced pile deflection (Test 3) and free-field lateral soil
movement at pile location (Test 1)
Axial Force (kN)
0

100

200

300

400

500

600

-5

Depth (m)

-10

-15

T unnel

-20

-25
ST (Test 3)

LT (Test 3)

ST (Test 4)

LT (Test 4)

-30

Figure 4.20 Variation of pile axial force with volume loss (Tests 3 and 4)

146

Chapter 4 Basic Tests on Volume Loss

Settlement (mm)
0

50

100

150

200

250

-5

Depth (m)

-10

-15

T unnel

-20

Test 3 Pile head settlement


Test 4 Pile head settlement

-25

Free-field soil settlement


ST (Test 1)

LT (Test 1)

ST (Test 2)

LT (Test 2)

-30

Figure 4.21 Variation of pile head settlement (Tests 3 and 4) and observed free-field
soil movement at pile location (Tests 1 and 2) with volume loss

147

Chapter 4 Basic Tests on Volume Loss

Bending Moment (kNm)


-150

-50

50

150

250

350

450

550

650

-5

Depth (m)

-10

-15

T unnel

-20

-25

ST (Test 3)

LT (Test 3)

ST (Test 4)

LT (Test 4)

-30

Figure 4.22 Variation of pile bending moment with volume loss (Tests 3 and 4)
Lateral deflection (mm)
0

10

15

20

25

30

35

40

45

-5

Depth (m)

-10

-15

T unnel

Free-field lateral soil movement


-20

-25

ST (Test 1)

LT (Test 1)

ST (Test 2)
Pile deflection

LT (Test 2)

ST (Test 3)

LT (Test 3)

ST (Test 4)

LT (Test 4)

-30

Figure 4.23 Variation of pile deflection profiles (Tests 3 and 4) and observed free-field
lateral soil movement at pile location (Tests 1 and 2) with volume loss

148

Chapter 4 Basic Tests on Volume Loss

500

60

400

LT

Pile head settlement (mm)

Maximum pile axial force (kN)

ST

300

200

100

55

ST

50

LT

45
40
35
30
25
20
15
10
5
0

0
Test 3 (VL=3%)

Test 3 (VL=3%)

Test 4 (VL=6.5%)

Test 4 (VL=6.5%)

(a)

(b)

36

350

ST

32

ST
LT

LT

300

Pile head deflection (mm)

Maximum pile bending moment (kNm)

400

250
200
150
100

28
24
20
16
12
8
4

50

0
Test 3 (VL=3%)

(c)

Test 4 (VL=6.5%)

Test 3 (VL=3%)

Test 4 (VL=6.5%)

(d)

Figure 4.24 Variation of (a) maximum pile axial force (b) pile head settlement (c) pile
bending moment (d) pile head deflection with volume loss (Tests 3 and 4)

149

Chapter 4 Basic Tests on Volume Loss

4
3.5
3

LT/ST ratio

2.5

Long-term
effect

2
1.5
1
0.5
0
Pile axial force

Pile head
settlement

Test 3 (VL=3%)

Pile bending
moment

Pile head
deflection

Test 4 (VL=6.5%)

Figure 4.25 Long-term to short-term ratio of pile responses for different volume losses
(Tests 3 and 4)

150

Chapter 5 Effects of Tunnelling on Single Piles

CHAPTER FIVE

EFFECTS OF TUNNELLING
ON SINGLE PILES
5.1

INTRODUCTION

In this chapter, the results of further centrifuge model tests conducted to investigate the
effects of tunnelling on single piles in clay are presented. The detailed test program
and configurations in prototype scale is given in Table 5.1 and the pile tip positions
investigated in the parametric studies are schematically illustrated in Figure 5.1.

5.2

TEST SERIES 2- EFFECTS OF PILE TIP & HEAD


CONDITIONS

5.2.1 Effects of Pile Tip Condition (Tests 3, 9 And 10)

Tests 3, 9 and 10 were performed to simulate three different pile tip conditions, namely
a floating pile in Test 3 (presented in Chapter 4), a socketed pile in Test 9 and an
end-bearing pile in Test 10. For the test on floating pile (Test 3), the 22-m long
pile is entirely embedded in the 24-m thick soft clay layer. The soft clay was underlain
by a layer of 3.5-m thick sand layer. For the test on socketed pile (Test 9), the soft
clay was underlain by a 8-m thick sand layer and the pile was embedded 3m into the
underlying sand layer. For Test 10, the underlying sand thickness of 3-m is the same as
that for Test 3 but the 27.5 m long pile is resting on the rigid base of the container.
151

Chapter 5 Effects of Tunnelling on Single Piles

Strictly speaking this pile is not a pure end-bearing pile, as there will be some load
transfer in the 3 m thick sand layer just above the pile tip. All tests have the same
tunnel volume loss of 3% and the same pile-to-tunnel distance, i.e. 1D or 6 m. The pile
heads are free.

Figures 5.2(a) & (b) show the induced pile axial force profile of Tests 3, 9
and 10 at the end of tunnel excavation (short- term at 2 days) and in the long-term (720
days), respectively. It can be seen that the induced axial force in all tests increases
downwards from the pile head. However, the respective maximum axial force occurs
approximately at the tunnel spring elevation for Test 3 but slightly lower than the
tunnel spring elevation for Tests 9 and 10. This is as expected because the pile in Tests
9 and 10 are socketed into the sand layer and hence the pile settlement is much smaller
than that for a floating pile (Test 3). As such, the larger soil movement relative to pile
settlement induces much larger drag loads on the socketed pile (Test 9). The negative
skin friction is highest for the end-bearing pile (Test 10) as the pile is resting on a rigid
base with negligible measured pile settlement. The elevation of neutral plane in the
socketed pile (Test 9) is lower than that of the floating pile (Test 3). The neutral plane
for end-bearing pile in Test 10 shifts even much lower as compared to Test 9. The
axial force is mostly transferred to the pile socket in Tests 9 and 10, as evidenced from
the steep gradient of positive skin friction shown in Figure 5.2. Intuitively, the neutral
plane for a pure end-bearing pile should be at the pile tip. This is not so for Test 10 as
positive skin friction is mobilised to transfer the load in the sand layer as shown in
Figure 5.2.

152

Chapter 5 Effects of Tunnelling on Single Piles

Owing to lower part of pile socketed in stiff soil, the pile settlement for Test
9 (socketed pile) is only 2 mm in the short- term and 3 mm in the long term. As
expected, the pile settlement for Test 10 (end-bearing pile) is negligible. Owing to the
floating condition of the pile tip, the settlement of the floating pile (Test 3) increases
by 183% from 6 mm in the short- term to 17 mm in the long-term. However, these
magnitudes are still much smaller than the observed ground surface settlement. The
above observations reveal that pile settlement is more critical for a floating pile while
induced negative skin friction is more critical for both socketed and end-bearing piles.

Figures 5.3(a) & (b) show the variation of induced pile bending moment
profiles of Tests 3, 9 and 10 in the ST and LT, respectively. The bending moment
profiles in Tests 3 and 9 share a similar trend as the maximum bending moment occurs
close to the tunnel axis with double curvature. Although the volume loss for both tests
is the same, the socketed pile in Test 9 exhibits a larger maximum bending moment, as
compared to a floating pile (Test 3). This is due to socketing of the pile into sand, thus
sufficiently restricting movement of the lower portion of the pile. The restraint at the
pile tip would restrict pile movement and hence result in slightly larger pile bending
moments. As a result, the induced bending moment on the socketed and end-bearing
pile is more critical than that on the floating pile. On the other hand, the end-bearing
pile (Test 10) exhibits triple curvatures profile with the maximum bending moment
occurring close to the tunnel axis as well. The maximum bending moment is the largest
among the three tests. This is probably because the longer span of pile in Test 10 is
exposed to more soil movements induced by tunnelling and hence the pile length effect
can be one of the factors contributing to the magnitude of induced pile bending
moment. This aspect will be further examined in Section 5.3.

153

Chapter 5 Effects of Tunnelling on Single Piles

Figures 5.4(a) & (b) show the pile deflection profiles for Tests 3, 9 and 10.
The pile head deflection for the socketed pile (Test 9) is only 2.1 mm and 3 mm in the
short- and long-term, respectively. This is much smaller than the corresponding pile
head deflection of the floating pile of 5 mm (ST) and 12.1 mm (LT). From the pile
deflection profile exhibited in the socketed pile, the pile head is noted to bend towards
the tunnel, but there is almost no deflection from the mid-pile shaft to the pile tip. This
is because the lower part of the pile is restrained and hardly moves. In contrast, much
larger deflection is noted for the floating pile. Nevertheless, the pile head deflection
for the end-bearing pile (Test 10) is slightly larger than that of socketed pile, with 2.8
mm and 6 mm in the short- and long-terms, respectively. This is likely due to the
higher elevation of the underlaying sand layer for the socketed pile. As such, the endbearing pile in Test 10 is exposed to more lateral soil movements and its lower
elevation of pile toe fixity point causes it to deflect more than the socketed pile in Test
9. Moreover, it is noted that the mid-pile shaft in end-bearing pile (Test 10) being
pushed away from the tunnel. This is probably due to the relatively large lateral soil
movement in the immediate shear zone pushing the pile head while the pile tip is being
restrained, and hence causing the mid-pile shaft being bent away from the tunnel. In
short, the induced pile deflection is more critical for a floating pile as compared to
socketed and end-bearing piles.

5.2.2 Effects of Pile Head Condition (Tests 10, 13)

Tests 10 and 13 were performed to study the effects of pile head condition. Test 13 is
modelled such that the pile head is totally fixed in position with no vertical or lateral
movements allowed. This simulates the condition where the pile cap is tied rigidly

154

Chapter 5 Effects of Tunnelling on Single Piles

with the ground beams. In both tests, the 27.5-m long model pile is fully embedded
into the 24-m thick soft clay layer and 3.5 m thick sand layer, and resting on the rigid
base of the model container, simulating an end-bearing pile.

The induced pile axial force profiles of Tests 10 and 13 are shown in Figure 5.2.
It is observed that the axial load transfer profiles are similar for both the free-head pile
and fixed-head pile, with the exception of the development of tensile force in fixedhead piles. In a completely fixed head condition, the pile is not allowed to settle,
resulting in tensile force induced along the upper portion of pile. The maximum
negative skin friction is observed at an elevation lower than the tunnel spring line or at
approximately 17.5 m (Short-term) and 20 m (Long-term). The general trend of
development of tensile force is observed by Mroueh and Shahrour (2002) and Pang
(2006) as well. Pang (2006) reported that the pile is significantly affected when the pile
cap is restrained. Tensile force is observed along the top 15 m of the pile while the
maximum drag load is reduced compared to the pile without pile cap restraint. For a
fixed-head pile, engineers may need to evaluate the connection between the pile and
the pile cap in resisting the tensile force.

In addition, it is noted that the reduction of maximum drag load in fixed-head


pile is approximately 36% and 33% in the short- and long-term, respectively. However,
the trade-off is that tension force would be induced near to the pile head. It might be
due to the load transfer curve essentially offset towards the tension side due to the total
restraint at pile head (fixed-head condition) applied on the pile. The measured
settlement of end-bearing pile is less than 2mm as the pile tip is rested on a rigid base.

155

Chapter 5 Effects of Tunnelling on Single Piles

Figure 5.3 shows the tunnelling-induced pile bending moment for the free- and
fixed-head end-bearing piles (Tests 10 & 13). The pile bending moment profile is
similar for both cases, where triple curvature is induced with negative bending
moments at the upper and lower portions of the pile body, whilst positive pile bending
moment occurs approximately at the tunnel spring line. As expected, no bending
moment is induced in the free pile head which is allowed to move, while relatively
large negative bending moment is induced at the fixed pile head with zero rotation and
displacement. It is worth noting that the observed negative bending moment at the pile
head is larger than the positive bending moment at the mid-pile shaft. For the fixedhead pile, the bending moment profile is offset towards the negative side as compared
to the free-head pile. The large magnitude of bending moments at the pile head needs
to be evaluated in practice.

As no pile head deflection is allowed for the fixed-head pile (Test 13), the
measured mid-pile shaft deflection of less than 0.2mm is much smaller than that for the
free-head pile, see Figure 5.4. This might be attributed to the fact that the pile head and
toe are fixed in placed and thus the movement of the pile at mid-pile shaft could be
purely due to the bending of the pile shaft due to the large rigidity of the pile. Since the
pile deflection is negligible, the effects of deflection induced by tunnelling toward the
existing pile in fixed-head pile are not a major concern as compared to the free-head
pile.

Figure 5.5 shows a summary of the variation of short-term and long-term pile
responses with tip and head conditions. As compared to floating pile, socketed and
end-bearing piles experienced smaller induced pile settlement and deflection; but

156

Chapter 5 Effects of Tunnelling on Single Piles

larger induced pile axial force and bending moment. Thus, in conclusion, a floating
pile (Test 3) will be mainly governed by pile settlement when tunnelling is carried out
nearby. For end-bearing piles and piles that are socketed into stiffer material (e.g.
dense sand), the piles will experience significant negative skin friction. In practice, it is
common that a pile is rested or socketed into hard strata and engineers should assess
whether the pile can resist the induced negative skin friction due to tunnelling.

The variation of pile responses for free- and fixed-head pile is now examined.
Since the comparison is made between end-bearing piles, pile movements (settlement
and deflection) are not a major concern, as any fixity in toe would substantially reduce
the pile movement as discussed before. However, the trade-off is the increase in pile
material stress (bending moment and axial force) due to fixity. Particularly, the
restriction of pile head which will cause negative bending moment and tension at the
pile head (Figure 5.5(c)), which may be detrimental to the pile (Figure 5.5(a)).

5.3 TEST SERIES 3 -EFFECTS OF PILE LENGTH (TESTS 3, 7, 8)


Test 7 was conducted using a short pile with embedment length of 11.4 m (ratio of pile
length over tunnel depth, L/H=0.76). The pile tip is located within the immediate shear
zone. In addition, Test 8 was conducted using a pile with embedment length of 15.6 m
(ratio of pile length over tunnel depth, L/H=1.04). The pile tip is located approximately
at the tunnel axis. The results of these 2 tests will be compared with the performance of
the long pile in Test 3, which has an embedment length of 22 m (pile tip below tunnel
invert) to study the effects of pile length. In this section, a short pile is referred to as a
pile with its tip elevation at or above the tunnel axis elevation. As such, Test 3 involves

157

Chapter 5 Effects of Tunnelling on Single Piles

a long pile case and Tests 7 and 8 involve short pile cases. The pile-to-tunnel distance,
tunnel volume loss and other parameters are kept constant in the tests.

Figure 5.6 shows the variation of induced pile axial force with pile length for
Tests 3, 7 and 8. The longest pile (22 m) in Test 3 experiences the largest negative skin
friction as compared with shorter piles in Tests 7 and 8. This is because the pile in test
3 has the greatest pile shaft area, thus allowing the development of a larger negative
skin friction. In addition, the pile tip is located far below the large soil displacement
immediate shear zone. The portion of the pile beneath the shear zone develops positive
skin friction and end bearing to resist the down drag force.

Figure 5.7 shows the variation of induced pile head settlement with pile
length for Tests 3, 7 and 8 and the free-field vertical soil movement at the respective
pile locations. The results demonstrate that the tunnelling-induced settlement on a short
pile is more critical than on a long pile for both short-term and long-term. Moreover,
the incremental short pile settlement in the long-term over short-term is also much
larger than that of the long-pile. The pile settlement in Test 7 is significantly larger
than the piles in Tests 3 and 8 whose tips are beneath the immediate shear zone. These
findings are consistent with the observed large settlement zone for tunnelling in sand
reported by Jacobsz (2002). However, as the trough width in clay is generally larger
than that in sand (Rankin, 1998; Ran, 2004), the large settlement zone in clay is noted
to be wider than that in sand.

Figure 5.8 shows the variation of induced bending moment with pile length.
The pile bending moment profile changes from double to single curvature as the pile
tip moves from L/H=1.5 to L/H=1.04 or at the tunnel axis level. It is observed that the

158

Chapter 5 Effects of Tunnelling on Single Piles

maximum bending moment increases with pile length in both short- and long-terms.
The maximum bending moment of the piles (L/H=1.04 and 0.76) occurs above the
elevation of the tunnel spring line and approximately at the middle of the piles, instead
of at the tunnel spring line in the case of long pile (L/H=1.5). In addition, the induced
pile bending moment is greater for a longer pile (pile base located outside the
immediate shear zone) than that of a shorter piles. This is due to restraint of the pile
beneath the immediate shear zone with relatively small or negligible soil movement.
The lateral soil movement profiles shown in Figure 5.9 provide clear evidence on the
changes of induced pile bending moment. As a result, a longer pile would induce a
larger bending moment as compared with a shorter pile.

Figure 5.9 shows the variations of pile deflection for Tests 3, 7 and 8 and the
corresponding free-field lateral soil movement profile obtained from PIV (Test 1). The
pile generally deflects in a similar fashion as the free-field soil displacement profile but
with much smoother and smaller movements. The restraint in movement can be
attributed to the large bending stiffness (EI) of the pile body. In contrast to the induced
pile bending moment, the short pile head deflection is significantly larger than that of
the long pile. In the case of long pile, its deflection is found to be the smallest among
the three cases. When the pile length is reduced, the magnitude of pile deflection also
increases. This is due to the fact that the long pile is partially embedded in the support
zone with smaller induced soil movements while the short pile is embedded entirely in
the immediate shear zone which induces relatively larger soil movements. In
consequence, the long pile bends towards the tunnel while for the short pile, it
shifts toward the tunnel with a maximum pile deflection at the pile head for both
cases.

159

Chapter 5 Effects of Tunnelling on Single Piles

Figure 5.10(a) shows the variation of maximum pile axial force with
normalised pile length over tunnel depth (L/H=0.76, 1.04, 1.5). The results clearly
show that the induced maximum pile axial force increases with pile length over tunnel
depth. It is noted that the maximum axial force increases linearly with normalised pile
length over tunnel depth, from 25 kN (L/H=0.76) to 92 kN (L/H= 1.04) and finally to
198 kN when the pile tip is located below the tunnel invert (L/H=1.5). However, in the
long-term, the linear increment of axial force with normalised pile length over tunnel
depth has a much steeper gradient than when compared with the short-term. The axial
force increases from 45 kN to 92 kN and then 198 kN when the normalised pile length
over tunnel depth (L/H) increases from 0.76, 1.04 to 1.5, respectively. This is because
the piles whose tips are located in the zone of immediate shear zone would be
intensively dragged down by the large soil movements, resulting in large pile
settlement. On the other hand, piles with tips located out of the shear zone would
experience less settlement as the lower part of the pile shift is socketed in stiff soil.

Figure 5.10(b) shows the variation of pile head settlement with normalised
pile length over tunnel depth (L/H=0.76, 1.04, 1.5). The data clearly show that when
the pile tip is located within the zone of large displacements, the pile would settle
excessively with the tip in the short-term with a similar magnitude to the soil
displacement at the location, see Figure 5.7.

Figure 5.10(c) shows the maximum induced pile bending moment with
normalized pile length over tunnel depth in the short and long-terms. It is observed that
the maximum bending moment of the long pile is larger than that of shorter piles due

160

Chapter 5 Effects of Tunnelling on Single Piles

to the restraining effect discussed earlier in this section. Similar findings are also
reported by Chen et al. (1999) and Pang (2006).

Figure 5.10(d) shows the maximum pile head deflection with normalized pile
length over tunnel depth in short and long-terms. The maximum induced pile head
deflection of a short pile also exhibits marked increase compared with a long pile. In
the short- term, the pile head deflection of the long pile (Test 3, L/H=1.5) is 5 mm and
increases to 8 mm and 11 mm when the pile length over tunnel depth reduces to 1.04
and 0.76, respectively. This can be attributed to the lateral soil movement profile (Fig.
5.9) induced by tunnelling, where a long pile is restrained if the pile length extends
beneath the immediate shear zone. Hence, the pile deflection is smaller when the pile
length is longer. Similar trend is observed in the long term as well, in which the pile
head deflection increases from 12.1 mm to 18 mm and 22 mm when the pile length
over tunnel depth reduces from 1.5 to 1.04 and 0.76 respectively.

To further assess the effect of pile length over tunnel depth due to tunnelling,
the key pile responses are summarized in Figure 5.11. In a short pile, especially those
located in the immediate shear zone, the pile structure is less vulnerable as compared
to the long pile. However, there will be excessive pile movements (settlement and
deflection) because of lack of anchorage of pile into the stable support zone. In this
respect, a longer pile with extension of pile length into stabilised support zone tends to
provide more resistance to the pile movements but will experiences larger structural
stress in term of bending moment and axial force.

161

Chapter 5 Effects of Tunnelling on Single Piles

The figures illustrate that a short pile is less vulnerable in terms of tunnellinginduced pile axial force and bending moment, while significant adverse effect is
observed in terms of pile head settlement and deflection for short pile. This
phenomenon is verified in the soil settlements and lateral soil movements shown in
Figures 5.7 and 5.9. It is demonstrated that relatively large soil movements are induced
within the immediate shear zone while relatively small or negligible soil movement
induced in the support zone, thus when the pile base is extended into the support zone,
smaller pile settlement and deflection are expected due to restraint. Paradoxically,
larger axial force and bending moment are induced for longer pile caused by the same
restraint. This finding provides evidence that tunnelling-induced displacements of
short piles (settlement and deflection) are more critical than that of long piles. In
contrast, the pile axial force and bending moment are crucial in the long pile upon
tunnel excavation. This implies that the effect of pile length over tunnel depth has
pronounced implications in tunnelling-induced pile responses.

5.4

EFFECTS OF DISTANCE OF PILE FROM TUNNEL

5.4.1 Test Series 4 - Free-Head Floating Piles (Tests 3, 5, 6 and 16)

In Test series 4, centrifuge model tests (Tests 3, 5, 6 and 16) were performed to study
the behaviours of long free-head floating piles with various pile-to-tunnels centre
distances. The pile-to-tunnel distance in Tests 3, 5, 16 and 6 is 6 m (or 1D), 9 m (or
1.5D), 10m (or 1.67D) and 12 (or 2D), respectively. Other parameters are kept
constant in all tests. For clarity, the results of Test 16 are not presented in Figures 5.12

162

Chapter 5 Effects of Tunnelling on Single Piles

to 5.15, but the maximum pile responses of this test are presented in Figure 5.21 for
comparison purpose.

Figure 5.12 shows the induced pile axial force due to tunnel excavation in
Tests 3, 5, 6 and 16. The test results reveal that all the induced axial load profiles are
similar with the maximum values taking place slightly higher than the tunnel spring
line in the short-term, with the neutral planes shifting lower over time. This implies
that the axial load transfer patterns of the piles are similar irrespective of the pile-totunnel distance for a long pile. A similar trend of axial load variation with pile-totunnel distance was also reported by Ran (2004) and Mroueh et al. (1999). It is
observed that the induced pile axial forces decrease with an increase in pile-to-tunnel
distance. Similar steady decrease is also observed in the long-term. This may be
attributed to the fact that the total contact area for piles in the immediate shear zone is
reduced when the distance of pile-to-tunnel increases. As such, a shorter portion of the
total pile length experiences negative skin friction as a pile is located further away
from the zone of large displacements. This can be attributed to the reduced shaft
contact area with the soils in the immediate shear zone when the distance of pile-totunnel increases. Figure 5.13

shows the variation of pile head settlement and the

free-field vertical soil movement at the respective pile location. Similar gradual
decreases in vertical soil settlement with depth and pile-to-tunnel distance are observed.
The magnitude of pile head settlement also decreases with increasing pile-to-tunnel
distance. This is consistent with the observed variations of pile axial forces from the
three tests. Besides, the smaller magnitudes of soil settlement is expected to induce less
negative skin friction (Figure 5.12) for the case in which pile tip is below the tunnel
invert. Once again, the pile head settlement exhibits time-dependent behaviour and

163

Chapter 5 Effects of Tunnelling on Single Piles

reaches its respective peak value after 720 days. The results suggest that the induced
axial pile responses are insignificant when the pile-to-tunnel distance is larger than 2D
in the present study.

