278
NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION LIU KAO XUE NATIONAL UNIVERSITY OF SINGAPORE 2003

NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

  • Upload
    others

  • View
    5

  • Download
    0

Embed Size (px)

Citation preview

Page 1: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

LIU KAO XUE

NATIONAL UNIVERSITY OF SINGAPORE

2003

Page 2: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

LIU KAO XUE B. Eng (Tsinghua), M. Eng (XAUT), M. Eng (NUS)

A THESIS SUBMITTED FOR

THE DEGREE OF DOCTOR OF PHILOSOPHY

DEPARTMENT OF CIVIL ENGINEERING

NATIONAL UNIVERSITY OF SINGAPORE

2003

Page 3: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Dedicated to my wife, Cao Qiong and my two daughters, Liu Qian

and Liu Jia Xin, Jasmine

Page 4: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Acknowledgment

The author is greatly indebted to his supervisors, Prof. K Y Yong and Assoc. Prof. F H

Lee, for their tremendous contributions of ideas to the thesis through many beneficial

discussions and suggestions. In addition, the helpful discussions with, and suggestions

from, Prof. Y K Chow and Dr S H Chew are appreciated.

The author also thank Mr. Peter Lee and Mr. Y S Yoong, of School of The Built

Environment and Design, Singapore Polytechnic, for their understanding,

encouragement and support throughout the course of this study.

Page 5: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Summary

Deep excavation is common in urban development especially in land scarce city like

Singapore. It is important to assess the adverse effects of deep excavations to nearby

structures and thus design adequate supporting system to minimize the damage to such

structures. Ground anchors and tiebacks are used as part of the supporting system in

deep excavations. A novel FEM element for modelling the anchor-soil interaction was

formulated and developed during the course of this research work. The proposed

element was constructed by wrapping interface element around a beam which

represents the solid inclusion. The element stiffness matrix was derived based on the

internal equilibrium of interface and beam with no prescribed shape function for nodes

on beam. This constitutes a major improvement to the conformity of the Linker

element which is frequently used for modelling soil nails.

Closed form solutions of elastic and elastoplastic anchor-soil interaction were

developed to verify the correctness of the theoretical frame work and the program

algorithm. The advantages of the proposed element over the conventional element,

such as bar element, and limitations of the proposed element for modelling solid

inclusions in soil were demonstrated through a case study of an axially loaded pile.

The capability of the proposed element in capturing the salient phenomena of

anchor-soil interaction was further explored through a numerical simulation of pull-out

test for anchor in residual soil and granite. It was found through this case study that the

hysteresis loop of the load-displacement curve from field test for the anchor can be

effectively replicated using the proposed element. The typical segments in the

load-displacement curve can be easily explained based on the anchor-soil interaction

Page 6: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

that takes place in both bonded length and the unbonded length.

The performance of the proposed element was further tested in an excavation in sand

supported with soldier piles, timber lagging and tiebacks. It was found that the

proposed element can reasonably model the mechanism of anchor-soil interaction and

achieved reasonable agreement with the measured deflection of a soldier pile. It was

also noted that the slippage in the interface between anchor and soil explained better

the performance of the anchor. The significance of modelling the anchor unbonded

length with the proposed element was also highlighted through this case study.

The proposed element was also applied to an analysis of deep excavation in soft

Singapore marine clay supported with heavy metal sheet-pile wall, three levels of

internal struts and three levels of ground anchors. The performance of the proposed

element in this full scale excavation was assessed and compared with the conventional

bar element. It was found that the drag force along the anchor unbonded length which

penetrated through soft marine clay had significant localized influence on the ground

movement and which cannot be modelled using the conventional friction free bar

element.

Proposals are also presented for further development and applications of the proposed

element to many other engineering problems, such as soil nail system, tunnels using

NATM with rock bolts in soft ground and reinforced earth structures.