Figure 5.14 shows the variation of induced bending moment with pile-to-tunnel
distance. Generally, the induced pile bending moment profiles are similar in all tests
with the maximum bending moment occurring approximately at the tunnel spring
elevation. This is consistent with the numerical predictions reported by Cheng (2003)
and field measurements reported by Pang (2006). As expected, the maximum induced
bending moment generally decreases with increasing pile-to-tunnel distance for both
short and long-terms. Hence, it would be reasonable to assume that induced bending
moment are generally small beyond a horizontal offset of 2D from the tunnel centre as
magnitudes are less than 50 kNm (long-term) even with a relatively large tunnel
volume loss of 3%. In addition, it is evident that regardless of pile-to-tunnel distance,
the induced maximum bending moment increases for some time after the completion
of tunnel excavation in all tests, exhibiting the time-dependent behaviours as described
earlier. The pile responses peak at 720 days after excavation. Figure 5.15 shows the
variations of pile deflection and free-field lateral soil movement profiles from these
three tests. As expected, the pile deflection decreases with increasing pile-to-tunnel
distance. Moreover, it is observed that the pile deflection drops rapidly from 1D to
1.5D, with a much smaller decrease from 1.5D to 2D. The pile lateral responses are
best explained by the soil deflection profiles obtained from Test 1 as shown in Figure
5.15. It is noted that the lateral soil movements in the three tests increase with time and
decrease with increasing distance of pile location to the tunnel, as the soil movement
decreases when the pile-to-tunnel distance increases. It is also noted that at the distance

164

Chapter 5 Effects of Tunnelling on Single Piles

of one tunnel diameter (1D), the lateral soil displacement is prominent at the tunnel
spring elevation and surface. However, when the distance is large enough, for instant
at a distance of 2D, the lateral soil displacement profile reveals significant soil
deflection at the ground surface while the soil movement at the tunnel spring elevation
becomes negligible. This can be related to the soil movement pattern as observed in the
immediate shear zone and thus results in the observed pile bending moment and
deflection shown in Figures 5.14 and 5.15. This finding demonstrates that when the
pile-to-tunnel distance increases, a shorter portion of the pile length is inside the
immediate shear zone, resulting in a smaller tunnel-pile interaction.

5.4.2 Test Series 5 - Free-Head End Bearing Piles (Tests 10, 11, 12)

In this test series, Tests 10, 11 and 12 were performed to study the effects of pile from
tunnel for free-head end bearing piles. The pile-to-tunnel distance in Tests 10, 11 and
12 is 6 m, 10 m and 14 m, respectively. Other parameters are kept constant in all tests.
Figure 5.16 shows the variation of induced free-head end-bearing pile axial force due
to tunnel excavation with pile-to-tunnel distance for Tests 10, 11 and 12. Generally,
the test results illustrate that all the induced axial load profiles are similar with the
maximum values taking place slightly lower than the tunnel spring line in the shortterm, with the neutral plane becomes deeper over time. This implies that the axial load
transfer patterns of the piles are essentially the same irrespective of pile-to-tunnel
distance for an end-bearing pile.

Similar to test series 4 (effects of distance of pile from tunnel for free-head
floating piles), the induced axial force decreases when the pile-to-tunnel distance
165

Chapter 5 Effects of Tunnelling on Single Piles

increases. This may be attributed to the fact that the total contact area for piles in the
immediate shear zone is lessened when the distance of pile-to-tunnel increases, as
noted in Section 5.4.1.

An important feature of an end-bearing pile is that the pile would undergo


minimal settlement as the pile tip is rested on very stiff ground. As such, pile
settlement is not a major concern. Although the pile may undergo elastic shortening,
the magnitude is small and thus negligible.

Figure 5.17 shows the variation of induced free-head end-bearing pile


bending moment with distance of pile from tunnel centre. The induced positive and
negative bending moments generally decrease, as expected, with increase in piletunnel distance. This trend is similar to that observed in test series 4 for free-head
floating piles.

Figure 5.18 illustrates the variation of induced pile deflection profiles for
Tests 10, 11 and 12. The magnitude of pile head deflection decreases when the piletunnel distance increases. Likewise, the mid-pile shaft also deflects in a similar trend.

5.4.3 Test Series 6 - Fixed-Head End Bearing Piles(Tests 13, 14A, 14B)

Tests 13, 14A and 14B aim to investigate the effects of pile location from tunnel for
fixed-head end bearing piles. The pile-to-tunnel distance in Tests 13, 14A and 14B is 6
m, 10 m and 14 m, respectively. Other parameters are kept constant in all tests. It
should be noted that pile axial force is only investigated in Test 13 (See Figure 5.2) but

166

Chapter 5 Effects of Tunnelling on Single Piles

not in Tests 14A and 14B, as the axial force profiles for pile-to-tunnel distance of 10 m
and 14 m are expected to be similar to that of Test 13 but with smaller magnitudes.
Nevertheless, the pile bending moment in Tests 14A and 14B are worthy for further
study as the changes are more significant and the results can also be served as the
bench mark for comparison with the pile group responses presented in Chapter 6.

Figure 5.19 shows the variation of induced fixed-head end-bearing pile bending
moment with pile distance from tunnel centre for Tests 13, 14 and 15. In all cases, the
negative bending moment is larger than the positive bending moment due to the
restraint at pile head. The results indicate that both positive and negative bending
moment decreases when the pile-tunnel distance increases.

Figure 5.20 shows the pile deflection profiles which are derived from the pile
bending moment profiles. It is shown that the pile deflection is generally very small
(less than 0.02 mm). The pile deflection profile reveals that the mid-pile shaft deflects
toward the tunnel, but the magnitude decreases when the pile distance to tunnel
increases.

5.4.4 Comparison of Results from Test Series 4, 5 and 6

Figure 5.21(a) shows the variation of maximum pile axial force with pile-to-tunnel
distance for Test Series 4, 5 and 6. It should be noted that for Test Series 6, there were
no axial piles monitored in Tests 14A and 14B. With the test range of pile location of 6
m to 14 m from the tunnel centre, the induced pile axial forces are observed to
decrease fairly linearly with an increase in pile-to-tunnel distance for Test Series 4 and

167

Chapter 5 Effects of Tunnelling on Single Piles

5. Similar steady decrease is also observed in the long-term. This may be attributed to
the fact that the total contact area for piles in the immediate shear zone becomes
smaller when the distance of pile-to-tunnel increases, as illustrated in Figure 5.22. As
such, a shorter portion of the total pile length experiences smaller negative skin friction
as the pile moves away from the zone of large displacements. It is also observed that
the induced axial force in Test Series 5 (end-bearing piles) is much larger than that of
Test Series 4 (floating piles) because the larger soil settlement relative to pile
settlement would induce a much larger axial force on the end-bearing pile. In addition,
the most significant difference in Test Series 5 (free-head) and Test Series 6 (fixedhead) is that tension force is induced in the fixed-head pile due to the total fixed
condition at the pile head.

Similarly to pile axial force, consistent responses in the corresponding soil


surface settlement (Test 1) and the maximum pile head settlement in Test Series 4
(Tests 3, 5, 6 and 16) are observed from Figure 5.21(b). The soil surface and pile
settlements are observed to decrease with an increase in pile-to-tunnel distance in both
short- and long-term. This is simply because when the pile-to-tunnel distance increases,
the soil settlements within the immediate shear zone decreases, as shown in Figure
5.13. The result seems to suggest that the pile settlement are insignificant for pile-totunnel distance larger than 2D in Test Series 4. It should be noted that the pile
settlements of end-bearing piles in Test Series 5 and 6 are practically negligible as
shown in the figure.

In contrast to the pile vertical responses, the pile lateral responses are different.
Figure 5.21(c) shows the induced maximum pile bending moment with pile-to-tunnel

168

Chapter 5 Effects of Tunnelling on Single Piles

distance in the short and long-terms for Test Series 4, 5 and 6. Generally, the
maximum induced bending moments decrease reasonably linearly with increasing pileto-tunnel distance when the magnitude is relatively small (see for instant pile bending
moment in the short-term and positive bending moment in series 6 (long-term)).
However, the bending moments decrease exponentially when the magnitude is
relatively large (for example long-term bending moment in series 4, 5 and negative
bending moment in series 6). Generally, the induced bending moments in end-bearing
piles (Test Series 5) are larger than the floating piles (Test Series 4). This is probably
because the restraint at the pile toe would restrict the pile lateral movement and induce
a larger bending moment. On the other hand, the data reveals that negative bending
moments are induced at pile head due to total fixity condition for Test Series 6 (fixedhead) as compared to the free-head piles in Test Series 5. Nevertheless, since the
bending moments are offset toward the negative bending moment, the trend reveals
that the positive bending moment in a fixed-head pile (Test Series 6) is consistently
smaller than that of the free-head pile (Test Series 5), regardless of pile-tunnel position.
In addition, it would be reasonable and safe to assume that induced bending moments
are generally small beyond a horizontal offset of 2D from the tunnel centre as
magnitudes are less than 50 kNm (long-term) even with a tunnel volume loss of 3%. It
is evident that regardless of pile-to-tunnel distance, both induced maximum bending
moments increase for some time after the completion of tunnel excavation in all the
tests, exhibiting the time-dependent behaviours described earlier. The pile responses
peak at 720 days after excavation. This further illustrates that the induced pile bending
moments are small as the lateral soil movements are not significant when the pile-totunnel distance increases beyond 2D, as illustrated in Figures 5.22 and 5.23. As
discussed in Section 4.3.1, the soil within the Immediate Shear Zone is unloaded

169

Chapter 5 Effects of Tunnelling on Single Piles

due to tunnel excavation and gradually deforms by arching, causing the radial stress in
the immediate shear zone to be reduced due to stress relief. This leads to the observed
soil movement pattern which subsequently affects the pile responses as observed in
this chapter. On the other hand, circumferential soil stresses increase within the
Support Zone to support the arches formed in the immediate shear zone. Thus, Figure
5.22 illustrates that the pile responses for different pile-tunnel distance is greatly
influenced by the changes of the stress in both the Immediate Shear Zone and Support
Zone.

Figure 5.21(d) illustrates the variations of pile head deflection for the floating
piles (Test Series 4) and end-bearing piles (Test Series 5). Generally, it is observed that
the pile deflection reduces rapidly from 1D to 1.5D, with a much smaller decrease
from 1.5D to 2D for both Series 4 and 5. This is probably because the lateral soil
movement decreases with increasing distance of pile from to the tunnel. Nevertheless,
it is expected that the pile head deflection for end-bearing piles (Test Series 5) is
smaller than that for floating piles (Test Series 4), regardless of pile-tunnel distance
because the lower portion of pile is restrained and cannot move. On the other hand, the
measured pile head deflection of Test Series 6 (fixed-head) is negligible due to totally
fixed condition at pile head.

The observed variation of pile bending moment and deflection with pile-totunnel distance can be explained by the soil movemnt profiles obtained from Test 1,
which was analysed by PIV (see Figure 5.23). The results reveal that at locations near
the tunnel, the lateral soil displacement is prominent at the tunnel spring elevation.
However, when the distance is large enough, the lateral soil displacement profile

170

Chapter 5 Effects of Tunnelling on Single Piles

reveals significant horizontal soil movement at the ground surface while the soil
movement at the tunnel spring elevation became negligible. This finding is consistent
with the relatively insignificant observed pile lateral responses when the pile-to-tunnel
distance is more than 2D. It thus further illustrates that when the pile-to-tunnel distance
increases, a shorter portion of the pile length is inside the immediate shear zone shown
in Figure 5.22.

5.5

EFFECTS OF TIME ON PILE RESPONSES IN SOFT CLAY

In order to further assess the long-term pile responses, Figure 5.24 shows a summary
of the ratio of long-term to short-term pile responses for all tests presented in this
chapter. The results show that the long-term to short-term pile responses (pile axial
force, pile bending moment, pile head settlement and pile head deflection) ranges from
1.34 to 3.5. It is noted that the soil movement is dominant in the vertical direction in
the LT. As such, the LT over ST ratio is comparatively important for pile settlement
which depends on the magnitude of downward soil movement before full pile slip. On
the other hand, the LT over ST ratio for pile axial force is comparatively small as the
axial forces might have been fully mobilized at an earlier stage. Although the
magnitude of lateral soil movement is much smaller than that of vertical soil
movement, the increase in lateral soil movement is significant and thus in turn, causing
both pile bending moment and pile deflection to increase over time. This study has
provided strong evidence that the long-term pile behaviour needs to be considered if
the volume loss due to tunnelling is large.

171

Chapter 5 Effects of Tunnelling on Single Piles

5.6

COMPARISON OF PILE BEHAVIOUR DUE TO INWARD


AND OUTWARD TUNNEL DEFORMATIONS

Ran (2004) carried out centrifuge model studies on the effect of tunnelling on single
piles in clay. In his study, outward tunnel deformation was simulated (ovalisation of
tunnel lining, i.e. tunnel springline moves outwards), as shown in Figures 5.25 and
5.26. Although outward tunnel deformation is not as commonly observed as inward
tunnel deformation, there have been some reports of outward tunnel deformation in the
field (George, 1981 and Yann and Alain, 1991). For example, the Handbook of Plastic
Pipe Institute (2003) reported that the deformation patterns of poly-material linings
under service load are of horizontal-oval shape. Furthermore, large deformations
experienced by some tunnel lining rings are encountered in some tunnelling projects.
In this section, the difference between inward and outward tunnel deformation patterns
and its influences to adjacent piles are investigated.

It should be noted that the volume loss for the outward tunnel deformation
simulation was 2%, while it is slightly larger at 3% for the inward tunnel deformation
simulation in the present study. Some similarities and differences can be drawn in the
behaviours of soil and single piles induced by both inward and outward tunnel
deformations in clay.

5.6.1 Tunnel-Soil Interaction

Figure 5.27 shows the development of subsurface soil movements at 2 days and 720
days after tunnel excavation for the case of outward tunnel deformation, which was
conducted by (Ran, 2004). Subsurface soil movement was traced from high resolution

172

Chapter 5 Effects of Tunnelling on Single Piles

photographs of the marker beads and analyzed using Computer Program OPTIMUS
instead of PIV used in this study.

5.6.1.1 Similarities (Tunnel-Soil Interaction)

Figures 5.28 and 5.29 show the variation of surface soil settlement troughs with tunnel
deformation over time. It is interesting to note that the measured short-term surface
settlement trough follows the Gaussian distribution curve fairly well with the inflection
point (i) at approximately 7.5 m for both cases, despite the difference in tunnel
deformation. This finding is consistent with the observation made by Verruijt and
Booker (1996). The void created by ovalisation deformation at the tunnel crown is
similar to the clearance at the upper half of the tunnel for the case of contraction
deformation. Hence the soil above the tunnel crown in both cases settles by a similar
vertical distance. However, Verruijt and Booker (1996), (2000) reported that for cases
with significant tunnel ovalisation, the surface settlement trough would be narrower
than the Gaussian curve.

The Gaussian curve is also found to be inappropriate in depicting the measured


long-term surface settlement troughs for both cases. In addition, it is observed that the
soil continues to settle with time and the settlement rate decreases with time. This
exhibits time-dependent behaviour of clay for both cases. An examination of soil
settlement at the pile location (i.e. 6m or 1D from tunnel centre-line) shown in Figure
5.30 reveals that the soil settlement profile is fairly similar for these 2 situations.

173

Chapter 5 Effects of Tunnelling on Single Piles

5.6.1.2 Differences (Tunnel-Soil Interaction)

The most distinct difference in the soil behaviour between the 2 different tunnel
deformations is that the soil moves intensely away from the tunnel due to spring line
expansion in the case of outward tunnel deformation and a much larger deformation
zone is formed. In the case of inward tunnel deformation, the soil moves towards the
tunnel and the immediate shear zone is defined as discussed in Section 4.3 which is
smaller than that of the deformation zone for the outward tunnel deformation. Thus,
qualitatively, the soil movement above the tunnel crown is prominent in the case of
inward tunnel deformation; whereas the lateral soil movement near tunnel axis is
prominent in the case of outward tunnel deformation. This is illustrated in Figure 5.31
that the soil moves in opposite directions at the pile location for these 2 situations due
to differences in the deformation of the tunnel lining. However, the magnitude is much
larger in the case of outward tunnel deformation, despite the volume loss is smaller at
only 2% as compared to 3% for the inward tunnel deformation case.

5.6.2 Tunnel-Pile Interaction

In both tunnel deformation studies, the pile-to-tunnel distance is similar, i.e. 6 m from
the tunnel centre-line. However, the pile length in the case of outward tunnel
deformation is 23.5 m whereas it is slightly shorter at 22 m in the case of inward tunnel
deformation. The tunnel depth remains at 15 m in both cases. The long-term pile axial
forces were not presented by Ran (2004) as the pile axial forces were measured by
quarter bridge strain gauge circuits. Strain gauge readings obtained with quarter

174

Chapter 5 Effects of Tunnelling on Single Piles

bridge circuits have a tendency to drift with time due to temperature changes, whereas
this does not happen for full bridge circuits in the present study. Similar to the
tunnel-soil interaction studies, the volume loss in the outward tunnel deformation and
inward tunnel deformation simulations is 2% and 3%, respectively.

5.6.2.1 Similarities (Tunnel-Pile Interaction)

For both tunnel lining deformations, the induced pile axial forces (Figure 5.32) exhibit
similar profiles, as the piles experience negative skin friction due to settling soil
around the pile shaft. Also, the neutral plane is found to be located approximately at
the tunnel axis for both cases. Despite the slightly lower volume loss in the case of
outward tunnel deformation, the pile axial forces in both cases have similar magnitudes.
This is because the relatively large amount of tunnel expansion at the springline for the
outward tunnel deformation causes large settlement, as clearly shown in Figure 5.30.
In the same way, Figure 5.33 reveals that the pile vertical settlement is also timedependent. As the piles are floating in the soft clay instead of being socketed into the
hard stratum for both cases, the pile settlement depends very much on the vertical soil
movement along the pile shaft. Hence, the piles continue to settle with the soil until
full pile slip, regardless of tunnel deformation patterns.

5.6.2.2 Differences (Tunnel-Pile Interaction)


It is worth noting that totally different pile bending moment profiles are induced under
the two different tunnel deformation patterns. It is observed that outward tunnel
deformation causes negative maximum induced bending moment (bending away from

175

Chapter 5 Effects of Tunnelling on Single Piles

tunnel) whereas inward tunnel deformation causes positive maximum induced bending
moment which bends toward tunnel (Figures 5.34 and 5.35).This is because in the case
of outward tunnel deformation, the soil moved away from the tunnel at the tunnel axis
due to the protrusion of the tunnel lining at the tunnel spring line, hence pushing the
pile away from the tunnel. On the other hand, in the case of inward tunnel deformation,
the soil moves towards the tunnel axis due to volume loss caused by excavation overcut, thus drawing the pile towards the tunnel. The relatively large magnitude of
induced pile bending moment in the outward tunnel deformation simulation is mainly
due to the large expansion of the tunnel lining at the springline. In the case of inward
tunnel deformation, the gap above the tunnel crown is the main cause of soil
movements (Leung, 2006). The induced pile deflection profiles shown in Figure 5.36
demonstrate the pushing of the mid-pile shaft away from the tunnel due to tunnel
protrusion at the tunnel spring elevation. In the case of inward tunnel deformation, the
pile moves towards the tunnel, with the pile head deflecting more than the pile tip.
These findings are consistent with the soil movements observed in Figure 5.31.

Similar to the study on deep excavations in clay at NUS (Leung et al.,


2006), the clay continues to move over time after the completion of tunnel excavation.
Thus, the induced pile bending moments, head settlement and deflection in both tunnel
deformation simulations also change with time, as illustrated by the long-term to shortterm ratio of pile responses in Figure 5.37.

The above comparison of pile behaviours due to inward and outward tunnel
deformations illustrate that the trend of pile axial force and pile settlement behaviour
and profile are essentially similar regardless of tunnel deformation pattern. However,

176

Chapter 5 Effects of Tunnelling on Single Piles

the outward tunnel formation would induce higher pile responses as compared to the
inward tunnel deformation under the same volume loss. On the other hand, the pile
lateral responses (bending moment and deflection) are totally opposite for both inward
and outward tunnel deformations, respectively, in terms of profiles and magnitude.
This is due to the fact that the different tunnel deformation patterns would induce
different soil movement profiles and patterns, which in turn changes the pile behaviour
significantly. Nevertheless, considerable engineering judgement and experiences are
required to first determine the tunnel deformation pattern, in a case by case basis.
Subsequently, the analysis of soil movements is necessary to evaluate the pile
responses due to tunnelling.

5.7

CONCLUDING REMARKS

A total of thirteen centrifuge model tests (Tests 3 to 16) have been performed to
examine the fundamental mechanisms behind the effects of tunnelling on single piles.
Three main items were investigated: (1) pile vertical and lateral responses in different
tunnel-pile configurations (six test series), (2) tunnelling-induced response of a single
pile in relation to tunnelling-induced response of free-field ground, and (3) comparison
of soil and single pile behaviours due to inward and outward tunnel deformations.

Three different pile tip conditions, namely floating pile, socketed pile and
end-bearing pile were investigated to study the effects of pile tip condition. It is
noted that a floating pile is mainly governed by pile settlement when tunnelling is
carried out adjacent to it. On the contrary, socketed piles are likely governed by the
material stress of the pile. On the other hand, some opposite trends are observed in the
fixed-head when compared to free-head. It is noted that tensile force and relatively

177

Chapter 5 Effects of Tunnelling on Single Piles

large negative bending moments are induced at the pile head due to total fixity.
Nevertheless, these responses have led to the reduction in drag load and positive
bending moment at the mid-pile shaft.

Different lengths of piles are deployed to further assess the effect of pile
length over tunnel depth due to tunnelling. In a short pile, especially those located in
the immediate shear zone, the pile structural responses are less vulnerable as compared
to those of long pile. However, there will be excessive pile movements (settlement and
deflection) because of lack of anchorage of pile into the stable support zone. In this
respect, a longer pile with pile length in the stabilised support zone tends to provide
more resistance to the pile movements but will attract more bending moment and axial
force.

Test Series 4, 5 and 6 examine the effects of pile-to-tunnel distance for


different pile head and tip conditions. Generally, it is observed that the pile responses
decrease with increase in pile-to-tunnel distance. The induced pile axial forces are
observed to decrease fairly linearly with increase in pile-to-tunnel distance for Test
Series 4 (floating piles) and 5 (end-bearing pile) and the axial force in Test Series 5 is
always much larger than that of Test Series 4 for all pile to tunnel distances. In
addition, the most significant difference in Test Series 5 (free-head) and Test Series 6
(fixed-head) is that tension force is induced in the fixed-head pile due to total fixed
condition at the pile head. Similarly to pile axial force, consistent responses are also
observed in the corresponding soil surface settlement (Test 1) and the maximum pile
head settlement in Test Series 4.

178

Chapter 5 Effects of Tunnelling on Single Piles

In contrast to the pile vertical responses, the pile lateral responses are different.
Generally, the maximum induced bending moments decrease fairly linear with
increasing pile-to-tunnel distance when the magnitude is relatively small (for instant
pile bending moment in the short-term and positive bending moment in series 6, but
the bending moments decrease exponentially when the magnitude is relatively large.
Generally, the induced bending moments in end-bearing piles (Test Series 5) are larger
than the floating piles (Test Series 4) due to the restraint at the pile toe would restrict
the pile lateral movement and induce a larger bending moment. On the other hand, the
results reveal that negative bending moments are induced at pile head due to total
fixity condition for Test Series 6 (fixed-head) as compared to the free-head piles in
Test Series 5. As the bending moment profile is offset toward the negative bending
moment for fixed-head pile, the positive bending moment in fixed-head pile (Test
Series 6) are consistently lower than that of free-head pile (Test Series 5). It can be
established from the test results that induced bending moments are generally small
beyond a horizontal offset of 2D from the tunnel centre. Generally, it is observed that
the induced pile deflection reduces rapidly from 1D to 1.5D, with a much smaller
decrease from 1.5D to 2D for both Series 4 and 5. This is because the lateral soil
movements decrease with increasing distance of pile location to the tunnel. The pile
head deflection for end-bearing piles (Test Series 5) is smaller that of floating piles
(Test Series 4), regardless of pile-tunnel distance, as the lower portion of the pile is
restrained and cannot move.

Some similarities and differences were drawn in the comparisons of soil and
single pile behaviour in the cases of both inward (present study) and outward (Ran,
2004) tunnel deformations. It is noted that the measured short-term surface settlement

179

Chapter 5 Effects of Tunnelling on Single Piles

trough follows the Gaussian distribution curve fairly well with the inflection point (i)
at approximately 7.5 m for both tunnel deformation cases. The most distinct difference
in the soil behaviour is that the soil moves significantly away from the tunnel in the
case of outward tunnel deformation, whereas in the case of inward tunnel deformation,
the soil moved towards the tunnel. It is revealed that the pile axial force and pile
settlement behaviour and profile are essentially similar regardless of tunnel
deformation pattern, but the outward tunnel formation would induce larger pile
responses as compared to the inward tunnel deformation under the same volume loss.
On the other hand, the pile lateral responses (bending moment and deflection) are
opposite in direction for both inward and outward tunnel deformations, respectively, in
terms of profiles and magnitude.

180

Chapter 5 Effects of Tunneling on Single Piles

Table 5.1 Test program and prototype parameters in Phase 2 study


Phase 2 -Effects of tunnelling on single piles
Test series 2 Effects of pile tip & head conditions
Test
Configuration
No.
Bending Pile

Common
parameters

Floating
free-head pile.
Thickness of sand,
S=3.5m
Pile embedment
length, L=22m

Axial Pile

3
Typical

Individual
parameters

24m
D
2m

Bending Pile

3.5m

Axial Pile

Socketed
free-head pile.
S=8.5m.
Pile socketed 3m
in sand
L=22m

19m

L
D
3m

8.5m

C = 12 m
D=6m

Bending Pile

X=6m

Axial Pile

Volume loss = 3 %
C
L

24m
D

10

End-bearing
free-head pile.
S=3.5m
L=27.5m

3.5m

Bending Pile Axial Pile


Tie
Beam

13

C
24m
D

24m
X

End-bearing
fixed-head pile.
S=3.5m
L=27.5m

3.5m

181

Chapter 5 Effects of Tunneling on Single Piles

Test series 3 Effects of pile length for free-head piles


Test
No.
3
Typical

Common
parameters

Configuration
Bending Pile

Individual
parameters

Axial Pile

L=22m
C
L

24m
D

C = 12 m

D=6m
3.5m

X=6m

L=11.4m

Volume loss = 3 %
8
L=15.6m

Test series 4 Effects of distance of pile from tunnel (a) free-head floating piles
Test
No.
3
Typical

Common
parameters

Configuration
Bending Pile

Individual
parameters

Axial Pile

X=6m
C
L

24m
D

X
3.5m

16

X=9m
C = 12 m
D=6m
L = 22 m
X = 10m
Volume loss = 3 %

X= 12 m

182

Chapter 5 Effects of Tunneling on Single Piles

Test series 5 Effects of distance of pile from tunnel (b) free-head end bearing piles
Test
No.

Common
parameters

Configuration

Individual
parameters

10
Bending Pile

X=6m

Axial Pile

11
C
L

24m
D

C = 12 m

X = 10 m

D=6m

X
3.5m

L = 27.5 m
Volume loss = 3 %

12

X= 14 m

`
Test series 6 Effects of distance of pile from tunnel (c) fixed-head end bearing piles
Test
No.
13

Common
parameters

Configuration

X=6m

Bending Pile Axial Pile


Tie
Beam
C

14A

C = 12 m
D=6m

24m
D

Individual
parameters

X = 10 m

L = 27.5 m

Volume loss = 3 %
14B

3.5m

X= 14 m

183

Chapter 5 Effects of Tunneling on Single Piles

Pile-tunnel configuration
(Pile base position)

Test 7

L/H=0.76

H=15m

Tunnel

Test 8

L/H=1.04

D=6m

Tests 3, 4, 9

Tests 5, 16

Tests 6
L/H=1.5

Kaolin Clay

X/D=1

X/D=1.5

X/D=2

Toyoura Sand
Free-head- Test 10
Fixed-head- Test 13

Test 11
Test 14A

Test 12
Test 14B

Notes:
Volume loss for all tests is 3%, except Test 4 (Vol. loss=6.5%)
L
H
X
D

=
=
=
=

Pile length
Tunnel depth
Distance between tunnel axis and centre of pile
Tunnel diameter

Figure 5.1 Pile base position investigated in the parametric studies (not to scale)

184

Chapter 5 Effects of Tunneling on Single Piles

Axial Force (kN)


-200

200

400

600

800

1000

Test 3-floating, free-head

-2.5

Test 9-socketed, free-head


Test 10-end-bearing, free-head
Test 13-end-bearing, fixed-head

Depth (m)

-7.5

-12.5
T unnel

-17.5

-22.5

-27.5

(a) Short-term
Axial Force (kN)
-200

200

400

600

800

1000

Test 3-floating, free-head

-2.5

Test 9-socketed, free-head


Test 10-end-bearing, free-head
Test 13-end-bearing, fixed-head

Depth (m)

-7.5

-12.5
T unnel

-17.5

-22.5

-27.5

(b) Long-term
Figure 5.2 Variation of pile axial force with tip condition in (a) Short-term (b) Longterm (Tests 3, 9, 10 and 13)
185

Chapter 5 Effects of Tunneling on Single Piles

Bending Moment (kNm)


-200 -150 -100 -50

50

200 250

300

Test 3-floating, freehead


Test 9-socketed, freehead
Test 10-end-bearing,
free-head
Test 13-end-bearing,
fixed-head

-2.5

-7.5

Depth (m)

100 150

-12.5
T unnel
-17.5

-22.5

-27.5

(a) Short-term
Bending Moment (kNm)
-200 -150 -100 -50

50

150 200

250 300

Test 3-floating, freehead


Test 9-socketed,
free-head
Test 10-end-bearing,
free-head
Test 13-end-bearing,
fixed-head

-2.5

-7.5

Depth (m)

100

-12.5
T unnel
-17.5

-22.5

-27.5

(b) Long-term
Figure 5.3 Variation of pile bending moment with tip condition (a) Short-term (b)
Long-term (Tests 3, 9, 10 and 13)
186

Chapter 5 Effects of Tunneling on Single Piles

Lateral deflection (mm)


-4

-2

10

12

14

-2.5

Depth (m)

-7.5

-12.5
T unnel
-17.5
Test 3-floating, free-head
Test 9-socketed, free-head

-22.5

Test 10-end-bearing, free-head


Test 13-end-bearing, fixed-head
-27.5

(a) Short-term
Lateral deflection (mm)
-4

-2

10

12

14

-2.5

Depth (m)

-7.5

-12.5
T unnel
-17.5
Test 3-floating, free-head
-22.5

Test 9-socketed, free-head


Test 10-end-bearing, free-head
Test 13-end-bearing, fixed-head

-27.5

(b) Long-term
Figure 5.4 Variation of pile deflection with tip condition (a) Short-term (b) Long-term
(Tests 3, 9, 10 and 13)

187

Chapter 5 Effects of Tunneling on Single Piles

24

800

ST

Pile head settlement (mm)

ST

600

LT

500
400
300
200
100
0
-100
-200

Test 3Test 9floating, socketed,


free-head free-head

Test 10- Test 13- Test 13endendendbearing, bearing, bearing,


free-head fixed-head fixed-head
(at pile
head)

20

LT

16
12
8
4
0
Test 3floating, freehead

(a)

Test 9socketed,
free-head

Test 10-end- Test 13-endbearing, free- bearing, fixedhead


head

(b)

16

200

LT

100
50
0
-50
-100
-150

Test 3Test 9floating, socketed,


free-head free-head

Test 10Test 13Test 13endendendbearing,


bearing,
bearing,
free-head fixed-head fixed-head
(at pile
head)

-200

(c)

Pile head deflection (mm)

ST
ST

150

Maximum pile bending moment (kNm)

Maximum pile axial force (kN)

700

LT

12

0
Test 3floating, freehead

Test 9socketed,
free-head

Test 10-end- Test 13-endbearing, free- bearing, fixedhead


head

(d)

Figure 5.5 Variation of (a) maximum pile axial force (b) pile head settlement (c) pile
bending moment (d) pile head deflection with tip and head conditions (Tests 3, 9, 10
and 13)

188

Chapter 5 Effects of Tunneling on Single Piles

Axial Force (kN)


0

100

200

300

400

500

600

700

800

-5

Depth (m)

-10

-15

T unnel

-20
ST (Test 3, L=22m)
LT (Test 3, L=22m)
ST (Test 7, L=11.4m)
LT (Test 7, L=11.4m)
ST (Test 8, L=15.6m)
LT (Test 8, L=15.6m)

-25

-30

Figure 5.6 Variation of pile axial force with pile length (Tests 3, 7 and 8)
Settlement (mm)
0

20

40

60

80

100

120

-5

Depth (m)

-10

-15

-20

-25

T unnel

ST (Test 1, Soil settlement)


LT (Test 1, Soil settlement)
ST (Test 3, L=22m)
LT (Test 3, L=22m)
ST (Test 7, L=11.4m)
LT (Test 7, L=11.4m)
ST (Test 8, L=15.6m)
LT (Test 8, L=15.6m)

-30

Figure 5.7 Variation of pile head settlement and soil settlement profile (Test 1) with
pile length (Tests 3, 7 and 8)

189

Chapter 5 Effects of Tunneling on Single Piles

Bending Moment (kNm)


-100

-50

50

100

150

200

250

300

-5

Depth (m)

-10

-15

T unnel

-20
ST (Test 3, L=22m)
LT (Test 3, L=22m)
ST (Test 7, L=11.4m)
LT (Test7, L=11.4m)
ST (Test 8, L=15.6m)
LT (Test 8, L=15.6m)

-25

-30

Figure 5.8 Variation of pile bending moment with pile length (Tests 3, 7 and 8)
Lateral deflection (mm)
0

10

20

30

40

-5

Depth (m)

-10

-15

-20

-25

T unnel

ST (Test 1, Soil lateral deflection)


LT (Test 1, Soil lateral deflection)
ST (Test 3, L=22m)
LT (Test 3, L=22m)
ST (Test 7, L=11.4m)
LT (Test 7, L=11.4m)
ST (Test 8, L=15.6m)
LT (Test 8, L=15.6m)

-30

Figure 5.9 Variation of pile head deflection and free- field lateral soil displacement
(Test 1) with pile length (Tests 3, 7 and 8)
190

Chapter 5 Effects of Tunneling on Single Piles

500

80

Pile head settlement (mm)

Maximum pile axial force (kN)

70
ST

400

LT

300

200

ST

60
LT

50
40
30
20

100

10
0
0

0.5

1.5

Normalised pile length over tunnel depth, L/H

0.5

1.5

(a)

(b)

200

30

ST

Pile head deflection (mm)

Maximum pile bending moment (kNm)

Normalised pile length over tunnel depth, L/H

150
LT

100

25

ST

20

LT

15

10

50
5
0
0

0.5

1.5

Normalised pile length over tunnel depth, L/H

(c)

0
0

0.5

1.5

Normalised pile length over tunnel depth, L/H

(d)

Figure 5.10 Variation of (a) maximum pile axial force (b) pile head settlement (c) pile
bending moment (d) pile head deflection with normalized pile length over tunnel depth
(Tests 3, 7 and 8)

191

Chapter 5 Effects of Tunneling on Single Piles

Short pile/Long pile (L=22m, Test 3) ratio

5
4.5
4
3.5
3

Negative
effect for
shorter pile

2.5
2
1.5

Positive
effect for
shorter pile

1
0.5
0
Pile axial force

Pile bending
moment

ST-Test 7 (L/H=0.76)
ST-Test 8 (L/H=1.04)

Pile head
settlement

Pile head
deflection

LT-Test 7 (L/H=0.76)
LT-Test 8 (L/H=1.04)

Figure 5.11 Short pile to long pile ratio of pile responses for different pile length over
tunnel depth (Tests 3, 7 and 8)

192

Chapter 5 Effects of Tunneling on Single Piles

Axial Force (kN)


0

100

200

300

400

500

600

-5

Depth (m)

-10

-15

T unnel

-20

ST (Test 3, X= 6m)

LT (Test 3, X= 6m)

ST (Test 5, X= 9m)

LT (Test 5, X= 9m)

ST (Test 6, X=12m)

LT (Test 6, X=12m)

-25

-30

Figure 5.12 Variation of pile axial force with pile-to-tunnel distance for free-head
floating piles (Tests 3, 5 and 6)
Settlement (mm)

20

40

60

80

100

120

-5

Depth (m)

-10

T unnel

-15

Test 3 Pile head settlement


-20

Test 5 Pile head settlement


Test 6 Pile head settlement

-25

Free-field soil settlement


ST (Test 1, X= 6m)

LT (Test 1, X= 6m)

ST (Test 1, X= 9m)

LT (Test 1, X= 9m)

ST (Test 1, X=12m)

LT (Test 1, X=12m)

-30

Figure 5.13 Variation of pile head settlement for free-head floating piles (Tests 3, 5
and 6) and free-field soil settlement (Test 1) with pile-to-tunnel distance

193

Chapter 5 Effects of Tunneling on Single Piles

Bending Moment (kNm)


-80

-40

40

80

120

160

200

240

-5

Depth (m)

-10

-15

T unnel

ST (Test 3, X= 6m)

-20

LT (Test 3, X= 6m)
ST (Test 5, X= 9m)
-25

LT (Test 5, X= 9m)
ST (Test 6, X=12m)
LT (Test 6, X=12m)

-30

Figure 5.14 Variation of pile bending moment for free-head floating piles with pile-totunnel distance (Tests 3, 5 and 6)
Lateral deflection (mm)
0

10

15

20

25

30

35

-5

Depth (m)

-10

-15

T unnel
Free-field lateral soil deflection

-20

-25

ST (Test 1, X= 6m)

LT (Test 1 X= 6m)

ST (Test 1, X= 9m)

LT (Test 1, X= 9m)

ST (Test 1, X=12m)
Pile head deflection
ST (Test 3, X= 6m)

LT (Test 1, X=12m)

ST (Test 5, X= 9m)

LT (Test 5, X= 9m)

ST (Test 6, X= 12m)

LT (Test 6, X=12m)

LT (Test 3, X= 6m)

-30

Figure 5.15 Variation of pile deflection for free-head floating piles (Tests 3, 5 and 6)
and free-field lateral soil displacement profile (Test 1) with pile-to-tunnel distance
194

Chapter 5 Effects of Tunneling on Single Piles

Axial Force (kN)


0

200

400

600

800

1000

-5

-10

Depth (m)

-15

T unnel

-20

-25

-30
Test 10, X= 6m (ST)

Test 10, X= 6m (LT)

Test 11, X=10m (ST)

Test 11, X=10m (LT)

Test 12, X=14m (ST)

Test 12, X=14m (LT)

Figure 5.16 Variation of pile axial force for free-head end bearing piles with pile-totunnel distance (Tests 10, 11 and 12)

195

Chapter 5 Effects of Tunneling on Single Piles

Bending Moment (kNm)


-100

-50

50

100

150

200

250

-5

Depth (m)

-10

-15

T unnel

-20
Test 10, X= 6m (ST)
Test 10, X= 6m (LT)
Test 11, X=10m (ST)
Test 11, X=10m (LT)
Test 12, X=14m (ST)
Test 12, X=14m (LT)

-25

-30

Figure 5.17 Variation of pile bending moment for free-head end bearing piles with
pile-to-tunnel distance (Tests 10, 11 and 12)

Lateral deflection (mm)


-2

-1

-5

Depth (m)

-10

`
-15

T unnel

-20

-25

ST (Test 10, X= 6m)


LT (Test 10, X= 6m)
ST (Test 11, X=10m)
LT (Test 11, X=10m)
ST (Test 12, X=14m)
LT (Test 12, X=14m)

-30

Figure 5.18 Variation of pile deflection for free-head end bearing piles with pile-totunnel distance (Tests 10, 11 and 12)

196

Chapter 5 Effects of Tunneling on Single Piles

Bending Moment (kNm)


-200

-150

-100

-50

50

100

150

200

-5

Depth (m)

-10

-15

T unnel

-20

-25

-30
Test 13, X= 6m (ST)

Test 13, X= 6m (LT)

Test 14A, X=10m (ST)

Test 14A, X=10m (LT)

Test 14B, X=14m (ST)

Test 14B, X=14m (LT)

Figure 5.19 Variation of pile bending moment for fixed-head end bearing piles with
pile-to-tunnel distance (Tests 13, 14A and 14B)
Lateral deflection (mm)
-0.01

0.01

0.02

0.03

-5

Depth (m)

-10

`
-15

T unnel

-20

-25

Test 13, X= 6m (ST)


Test 13, X= 6m (LT)
Test 14A, X=10m (ST)
Test 14A, X=10m (LT)
Test 14B, X=14m (ST)
Test 14B, X=14m (LT)

-30

Figure 5.20 Variation of pile deflection for fixed-head end bearing piles with pile-totunnel distance (Tests 13, 14A and 14B)

197

Chapter 5 Effects of Tunneling on Single Piles

100

800
Series 4 (ST)

700

Series 4 (LT)

Series 4 (LT)

80

Series 5 (ST)

Series 5 (ST)

Series 5 (LT)

Series 5 (LT)

400

Series 6 (ST)

300

Series 6 (LT)

200

Series 6,
tension (ST)
Series 6,
tension (LT)

100

70

Settlement (mm)

500

Series 6 (ST)

60

Series 6 (LT)

50

Soil surface settlement (LT), Test 1


Soil surface settlement (LT), Test 1

40
30
20

0
0

10

12

14

10

-100

-200

Distance between tunnel centre and pile (m)

10

12

14

Distance between tunnel centre and pile (m)

(a)

(b)

160

Series 4 (ST)

16

120

Series 4 (LT)

14

Series 4 (ST)
Series 4 (LT)

80

Series 5 (ST)

40

-80
-120

Series 5 (ST)

Series 5 (LT)

0
-40

12

Series 6 (ST)

10

12

14

Series 6 (LT)

Deflection (mm)

Maximum pile bending moment (kNm)

Maximum pile axial force (kN)

600

Series 4 (ST)

90

10
8

Series 6, head BM
(LT)

(c)

Series 6 (LT)

4
2

Distance between tunnel centre and pile (m)

Series 6 (ST)

Series 6, head BM
(ST)

-160

Series 5 (LT)

10

12

Distance between tunnel centre and pile (m)

(d)

Note:Test Series 4 Free-head floating piles


Test Series 5 Free-head end-bearing piles
Test Series 6 Fixed-head end-bearing piles

Figure 5.21 Variation of (a) maximum pile axial force (b) maximum pile head
settlement and soil surface settlement (Test 1) (c) pile bending moment (d) maximum
pile head deflection for Test Series 4, 5 and 6 with pile-to-tunnel distance

198

14

Chapter 5 Effects of Tunneling on Single Piles

Immediate
Shear Zone

Tunnel

Support Zone

Kaolin Clay
Toyoura Sand

Figure 5.22 Assessment of pile responses for different pile-to-tunnel distance


(Tests 3, 5 and 6)
6m from tunnel
Soil movement (mm)

4m from tunnel
Soil movement (mm)

-40

-30

-20

-10

-40

-30

-20

-10

12m from tunnel


Soil movement (mm)
0

-40

-30

-20

-10

15m from tunnel


Soil movement (mm)
0

-40

-30

-20

-10

-5

-5

-5

-5

-5

-10

-10

-10

-10

-10

-15

Tunnel

-15

-20

-20

-25

-25

-30

-30

-15

-15

Depth below GL (m)

Depth below GL (m)

Depth below GL (m)

Depth below GL (m)

-40 -30 -20 -10

9m from tunnel
Soil movement (mm)

-15

-20

-20

-25

-25

-30

-30

-20

-25

2 days
Loganathan
et al 1998

-30

720 days

Figure 5.23 Lateral soil displacement profiles at different pile-to-tunnel distance (Tests
3, 5 and 6)

199

Chapter 5 Effects of Tunneling on Single Piles

4
3.5
3

Long-term/ Short-term ratio

2.5
2

Long-term
effect

1.5
1
0.5
0
Pile axial force

Pile head settlement

Test 3 (Typical)
Test 6 (X=12m)
Test 9 (Socketed)
Test 12 (free-head, X=14m)
Test 14B (fixed-head, X=14m)

Pile bending moment

Test 4 (VL=6.5%)
Test 7 (L/H=0.76)
Test 10 (free-head, X=6m)
Test 13 (fixed-head, X=6m)
Test 16 (X=10m)

Pile head deflection

Test 5 (X=9m)
Test 8 (L/H=1.04)
Test 11 (free-head, X=10m)
Test 14A (fixed-head, X=10m)

Figure 5.24 Long-term to short-term ratio of pile responses for all tests
(Tests 3, 4, 5, 6, 7, 8, 9, 10, 11, 12, 13, 14A, 14B and 16)

200

Chapter 5 Effects of Tunneling on Single Piles

Figure 5.25 Comparison of (a) ovalisation of tunnel lining by Ran (2004); and
(b) over-cut of tunnel in the present study

Figure 5.26 Simplified tunnel lining ovalisation with time (not to scale) (Test 1)
(after Ran, 2004)

201

Chapter 5 Effects of Tunneling on Single Piles


Distance from tunnel central line (m)
0-20

-18

-16

-14

-12

-10

-8

-6

Ground surface
-4

-2

Depth below gr ound level (m)