Key Words : Finite element, numerical model, anchor, soil, interaction,

excavation

Page 7: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

TABLE OF CONTENTS

Title page

Dedication

Acknowledgement

Summary

Table of Contents i

List of Tables viii

List of Figures ix

Notations xv

CHAPTER 1 – INTRODUCTION

1.1 Background 1

1.2 Objectives of the research 2

1.3 Scope of the study 3

CHAPTER 2 – LITERATURE REVIEW

2.1 Overview 6

2.2 Construction aspects of deep excavation 6

2.3 Design considerations and design methods 8

2.3.1 Empirical and semi-empirical approaches 9

2.3.2 Numerical analyses of supported excavation 12

2.3.3 The 3D numerical analysis of excavation 14

2.3.4 Drainage conditions and ground water draw down 15

2.4 Experimental methods 17

i

Page 8: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

2.5 Soil-structure interface 17

2.6 Numerical modeling of soil-structure interface 18

2.7 Numerical modeling of soil reinforcement / soil nail 21

2.8 FEM model for modeling of bond-slip in reinforced concrete 24

2.9 Interface properties, Ks and Kn 27

2.10 Soil-pile interaction during supported excavations 28

CHAPTER 3 – DEVELOPMENT OF A FIVE NODED ANCHOR-INTERFACE

ELEMENT

3.1 Background 31

3.2 The construction and the formulation of the anchor-interface element 31

3.2.1 The Equilibrium in axial direction 32

3.2.2 The Elastic formulation of displacement of anchor in axial direction

33

3.2.3 Element stiffness for axial displacement 35

3.2.4 The Equilibrium and anchor displacement in transverse

direction 37

3.2.5 Element stiffness for flexural degrees of freedom 40

3.3 Global stiffness matrix--coordinate system transformation 42

3.4 Body forces on anchor-interface element 44

3.5 Stresses, strain, axial force, bending moment and shear force in

an element 46

3.6 Verification studies 47

3.6.1 Anchor displacement in axial direction 48

3.6.2 Anchor displacement in flexure 49

ii

Page 9: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

3.7 Verification of the model against existing bar element 51

3.8 Summary of the chapter 52

CHAPTER 4 – NON-LINEAR ALGORITHMS FOR ANCHOR-INTERFACE

ELEMENT

4.1 Background 53

4.2 Proposed non-linear algorithms for anchor-interface element 53

4.2.1 Incremental tangent stiffness (ITS) approach 53

4.2.2 Total load secant iteration (TLI) approach. 55

4.3 Secant stiffness matrix for elastic perfectly-plastic

anchor-interface element 57

4.3.1 Axial stiffness 57

4.3.2 Flexural stiffness 60

4.4 Stresses, strains, internal forces and bending moment 62

4.5 Validation with idealised problems 63

4.5.1 Axially loaded anchor in rigid medium 63

4.5.1.1 Closed-form solution 63

4.5.1.2 Comparison of FE and closed-form solutions 66

4.5.2 Lateral loading on an anchor in a rigid medium 68

4.5.2.1 Closed-form solution 70

4.5.2.2 Comparison of FE and closed-form solution 72

4.6 Case study for an instrumented pile subject to axial load 73

4.6.1 Calibration of the element model and mesh accuracy 73

4.6.1.1 The effects of Ks 74

4.6.1.2 The mesh independency 75

iii

Page 10: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

4.6.2 Nonlinear analyses of an instrumented pile subject to axial load 77

4.7 Summary of the chapter 79

CHAPTER 5 – CASE STUDY 1 – ANCHORED PULL-OUT TEST

5.1 Introduction 81

5.2 Background and site geological conditions 81

5.3 Finite element model 83

5.4 Results and discussions 87

5.4.1 Effects of different methods of modelling anchors 89

5.4.2 Parametric study 93

5.4.2.1 The influence of in-situ stresses 93

5.4.2.2 Influence of rock/soil properties 94

5.4.2.3 Influence of interface properties 96

5.5 Summary of the results and discussions 98

CHAPTER 6 – CASE STUDY 2 – AN ANCHORED RETAINING WALL IN SAND

6.1 Introduction 99

6.2 Finite element mesh 100

6.3 Boundary conditions 101

6.4 Material properties and calculation parameters 102

6.5 Modelling of construction sequence 107

6.6 Results and discussions 109

6.6.1 Calibration of the FEM models 112

6.6.2 Comparison with published results 115