-5

-10

Tunnel depth 15m

-15

60 ( mm)

50

40

-20

30

20

(a)

10

-25

Distance from tunnel central li ne (m)


0-20

-18

-16

-14

-12

-10

-8

-6

Ground surface
-4

-2

Depth below ground level (m)

-5

-10

1 30

(mm)

-15
1 10
90
70

-20

50
30

(b)

10

-25

Figure 5.27 Development of subsurface soil movements at (a) 2 days and (b) 720 days
after tunnel excavation (after Ran, 2004)

202

Chapter 5 Effects of Tunneling on Single Piles

Distance from tunnel centre-line (m)


0

10 12

14 16 18 20 22 24 26

0
-0.1
-0.2

inflection
point, 'i"

-0.3

Sv/Smax

-0.4
-0.5
-0.6
-0.7
-0.8

Test 3, 2 days (VL=3%)


Test 3, 2 days, Gaussian curve (VL=3%)
Test 3, 720 days(VL=3%)
Ran (2004), 2 days (VL=2%)
Ran (2004), 720 days (VL=2%)
Ran (2004), 2 days, Gaussion Curve (VL=2%)

-0.9
-1

Figure 5.28 Variation of surface soil settlement troughs with tunnel deformation
Time (days)
0

180

360

540

720

900

1080 1260 1440

0
-0.1
-0.2

Sv/ Smax

-0.3
-0.4

Inward tunnel deformatio (Test 3, VL=3%)

-0.5
-0.6

Outward tunnel deformation (Ran, 2004, VL=2%)

-0.7
-0.8
-0.9
-1

Figure 5.29 Variation of maximum surface soil settlement at tunnel central line with
tunnel deformation

203

Chapter 5 Effects of Tunneling on Single Piles

Settlement (mm)
0

20

40

60

80

100

120 140

160 180

200

-5

Depth (m)

-10

-15

T unnel

Free-field soil settlement

-20

Inward tunnel def ormation, Test 3, 2 day s

Inward tunnel def ormation, test 3, 720 day s

-25

Outward tunnel def ormation, Ran (2004), 2 day s

-30

Figure 5.30 Variation of vertical soil settlement at pile location with tunnel
deformation
Lateral deflection (mm)
-40

-30

-20

-10

10

20

30

40

50

-5

Depth (m)

-10

-15

T unnel

-20

Inward tunnel deformation, Test 3, 2 days (VL=3%)

-25

Inward tunnel deformation, Test 3, 720 days (VL=3%)


Outward tunnel deformation, Ran (2004), 2 days (VL=2%)
Outward tunnel deformation, Ran (2004), 720 days (VL=2%)

-30

Figure 5.31 Variation of soil deflection at pile location with tunnel deformation
204

Chapter 5 Effects of Tunneling on Single Piles

Axial Force (kN)


0

100

200

300

400

500

600

700

800

-5

Depth (m)

-10

-15

T unnel

Approximate
Neutral Plane

-20

-25

Inw ard tunnel deformation, Test 3, 2 days


Inw ard tunnel deformation, Test 3, 720 days
Outw ard tunnel deformation, Ran (2004), 2 days

-30

Figure 5.32 Variation of pile axial force with tunnel deformation

Pile head settlement (mm)

30
25
20
15
10
5
0
Inw ard Tunnel Deformation
(Test 3)

ST

Outw ard Tunnel Deformation


(Ran,2004)

LT

Figure 5.33 Variation of pile head settlement with tunnel deformation

205

Chapter 5 Effects of Tunneling on Single Piles

Bending Moment (kNm)


-350

-250

-150

-50

50

150

250

350

-5

Depth (m)

-10

-15
T unnel

-20

Inw ard tunnel deformation, Test 3, 2 days


-25

Inw ard tunnel deformation, Test 3, 720 days


Outw ard tunnel deformation, Ran (2004), 2 days
Outw ard tunnel deformation, Ran (2004), 720 days

-30

Figure 5.34 Variation of pile bending moment with tunnel deformation

Maximum pile bending moment (kN)

300
250
200
150
100
50
0
-50 0
-100

180

360

540

720

900

1080

1260

1440

Time (days)

-150
-200
-250
-300

Inward tunnel deformation, Test 3


Outward tunnel deformation, Ran (2004)

Figure 5.35 Variation of tunnelling-induced maximum pile bending moment


over time for different tunnel deformation

206

Chapter 5 Effects of Tunneling on Single Piles

Lateral deflection (mm)


-5

10

15

20

-5

Depth (m)

-10

-15

T unnel

-20
Pile head deflection, 2 days
Pile head deflection, 720 days

-25
Ran (2004), Pile head deflection, 2 days
Ran (2004), Pile head deflection, 720 days

-30

Figure 5.36 Variation of pile deflection with tunnel deformation


4
3.5
3

LT/ST ratio

2.5

Long-term
effect

2
1.5
1
0.5
0
Pile bending moment Pile head settlement

Pile head deflection

Inward tunnel deformation, Test 3


Outward tunnel deformation, Ran (2004)

Figure 5.37 Long-term to short-term ratio of pile responses over time for different
tunnel deformation

207

Chapter 6 Effects of Tunnelling on Pile Groups

CHAPTER SIX

EFFECTS OF TUNNELLING
ON PILE GROUPS

6.1

INTRODUCTION

The results on the effects of tunnelling on single piles presented in Chapters 4 and 5
provide valuable insights. As piles are commonly installed in groups in practice, the
centrifuge model study on single piles is extended to pile groups in the same soil
conditions. The study aims to address the following issues:

1. Effects of tunnelling on a floating pile group as compared to single floating


pile.
2. Effects of tunnelling on an end-bearing pile group as compared to single
end-bearing pile with capped-head and fixed-head conditions.
3. Effects of size of pile group due to tunnelling.
4. Effects of tunnelling on pile groups with capped-head and fixed-head
conditions.
5. Effects of time on pile group responses in soft clay.

The test procedure for pile group is similar to that for single piles and the
schematic plan and elevation views for the five pile group tests (Tests PG1 to 5) are
shown in Table 6.1. In conjunction with the results of Tests 3 and 16 reported in the
previous chapter, the responses of a floating capped-head 2-pile group obtained from

208

Chapter 6 Effects of Tunnelling on Pile Groups

Test PG1 are compared with a single floating free-head pile. Test PG2 investigates the
responses of an end-bearing capped-head 2-pile group and the results are compared
with those of a single end-bearing free-head pile obtained from Tests 10 and 11. Test
PG 3 evaluates the responses of an end-bearing fixed-head 2-pile group and the results
are compared with those of a single end-bearing fixed-head pile obtained from Tests
13 and 14.

In order to study the effects of size of pile group due to tunnelling, the
behaviours of 2-pile group and 6-pile group are compared (for capped-head - Tests
PG2 and 4; for fixed-head - Tests PG3 and 5). Finally, the effects of tunnelling on pile
groups with capped-head and fixed-head conditions are investigated (for 2-pile group Tests PG2 and 3; for 6-pile group - Tests PG4 and 5).

6.2

FLOATING PILE GROUP

Test PG1 was carried out on capped-head floating 2-pile group with the front pile at 6
m and the rear pile at 10 m from the tunnel centreline. The centre-to-centre pile
spacing is hence approximately three times pile diameter, similar to that recommended
by BS8004 (BSI, 1986). The results of Test PG1 are compared with those of single
free-head floating piles Tests 3 and 16 (presented in Chapters 4 and 5). Unfortunately
it was not possible to conduct single capped-head floating pile in the present centrifuge
model setup. In the capped-head pile groups, the cap is connected to the individual pile
heads at about 200 mm above the ground level to avoid interaction between the pile
cap and the soil following the approach adopted by Bransby and Springman (1997).

209

Chapter 6 Effects of Tunnelling on Pile Groups

Leung et al. (2003) established that the individual pile responses are similar if the 2
piles are aligned parallel to the induced soil movement direction. As such, the 2 piles
are aligned perpendicular instead of parallel to the tunnel for Test PG1.

6.2.1 Induced Axial Force and Settlement


Figure 6.1 shows the tunnelling-induced axial force for the front and rear piles in a
capped-head 2-pile group. It is observed that the front pile in the pile group
experiences a larger pile axial force than that of the rear pile which is further away
from the tunnel. The results reveal that the trend of axial load transfer along the front
and rear piles are similar. Figure 6.2 compares the short-term (ST, 2 days after tunnel
of excavation) and long-term (LT, 720 days) axial load transfer of the front pile of the
2-pile group located at 6 m from tunnel centre and that of a single free-head pile
located at the same distance. It is evident that the magnitude of the induced axial forces
for capped-head pile is significantly reduced. Similar observation is noted when
comparing the axial pile load profiles of the rear pile of the 2-pile group located at 10
m from tunnel centre and that of a single pile at the same location, see Figure 6.3. As
the front and rear piles are connected by a rigid pile cap, the 2 piles are forced to act in
unison when subjected to tunnelling induced soil movement. The induced axial force
on the front pile, which experiences larger soil movement as presented in chapter 4,
would be moderated by the rear pile via the rigid pile cap. The reduction of negative
skin friction for the front pile is about 33 % (ST) and 28 % (LT) of that of single pile at
the same location. On the other hand, the corresponding reduction in negative skin
friction for the rear pile is 33 % (ST) and 27 % (LT). Such positive pile group effect is

210

Chapter 6 Effects of Tunnelling on Pile Groups

consistent with the findings by Kuwabara and Poulos (1989) and Loganathan et al.
(2001).

Figure 6.4 compares the tunnelling-induced pile head settlement for the front
and rear piles of the capped-head 2-pile group and that of corresponding single piles.
The front pile settlement is 53% and 56% smaller than that of corresponding single
piles at 6 m from tunnel centre, in the short-and long-terms, respectively. However, it
is worth noting that the rear pile settlement is slightly larger than that of the
corresponding single pile. The rear pile settlements are 2.2 mm (short-term) and 7.1
mm (long-term), as compared to the 2 mm (short-term) and 6.9 mm (long-term) for the
corresponding single pile. This can be explained by the interaction between the pile
and pile cap, in which the front pile in the rigidly connected pile group is moderated by
the rear pile and hence the pile settlement is smaller than that of the corresponding
single pile. However, as the rear pile is being dragged by the front pile via the rigid
pile cap, the rear pile settlement becomes slightly higher than that of a single pile at the
same location. It is noted that since the front pile settlement is larger than that of rear
pile, the pile cap has tilted slightly. This is evident by the measured pile cap deflection
which will be further discussed later.

6.2.2 Induced bending moment and deflection


Figure 6.5 shows the tunnelling-induced pile bending moment for the front and rear
piles in a capped-head 2-pile group. It was postulated by Ong (2005) that a relaxation
in the fixity of the pile cap may have occurred with the fabricated rigid pile cap used in
his study of excavation-induced soil movement on piles. Hence, for this study, a new

211

Chapter 6 Effects of Tunnelling on Pile Groups

and improved rigid pile cap has been fabricated to include double layers of bolts
instead of one as used by Ong (2005). With this new rigid pile cap, an almost perfect
fixity can be provided between the pile and the pile cap.

The results shown in Figure 6.5 demonstrate that positive and negative bending
moments were induced for both the front and rear piles. The maximum pile bending
moment occurs approximately at the tunnel spring line for the front pile while the
magnitude of maximum positive or negative bending moments for the rear pile is
similar. The rear pile generates a larger negative bending moment at the pile cap level
as compared to the front pile. As before, since the pile cap is tilted, the pile-cap-pile
interaction causes the front pile responses to be moderated by the rear pile via the rigid
connecting pile cap. Thus, the upper bending moment profiles are different between
the front and rear pile because backward dragging force is exerted on the front pile by
the rear pile through the tie beam. This finding is consistent with the observation of
pile group due to excavation-induced soil movements as reported by Leung et al. (2003)
and Ong et al. (2009).

Figures 6.6 and 6.7 compare the front and rear pile responses with the
corresponding free-head single piles at the same location. Although the trend of the
pile bending moment profiles of 2-piles group and the corresponding free-head single
piles are similar, both front and rear piles of the capped-head pile group demonstrate
negative bending moment due to the presence of pile cap while zero bending moment
is recorded for single free-head pile whose head can move freely. The results further
reveal the shadowing effects of the front pile over the rear pile from the soil movement,
resulting in a smaller measured positive bending moment along the rear pile. On the

212

Chapter 6 Effects of Tunnelling on Pile Groups

contrary, Loganathan et al. (2001) observed that the bending moment profiles for piles
in a group and single free-head pile are almost the same, except for a small difference
at the pile cap location due to fixity condition.

Since the pile cap is tilted due to differential pile settlement, the pile deflection
profiles for the front and rear piles in a capped-head 2-pile group (Test PG1) is
different, as shown in Figure 6.8. The deflection of each individual pile head is
identical as the pile groups are capped. The pile cap deflection is 1.9 mm in the shortterm and increases continuously with time until the end of the tests with a final
deflection of 3.5 mm. In addition, the pile deflection profiles are slightly different for
the front and rear piles. The front pile, which is subjected to a larger soil movement, is
dragged back towards the rear pile due to the connecting pile cap. For the rear pile, the
lateral soil movement on the pile is smaller but the rigid pile cap drags the rear pile
deflecting towards the tunnel resulting in a different pile deflection profile. The
observed lateral pile deflection profiles are similar to those reported by Leung et al.
(2003) on pile groups subject to excavation-induced soil movement. This demonstrates
that the interaction of pile-cap-pile has a significant effect on pile group for excavation
and tunnelling works.

Figures 6.9 and 6.10 compare the pile deflection profile of the front and rear
pile, respectively, with their corresponding single piles. Since the two piles at various
distances are being capped by a rigid pile cap, considerable interaction between two
piles is expected through the rigid pile cap. The results illustrate that the magnitudes of
the front pile deflection of 3.2 mm (short-term) and 6.1 mm (long-term) for the 2-pile
group (Test PG1) are smaller than that of a corresponding single pile at 6 m (Test 3)

213

Chapter 6 Effects of Tunnelling on Pile Groups

from tunnel centre with pile deflection of 5 mm (short-term) and 12.1 mm (long-term).
However, they are larger that that of a corresponding single pile at 10 m (Test 16) from
tunnel centre with pile deflection of 2.8 mm (short-term) and 5.4 mm (long-term).
Nonetheless, the magnitude of the deflection of the capped-head pile group is smaller
than the average of the deflections of the two single piles at 6 m and 10 m from the
tunnel centre. This data further suggests that pile cap plays a vital role in the pile group
in resisting the lateral deflection induced by soil movement. In addition, the results
indicate that the pile deflection profile in the capped pile group is different from that of
free-head single pile, especially at the pile head. The front pile in the 2-pile group,
which is subjected to larger soil movements, is restrained at the pile head via the rigid
pile cap and hence the upper portion of the front pile is being dragged back as oppose
to the rigid body translation of pile deflection in the corresponding single pile due to
free head condition. On the other hand, the deflection profile of the rear pile in the 2pile group is similar to the corresponding single pile, which the pile tends to bend
towards the tunnel due to the unloading process of tunnel excavation.

Figures 6.12 and 6.13 show the ratio of pile responses for a single pile over pile
group ratio for the front and rear piles in capped-head pile and the corresponding
single pile. Generally, the group effect is beneficial to the front pile in all aspects
except for the negative bending moments. Figure 6.12 provides evidence that the pile
group effect is significant for the front pile, which is subjected to larger soil movement,
particularly bending moment, settlement and deflection. The induced substantial
bending moment at the pile head in the 2-pile group is mainly due to the restraint
provided by the pile cap. However, for the rear pile, the group effect is only beneficial
in axial force and positive bending moment. Adverse effects are observed in the

214

Chapter 6 Effects of Tunnelling on Pile Groups

negative bending moment, pile head settlement and pile head deflection. The pile
group effect can be attributed to the significant pile-cap-pile interaction as the
individual piles in a group are forced to act in unison when subject to different
magnitudes of soil movement (Ong, 2005). As such, the front pile would be moderated
by the rear pile via the rigid pile cap. In addition, the shadowing effect of the front pile
on the rear pile reduces the detrimental effects experienced by the rear pile, thus
resulting in an overall positive effect for the pile group.

6.3

END-BEARING PILE GROUP

6.3.1 Capped-Head
The configuration of Test PG2 is similar to that of Test PG1 except the pile length is
increased from 22 m to 27.5 m and the piles are rested on the base of the strong box to
simulate end-bearing piles (see Table 6.1).

Figure 6.14 shows the tunnelling-induced axial force for the front piles in a
capped-head end-bearing 2-pile group. The results show that the front pile experiences
larger pile axial force than that of the rear pile, similar to that observed for the cappedhead floating 2-pile group. The results of the respective free-head single piles at the
same location from Tests 10 and 11 are compared with the 2 piles from Test PG2 in
Figures 6.15 and 6.16. Although the induced axial pile profiles are similar for single
piles and the 2-pile group, the magnitude of the induced axial forces for the cappedhead piles is significantly reduced in both short-and long-terms, respectively. This
reduction is caused by the presence of pile cap, as explained in Section 6.2.1. Unlike
the floating pile group (PG1) which experiences long-term pile settlement of 7.5 mm
215

Chapter 6 Effects of Tunnelling on Pile Groups

and 4.8 mm for the front and rear pile, respectively, the measured pile settlement of
end-bearing pile (PG2) is negligible as the pile tip is rested on very stiff ground.

Figure 6.17 shows the tunnelling-induced pile bending moment for both front
and rear piles in a capped-head end-bearing 2-pile group. The results demonstrate
triple curvature in the induced bending moment, whereby negative bending moments
are induced at the pile upper and lower portions of the pile shaft, whilst positive pile
bending moment occurs approximately at the tunnel spring line. Three major trends
have been observed in this study. Firstly, the bending moment profile for the front and
rear pile are different, especially at the upper part of the pile. Secondly, the induced
bending moment increases over time for both front and rear piles. Finally, the induced
maximum positive bending moments are always larger than the maximum negative
bending moments.

The bending moment profiles of end-bearing single piles in Tests 10 & 11.
(presented in Chapter 5) and the pile group test (Test PG2) are plotted in Figures 6.18
and 6.19. The trend of pile bending moment profiles of Test PG2 and those of the
corresponding free-head single piles are similar. However, positive bending moment is
induced near the tunnel axis and negative bending moment induced at the upper and
lower parts of the pile for the pile group. Owing to fixity for capped pile in Test PG2,
there is a significant difference between the capped-head pile group and free-head
single pile at the pile top. For the capped-head pile group, both the front and rear piles
experience negative bending moment due to presence of the pile cap, while zero
bending moment is recorded for the single free-head pile as the pile head can move
freely without any restraint. This is consistent with the capped-head floating pile

216

Chapter 6 Effects of Tunnelling on Pile Groups

responses in Test PG1. When compared to the corresponding single pile, the
magnitude of maximum induced positive bending moment for capped-head piles is
reduced by about 23% in the short-term and approximately 38 to 46% in the long-term,
demonstrating the positive pile group effect.

Figure 6.20 shows the tunnelling-induced pile deflection profiles for the front
and rear piles in a capped-head end-bearing 2-pile group (Test PG2). The deflection of
each individual pile head is essentially identical as the pile groups are capped. The
induced pile deflection is 1.9 mm in the short-term and increases to 3.5 mm in the
long-term. The pile deflection profiles along the upper portion is similar to the cappedhead floating 2-pile group (Test PG1) due to the capping and dragging effect as
explained in Section 6.2.2. As expected, the pile deflection profile at the mid-pile shaft
for end-bearing pile group is different from the floating pile group. The induced pile
deflection profiles shown in Figure 6.20 demonstrate the pushing of the mid-pile shaft
away from the tunnel due to underlying sand layer that restrains the pile toe movement
and lateral soil movement at pile head, as explained in Section 5.2.1.

Figures 6.21 and 6.22 compare the pile deflection of the front and rear pile with
their corresponding single piles. Two main findings are observed. Firstly, the pile
deflection profile along the upper portion for front pile is very different from the single
free-head piles due to presence of the pile cap. The front pile is dragged back by rear
pile via rigidly connected pile cap. For the rear pile, the profile is similar to the single
pile, which moves in rigid body translation mode. Secondly, the magnitude of the front
pile head deflection for the 2-pile group (Test TG2) of 1.9 mm (short-term) and 3.5
mm (long-term) is smaller than that of a corresponding single pile (Test 10, X = 6 m)

217

Chapter 6 Effects of Tunnelling on Pile Groups

of 2.8 mm (short-term) and 6 mm (long-term). The rear pile deflection is bigger when
compared to the corresponding single pile (Test 11, X = 10 m) of 1.5 mm (short-term)
and 3 mm (long-term) due to pile group effect (See Figure 6.23). Nevertheless, the
deflection of the pile group (Test PG2) again lies in between and smaller than the
average of deflections of two single piles at the same location, similar to the findings
for the floating 2-pile group (Test TG1).

Figures 6.24 and 6.25 show the single pile over pile group ratio for the front
and rear piles in capped-head condition with the corresponding single pile. The results
evidently reveal that the group effect is beneficial to the front in all aspects, except
negative bending moment which is induced due to pile cap condition. On the other
hand, positive group effects are only observed in the axial force and bending moment
for the rear pile. The pile negative bending moment and pile head deflection are found
to increase when the piles are capped in a group. This is consistent with the findings
for the floating capped-head 2-pile group condition.

6.3.2 Fixed-Head

Test PG3 was conducted with the pile head totally fixed in position having zero
vertical or lateral movements. This simulates the condition where the pile cap is tied
with a rigid pile cap and very strong/stiff ground beams. The test results are compared
with corresponding fixed-head single pile (Tests 13, 14A). However, as noted in
Chapter 5, the pile axial force response in Test 14A is not recorded.

218

Chapter 6 Effects of Tunnelling on Pile Groups

Figure 6.26 shows the tunnelling-induced axial force for the front and rear piles
of a fixed-head end-bearing 2-pile group. As discussed in Chapter 5, the upper pile
shaft of a fixed-head end-bearing pile is subjected to tensile force while compression
force is induced along the lower pile shaft. Since Test PG3 is a 2-pile group, the front
pile experiences higher tensile and compression forces than that of rear pile as it is
closer to the tunnel. When compared to single pile (Test 13), it is noted that the front
pile axial force responses are smaller than that of a single pile due to the group effect
of the pile, as shown in Figure 6.27. It is noted that pile axial force is not measured in
Test 14A, thus the comparison of rear pile in Test PG3 with a single pile at the same
location is not possible.

Similarly for pile bending moment responses (Figure 6.28), the rear pile
bending moment is smaller than that of front pile as the lateral soil movement acting
on the pile is much smaller. The trend of front and rear pile bending moment profiles
are similar, suggesting that pile-cap-pile interaction is less severe. Owing to the group
and shadowing effects, the rear pile experiences smaller bending moment.

Figures 6.29 and 6.30 compare the front and rear pile bending moments with
the corresponding single piles (Tests 13 & 14A) at the same location. The results
indicate similar pile bending moment profiles for all cases with a smaller maximum
bending moment for the 2-pile group due to shadowing effects.