6.6.3 Sensitivity study-- The influence of the interface

properties of anchor 118

iv

Page 11: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

6.6.4 The 3D effects and 2D equivalency 119

6.6.5 Influence of different models for anchor unbonded length 126

6.7 Summary of the chapter 128

CHAPTER 7 – CASE STUDY 3 – ANCHORED EXCAVATION IN SOFT CLAY

7.1 Introduction 130

7.2 Site conditions and site geology 130

7.3 FEM mesh for the analyses 136

7.4 Material properties for the FE analyses 138

7.5 Simulation of construction sequences and activities 141

7.6 Calibration of FEM model and discussion on the results 143

7.6.1 Deformation 143

7.6.2 Excess pore water pressure 144

7.6.3 Ground anchors 145

7.7 Proposed element vs bar element representation of anchor 147

7.8 Summary of the chapter 149

CHAPTER 8 – CONCLUSIONS

8.1 Summary of research findings 151

8.2 Recommendations for future research 156

REFERENCES 159

v

Page 12: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

LIST OF TABLES

Table 4.1 Results from mesh C1A3: 8-noded element + slip elements Table 4.2 Results from mesh C1B using the proposed elements Table 4.3 Ground profile Table 4.4 Pile settlement along the shaft for coarse meshes Table 4.5 Pile settlement along the shaft for fine meshes Table 4.6 Soil properties Table 5.1 Summary of in-situ soil properties from soil investigation test Table 5.2 Material properties used for the analyses Table 5.3 List of the cases studied Table 5.4 In-situ stresses Table 6.1 Soil properties Table 6.2 Comparison of material properties used in the present study Table 6.3 The geometric/section properties of anchors Table 6.4 The material properties of anchors for 3D analyses Table 6.5 The material properties of anchors for 2D analyses Table 6.6 Interface properties (after Schnabel, 1982) Table 6.7 List of cases investigated in present study Table 7.1 Typical subsoil properties (After Parnploy, 1990) Table 7.2 Properties of internal struts (After Parnploy, 1990) Table 7.3 Properties of ground anchors (After Parnploy, 1990) Table 7.4 Construction sequences used in the numerical simulation Table 7.5 Soil properties from site investigation report (After Parnploy,1990) Table 7.6 Soil properties used in present study Table 7.7 Strut properties used in present study Table 7.8 Anchor/interface properties used in present study

vi

Page 13: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

LIST OF FIGURES Fig. 2.1 Apparent pressure diagrams for computing strut load (Peck, 1969) Fig. 2.2 Empirical chart of settlement adjacent to strutted excavation (O’Rouke

et. al., 1976) Fig. 2.3 Element model for soil nail proposed by Herrman and Zaynab (1978) Fig. 2.4 Element model for soil nail in FLAC (Itasca, 1996) Fig. 2.5 Element model for rock anchor interface by Kawamoto et. al. (1994) Fig. 2.6 Illustration of linker element and the application in RC modelling.

(Adapted from ASCE report, 1982) Fig. 3.1 The proposed anchor-interface element Fig. 3.2 Coordinate system transformation Fig. 3.3 Variation of axial displacement along an anchor Fig. 3.4 Variation of skin friction along an anchor Fig. 3.5 Model for beam on elastic medium Fig. 3.6 Finite element mesh of elastic beam on stiff soil foundation. Fig. 3.7 Comparison of lateral deflection with different relative anchor stiffness

ratio Fig. 3.8 FEM mesh for comparison of the anchor-interface element with bar

element Fig. 3.9 Distorted FEM mesh from analysis using the anchor-interface elements. Fig. 3.10 Distorted FEM mesh from analysis using 3-D Bar elements. Fig. 3.11 Relative displacement vs eigen value µ. Fig. 4.1 Typical model for stress-strain relationship of interface Fig. 4.2 Anchor in rigid ground subjected to axial load.

Fig. 4.3 Variation of anchor settlement along anchor height from closed-form solutions.

Fig. 4.4 Comparison of anchor displacement between proposed FEM and closed-form solutions.

Fig. 4.5 Comparison of skin friction between proposed FEM and closed-form solutions.

Fig. 4.6 Comparison of axial force between proposed FEM and closed form solutions.

Fig. 4.7 Anchor in rigid ground subjected to lateral loading. Fig. 4.8 Variation of deflection along anchor fixed length. Fig. 4.9 Variation of bending moment along fixed length of an anchor. Fig. 4.10 Variation of shear force along anchor fixed length Fig. 4.11 Comparison of pile head settlement from conventional FEM (C1A3)

and proposed Element (C1B) Fig. 4.12 Deformed mesh from FEA using different types of elements Fig. 4.13 Deformed meshes Fig. 4.14 Pile shaft settlement Fig. 4.15 Deformed meshes Fig. 4.16 Pile shaft settlement Fig. 4.17 Finite element mesh used for comparison with result from Cheung et.

al. Fig. 4.18 The load settlement curve for an instrumented pile Fig. 5.1 Site plan of the slope protection work (Courtesy, L&M) Fig. 5.2 Detail of the selected section. Fig. 5.3 Field measured load vs displacement curve from pull-out test

vii

Page 14: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Fig. 5.4 Mesh for analyses of pull-out test (Mesh 1) Fig. 5.5 Comparative mesh (Mesh 2) fro mesh dependency study Fig. 5.6 Deformed meshes at load=300kN (Displacement x 100) Fig. 5.7 In-situ stresses Fig. 5.8 Calculated anchor head displacement vs load curve where unbonded

length was modelled using friction free bar element. Fig. 5.9 Comparison of calculated anchor head displacement vs load curves

between analyses using the friction free bar element and those using proposed anchor-interface element for anchor unbonded length.

Fig. 5.10 Comparison with in-situ pull out test Bonded length: Ks =1.8x 104 kPa/m, Unloading Ks =1.5x 104 kPa/m, c=200kPa, φ=30°. Unbonded length: Ks =6.0x 103 kPa/m, Unloading Ks =5x 103 kPa/m, c=1.0kPa, φ=10°

Fig. 5.11 Typical load displacement curve from pull-out test. Fig. 5.12 Simulated hysteresis loop (load to 610kN before unloading)

Bonded length: Ks =5.0x 105 kPa/m, Unloading Ks =1.8x 106 kPa/m, c=85kPa, φ=20°. Unbonded length: Ks =5.0x 104 kPa/m, Unloading Ks =6x 104 kPa/m, c=10 kPa, φ=10°

Fig. 5.13 Friction along anchor Fig. 5.14 Influence of in-situ stress Fig. 5.15 Influence of elastic modulus of rock Fig. 5.16 Influence of Elastic modulus of soil Fig. 5.17 Deformation of soil at low elastic modulus Fig. 5.18 Influence of Ks for anchor bonded length Fig. 5.19 Influence of Ks for anchor unbonded length Fig. 5.20 Effects of c in unbonded length. Bonded length: Ks =1.8x 104 kPa/m,

Unloading Ks =1.5x 104 kPa/m, c=200kPa, φ=20°. Unbonded length: Ks =7.0x 103 kPa/m, Unloading Ks =5x 103 kPa/m, c=10 and15kPa, φ=1°

Fig 6.1 Elevation view of the Texas A & M University tieback wall

(after Briaud & Lim, 1999) Fig 6.2 Section view of the Texas A & M University tieback wall (after Briaud & Lim, 1999) Fig 6. 3 Soil profile Fig 6.4 Initial mesh Fig 6.5 Boundary conditions Fig. 6.6 Variation of elastic modulus of soil with depth Fig. 6.7 Installation of steel H soldier Pile Fig. 6.8 Zoomed-in view of soldier pile elements Fig. 6.9 Excavation of 1st Lift Fig. 6.10 Installation of timber lagging for the 1st lift Fig. 6.11 Zoomed-in view of timber lagging and soldier pile elements Fig. 6.12 Installation and pre-stressing of anchor row 1 Fig. 6.13 Zoomed-in view of the locked-in elements

Fig. 6.14 Excavation of 2nd Lift

Fig. 6.15 Installation of timber lagging for the 2nd lift

Fig. 6.16 Installation and pre-stressing of anchor row 2 anchor elements

Fig. 6.17 Activate the lock-in element

viii

Page 15: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Fig. 6.18 Zoomed-in view of the lock-in elements

Fig. 6.19 Excavation of the 3rd lift

Fig. 6.20 Installation of timber lagging for the 3rd lift

Fig. 6.21 Analysis area used by Briaud and Lim (1999)

Fig. 6.22 Mesh configuration and boundary conditions in Briaud and Lim’s study

Fig. 6.23 Mesh configurations used in calibration study (B3D1) based on the study of Briaud and Lim (1999)