The single pile over pile group ratio for the front and rear piles in fixed-head
condition (Test PG2) with the corresponding single piles (Tests 13 & 14A) are plotted
in Figures 6.31 and 6.32. The comparisons summarise the positive effects of the pile

219

Chapter 6 Effects of Tunnelling on Pile Groups

group, with improvement ratio of 1.05 to 1.5. In general, the reduction in pile axial
tensile force and pile head bending moment are more significant than that in pile axial
compression force and mid-pile shaft bending moment. This suggests that the restraint
due to fixed-head condition is dominant.

6.4

PILE GROUP SIZE

6.4.1 Capped-Head
In order to evaluate the effect of pile group size, capped-head end-bearing 6-pile group
(Test PG4, 2x3 configurations) was performed and compared with capped-head endbearing 2-pile group (Test PG2, 1x2 configuration). Following the finding by Leung et
al. (2003) on excavation-induced soil movement on pile group, it is expected that a 2x2
pile group would have similar behaviour with those of a 1x2 pile group (Test PG2)
except that the responses might be smaller due to greater number of piles. It is thus
decided to investigate the 2x3 configuration (Test PG4) as the effect of pile group size
is expected to be more significant and hence providing further insight on pile group
size effect. In Test PG4, two rows of piles were arranged in three columns at 6 m
(front), 10 m (middle) and 14 m (rear) perpendicular to the tunnel. Owing to the
symmetrical arrangement of the 2x3 pile group configuration, only one of each of the
three pairs of was instrumented. The minimum boundary clearance for this pile group
is 12 m to the edge of the container in the direction perpendicular to the tunnel, and 8
m to the edge of container in the direction parallel to the tunnel. Randolph and Wroth
(1978) established that it is reasonable to assume a maximum shear stress at the pilesoil interface and the shear stress decreases with increasing distance from the pile.
Shen (2008) reported a minimum boundary clearance of 8 m would not cause
220

Chapter 6 Effects of Tunnelling on Pile Groups

significant container boundary effect. Thus the container boundary effect should not be
significant for the present study. As the pile settlement for end-bearing pile is
negligible as presented in Chapter 5, the pile settlement of Test PG4 will not presented
here.

Figure 6.33 shows the tunnelling-induced axial force for the front, middle and
rear piles in the fixed-head end-bearing 6-pile group (Test PG4). The results reveal that
as the distance between the pile and tunnel increases, the pile axial force decreases as
the magnitude of soil movement becomes smaller away from the tunnel. In the shortterm, the induced maximum drag load reducing from 321 kN (front pile) to 161 kN
(middle pile), and finally to 109 kN for the rear pile. The axial forces increase over the
time for all cases. The measured maximum drag load in the long-term are 510 kN
(front pile), followed by 261 kN (middle pile) and 186 kN (rear pile). The observed
trend appears consistent with the corresponding single piles reported in Chapter 5. This
finding suggests that pile-cap-pile interaction on the pile axial force is not significant.
The changes of the induced axial forces are mainly affected by the magnitude of the
tunnelling-induced soil movement at various locations from the tunnel.

To further compare the group effects, the results of front and middle piles in the
6-pile group are compared with the corresponding capped-head end-bearing 2-pile
group (front and rear piles) in Figures 6.34 and 6.35. In general, the piles in a bigger
pile group would be subjected to lower induced axial forces. This is reasonable as a
bigger pile group is likely to provide more shadowing effect on the piles behind the
front piles and reinforcing effects to other piles in the group. The reduction in the
maximum drag load for the front pile of the 6-pile group (Test PG4) is about 10% to

221

Chapter 6 Effects of Tunnelling on Pile Groups

11% while the corresponding reduction for the middle pile is slightly higher of
approximately 13% to 17%.

Figures 6.36(a) and (b) show the tunnelling-induced pile bending moment for
the front, middle and rear piles in a capped-head end-bearing 6-pile group, in the shortand long-term, respectively. The bending moment profiles generally display triple
curvatures with negative bending moments at the upper and lower portions of the pile
body, whilst positive pile bending moments occur approximately at the tunnel spring
line. It is observed that the middle and rear piles have similar induced bending moment
profiles while the front pile shows a markedly different profile. Since the front, middle
and rear piles are connected via a rigid pile cap, the front pile is being dragged back
while the middle and rear piles being pulled by front pile toward the tunnel through the
connecting rigid pile cap.

The pile-cap-pile interaction would moderate the induced pile bending


moments among the piles within a pile group, as a result, the induced pile bending
moments in the middle row is smaller than that of rear row. This is contrary to the
induced lateral soil movements, in which the corresponding lateral soil movement on
the middle row of piles is larger than the corresponding lateral soil movement on the
rear row of pile. The observation is consistent with finding reported by Leung et al.
(2003), which suggesting that part of the bending moments of the pile in the middle
row is transferred to the rear piles due to the interaction through pile cap. Similarly, the
long-term pile bending moments are showing the consistent behaviours except the
magnitudes are increased over time.

222

Chapter 6 Effects of Tunnelling on Pile Groups

The comparison of bending moment responses of respective piles at the same


location for the 2-pile group (Test PG2) and 6-pile group (Test PG4) are plotted in
Figures 6.37 and 6.38. The shape of bending moment profile in the front pile is similar
for both 2-pile and 6-pile group. In addition, the bending moment profiles of rear pile
in 2-pile group are similar to that of middle row and rear row of piles in 6-pile group.
Furthermore, the reduction in maximum bending moment in 6-pile group (Test PG4)
as compared to 2-pile group (Test PG2) is more significant for the middle row piles
than the front row piles in 6-pile group, which is illustrated earlier that the part of
bending moments in the middle row piles are shared by the rear row of piles. The data
demonstrated that the induced bending moment is reduced by 21% to 30% for the front
row of piles and by a considerable reduction for the middle row of piles at 27% to 54%
from the above finding. As such, the pile-cap-pile interaction is more significant in
lateral pile responses for a bigger pile group. With more piles in a bigger pile group,
the larger pile-cap-pile interaction and shadowing of front piles over rear piles
significant affect the performance of the pile group.

Figures 6.39 (a) & (b) show the tunnelling-induced pile deflection profiles for
the front, middle and rear piles in a capped-head 6-pile group (Test PG4) in the shortand long-term, respectively. Since all 6 piles at different position are capped by a rigid
pile cap, the pile head deflection is forced to act in unison and thus the lateral
deflection at pile head moves in a translation mode with the same deflection magnitude.
The measured pile head deflection increases from 1.3 mm (short-term) to 2.7 mm
(long-term). The data reveal that the pile deflection profiles demonstrate some
differences between the front, middle and rear piles. For the front pile, the pile head is
dragged back towards the rear pile similar to the Test PG2 (see Fig. 6.40). Moreover,

223

Chapter 6 Effects of Tunnelling on Pile Groups

the middle pile in the 6-pile (PG4) group shares the same shape of the deflection
profile with the rear pile in the 2-pile group (PG2) (Fig. 6.41). The comparison of pile
head deflection for different pile groups is shown in Figure 6.23. The pile head
deflection reduces from 1.9 mm to 1.3 mm (short-term) and 3.5 mm to 2.7 mm (longterm) when the pile group size increases from 2 to 6 piles. This further demonstrates
the positive effect of pile group increases with group size.

It is acknowledged that the pile head deflection is directly proportional to the


bending moment. In other words, if the pile head deflection increases, so would the
maximum negative bending moment at the pile head. The consistent responses of
deflection with bending moment observed in the present study show that the improved
pile cap has served its intended function of providing strong pile cap fixity, with
considerably less pile cap relaxation reported by Ong (2005).

Figures 6.42 and 6.43 show the ratio of the pile group responses of the 2-pile
group over 6-pile group for the respective piles at 6 m (front) and 10 m (middle piles in
6-pile group and rear pile in 2-pile group) in order to evaluate the effect of pile group
size. The pile group effect is positive if the ratio exceeds one as there is a larger
reduction in pile responses for the larger pile group. The measured ratio of between 1.1
and 2.1 clearly shows that a larger pile group would experience a greater positive
group effect.

224

Chapter 6 Effects of Tunnelling on Pile Groups

6.4.2 Fixed-Head
Test PG5 with a fixed-head end-bearing 6-pile group (2x3 configurations) was
conducted and compared with the fixed-head end-bearing 2-pile group (Test PG3, 1x2
configurations). The configurations of Test PG5 are similar to Test PG4 except for the
pile head condition.

Figure 6.44 shows the tunnelling-induced axial force for the front, middle and
rear piles for the fixed-head end-bearing 6-pile group (Test TG5). It is observed that
the shapes of pile axial force profiles are similar to those of the fixed-head 2-pile group.
The results reveal that as the distance of pile from the tunnel increases, both the
induced pile tensile and compression forces reduce. This is because a smaller soil
movement is induced at location further away from the tunnel. It is also noted that the
pile-cap-pile interaction in a totally fixed-head condition is not significant. The bigger
pile group is able to provide more significant reinforcing and shadowing effects. The
results of front and middle piles in the 6-pile group are compared with the
corresponding piles in 2-pile group (Test PG3) in Figures 6.45 and 6.46. Similar to the
capped-head, the fixed-head pile group also exhibit the same positive group effect.

For the front pile, the reduction in the maximum tensile force in the short-term
for the 6-pile group (Test PG5) is about 25% with a smaller reduction of 14% for the
maximum compression force when compared with the 2-pile group (Test PG3). On the
other hand, the reduction in tensile and compression force for the middle pile is about
the same, i.e. 13% to 14% when the size of pile group increases from 2 to 6 piles.

225

Chapter 6 Effects of Tunnelling on Pile Groups

Figures 6.47(a) and (b) show the tunnelling-induced pile bending moment for
the front, middle and rear piles in a fixed-head end-bearing 6-pile group (Test PG5) in
the short-and long-terms, respectively. The results reveal that the bending moment
profiles display triple curvatures with negative bending moments along the upper and
lower portions of the pile body, whilst positive pile bending moment occur
approximately at the tunnel spring line. The shape of profile is similar to the fixedhead end-bearing 2-pile group (Test PG3) as well as fixed-head single piles (Tests 13,
14A & 14B). The results reveal that the absolute magnitude of the negative bending
moment at the pile head is larger than the positive bending moment at mid-pile shaft
suggesting a significant fixed-head effect. As before, the induced maximum positive
bending moment reduces from the front to the middle finally to the rear pile as the
lateral soil movement reduces with increasing distance between the pile and tunnel. In
addition, the front row of pile provides shadowing to the trailing middle and rear rows
pile during tunnel excavation, resulting in smaller bending moment on the trailing piles.

A comparison of bending moment profiles of respective piles at the same


location for the 2-pile group (Test PG3) and 6-pile group (Test PG5) are shown in
Figures 6.48 and 6.49. As expected, a bigger pile group provides more shadowing and
reinforcing effect to other piles within the group, resulting an average smaller pile
response due to tunnelling. The comparison reveals that the reduction in maximum
negative bending moment at the pile head is more significant than the reduction in
maximum positive bending moment at mid-pile shaft for the 6-pile group. The data
illustrates that in the short-term; the induced negative bending moment reduces by
34% and 41% for the front and middle piles, respectively. On the other hand, the
corresponding reduction of 28% and 29% is smaller for the positive bending moment.

226

Chapter 6 Effects of Tunnelling on Pile Groups

Figures 6.50 and 6.51 show the ratio of pile responses of the 2-pile group over
6-pile group for the respective piles at 6 m (front) and 10 m (middle piles in 6-pile
group and rear pile in 2-pile group). It is noted that the positive group effect is
generally higher in pile bending moment than axial force, being between 1.02 and 1.33
for axial force, between 1.3 and 1.7 for bending moment.

6.5

PILE CAP CONDITIONS

6.5.1 2-Pile Group

The effect of fixity between the pile and pile cap has been studied by many researchers.
However, the effect of fixity of pile cap itself has been rarely been studied before. In
reality, the behaviours of pile group heavily depend on the pile cap fixity. Hence, Tests
PG2 (capped-head) and PG3 (fixed-head) have been conducted to study the effects of
tunnelling on 2-pile groups with capped-head and fixed-head condition.

Figures 6.52 and 6.53 show the tunnelling-induced axial force respectively for
the front and rear piles of a fixed-head and capped head end-bearing 2-pile groups. The
most distinct difference observed is that both compression and tensile forces are
induced in fixed-head pile group (PG3), while only compression force (drag load) is
induced in capped-head pile group (PG2). This is due to the total restraint provided
by the pile head in fixed-head condition, in which no vertical or lateral pile head
movement is allowed, as explained in Chapter 5. As tensile force is induced along the
upper pile shaft due to total pile cap fixity in Test PG3, there is a reduction in the
maximum drag load of about 20% in the front pile and about 15% in the rear pile. Thus,

227

Chapter 6 Effects of Tunnelling on Pile Groups

Figures 6.56 and 6.57 show the induced pile responses of capped-head pile over fixedhead pile ratio for the front and rear piles. The figures reveal that a fixed-head is
beneficial for induced pile axial force, or reduction in maximum drag load, with the
ratio of axial force of capped-head pile over fixed-head pile of 1.15 to 1.24.

The induced 2-pile group bending moments in capped-head (Test PG2) are
compared with those in fixed-head (Test PG3) in Figures 6.54 and 6.55. For the front
pile, it is noted that the pile bending moment profiles for the 2 pile cap conditions are
not similar, particularly at the pile head. In fixed-head condition (Test PG3), the profile
is very much similar to those single fixed-head piles with the maximum bending
moment occurring at the pile head, but it is different for capped-head condition (Test
PG2) due to tilting of cap resulting in a downward shift of the location of maximum
bending moments. This is probably due to the different pile cap fixity condition,
whereby capped-head demonstrates a significant pile-cap-pile interaction while for
fixed-head condition, the pile-soil-pile interaction is dominant and thus suggesting that
pile-cap-pile interaction is less severe as the cap is totally fixed in the position. Owing
to total restraint at the pile head, the induced negative bending moment of fixed-head
pile group is much larger than that of capped-head pile group. The capped-head pile is
allowed to deflect freely and interact with other piles in a group and thus inducing a
smaller negative bending moment compared to the fixed-head pile. On the other hand,
the profiles of rear pile bending moments for capped- and fixed-pile are similar. It is
also worth noting that for fixed-head, the negative bending moment induced at pile
head is always larger than the positive bending moment induced at the mid-pile shaft
because the fixed-head condition has provided a very rigid restraint at the pile head
which dominates the lateral pile responses.

228

Chapter 6 Effects of Tunnelling on Pile Groups

As shown in Figures 6.56 and 6.57, the maximum positive and negative
bending moments of capped-head pile over fixed-head pile ratio for the front and rear
piles experiencing opposite trends. Generally, the induced positive pile bending
moments near mid-pile shaft is beneficial to a fixed-head pile group, with the cappedhead pile over fixed-head pile ratio of 1.06 to 1.28. Paradoxically, at the pile head, the
negative pile bending moment ratio of capped-head pile over fixed-head pile is less
than 1, at about 0.4 to 0.5 for front pile and 0.7 for rear pile. This indicates an adverse
effect on the fixed-head pile group, which can be explained by the totally fixed-head
condition, which restraints the pile movement and thus induces the high bending
moment. It is thus important to check the adequacy of the steel reinforcement for the
different pile cap fixity, i.e. it is critical at the pile head for fixed-head pile group and
at the mid-pile shaft for the capped-head pile group.

6.5.2 6-Pile Group


The results of 6-pile group, i.e. Tests PG4 (capped-head) and PG5 (fixed-head) are
compared in this section. Figures 6.58, 6.59 and 6.60 show the tunnelling-induced
axial force for the front, middle and rear piles of a fixed-head and capped head endbearing 6-pile groups, respectively. Similar to 2-pile group, compression and tensile
forces are induced in fixed-head pile (PG5). Likewise, the compression forces induced
in the fixed-head pile (PG5) are always smaller than that of capped-head pile (PG4)
with a reduction ranging from 10% to 22%. Nevertheless, the trade-off for such
reduction in drag load is that tensile force is induced near to the pile head due to totally
fixed head condition.

229

Chapter 6 Effects of Tunnelling on Pile Groups

Unlike axial force, the behaviour of bending moment is much more


complicated. The results of induced 6-pile group bending moments in capped-head
(Test PG4) are compared with those in fixed-head (Test PG5) in Figures 6.61, 6.62 and
6.63. The front pile (Fig. 6.61) exhibits different bending moment profiles for both
capped-head and fixed-head conditions, similar to the comparison of 2-pile group
discussed earlier. It is observed the upper portion of the front pile is being dragged
backward in Test PG4 (capped-head) while the profile of the front pile in Test PG5
(fixed-head) is similar to those single fixed-head piles. On the other hand, all of the
middle and rear row piles (Figs. 6.62 and 6.63) are noted to bend toward the tunnel
(similar to the front pile in the fixed-head pile group) regardless of pile cap conditions.
Despite the shape of bending moment profiles are similar in middle and rear row piles,
the comparison of capped- and fixed-head conditions illustrates a very different
behavior in transferring the induced bending movements within a bigger group of 6pile, as oppose to 2-pile group. Intuitively the induced lateral soil movement is largest
at the position of front pile, followed by middle and rear pile. It is thus expected that
the magnitude of the induced pile bending moment should follow a similar trend of
soil movements. Nevertheless, this trend is only observed in the fixed-head 6-pile
group (Test PG5) but not the capped-head 6-pile groups (Test PG4) because
moderation effect is negligible for fixed-head pile group and the profiles follow those
of single fixed-head pile with the magnitudes reduce with increase of tunnel-pile
distance. However, as discussed in Section 6.4.1, for Test PG4 (capped-head 6-pile
group), the induced pile bending moments in the middle row is smaller than that of
rear row. It is observed that since the pile cap is tilted slightly in the capped-head pile
group, the pile-cap-pile interaction would moderate the induced pile bending moments
among the piles within a pile group. As a result, part of the bending moments of the

230

Chapter 6 Effects of Tunnelling on Pile Groups

pile in the middle row is transmitted to the rear piles due to interaction through pile cap.
In addition, the middle row piles are shielded in the 6-pile group and experiencing the
least induced pile bending moment.

The above-mentioned behaviors can be further illustrated in Figures 6.64 &


6.65. It appears that the maximum positive and negative pile bending moments reduce
almost linearly from the front row piles to the middle row piles and finally to the rear
row piles in Test PG5 (fixed-head). This is consistent with the trend of observed lateral
soil movement which reduces when the distance to tunnel increases. This suggests that
in a fixed-head 6-pile group (Test PG5), the pile-soil-pile interaction is dominant and
largely depends on the soil movement instead of pile-cap-pile interaction. Nonetheless,
for the capped-head 6-pile group in Test PG4, the front pile registered the highest
induced maximum bending moment, followed by the rear pile and finally the smallest
bending moment was noted for the rear pile. This observation further confirms that
pile-cap-pile interaction is significant in capped-head fixity, especially when the pile
group increases from 2-pile (Test PG-2) to 6-pile (Test PG4). The interaction among
the piles in a bigger group would moderate, transfer and share the induced responses
among the piles particularly for the pile bending moments.

Figures 6.66, 6.67 and 6.68 show the capped-head pile over fixed-head pile
ratio. It is revealed that maximum positive bending moments induced at the mid-pile
shaft for capped-head (PG4) is always larger than that of fixed-head (PG5), thus
registering a positive effect for fixed-head pile with the ratio ranging from 1.1 to 1.2
for the front and middle piles and a higher ratio ranging from 1.9 to 2.6 for the rear pile
due to the moderating effects in capped-head pile, whereby the rear pile in capped-

231

Chapter 6 Effects of Tunnelling on Pile Groups

head is sharing some bending moment from middle pile while rear pile in the fixedhead induced the smallest bending moment due to distance effect. Paradoxically, the
induced maximum negative bending moments induced near to pile head are expected
to be larger in the fixed-head pile (PG5) than capped-head pile (PG4) due to the
restraint enforced at the pile head. Thus the ratio of capped-head over fixed-head is
about 0.5 to 0.75. However, it is contrary for the rear piles, whereby a ratio of 1.5 to
1.6 is observed. This means that the induced negative bending moment at capped-head
is larger than that in fixed-head. It is attributed to the fact that the rear pile is inducing
higher bending moment than middle pile in the capped-head piles (PG4).

From the above findings, it is postulated that the bending moment transfer
mechanism for a capped-head 6-pile group (Test PG4) is very different from the fixedhead pile. When the capped-head pile group is tilted due to tunnelling, the front row
piles tend to bend the most toward tunnel direction but the upper portion of the
bending moment profile is being dragged back by the by middle and rear row piles via
the connecting pile cap. It is postulated that the rear row piles behave like passive
pile when pile cap tilts and bends toward the tunnel. Thus, a pile cap tends to transmit
more bending moment from the middle pile to the rear pile which would otherwise be
less affected by the tunnelling-induced soil movement. Contrary, all of the piles in
fixed-head (Test PG5) 6-pile group behave like single piles standing side by side
without direct pile-cap-pile interaction, except that the magnitude is affected by the
total number of piles because the behavior is largely governed by the pile-soil-pile
interaction.

232

Chapter 6 Effects of Tunnelling on Pile Groups

Figure 6.69 shows the summary of the ratio of pile responses of long-term to
short-term pile group responses for all tests presented in this chapter. Similar plot for
single pile responses have been presented in Section 5.5. The data showing the longterm effects with the long-term over short-term ratio of 1.32 to 2.4 as oppose to the
single pile long-term over short-term ratio of wider range of 1.34 to 3.5. This
comparison again confirm the long-term time effects of pile responses due to
tunnelling, regardless of single or group, pile cap condition or group size.

6.6 CONCLUDING REMARKS

This chapter presents the results of five centrifuge model tests on pile groups with
different number of piles, pile cap and pile tip condition.

In the case of a floating capped-head pile group (Test PG1), the pile group is
generally beneficial as the average pile group responses are smaller than the average of
those of single piles at the same locations. This is because more efforts are required to
drag or bend the entire pile group including the pile cap. When a pile group gets larger,
the induced pile responses become smaller, in which shadowing and reinforcing effects
are dominant, thus diminishing the effects of induced soil movements acting on the
piles. When the pile toe condition changes to end-bearing (Test PG2) with a short
socket, the behavior is totally different. The pile movements reduce substantially, but
the trade-off is that intensive structural responses are induced in term of axial forces
and bending moments at the upper shaft and pile cap.

233

Chapter 6 Effects of Tunnelling on Pile Groups

The scenario becomes more complicated if different pile cap conditions are
modeled, as the head conditions play vital roles in dictating the pile responses.
Generally, capped-head piles (Test PG2 (2-pile group) & Test PG4 (6-pile group))
demonstrate significant pile-cap-pile interaction among the piles. Contrary, the fixedhead piles (Test PG3 (2-pile group) and Test PG5 (6-pile group)) behave like single
piles standing side by side without direct pile-cap-pile interaction; expect that the
magnitude is affected by the total number of piles because the behavior is largely
governed by the pile-soil-pile interaction.