Fig. 6.24 Pile deflection from calibration analysis

Fig. 6.25 Effects of mesh configurations

Fig. 6.26 Effects of element types for anchors

Fig. 6.27 Summary of pile deflections from calibration analyses

Fig. 6.28 Effects of coordinates updating on numerical results

Fig. 6.29 Effects of soil stiffness on pile deflections

Fig. 6.30 Parametric study on the influence of Ks

Fig. 6.31 Influence of interface cohesion

Fig. 6.32 Soldier pile deflection from 2D and 3D analyses

Fig. 6.33 Soldier pile and wall deflection from 2D and3D analyses

Fig. 6.34 Deflection profiles for 2D and 3D analyses at end of 1st lift excavation

Fig. 6.35 Deflection profiles for 2D and 3D analyses at end of stressing row 2 anchor

Fig. 6.36 Displacement vector fields at the end of stressing the row 1 anchor from 2D analysis using proposed element model (Displacement x 50).

Fig. 6.37 Displacement vector fields at the end of stressing the row 1 anchor from 3D analysis using proposed element model (Displacement x 50).

Fig. 6.38 Deformed mesh at the end of stressing the Row 1 anchor from 2D analysis using proposed element model (Displacement x 50).

Fig. 6.39 Deformed mesh at the end of stressing the Row 1 anchor from 3D analysis using proposed element model (Displacement x 50).

Fig. 6.40 Displacement vector fields at the end of final excavation from 2D analysis using proposed element model (Displacement x 50).

Fig. 6.41 Deformed mesh at the end of final excavation from 2D analysis using proposed element model (Displacement x 50).

Fig. 6.42 Displacement vector fields at the end of final excavation from 3D analysis using proposed element model (Displacement x 50).

Fig. 6.43 Deformed mesh at the end of final excavation from 3D analysis using proposed element model (Displacement x 50).

Fig. 6.44 Plan view of principle stresses at end of excavation on plane Yo=-1.5 Fig. 6.45 Plan view of principle stresses at end of excavation on plane Yo=-3 Fig. 6.46 Plan view of principle stresses at end of excavation on plane Yo=-4.5 Fig. 6.47 Plan view of principle stresses at end of excavation on plane Yo=-6 Fig. 6.48 Plan view of principle stresses at end of excavation on plane Yo=-7.6 Fig. 6.49 The distribution of skin shear along row 1 at end of stressing row 1

anchor.

ix

Page 16: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Fig. 6.50 The distribution of skin shear along row 1 at different construction stages (3-D).

Fig. 6.51 The distribution of skin shear along row 1 at different construction stages(2-D).

Fig. 6.52 Deflection profiles from analyses using different model in 3D analyses Fig. 6.53 Deflection profiles at the end of 1st lift excavation Fig. 6.54 Deflection profiles at the end of stressing row 1 anchor Fig. 6.55 Deflection profiles at the end of 2nd lift excavation Fig. 6.56 Deflection profiles at the end of stressing row 2 anchor Fig. 6.57 Displacement vector fields at the end of final excavation from 3D

analysis using bar element model (Displacement x 50). Fig. 6.58 Ground settlement Fig. 6.59 Plan view of principle stresses from analysis using bar element on

yo=-1.5 Fig. 6.60 Plan view of principle stresses from analysis using bar element on

yo=-3.0 Fig. 6.61 Plan view of principle stresses from analysis using bar element on

yo=-4.5 Fig. 6.62 Plan view of principle stresses from analysis using bar element on

yo=-6.0 Fig. 6.63 Plan view of principle stresses from analysis using bar element on

yo=-7.6 Fig. 7.1 Location map of the site (After Parnploy,1990) Fig. 7.2 Site plan (After Parnploy,1990) Fig. 7.3 Ground profile (After Parnploy,1990) Fig. 7.4 The 2D mesh at initial stage (AX1CA)