In addition, once the pile group size increases from 2 to 6 piles, the position of
the pile within a group demonstrates a totally different transfer mechanism in the
lateral pile responses, as compared to that of a single pile with responses reducing with
increasing distance of pile to tunnel. The pile-cap-pile interaction in capped-head 6pile group (Test PG4) would moderate the induced pile bending moments among the
piles within a pile group. As a result, the induced pile bending moments in the middle
row is smaller than that of rear row. This is contrary to the induced lateral soil
movements, in which the corresponding lateral soil movement on the middle row piles
is larger than the corresponding movement on the rear row piles. This suggests that
part of the bending moments of the middle row piles is transferred to the rear piles due
to the interaction through the pile cap. It is worth noting that the axial forces reduce
when the distance between the pile and tunnel increases. It is thus suggested that the
pile-cap-pile interaction in capped-head 6-pile group (Test PG4) is less significant in
the axial force as compared to bending moment and the induced pile axial forces are
mainly influenced by the soil settlement and the distance between tunnel and pile.

234

Chapter 6 Effects of Tunnelling on Pile Groups

On the other hand, for the piles in fixed-head 6-pile group (Test PG5), the piles
behave like single piles in term of axial force and bending moment, except that the
magnitude is affected by the total number of piles. Moreover, the results show that the
fixed-head condition reduces the pile axial force, or in other words, a reduction in the
maximum drag load, but with tensile forces induced along the upper pile shaft due to
the total fixity at the pile cap. In contrast, there is an increment in the pile bending
moment. This is due to the totally fixed-head condition, which restrains the pile
movement causing a high pile bending moment.

A common trend has been observed for the long-term over short-term ratio of
pile responses for both single pile and pile group. The results reveal that soil and pile
responses increase over time with long-term over short-term pile responses ratio
ranging from 1.32 to 2.4, regardless of pile size, pile head and toe conditions.

235

Chapter 6 Effects of Tunneling on Pile Groups

Table 6.1 Test program and prototype parameters for pile group tests
Test
Ref.

Plan View

Elevation View

PG1

Bending Pile

Parameter

A=6m
B=10m
C=12m
D=6m
L=22m

Axial Pile

A
T
u
n
n
e
l

FP RP

C
L
D

PG2

Bending Pile

A
B

FP RP

A=6m
B=10m
C=12m
D=6m
L=27.5m

Axial Pile

C
L
D

PG3

A
B

Bending Pile

FP RP

Axial Pile

PG4
T
u
n
n
e
l

Bending Pile

A=6m
B=10m
C=14m
D=6m
L=27.5m

Axial Pile

L
D

Bending Pile

A
B
E

Tie
Beam
C
L
D

Tie
Beam

A
B
E

Fixed
head
Endbearing
pile

Capped
head
Endbearing
pile
6-pile
group

Axial Pile

FP MP RP

B
C

Endbearing
pile

2-pile
group

B
C

A=6m
B=10m
C=12m
D=6m
L=27.5m

A
B

FP MP RP

PG5
T
u
n
n
e
l

Tie
Beam

Tie
Beam

Capped
head

2-pile
group

A
T
u
n
n
e
l

Floating
pile
2-pile
group

A
T
u
n
n
e
l

Pile head
& toe
condition
Capped
head

A=6m
B=10m
C=14m
D=6m
L=27.5m

Fixed
head
Endbearing
pile
6-pile
group

236

Chapter 6 Effects of Tunneling on Pile Groups

Axial Force (kN)


0

100

200

300

400

500

600

-5

Depth (m)

-10

-15

T unnel

-20
Test PG1, Front (ST)
-25

Test PG1, Front (LT)


Test PG1, Rear (ST)
Test PG1, Rear (LT)

-30

Figure 6.1 Tunnelling-induced pile axial force (Test PG1)

237

Chapter 6 Effects of Tunneling on Pile Groups

Axial Force (kN)

100

200

300

400

500

600

Distance of pile from


tunnel centre = 6 m

-5

Depth (m)

-10

-15

T unnel

-20
Test PG1, Front (ST)
Test PG1, Front (LT)

-25

Test 3, Single, X=6m(ST)


Test 3, Single, X=6m (LT)
-30

Figure 6.2 Tunnelling-induced front pile (Test PG1) and corresponding single pile
(Test 3) axial force
Axial Force (kN)
0

100

200

300

400

500

600

Distance of pile from


tunnel centre = 10 m
-5

Depth (m)

-10

-15

-20

T unnel

Test PG1, Rear (ST)


Test PG1, Rear (LT)

-25

Test 16, Single, X=10m (ST)


Test 16, Single, X=10m (LT)

-30

Figure 6.3 Tunnelling-induced rear pile (Test PG1) and corresponding single pile (Test
16) axial force

238

Chapter 6 Effects of Tunneling on Pile Groups

Pile head settlement (mm)

20

16

12

0
Single pile, front,
Test 3 (X=6m)

Single pile, rear,


Test 16 (X=10m)

ST (2 days)

2-pile group, front, 2-pile group, rear,


Test PG1 (X=6m) Test PG1(X=10m)

L (720 days)T

Figure 6.4 Tunnelling-induced pile head settlement (Tests PG1, 3, 16)


Bending Moment (kNm)
-100

-50

50

100

150

200

250

-5

Depth (m)

-10

-15

T unnel

-20

-25

Test PG1, Front (ST)


Test PG1, Front (LT)
Test PG1, Rear (ST)
Test PG1, Rear (LT)

-30

Figure 6.5 Tunnelling-induced pile bending moment (Test PG1)

239

Chapter 6 Effects of Tunneling on Pile Groups

Bending Moment (kNm)


-100

-50

50

100

150

200

250

Distance of pile from


tunnel centre = 6 m

-5

Depth (m)

-10

-15

T unnel

-20

Test PG1, Front (ST)


Test PG1, Front (LT)
Test 3, Single, X=6m (ST)
Test 3, Single, X=6m (LT)

-25

-30

Figure 6.6 Tunnelling-induced front pile (Test PG1) and corresponding single pile
(Test 3) bending moment
Bending Moment (kNm)
-100

-50

50

100

150

200

250

-5

Distance of pile from


tunnel centre = 10 m

Depth (m)

-10

-15

T unnel

-20

-25

Test PG1, Rear (ST)


Test PG1, Rear (LT)
Test 16, Single, X=10m (ST)
Test 16, Single, X=10m (LT)

-30

Figure 6.7 Tunnelling-induced rear pile (Test PG1) and corresponding single pile (Test
16) bending moment

240

Chapter 6 Effects of Tunneling on Pile Groups

Pile deflection (mm)


0

Depth (m)

-5

-10

-15

-20

T unnel

PG1, Front (ST)


PG1, Front (LT)

-25

PG1, Rear (ST)


PG1, Rear (LT)

-30

Figure 6.8 Tunnelling-induced pile deflection (Tests PG1)

241

Chapter 6 Effects of Tunneling on Pile Groups

Pile deflection (mm)


0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16
0

-5

Distance of pile from


tunnel centre = 6 m
Depth (m)

-10

-15

-20

T unnel

PG1, Front (ST)


PG1, Front (LT)

-25

Test 3, Single, X=6m (ST)


Test 3, Single, X=6m (LT)

-30

Figure 6.9 Tunnelling-induced front pile (Test PG1) and corresponding single pile
(Test 3) deflection
Pile deflection (mm)
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16
0

-5

Distance of pile from


tunnel centre = 10 m
Depth (m)

-10

-15

-20

T unnel

PG1, Rear (ST)


PG1, Rear (LT)

-25

Test 16, Single, X=10m (ST)


Test 16, Single, X=10m (LT)

-30

Figure 6.10 Tunnelling-induced rear pile (Test PG1) and corresponding single pile
(Test 16) deflection

242

Chapter 6 Effects of Tunneling on Pile Groups

16

Pile head deflection (mm)

14
12
10
8
6
4
2
0
Single pile, front, Test 3

Single pile, rear, Test 16

ST (2 days)

2-pile group,Test PG1

LT (720 days)

Figure 6.11 Tunnelling-induced pile head deflection (Test PG1, 3 & 16)

243

Chapter 6 Effects of Tunneling on Pile Groups

Single pile/ pile group ratio

3
2.5

Positive
effect of
pile group

2
1.5
1
0.5
0
Pile axial
force

Pile head Pile negative Pile postive


settlement
bending
bending
moment
moment

ST (2 days)

Pile head
deflection

LT (720 days)

Figure 6.12 Single pile over pile group ratio for front pile (Test 3/ PG1)

Single pile/ pile group ratio

3
2.5

Positive
effect of pile
group

2
1.5
1
0.5
0
Pile axial
force

Pile head
settlement

ST (2 days)

Pile negative Pile postive


bending
bending
moment
moment

Pile head
deflection

LT (720 days)

Figure 6.13 Single pile over pile group ratio for rear pile (Test 16/ PG1)

244

Chapter 6 Effects of Tunneling on Pile Groups

Axial Force (kN)


0

200

400

600

800

1000

1200

-5

-10

Depth (m)

-15

Tunnel

-20

-25

-30
Test PG2, Front (ST)
Test PG2, Front (LT)
Test PG2, Rear (ST)
Test PG2, Rear (LT)

Figure 6.14 Tunnelling-induced pile axial force (Test PG2)

245

Chapter 6 Effects of Tunneling on Pile Groups

Axial Force (kN)


0

200

400

600

800

1000

1200

0
Distance of pile from
tunnel centre = 6 m

-5

-10

Tunnel

Depth (m)

-15

-20

-25

-30
Test PG2, Front (ST)
Test PG2, Front (LT)
Test 10, Single, X=6m(ST)
Test 10, Single, X=6m (LT)

Figure 6.15 Tunnelling-induced front pile (Test PG2) and corresponding single pile
(Test 10) axial force
Axial Force (kN)
0

200

400

600

800

1000

1200

-5

Distance of pile from


tunnel centre = 10 m

-10

Tunnel

Depth (m)

-15

-20

-25

-30
Test PG2,Rear (ST)
Test PG2, Rear (LT)
Test 11, Single, X=10m(ST)
Test 11, Single, X=10m (LT)

Figure 6.16 Tunnelling-induced rear pile (Test PG2) and corresponding single pile
(Test 11) axial force

246

Chapter 6 Effects of Tunneling on Pile Groups

Bending Moment (kNm)


-80

-40

40

80

120

160

-5

Depth (m)

-10

-15

T unnel

-20
Test PG2, Front (ST)
Test PG2, Front (LT)
-25
Test PG2, Rear (ST)
Test PG2, Rear (LT)
-30

Figure 6.17 Tunnelling-induced pile bending moment (Test PG2)

247

Chapter 6 Effects of Tunneling on Pile Groups

Bending Moment (kNm)


-100

-50

50

100

150

200

250

Distance of pile from


tunnel centre = 6 m
-5

Depth (m)

-10

-15

T unnel

-20

Test PG2, Front (ST)


-25

Test PG2, Front (LT)


Test 10, Single, X=6m (ST)
Test 10, Single, X=6m (LT)

-30

Figure 6.18 Tunnelling-induced front pile (Test PG2) and corresponding single pile
(Test 10) bending moment
Bending Moment (kNm)
-100

-50

50

100

150

200

250

Distance of pile from


tunnel centre = 10 m
-5

Depth (m)

-10

-15

T unnel

-20

Test PG2, Front (ST)


-25

Test PG2, Rear (LT)


Test 11, Single, X=10m (ST)
Test 11, Single, X=10m (LT)

-30

Figure 6.19 Tunnelling-induced rear pile (Test PG2) and corresponding single pile
(Test 11) bending moment

248

Chapter 6 Effects of Tunneling on Pile Groups

Lateral deflection (mm)


-2

-1

-5

Depth (m)

-10

-15

T unnel

-20
Test PG2, Front (ST)
Test PG2, Front (LT)

-25

Test PG2, Rear (ST)


Test PG2, Rear (LT)

-30

Figure 6.20 Tunnelling-induced pile deflection (Test PG2)

249

Chapter 6 Effects of Tunneling on Pile Groups

Lateral deflection (mm)


-2

-1

-5

Distance of pile from


tunnel centre = 6 m

Depth (m)

-10
`

T unnel

-15

-20
Test PG2, Front (ST)
Test PG2, Front (LT)

-25

Test 10, Single, X=6m(ST)


Test 10, Single, X=6m(LT)

-30

Figure 6.21 Tunnelling-induced front pile (Test PG2) and corresponding single pile
(Test 10) deflection
Lateral deflection (mm)
-2

-1

-5

Depth (m)

Distance of pile from


tunnel centre = 10 m

-10
``

-15

T unnel

-20
Test PG2, Rear (ST)
Test PG2, Rear (LT)

-25

Test 11, Single, X=10m(ST)


Test 11, Single, X=10m(LT)

-30

Figure 6.22 Tunnelling-induced rear pile (Test PG2) and corresponding single pile
(Test 11) deflection

250

Chapter 6 Effects of Tunneling on Pile Groups

Pile head deflection (mm)

6
5
4
3
2
1
0
Single pile,
Test 10 (X=
6m)

Single pile,
Test 11
(X=10m)

Single pile,
Test 12
(X=14m)

2-pile
group,Test
PG2

6-pile
group,Test
PG4

2-pile
group,Test
PG2

6-pile
group,Test
PG4

ST

(a) Short-term

Pile head deflection (mm)

6
5
4
3
2
1
0
Single pile,
Test 10 (X=
6m)

Single pile,
Test 11
(X=10m)

Single pile,
Test 12
(X=14m)

LT

(b) Long-term
Figure 6.23 Tunnelling-induced pile head deflection in the (a) short-term (b) long-term
(Tests PG2, 10 & 11)

251

Chapter 6 Effects of Tunneling on Pile Groups

Single pile/ pile group ratio

Positive
effect of
pile group

2.5
2
1.5
1
0.5
0
Pile axial force

ST

Pile negative
Pile positive
bending moment bending moment

Pile head
deflection

LT

Figure 6.24 Single pile over pile group ratio for front pile (Test 10/ PG2)

Single pile/ pile group ratio

3
2.5

Positive
effect of pile
group

2
1.5
1
0.5
0
Pile axial force

ST

Pile negative
Pile positive
bending moment bending moment

Pile head
deflection

LT

Figure 6.25 Single pile over pile group ratio for rear pile (Test 11/ PG2)

252

Chapter 6 Effects of Tunneling on Pile Groups

Axial Force (kN)


-200 -100

100 200 300 400 500 600 700 800

-5

Depth (m)

-10

-15

T unnel

-20

-25

-30
Test PG3, Front (ST)

Test PG3, Front (LT)

Test PG3, Rear (ST)

Test PG3, Rear (LT)

Figure 6.26 Tunnelling-induced pile axial force (Test PG3)

253

Chapter 6 Effects of Tunneling on Pile Groups

Axial Force (kN)


-200 -100

100 200 300 400 500 600 700 800

Distance of pile from


tunnel centre = 6 m

-5

-10

T unnel

Depth (m)

-15

-20

-25

-30
Test
Test
Test
Test

PG3, Front (ST)


PG3, Front (LT)
13, Single, X=6m (ST)
13, Single, X=6m (LT)

Figure 6.27 Tunnelling-induced front pile (Test PG3) and corresponding single pile
(Test 13) axial force

254

Chapter 6 Effects of Tunneling on Pile Groups

Bending Moment (kNm)


-200

-150

-100

-50

50

100

150

200

-5

Depth (m)

-10

-15

T unnel

-20
Test PG3, Front (ST)
-25

Test PG3, Front (LT)


Test PG3, Rear (ST)
Test PG3, Rear (LT)

-30

Figure 6.28 Tunnelling-induced pile bending moment (Test PG3)

255

Chapter 6 Effects of Tunneling on Pile Groups

Bending Moment (kNm)


-200

-150

-100

-50

50

100

150

200

Distance of pile from


tunnel centre = 6 m
-5

Depth (m)

-10

-15

T unnel

-20

Test PG3, Front


(ST)
Test PG3, Front
(LT)
Test 13, Single,
X=6m (ST)
Test 13, Single,
X=6m (LT)

-25

-30

Figure 6.29 Tunnelling-induced front pile (Test PG3) and corresponding single pile
(Test 13) bending moment
Bending Moment (kNm)
-200

-150

-100

-50

50

100

150

200

-5

Distance of pile from


tunnel centre = 10 m

Depth (m)

-10

-15

T unnel

-20

Test PG3, Rear


(ST)
Test PG3, Rear
(LT)
Test 14, X=10m
(ST)
Test 14, Single,
X=10m (LT)

-25

-30

Figure 6.30 Tunnelling-induced rear pile (Test PG3) and corresponding single pile
(Test 14A) bending moment
256

Chapter 6 Effects of Tunneling on Pile Groups

Positive
effect of pile
group

1.8

Single pile/pile group ratio

1.6
1.4
1.2
1
0.8
0.6
0.4
0.2
0
Pile axial tensile
force

Pile axial
compression
force

ST

Pile negative
Pile positive
bending moment bending moment

LT

Figure 6.31 Single pile over pile group ratio for front pile (Test 13/ PG3)

Positive
effect of pile
group

Single pile/pile group ratio

1.8
1.6
1.4
1.2
1
0.8
0.6
0.4
0.2
0
Pile negative bending moment

ST

Pile positive bending moment

LT

Figure 6.32 Single pile over pile group ratio for rear pile (Test 14A/ PG3)

257

Chapter 6 Effects of Tunneling on Pile Groups

Axial Force (kN)


0

100

200

300

400

500

600

-5

Depth (m)

-10

-15

T unnel

-20

-25

-30
Test PG4, Front (ST)

Test PG4, Front (LT)

Test PG4, Middle (ST)

Test PG4, Middle (LT)

Test PG4, Rear (ST)

Test PG4, Rear (LT)

Figure 6.33 Tunnelling-induced pile axial force (Test PG4)

258

Chapter 6 Effects of Tunneling on Pile Groups

Axial Force (kN)


0

100

200

300

400

500

600

700

800

0
Distance of pile from
tunnel centre = 6 m

-5

-10

Tunnel

Depth (m)

-15

-20

-25

-30
Test PG2, Front (ST)

Test PG2, Front (LT)

Test PG4, Front (ST)

Test PG4, Front (LT)

Figure 6.34 Tunnelling-induced front pile in 2-pile group (Test PG2) and
corresponding front pile in 6-pile group (Test PG4) axial force
Axial Force (kN)
0

100

200

300

400

500

600

700

800

-5

Distance of pile from


tunnel centre = 10 m

-10

Depth (m)

-15

Tunnel

-20

-25

-30
Test PG2,Rear (ST)

Test PG2, Rear (LT)

Test P4, Middle (ST)

Test PG4, Middle (LT)

Figure 6.35 Tunnelling-induced rear pile in 2-pile group (Test PG2) and corresponding
middle pile in 6-pile group (Test PG4) axial force
259

Chapter 6 Effects of Tunneling on Pile Groups

Bending Moment (kNm)


-60

-40

-20

20

40

60

80

100

-5

Depth (m)

-10

-15

T unnel

-20
Test PG4, Front (ST)
Test PG4, Middle (ST)

-25

Test PG4, Rear (ST)


-30

(a) Short-term
Bending Moment (kNm)
-60

-40

-20

20

40

60

80

100

120

-5

Depth (m)

-10

-15

T unnel

-20
Test PG4, Front (LT)
-25

Test PG4, Middle (LT)


Test PG4, Rear (LT)

-30

(b) Long-term
Figure 6.36 Tunnelling-induced pile bending moment (a) in the short-term (b) in the
long-term (Test PG4)

260

Chapter 6 Effects of Tunneling on Pile Groups

Bending Moment (kNm)


-100

-50

50

100

150

200

Distance of pile from


tunnel centre = 6 m

-5

Depth (m)

-10

-15

T unnel

-20

Test PG4, Front (ST)


-25

Test PG4, Front (LT)


Test PG2, Front (ST)
Test PG2, Front (LT)

-30

Figure 6.37 Tunnelling-induced front pile in 2-pile group (Test PG2) and
corresponding front pile in 6-pile group (Test PG4) bending moment
Bending Moment (kNm)
-100

-50

50

100

150

200

-5

Distance of pile from


tunnel centre = 10 m

Depth (m)

-10

-15

T unnel

-20

Test PG4, Middle (ST)


-25

Test PG4, Middle (LT)


Test PG2, Rear (ST)
Test PG2, Rear (LT)

-30

Figure 6.38 Tunnelling-induced rear pile in 2-pile group (Test PG2) and corresponding
middle pile in 6-pile group (Test PG4) bending moment
261

Chapter 6 Effects of Tunneling on Pile Groups

Lateral deflection (mm)


-1.5

-1

-0.5

0.5

1.5

2.5

-5

Depth (m)

-10

-15

T unnel

-20
Test PG4, Front (ST)

-25

Test PG4, Middle (ST)


Test PG4, rear (ST)

-30

(a) Short-term
Lateral deflection (mm)
-1.5

-1

-0.5

0.5

1.5

2.5

-5

Depth (m)

-10

-15

T unnel

-20
Test PG4, Front (LT)

-25

Test PG4, Middle (LT)


Test PG4, rear (LT)

-30

(b) Long-term
Figure 6.39 Tunnelling-induced pile deflection in the (a) short-term (b) long-term
(Test PG4)

262

Chapter 6 Effects of Tunneling on Pile Groups

Lateral deflection (mm)


-2

-1

-5

Depth (m)

-10
`
Tunnel

-15

Test PG2, Front (ST)

-20

Test PG2, Front (LT)


Test PG4, Front (ST)
Test PG4, Front (LT)

-25

Distance of pile from


tunnel centre = 6 m
-30

Figure 6.40 Tunnelling-induced pile deflection (Tests PG2 and PG4)


Lateral deflection (mm)
-2

-1

Depth (m)

-5

-10

-15

`
Tunnel
Test PG2, Rear (ST)

-20

Test PG2, Rear (LT)


Test PG4, Middle (ST)

-25

Test PG4, Middle (LT)

Distance of pile from


tunnel centre = 10 m
-30

Figure 6.41 Tunnelling-induced pile bending moment (Tests PG2 and PG4)

263

Chapter 6 Effects of Tunneling on Pile Groups

2-Pile Group/ 6-Pile Group ratio

2.4
2.2

Positive
effect of
pile group

2
1.8
1.6
1.4
1.2
1
c

0.8
0.6
0.4
0.2
0
Pile axial force

ST

Pile negative bending Pile positive bending


moment
moment

LT

Figure 6.42 2-pile over 6-pile group ratio for front pile (Test PG2/PG4)

2-Pile Group/ 6-Pile Group ratio

2.4
2.2
2
1.8

Positive
effect of
pile group

1.6
1.4
1.2
1
0.8
0.6
0.4
0.2
0
Pile axial force

ST

Pile negative bending


moment

Pile positive bending


moment

LT

Figure 6.43 2-pile over 6-pile group ratio for middle pile (Test PG2/PG4)

264

Chapter 6 Effects of Tunneling on Pile Groups

Axial Force (kN)


-200

-100

100

200

300

400

500

600

-5

-10

Depth (m)

-15

Tunnel

-20

-25

-30
Test PG5, Front (ST)

Test PG5, Front (LT)

Test PG5, Middle (ST)

Test PG5, Middle (LT)

Test PG5, Rear (ST)

Test PG5, Rear (LT)

Figure 6.44 Tunnelling-induced pile axial force (Test PG5)

265

Chapter 6 Effects of Tunneling on Pile Groups

Axial Force (kN)


-200 -100

100

200

300

400

500

600

Distance of pile from


tunnel centre = 6 m

-5

-10

Depth (m)

-15

Tunnel

-20

-25

-30
Test PG5, Front (ST)

Test PG5, Front (LT)

Test PG3, Front (ST)

Test PG3, Front (LT)

Figure 6.45 Tunnelling-induced front pile in 2-pile group (Test PG3) and
corresponding front pile in 6-pile group (Test PG5) axial force
Axial Force (kN)
-200 -100

100

200

300

400

500

600

-5

Distance of pile from


tunnel centre = 10 m

-10

Depth (m)

-15

Tunnel

-20

-25

-30
Test PG5, Middle (ST)

Test PG5, Middle (LT)

Test PG3, Rear (ST)

Test PG3, Rear (LT)

Figure 6.46 Tunnelling-induced rear pile in 2-pile group (Test PG3) and corresponding
middle pile in 6-pile group (Test PG5) axial force

266

Chapter 6 Effects of Tunneling on Pile Groups

Bending Moment (kNm)


-80

-40

40

80

120

160

-5

Depth (m)

-10

-15

T unnel

-20
Test PG5, Front (ST)
Test PG5, Middle (ST)

-25

Test PG5, Rear (ST)


-30

(a) Short-term
Bending Moment (kNm)
-80

-40

40

80

120

160

-5

Depth (m)

-10

-15

T unnel

-20
Test PG5, Front (LT)
-25

Test PG5, Middle (LT)


Test PG5, Rear (LT)

-30

(b) Long-term
Figure 6.47 Tunnelling-induced pile bending moment in the (a) short-term (b) longterm (Test PG5)

267

Chapter 6 Effects of Tunneling on Pile Groups

Bending Moment (kNm)


-120

-80

-40

40

80

120

160

200

Distance of pile from


tunnel centre = 6 m

-5

Depth (m)

-10

-15

T unnel

-20

Test PG5, Front (ST)


Test PG5, Front (LT)

-25

Test PG3, Front (ST)


Test PG3, Front (LT)
-30

Figure 6.48 Tunnelling-induced front pile in 2-pile group (Test PG3) and
corresponding front pile in 6-pile group (Test PG5) bending moment
Bending Moment (kNm)
-120

-80

-40

40

80

120

160

200

-5

Distance of pile from


tunnel centre = 10 m

Depth (m)

-10

-15

T unnel

-20
Test PG5, Middle (ST)
-25

Test PG5, Middle (LT)


Test PG3, Rear (ST)
Test PG3, Rear (LT)

-30

Figure 6.49 Tunnelling-induced rear pile in 2-pile group (Test PG3) and corresponding
middle pile in 6-pile group (Test PG5) bending moment
268

2-Pile Group/ 6-Pile Group ratio

Chapter 6 Effects of Tunneling on Pile Groups

Positive
effect of
pile group

2
1.8
1.6
1.4
1.2
1
0.8
0.6
0.4
0.2
0
Pile axial tensile
force

Pile axial
compression
force

ST

Pile negative
Pile positive
bending moment bending moment

LT

Figure 6.50 2-pile over 6-pile group ratio for front pile (Test PG3/PG5)

2-Pile Group/ 6-Pile Group ratio

Positive
effect of pile
group

1.8
1.6
1.4
1.2
1
0.8
0.6
0.4
0.2
0
Pile axial tensile
force

Pile axial
compression
force

ST

Pile negative
Pile positive
bending moment bending moment

LT

Figure 6.51 2-pile over 6-pile group ratio for middle pile (Test PG3/PG5)

269

Chapter 6 Effects of Tunneling on Pile Groups

Axial Force (kN)


-200 -100

100 200 300 400 500 600 700 800

-5

Distance of pile from


tunnel centre = 6 m

-10

Depth (m)

-15

Tunnel

-20

-25

-30
Test PG3, Front (ST)

Test PG3, Front (LT)

Test PG2, Front (ST)

Test PG2, Front (LT)

Figure 6.52 Tunnelling-induced front pile in capped-head 2-pile group (Test PG2) and
corresponding front pile in fixed-head 2-pile group (Test PG3) axial force

270

Chapter 6 Effects of Tunneling on Pile Groups

Axial Force (kN)


-200 -100

100 200 300 400 500 600 700 800

-5

Distance of pile from


tunnel centre = 10 m

-10

Depth (m)

-15

Tunnel

-20

-25

-30
Test PG3, Rear (ST)

Test PG3, Rear (LT)

Test PG2, Rear (ST)

Test PG2, Rear (LT)

Figure 6.53 Tunnelling-induced rear pile in capped-head 2-pile group (Test PG2) and
corresponding rear pile in fixed-head 2-pile group (Test PG3) axial force

271

Chapter 6 Effects of Tunneling on Pile Groups

Bending Moment (kNm)


-150

-100

-50

50

100

150

200

Distance of pile from


tunnel centre = 6 m

-5

Depth (m)

-10

T unnel

-15

-20
Test PG3, Front (ST)
Test PG3, Front (LT)

-25

Test PG2, Front (ST)


Test PG2, Front (LT)
-30

Figure 6.54 Tunnelling-induced front pile in capped-head 2-pile group (Test PG2) and
corresponding front pile in fixed-head 2-pile group (Test PG3)bending moment
Bending Moment (kNm)
-150

-100

-50

50

100

150

200

-5

Distance of pile from


tunnel centre = 10 m

Depth (m)

-10

-15

T unnel

-20

Test PG3, Rear (ST)


-25

Test PG3, Rear (LT)


Test PG2, Rear (ST)
Test PG2, Rear (LT)

-30

Figure 6.55 Tunnelling-induced rear pile in capped-head 2-pile group (Test PG2) and
corresponding rear pile in fixed-head 2-pile group (Test PG3) bending moment

272

Capped-head pile/ Fixed-head pile

Chapter 6 Effects of Tunneling on Pile Groups

Positive
effect of
fixed-head
pile

1.8
1.6
1.4
1.2
1
0.8
0.6
0.4
0.2
0
Pile compression
force

ST

Pile negative bending Pile positive bending


moment
moment

LT

Capped-head pile/ Fixed-head pile

Figure 6.56 Capped-head pile over fixed-head pile ratio (front pile, Test PG2/PG3)

Positive
effect of
fixed-head
pile

2
1.8
1.6
1.4
1.2
1
0.8
0.6
0.4
0.2
0
Pile compression
force

ST

Pile negative bending


moment

Pile positive bending


moment

LT

Figure 6.57 Capped-head pile over fixed-head pile ratio (rear pile, Test PG2/PG3)

273

Chapter 6 Effects of Tunneling on Pile Groups

Axial Force (kN)


-200

-100

100

200

300

400

500

600

-5

Distance of pile from


tunnel centre = 6 m

-10

Depth (m)

-15

Tunnel

-20

-25

-30
Test PG4, Front (ST)

Test PG4, Front (LT)

Test PG5, Front (ST)

Test PG5, Front (LT)

Figure 6.58 Tunnelling-induced front pile in capped-head 6-pile group (Test PG4) and
corresponding front pile in fixed-head 6-pile group (Test PG5) axial force

274

Chapter 6 Effects of Tunneling on Pile Groups

Axial Force (kN)


-200

-100

100

200

300

400

500

600

Distance of pile from


tunnel centre = 10 m

-5

-10

Depth (m)

-15

Tunnel

-20

-25

-30
Test PG4, Middle (ST)

Test PG4, Middle (LT)

Test PG5, Middle (ST)

Test PG5, Middle (LT)

Figure 6.59 Tunnelling-induced middle pile in capped-head 6-pile group (Test PG4)
and corresponding middle pile in fixed-head 6-pile group (Test PG5) axial force
Axial Force (kN)
-200

-100

100

200

300

400

500

600

-5

Distance of pile from


tunnel centre = 14 m

-10

Depth (m)

-15

Tunnel

-20

-25

-30

Test PG4, Rear (ST)

Test PG4, Rear (LT)

Test PG5, Rear (ST)

Test PG5, Rear (LT)

Figure 6.60 Tunnelling-induced rear pile in capped-head 6-pile group (Test PG4) and
corresponding rear pile in fixed-head 6-pile group (Test PG5) axial force

275

Chapter 6 Effects of Tunneling on Pile Groups

Bending Moment (kNm)


-80

-40

40

80

120

160

-5

Distance of pile from


tunnel centre = 6 m

Depth (m)

-10

-15

T unnel

-20

Test PG4, Front (ST)


-25

Test PG4, Front (LT)


Test PG5, Front (ST)
Test PG5, Front (LT)

-30

Figure 6.61 Tunnelling-induced front pile in capped-head 6-pile group (Test PG4) and
corresponding front pile in fixed-head 6-pile group (Test PG5) bending moment

276

Chapter 6 Effects of Tunneling on Pile Groups

Bending Moment (kNm)


-80

-40

40

80

120

160

Distance of pile from


tunnel centre = 10 m
-5

Depth (m)

-10

-15

T unnel

-20
Test PG4, Middle (ST)
Test PG4, Middle (LT)

-25

Test PG5, Middle (ST)


Test PG5, Middle (LT)
-30

Figure 6.62 Tunnelling-induced middle pile in capped-head 6-pile group (Test PG4)
and corresponding middle pile in fixed-head 6-pile group (Test PG5) bending moment
Bending Moment (kNm)
-80

-40

40

80

120

160

-5

Distance of pile from


tunnel centre = 14 m

Depth (m)

-10

-15

T unnel

-20
Test PG4, Rear (ST)
-25

Test PG4, Rear (LT)


Test PG5, Rear (ST)
Test PG5, Rear (LT)

-30

Figure 6.63 Tunnelling-induced rear pile in capped-head 6-pile group (Test PG4) and
corresponding rear pile in fixed-head 6-pile group (Test PG5) bending moment

277

Chapter 6 Effects of Tunneling on Pile Groups

40
30
20
10
0

Front

Middle

Rear

-10
-20
-30
-40
Capped-head (PG4), ST

Fixed-head (PG5), ST

Capped-head (PG4), ST

Fixed-head (PG5), ST

Figure 6.64 Variation of maximum bending moment for front, middle and rear pile in
the short-term (Tests PG4 and PG5)
80
70
60
50
40
30
20
10
0
-10
-20
-30
-40
-50
-60

Front

Middle

Rear

Capped-head (PG4), LT

Fixed-head (PG5), LT

Capped-head (PG4), LT

Fixed-head (PG5), LT

Figure 6.65 Variation of maximum bending moment for front, middle and rear pile in
the long-term (Tests PG4 and PG5)

278

Chapter 6 Effects of Tunneling on Pile Groups

Capped-head pile/ Fixed-head pile ratio

3
2.5

Positive effect of
fixed-head

2
1.5
1
0.5
0

Pile axial tensile


force

Pile axial
compression
force

Pile negative
Pile positive
bending moment bending moment

ST

LT

Capped-head pile/ Fixed-head pile rati

Figure 6.66 Capped-head pile over fixed-head pile ratio (front pile, Test PG4/PG5)

3
2.5
2

Positive effect of
fixed-head

1.5
1
0.5
0
Pile axial tensile
force

Pile axial
compression
force

ST

Pile negative
Pile positive
bending moment bending moment

LT

Capped-head pile/ Fixed-head pile ratio

Figure 6.67 Capped-head pile over fixed-head pile ratio (middle pile, Test PG4/PG5)

3
2.5
2

Positive effect of
fixed-head

1.5
1
0.5
0
Pile axial tensile
force

ST

Pile axial
compression
force

Pile negative
Pile positive
bending moment bending moment

LT

Figure 6.68 Capped-head pile over fixed-head pile ratio (rear pile, Test PG4/PG5)
279

Chapter 6 Effects of Tunneling on Pile Groups

2.5

LT/ST ratio

Long-term
effect

1.5

0.5

0
Pile axial tensile
force

Pile axial
compression
force

Pile negative
Pile positive
bending moment bending moment

Pile head
deflection

PG1, front

PG1, rear

PG2, front

PG2, rear

PG3, front

PG3, rear

PG4, front

PG4, middle

PG4, rear

PG5, front

PG5, middle

PG5, rear

Figure 6.69 Long-term over short-term ratio (Tests PG1, 2, 3, 4 & 5)

280

Chapter 7 Conclusions

CHAPTER SEVEN

CONCLUSIONS

7.1

CONCLUDING REMARKS

The overall purpose of this research study was to investigate tunnel-soil-pile


interaction in soft clay. The objectives were achieved through (a) the development of a
centrifuge tunnel excavation technique to simulate the inward tunnel deformation
pattern commonly observed in practice (b) a centrifuge model study of tunnellinginduced soil movements in free-field analysed using Particle Image Velocimetry (PIV)
technique (c) a centrifuge model study of tunnelling-induced single pile responses, and
(d) a centrifuge model study of tunnelling-induced pile group responses.

In this study, a series of centrifuge model tests were conducted to investigate


the effects of tunnelling on soft clay, single piles and pile groups in clay. A total of
twenty one centrifuge model tests - Tests 1 and 2 (tunnel-soil interaction), Tests 3 to
16 (tunnel-single pile interaction) and Tests PG1 to PG5 (tunnel-pile group interaction)
- were performed. The effects of factors such as volume loss, pile tip and head
condition, pile length, pile-to-tunnel distance, floating and end-bearing pile groups,
size of pile group and pile groups with capped-head and fixed-head conditions of pile
groups on pile due to tunneling were examined. In addition, the observed pile
behaviours are evaluated against the measured free-field soil movements due to
tunnelling.

281

Chapter 7 Conclusions

7.1.1 Technique for Simulation of Tunnelling


In the present centrifuge model study, an innovative tunnelling simulation technique
was developed to simulate the inward tunnel deformation due to over-excavation
commonly observed in practice. An oval-shape ground deformation pattern is imposed
as the boundary condition and the gap parameter (GAP) proposed by Lee et al. (1992)
is used to quantify the amount of tunnel over-cut. The experimental results from
tunnel-soil interaction presented in Chapter 4 provided clear evidence that the present
model tunnel was able to simulate the precise volume loss during the tunnelling
process. Moreover, this model could provide a very uniform oval-shape of the GAP
throughout the entire length of the model tunnel. This ensured that the volume loss was
constant along the model tunnel. Such simulation technique provides flexibility and
reproducibility to study the tunnel-soil-pile interaction in a centrifuge model that could
not be measured in field tests, particularly the time effects of soft clay.

However, the simplified two-dimensional model tunnel was unable to simulate


the complex three-dimensional effects of tunnelling before and after the passing of the
tunnel boring machine (TBM). Nevertheless, in a situation where the tunnel excavation
has passed a particular section, the vectors of the ground movement developed will be
more or less in the plane perpendicular to the tunnel axis. Consequently it is reasonable
to assume that a plane strain model of long tunnel section would be a reasonable
representation of tunnelling-induced soil movements; this is usually referred to as a
two-dimensional simulation (Taylor, 1998). Moreover, the results from Taylor (1998)
and the present study showed that the two-dimensional model tunnel was able to
measure the transverse responses of soils and piles.

282

Chapter 7 Conclusions

7.1.2 Tunnel-Soil Interaction


The centrifuge model tests with the application of PIV had provided a considerable
body of data to examine the patterns of soil movement induced by tunnelling in soft
clay. The surface settlement trough in clay generally follows the Gaussian distribution
curve in the short term. The magnitude of maximum ground surface settlement
increases with time and tunnel volume loss. The settlement magnitude is larger in the
long-term and the settlement trough is wider as compared to that in the short-term. The
observed data confirmed that the empirical equation proposed by Mair et al (1993) is
applicable in the prediction of the subsurface settlement troughs in clay in the shortterm. Empirical equations in the short-term and long-term were proposed for the
distribution of inflection point in soft clay.

In the short-term, an Immediate Shear Zone with large soil movement above
the tunnel can be identified, while the zone outside the immediate shear zone may be
identified as the Support Zone. In the long term, the significant soil movement zone
extends much wider. In addition, soil settlement was noted to be more dominant than
lateral soil movement in the long term.

7.1.3 Tunnel-Single Piles Interaction


During tunnelling, the ground around the tunnel often moves towards the tunnel
opening. The resulting ground movements induce additional axial (settlement and axial
force) and lateral (deflection and bending moment) responses on adjacent pile
foundations.

283

Chapter 7 Conclusions

Test Series 1 studied the effects of volume loss on pile performances. It was
found that the induced pile bending moment triples and the pile settlement and
deflection increase by almost 2.5 times when volume loss increases from 3% to 6.5%,
in this particular case.

For Test Series 2, three different pile tip conditions, namely floating pile,
socketed pile and end-bearing pile were investigated to study the effects of pile tip
condition. It is noted that a floating pile is mainly governed by pile settlement when
tunnelling is carried out adjacent to it whereas socketed piles are likely to be governed
by the material stress of the pile. On the other hand, some opposite trends were
observed in the fixed-head when compared to free-head. It is noted that tensile force
and relatively large negative bending moments were induced at the pile head due to
total fixity. Nevertheless, these responses had led to the reduction in drag load and
positive bending moment at the pile waist.

In a short pile, with pile base at or above the tunnel crown, especially those
located in the immediate shear zone, the pile structural responses (axial force and
bending moment) were of less significant if compared to those of long pile. However,
there will be excessive pile movements (settlement and deflection) because of lack of
anchorage of pile into the stable support zone. In this respect, a longer pile with pile
length in the stabilised support zone tends to provide more resistance to the pile
movements but will attract more bending moment and axial force.

Test Series 4, 5 and 6 examined the effects of pile-to-tunnel distance for


different pile head and tip conditions. Generally, it was observed that the pile

284

Chapter 7 Conclusions

responses decrease with increase in pile-to-tunnel distance. The induced pile axial
forces were observed to decrease fairly linearly with increase in pile-to-tunnel distance
for Test Series 4 (floating piles) and 5 (end-bearing pile) and the axial force in Test
Series 5 is always much higher than that of Test Series 4 for all pile to tunnel distance.
In addition, the most significant difference in Test Series 5 (free-head) and Test Series
6 (fixed-head) is that tension force is induced in the fixed-head pile due to total fixed
condition at the pile head. Generally, the maximum induced bending moments
decrease fairly linearly with increasing pile-to-tunnel distance when the magnitude is
relatively small as in the case of pile bending moment in the short-term and positive
bending moment observed in Series 6. However, the bending moments decrease
exponentially when the magnitude is relatively large. Generally, the induced bending
moments in end-bearing piles (Test Series 5) are larger than the floating piles (Test
Series 4) as the restraint at the pile toe would restrict the pile lateral movement and
induce a larger bending moment. On the other hand, the results revealed that negative
bending moments were induced at pile head due to total fixity condition for fixed-head
piles in Test Series 6 as compared to the free-head piles in Test Series 5. As the
bending moment profile was offset toward the negative bending moment for fixedhead pile, the positive bending moment in fixed-head pile (Test Series 6) are
consistently lower than that of free-head pile (Test Series 5). It can be established from
the test results that induced bending moments are generally small beyond a horizontal
offset of 2D from the tunnel centre. Generally, it was observed that the pile deflection
dropped rapidly from 1D to 1.5D, with a much smaller decrease from 1.5D to 2D for
both Series 4 and 5. This is because the lateral soil movements decrease with
increasing distance of pile location to the tunnel. The pile head deflection for endbearing piles (Test Series 5) was smaller that of floating piles (Test Series 4),

285

Chapter 7 Conclusions

regardless of the pile-tunnel distance, as the lower portion of the pile was restrained
and would not moves.

Some similarities and differences were drawn in the comparisons of soil and
single pile behaviour in the cases of both inward (present study) and outward (Ran,
2004) tunnel deformations. It is noted that the measured short-term surface settlement
trough follows the Gaussian distribution curve fairly well with the inflection point (i)
at approximately 7.5 m for both tunnel deformation cases. The most distinct difference
in the soil behaviour was that the soil moved significantly away from the tunnel in the
case of outward tunnel deformation, whereas in the case of inward tunnel deformation,
the soil moved towards the tunnel. It was revealed that the pile axial force and pile
settlement behaviour and profile were fairly similar regardless of the deformation
pattern but the outward tunnel formation would induce larger pile responses as
compared to the inward tunnel deformation under the same volume loss. On the other
hand, the pile lateral responses (bending moment and deflection) were opposite in
direction for both inward and outward tunnel deformations, respectively, in terms of
profiles and magnitude.

7.1.4 Tunnel-Pile Groups Interaction


In the case of a capped-head floating pile group (Test PG1), the pile group was
generally beneficial as the average pile group responses (bending moments, axial,
settlement and lateral deflection) are smaller than the average of those of single piles at
the same locations. This is because the rigidity of a pile group provides more resistance
to the tunnelling-induced soil movements.

286

Chapter 7 Conclusions

The scenario becomes more complicated when different pile cap conditions
were modeled, as the head conditions played a vital role in dictating the pile responses.
Generally, capped-head piles (Test PG2 (2-pile group) and Test PG4 (6-pile group))
demonstrate significant pile-cap-pile interaction among the piles. On the other hand,
the fixed-head piles (Test PG3 (2-pile group) and Test PG5 (6-pile group)), behaved
like single piles standing side by side without direct pile-cap-pile interaction, except
that the magnitude was affected by the total number of piles because the behaviors was
largely governed by the pile-soil-pile interaction.

When the pile group size increased from 2-pile to 6-pile, the position of the pile
within a group demonstrated a totally different transfer mechanism in lateral pile
responses, as compared to the single pile where the responses reduced consistently
when the distance of pile-tunnel increases. The pile-cap-pile interaction in capped-head
6-pile group (Test PG4) would moderate the induced pile bending moments among the
piles within a pile group. As a result, the induced pile bending moments in the middle
row was smaller than that of rear row. This is contrary to the induced lateral soil
movements, in which the corresponding lateral soil movement at the middle row of
piles was larger than the corresponding lateral soil movement at the rear row of pile.
This suggests that part of the bending moments of the pile in the middle row was
transferred to the rear piles due to the interaction through the pile cap. It is worth
noting that the axial forces reduce when the position of the pile-tunnel increases. Thus
the pile-cap-pile interaction in capped-head 6-pile group (Test PG4) has less influence
on the axial force as compared to bending moment and the induced pile axial forces
were mainly influenced by the soil settlement and distance effects.

287

Chapter 7 Conclusions

On the other hand, the piles in fixed-head 6-pile group (Test PG5) behaved like
single piles in term of axial force and bending moment, except that the magnitude is
affected by the total number of piles. In contrast, there was an increment in the pile
bending moment due to the totally fixed-head condition, which restrained the pile
movement that resulted in high bending moment.

7.2

RECOMMENDATIONS FOR FUTURE STUDIES

The findings from the centrifuge experiments in the present study provided the basis
for the understanding of tunnel-soil-pile interaction. Some possible areas that could be
explored further are discussed here:

In the present study, the smallest volume loss that was modeled in the
centrifuge test was 3%. However, with recent advancement in tunnelling
technology, the volume loss can be controlled to less than 1%. Hence, an
improvement to the current model tunnel to a smaller volume loss is
recommended.

Future work is needed to study three-dimensional tunnel excavation in order to


study the longitudinal effects of tunnelling. To achieve this, modification of the
present two-dimensional model tunnel is needed. It is also recommended that
mechanical model tunnel with several small segments be developed to simulate
three-dimensional tunnel excavation.

288

Chapter 7 Conclusions

It is proposed that numerical analysis could be performed to validate the


centrifuge experimental results and parametric studies could be carried out to
improve the understanding of various effects of tunnelling.

The effects of soil strength on tunnelling-induced soil movement and pile


responses could be further explored. In the present study, normally
consolidated clay used was relatively soft, and it would be interesting to study
the responses of stiffer clays.

289

References

REFERENCES
Augarde, C. E. and Burd, H. J. 2001. Three-dimensional finite element analysis of
lined tunnels. International Journal for Numerical and Analytical Methods in
Geomechanics, 25, pp. 243-262.
Asoaka, A. 1978. Observational procedure of settlement prediction. Soils and
Foundations, Vol. 18 (4), pp. 87-101.
Balasubramanian, 1987. Construction of effluent outfall pipeline in tunnel, using earth
pressure balanced shield. Case Histories in Soft Clay, 5th Int. Geo.Seminar. pp17-36.
Bezuijen, A. and Schrier, J. V. D. 1994. The influence of a bored tunnel on pile
foundations. CENTRIFUGE 94, Singapore. Leung, Lee and Tan (eds)., pp.681-686.
Bilotta, E. and Taylor, R.N. 2005. Centrifuge modelling of tunnelling close to a
diaphragm wall.International Journal of Physical Modelling in Geotechnics , 27-41.
Bowers, K.H., Hiller, D.M. and New, B.M. 1996. Ground movements over three years
at the Heathrow Express Trial Tunnel. Proc. Int. Symposium on Geotechnical Aspects
of Underground Construction in Soft Ground, London (eds. R. J. Mair and R. N.
Taylor), Balkema, pp. 647-652.
Bransby, M. F. and Springman, S. M. 1997. Centrifuge modelling of pile groups
adjacent to surcharge loads. Soils and Foundations, Vol. 37, No.2, pp. 39-49.
BS8004. 1986. Code of practice for foundations. British Standards Institution, London.
Cham, W. M. 2007. The Response Of Piles To Tunnelling. Master of Science thesis
Imperial College of Science, Technology and Medicine, University of London.
Chambon, P., Corte, J.F., Ganier, J. and Konig, D. 1991. Face stability of shallow
tunnels in granular soils. Proc. Int. Conf. Centrifuge 1991, Boulder/Colorado, Ko, H.
Y., Mc Leau, F.G. (edit), Balkema, pp.99-105
Chen, L. T., Poulos, H. G. and Loganathan, N. 1999. Pile responses caused by
tunneling. Journal of Geotechnical and Geoenvironmental Engineering, Vol. 125, No.
3, pp. 207-215.
Cheng, C. Y. 2003. Finite element study of tunnel-soil-pile interaction. M Eng thesis,
National University of Singapore.
Chow, Y. K. and Yong, K. Y. 1996. Analysis of piles subject to lateral soil movements.
Journal of The Institution of Engineers, Singapore, Vol. 36, No. 2, pp. 43-49.
Dyer, M. R., Hutchinson, M. T. and Evans, N. 1996. Sudden Valley Sewer: a case
history. Proc. Int. Symposium on Geotechnical Aspects of Underground Construction
in Soft Ground, London (eds. R. J. Mair and R. N. Taylor), Balkema, pp. 671-676.

290

References

Emeriault, F., Bonnet-Eymard, T., Kastner, E.R., Vanoudheusden, Petit, G., Robert
J.Y., de Lamballerie and Reynaud, B. 2005. Ground movements induced by EarthPressure Balanced, Slurry Shield and Compressed-Air tunneling techniques on the
Toulouse subway line B, Underground Space Use: Analysis of the Past and Lessons
for the Future (eds. Erdem and Solak), Taylor and Francis Group, pp 841
Feng, S. H. 2003. Centrifuge modeling of tunnel-pile interaction. M.Eng thesis,
National University of Singapore.
Fujita, K. 1981. On the surface settlements caused by various methods on shield
tunneling. Proc. XIth Int. Conf. On Soil Mechanics and Foundations Engineering, Vol.
4, pp. 609-610.
George. 1981. Lost-ground subsidence in two shallow tunnels, Soft-Ground Tunneling,
Failures and Displacement.
Ghahremannejad, B.,Surjadinata, J.,Poon, B. and Carter, J.P. 2006. Effects of
tunnelling on model pile foundations. Physical Modelling in Geotechnics 6th
ICPMG 06 Ng, Zhang and Wang (eds) Taylor and Francis Group, London. pp.
1157-1162
Goh, T.L. 2003. Stabilisation of an excavation by an embedded improved soil layer.
PhD thesis, National University of Singapore.
Glossop, N. H. 1978. Ground movements caused by tunneling in soft soils. PhD thesis,
University of Durham.
Grant, R. J. and Taylor, R. N. 2000. Tunneling-induced ground movements in clay.
Geotechnical Engineering, Proc. Institutions of Civil Engineers, Vol. 143, No. 1, pp.
43-55.
Hergarden, H. J. A. M., Poel, J. T. and Schrier, J. S. 1996. Ground movements due to
tunnelling:Influence on pile foundations. 2nd International Symposium on
Geotechnical Aspects of Underground Construction in Soft Ground, Mair and Taylor
(eds), pp. 519-524.
Howland, A. F. 1980. The prediction of the settlement above soft ground tunnels by
considering the groundwater response with the aid of flow net constructions. Ground
Movements and Structures, Proc. 2nd International Conference, The University of
Wales Institute of Science and Technology, (eds. J. D. Geddes et al), Pentech Press
1981, 345-358.
Jacobsz, S.W. 2002. The effects of tunnelling on pile foundations, PhD thesis,
University of Cambridge.
Jacobsz, S. W., Standing, J. R., Mair, R. J., Soga, K., Hagiwara, T. and Sugiyama, T.
2001. The effects of tunneling near single driven piles in dry sand. Proc. of Asian
Regional Conf. on Geotechnical Aspects of Underground Construction in Soft Ground,
Tongji University Press, Shanghai, pp. 29-35.

291

References

Jacobsz, S.W., Standing, J.R. and Mair, R.J. 2004. Tunnelling effects on pile groups in
sand. Proc. Advances in geotechnical engineering: The Skempton Conference, ICE,
Vol 2, 1056-1067.
Jacobsz, S. W., Bowers,K.H. and Moss, N.A. and Zanardo,G. 2005. The effects of
tunnelling on piled structures on the CTRL, Proc 5th International Symposium on
Geotechnical Aspects of Underground Construction in Soft Ground. Balkema,
Amsterdam.
Kaalberg,F.J. and Teunissen,E.A.H., Tol,A.F.van and Bosch, J.W. 2005. Dutch
research on the impact of shield tunnelling on pile foundations. 16th International
Conference on Soil Mechanics and Geotechnical Engineering, Osaka, pp 1615-1620
Kimura, T. 1998. Development of Geotechnical centrifuge in Japan. Proc. Centrifuge
98, Tokyo, Pre-print volume, pp. 23-32.
Komiya, K., Soga, K., Akagi H, Hagiwara, T. and Bolton, M. 1999. Finite element
modeling of excavation and advancement processes of a shield tunnlling machine.
Soils and Foundations, Vol. 39, No. 3, pp.37-52.
Konig, D.,Guttler, U. and Jessberger, H.L. 1991. Stress redistributions during tunnel
and shaft constructions. Proc.of the Int. Conf.Centrifuge 1991, Boulder/Colorado, Ko,
H.Y., Mc Leau, F.G.(edit), Balkema, pp.129-135.
Konig, D. 1998. An inflight excavator to model a tunnelling process. Proceeding of
Centrifuge 98, 1998, Rotterdam. Pp707-712.
Lade, P.V., Jessberger, H.L., Kakowski, E. and Jordan, P. 1981. Modelling of deep
shafts in centrifuge tests. Proc. 10th ICSMFE, Stockholm, Vol.1, pp.683-692.
Lake, L. M., Rankin, W. J. and Hawley, J. 1992. Prediction and effects of ground
movements caused by tunneling in soft ground beneath urban areas. CIRIA Project
Report 30, Construction Industry Research and Information Association, London.
Lee, C. J. and Chiang, K. H. 2004. Load transfer on single pile near new tunnelling in
sandy ground. Engineering Practice and Performance of Soft Deposits, IS-OSAKA, pp.
501-506.
Lee, C.J., Wi, B.R., and Chiou, S.Y. 1999. Soil movements around a tunnel in soft
soils. Proc. Natl.Sci.Counc. ROC(A), Vol.23, No.2, pp. 235-247.
Lee, F.H. 1992. The National University of Singapore Geotechnical Centrifuge-users
manual. Research Report No.CE001. Department of civil enginnering, National
University of Singapore.
Lee, F. H. 1992. The National University of Singapore Geotechnical Centrifuge-Users
Manual. Research Report No. CE001. Department of Civil Engineering, National
University of Singapore.

292

References

Lee, F. H., Phoon, K. K., and Lim, K. C. 2006. Large Scale Three-Dimensional Finite
Element Analysis of Underground Construction. Numerical Modelling of Construction
Processes in Geotechnical Engineering for Urban Environment. pp.155-163.
Lee, F. H., Tan, T. S., Leung, C.F., Yong, K.Y., Karunaratue, G. P. and Lee, S. L.
1991. Development of geotechnical centrifuge facility at the National University of
Singapore. Proc., Int. Conf. Centrifuge 91, Boulder, USA, 11-17.
Lee, G. T. K. and Ng, C. W. W. 2005. Effects of advancing open face tunneling on an
existing loaded pile. Journal of Geotechnical and Geoenvironmental Engineering,
ASCE, Vol. 131, No. 2, pp. 193-201.
Lee, K. M., Rowe, R. K. and Lo, K. Y. 1992. Subsidence due to tunneling: Part I
Estimating the gap parameter. Canadian Geotechnical Journal, Vol. 29, No. 5, pp. 929940.
Lee, Yongjoo and Yoo, Chungsik. 2006. Behaviour of a Bored Tunnel adjacent to a
Line of Loaded Piles Proceedings of the World Tunnel Congress and 32nd ITA
Assembly, Seoul, Korea, 2227 April 2006.
Leung, C.F., Lee, F.H. and Tan, T.S. 1991. Principles and application of geotechnical
centrifuge model testing. Journal of Institution of Engineers, Singapore, Vol.31, No.4,
1991, pp39-45.
Leung, C. F., Lim, J. K., Shen, R. F. and Chow, Y. K. 2003. Behaviour of pile groups
subject to excavation-induced soil movement. Journal of Geotechnical and
Geoenvironmental Engineering, Vol. 129, No. 1, pp. 58-65.
Leung, C. F 2006 Performance of piles subject to soil movements. Physical Modelling
in Geotechnics 6th ICPMG 06 Ng, Zhang and Wang (eds) Taylor and Francis
Group, London. pp. 87-98.
Leung, C. F., Ong, D.E.L. F. and Chow, Y. K. 2006. Pile behavior due to excavationinduced soil movement in clay II: Collapsed Wall. Journal of Geotechnical and
Geoenvironmental Engineering, Vol. 132, No. 1, pp. 45-53.
Lim K. C. 2003. Three-dimensional finite element analysis of earth pressure balance
tunnelling. PhD thesis, National University of Singapore.
Lim, J. K. 2001. Behaviour of piles subject to excavation-induced soil movement.
M.Eng Thesis, National University of Singapore.
Lin, D. G., Tseng, C. T., Phienwej, N. and Suwansawat, S. 2002. 3-D deformation
analysis of earth pressure balance shield tunneling in Bangkok subsoil. Journal of
Southeast Asian Geotechnical Society, April 2002, pp.13-27.
Loganathan, N. 1999. Effect of tunnelling adjacent to pile foundation. PhD thesis, The
University of Sydney.

293

References

Loganathan, N. and Poulos, H. G. 1998. Analytical prediction for tunneling-induced


ground movements in clays. Journal of Geotechnical and Geoenviromental
Engineering, Vol. 124, No. 9, pp. 846-856.
Loganathan, N., Poulos, H. G. and Stewart, D. P. 2000. Centrifuge model testing of
tunneling induced ground and pile deformations. Geotechnique, Vol. 50, No. 3, 283294.
Loganathan, N., Poulos, H. G. and Xu, K. J. 2001. Ground and pile-group response
due to tunneling. Soils and Foundations, Vol. 41, No. 1, pp. 57-67.
LTA. 2009. Civil design criteria for road and rail transit systems revision A7, Rail and
Engineering Group, Land Transport Authority, Singapore.
Mair, R. J. 1979. Centrifugal modeling of tunneling construction in soft clay. PhD
Thesis, University of Cambridge.
Mair, R. J., Philips, P., Schofield, A.N. and Taylor, R.N. 1984. Application of
centrifuge modeling to the design of tunnels and excavations in soft clay.
Proc.Int.Symposium on application of centrifuge modeling to geotechnical design.
Manchester, Craig, W.H. (edit),pp.356-366.
Mair, R. J., Taylor, R. N. and Bracegirdle, A. 1993. Subsurface settlement profiles
above tunnels in clay. Geotechnique, Vol. 43, No. 2, pp. 315-320.
Mair, R.J. and Taylor R.N. 1997. Bored tunnelling in the urban environment. State-ofthe-art Report and Theme Lecture, Proceedings of 14th International Conference on
Soil Mechanics and Foundation Engineering, Hamburg, Balkema, Vol.4., 2353-2385.
Melis, M., Medina, L. and Rodriguez, J. 2002. Prediction and analysis of subsidence
induced by shield tunneling in the Madrid Metro extension. Canadian Geotechnical
Journal, 39, pp.1273-1287.
Moh, Z-C., Ju, D. H. and Hwang, R. N. 1996. Ground movements around tunnels in
soft ground. Proc. Int. Symposium on Geotechnical Aspects of Underground
Construction in Soft Ground, London (eds. R. J. Mair and R. N. Taylor), Balkema, pp.
725-730.
Mhroueh H and Shahrour I. 2002. Three-dimensional finite element analysis of the
interaction between tunneling and pile foundations, International Journal for
Numerical and Analytical Methods in Geomechanics, Vol.26, pp.217-230.
Ng, C.W.W. and Lee, G.T.K. 2005. Three Dimensional Ground Settlements and Stress
Transfer Mechnisms due to Open Face Tunnelling. Canadian Geotechnical Journal.
Vol. 42, No. 4, pp. 1015-1029.

Ng, C.W.W. , Lee, G.T.K. and Tang, D.K.W. 2004. 3D Numerical Investigations of
NATM Twin Tunnel Interactions. Canadian Geotechnical Journal. Vol. 41, No. 3, pp.
523-539.

294

References

Ng, C.W.W., Springman, S.M. and Norrish, A.R.M. 1998, Centrifuge modelling of
spread-base ntegral bridge abutments, Journal of Geotechnical and Environmental
Engineering, ASCE, Vol. 124, No. 5, pp. 376-388.
Ng, R.M.C., Lo, K.Y., and Rowe, R.K. 1986. Analysis of Field Performance The
Thunder Bay Tunnel Canadian Geotechnical Journal Vol. 23, pp 30-50.

Nomoto, T., Mito, K., Imamura, S., Ueno, K. and Kusakabe, O. 1994. A miniature
shield tunneling machine for a centrifuge. Proc. Int. Conf. Centrifuge 1994, Singapore,
Leung, C. F., Lee, F. H., Tau, ET. S (edit) Balkema, pp. 699-704.
Ong, D.E.L. 2005. Pile behaviour subject to excavation-induced soil movement in clay.
PhD thesis, National University of Singapore.
OReilly, M. P. and New, B.M. 1982. Settlements above tunnels in the United Kindom
their magnitude and prediction. Tunneling 82, London, IMM, pp 173-181.
OReilly, M. P., Mair, R. J. and Alderman, G. H. 1991. Long-term settlements over
tunnels; an eleven year study at Grimsby, Tunneling 91, London, IMM, pp. 55-64.
Osman, A. S., Mair, R. J. and Bolton, M. D. 2006a. On the kinematics of 2D tunnel
collapse in undrained clayGeotechnique 56, No. 9, 585595.
Osman, A. S., Bolton, M. D. and Mair, R. J. 2006b. Predicting 2D ground movements
around tunnels in undrained clay Geotechnique 56, No. 9, 597604.
Pang, C.H. 2006. The Effects Of Tunnel Construction On Nearby Pile Foundation PhD
thesis, National University of Singapore.
Pang, C.H., Yong, K.Y., Chow,Y.K. and Wang, J. 2005a. The response of pile
foundations subjected to shield tunnelling. 5th International Symposium Geotechnical
Aspects of Underground Construction in Soft Ground, 15-17 June 2005, Amsterdam.
Pang, C.H., Yong, K.Y., and Chow,Y.K. 2005b. Three-dimensional numerical
simulation of tunnel advancement on adjacent pile foundation. Proceedings of the 31st
ITA-AITES World Tunnel Congress, Underground Space Use: Analysis of the Past
and Lessons for the Future, 7-12 May 2005, Istanbul, Turkey .
Park, K. H. 2004. Elastic solution for tunneling-induced ground movements in clays.
International Journal of Geomechanics Vol. 4, No. 4, pp. 310-318.
Park, K. H. 2005. Analytical solution for tunneling-induced ground movement in clays.
Tunneling and Underground Space Technology Vol. 20, pp. 249-261.
Peck, R. B. 1969. Deep excavations and tunneling in soft ground. Proc. 7th
International Conference Soil Mechanics and Foundation Engineering, Mexico City,
State of the Art Volume, pp. 225-290.

295

References

Phoon, K. K., Lee, F. H. and Chan, S. H. 2006. Iterative Solution of Intersecting


Tunnels Using The Generalised Jacobi Preconditioner. Numerical Modelling of
Construction Processes in Geotechnical Engineering for Urban Environment. pp.155163.
Potts, D. M. 1976. Behaviour of lined and unlined tunnels in sand. PhD Thesis,
University of Cambridge.
Randolph, M.F. and Wroth, C.P. 1978 Analysis of deformation of vertiocally loaded
piles. J.Geot.Eng.Div.,ASCE, Vol.104, No.12, pp.1465-1488.
Ran, X., Leung, C. F. and Chow, Y.K. 2003. Centrifuge modelling of tunnel-pile
interaction in clay. Proc. Underground Singapore, Singapore, 256-263.
Ran, X. 2004. Tunnel pile interaction in clay. M Eng thesis, National University of
Singapore.
Rowe, R. K., Lo, K. Y. and Kack, G. J. 1983. A method of estimating surface
settlement above tunnels constructed in soft ground. Canadian Geotechnical Journal,
Vol. 20, No. 8, pp. 11-22.
Sagaseta, C. 1987. Analysis of undrained soil deformation due to ground loss.
Geotechnique, Vol. 37, No. 3, pp. 301-320.
Selemetas, D., Standing, J.R. and Mair, R.J. 2005. The responss of full-scale piles to
tunnelling, Proc 5th International Symposium on Geotechnical Aspects of Underground
Construction in Soft Ground. Balkema, Amsterdam.
Schmidt, B. 1969. Settlements and ground movements associated with tunnelling in
soil. PhD thesis, University of Illinois.
Schmidt, B. 1989. Consolidation settlement due to soft ground tunnelling Proc. XII Int.
Conf. On Soil Mechanics and Foundations Engineering, pp. 797-800.
Schofield, A.N. 1980. Cambridge geotechnical centrifuge operations. Geotechnique,
Vol. 30(3), pp.227-268.
Schofield, A.N. 1998, Geotechnical centrifuge development can correct a soil
mechanic error. Proc. Centrifuge 98, Tokyo, Preprint volume, pp. 1-8.
Sharma, J. S., Bolton, M. D. and Boyle. R. E. 2001. A new technique for simulation of
tunnel excavation in a centrifuge. Geotechnical Testing Journal, Vol. 24, No. 4, pp.
343-349.
Shen R.F. 2008, Negative skin friction on single piles and pile groups. PhD thesis,
National University of Singapore.
Shirlaw, J. N. 1993, Pore pressures around tunnels in clay: Discussion. Canadian
Geotechinical Journal. Vol. 30, pp. 1044-1046.

296

References

Shirlaw, J. N. 1995. Observed and calculated pore pressures and deformations induced
by an earth pressure balance shield: Discussion. Canadian Geotechinical Journal. Vol.
32, pp. 181-189.
Shirlaw, J.N., Busbridge,J.R. and Yi.,X. 1994. Consolidation settlements over tunnels :
A review. Canadian Geotechnical Journal, 253-265.
Shirlaw, J. N., Ong, J. C. W, Rosser, H. B., Tan, C. G., Osborne, N. H. and Heslop, P.
E. 2003. Local settlements and sinkholes due to EPB tunnelling. Proceedings of the
Institution of Civil Engineers: Geotechnical Engineering. Issue GE4, pp.193-211.
Stewart, D. P. and Randolph, M. F. 1991. A new site investigation tool for the
centrifuge. Proc. Int. Conf. Centrifuge 91, Colorado, pp. 531-538.
Sven Moller. 2006. Tunnel Induced Settlement and Structural Forces in Lining.. PhD
thesis, University Stuttgart.
Tan, T.S. and R.F. Scott .1985. Centrifuge scaling considerations for fluid-particle
systems. Geotechnique, Vol. 35(4), pp. 461-470.
Take, W.A. and Bolton, M.D., 2004. Identification of seasonal slope behaviour
mechanisms from centrifuge case studies. Proc. Skempton Memorial Conference,
London. Vol. 2 pp. 992-1004.
Taylor, R. N. 1995. Centrifuges in modelluing: principles and scale effects.
Geotechnical Centrifuge Technology, Blackie Academic and Professional, London, pp.
19-59.
Taylor, R. N. 1998. Modelling of tunnel behaviour.Proc. Institutions of Civil Engineers,
Vol. 131, pp. 127-132.
Verrujit, A. and Booker, J. R. 1996. Surface settlements due to deformation of a tunnel
in an elastic half plane. Geotechnique, Vol. 46, No. 4, pp. 753-756.
Verrujit, A. and Booker, J. R. 2000. Complex variable analysis of Mindlins tunnel
problem. Proceedings of the Developments in Theoretical Geomechanics, 2000, pp. 322.
White, D.J. and Take, W.A., 2002. GeoPIV: Particle Image Velocimetry (PIV)
software for use in geotechnical testing. Cambridge University Engineering
Department Technical Report D-SOILS-TR322.
White, D.J., Take, W.A. and Bolton, M.D.2003.Soil deformation measurement using
particle image velocimetry (PIV) and photogrammetry. Geotechnique ,
Vol.53,No.7,619-631.
Xu, K.J. and Poulos, H.G. 2000 General elastic analysis of piles and pile group, Int.J.
For Numerical and Analytical Methods in Geomechanics Vol.24 Issue15 pp 1109.

297

References

Yann Leblais and Alain Bochon. 1991. Villejust Tunnel: Slurry shield effects on soil
and lining behaviour and comments on monitoring requirement. Tunneling 91, 1991.
Yashuhiro Katoh, Michio Miyake and Masato Wada. 1998. Ground deformation
around shield tunnel. Proceeding of Centrifuge 98, 1998, Rotterdam.
Zhang DM, Huang HW Hicher PY 2004 Numerical Prediction of Long-term
settlements over Tunnels in Clay, ITA 2004
Zhang,Y.D., Tan,T.S. and Leung, C.F. 2005.Application of particle imaging
velocimetry (PIV) in centrifuge testing of uniform clay. International Journal of
Physical Modelling in Geotechnics , 15-26.

298

You might also like