Fig. 7.5 The comparative mesh (AX1CB) for determining stand off distance

Fig. 7.6 Comparison of wall defection for different stand-off distances

Fig. 7.7 The comparative mesh (AX1CC) for mesh dependency study

Fig. 7.8 Mesh AX1CD for mesh dependency study

Fig. 7.9 Wall defection at the end of excavation

Fig. 7.10 3D mesh for present study

Fig. 7.11 Uneven excavation of 3D Mesh

Fig. 7.12 The 3-D mesh at the end of the final excavation Fig. 7.13 The 3-D mesh at the end of the final excavation with fixed boundary

conditions shown Fig. 7.14 Fluid flow boundary condition at initial stage Fig. 7.15 Fluid flow boundary after excavation of 1st lift Fig. 7.16 Fluid flow boundary after 2nd lift excavation Fig. 7.17 Fluid flow boundary after at the last stage of construction Fig. 7.18 in-situ stress profile assumed in present study Fig. 7.19 Simulated construction sequences Fig. 7.20 The 3D mesh after introducing the overburden of river bank Fig. 7.21 The 3D mesh after installation of anchors Fig. 7.22 Wall deflection profile from proposed analysis Fig. 7.23 Displacement vector field (Scaled by 20 times) and excess pore water

pressure contours after the pumping of water within cofferdam

x

Page 17: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Fig. 7.24 Displacement vector field (Scaled by 20 times) and excess pore water pressure contours after the excavation of top two lifts.

Fig. 7.25 Displacement (Scaled by 20 times) after excavation of top 3 lifts Fig. 7.26 Displacement (Scaled up by 20 times) at end of construction Fig. 7.27 The excess pore water pressure contours after introducing the

overburden of river bank Fig. 7.28 The excess pore water pressure contours after pumping water within

cofferdam Fig. 7.29 The excess pore water pressure contours after excavation of top two

lifts Fig. 7.30 The excess pore water pressure contours after excavation of top three

lifts Fig. 7.31 The excess pore water pressure contours after excavation of top 5-th

lifts Fig. 7.32 The excess pore water pressure contours at the end of analysis Fig. 7.33 Development of anchor head deformation (front column anchors) Fig. 7.34 Development of anchor head deformation (back column anchors) Fig. 7.35 Skin friction along anchor tendons Fig. 7.36 Comparison of skin friction along anchor tendons for different element

models Fig. 7.37 Wall deflection profiles from analysis using different element types Fig. 7.38 Comparison of wall deflection profiles at key construction stages (right

bank) Fig. 7.39 Comparison of wall deflection profiles at key construction stages (Left

bank) Fig. 7.40 Incremental displacement field by analysis using bar element

(Displacement x 100 at t=211 days) Fig. 7.41 Incremental displacement field by analysis using proposed element

(Displacement x 100 at t=211 days) Fig. 7.42 Ground movement

xi

Page 18: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Notation A = Cross-section area.

[A] = A coefficients matrix for the displacement shape function.

b = Breadth of the cross-section of a beam.

[B] = Strain transforming matrix for displacement.

[Bp] = Gradient Transforming matrix for excessive pore pressure.

c or c’ = Cohesion (effective cohesion).

[C] = A coefficients matrix for the displacement shape function.

[D] = Matrix for Elastic stress-strain relationship (Hooker's Law).

[D]ep = Matrix for Elastoplastic stress-strain relationship.

E = Young's modulus.

e = Voids ratio of soil.

[F] = A constant matrix as a function of eigen values.

f = Yield function.

[G] = A variable matrix as a derived shape function.

g = Gravitational acceleration.

[H] = A variable matrix as a function of [L], [F] and [C].

I = Moment of inertia of a beam section.

[J] = A variable matrix as a function of [F] and [C].

[K] = Stiffness matrix for displacement analysis.

Ko = Coefficient of earth pressure at rest

Ks = Tangential (Shear) stiffness of Interface.

Ksur = Tangential (Shear) stiffness of Interface for unloading.

Ksre = Residual tangential (Shear) stiffness of Interface.

xii

Page 19: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Kn = Normal stiffness of Interface.

l = Length

[L] = A variable matrix as a function of x.

[N] = Shape function matrix for displacement.

M = Bending moment in a beam.

[M] = A constant matrix as a function of eigen values.

P = Load, or, axial force in a beam.

Q = Shear force in a beam.

R = A column matrix (vector) for system load.

r = radius of the cross-section of a beam.

[S], [Π] = A constant matrix as a function of eigen values.

[T] = Matrix for coordinate system transformation.

[V] = A variable matrix as a function of eigen value and x.

[X] = A constant matrix as quadratic function of x.

u = Displacement in X direction.

v = Displacement in Y direction.

w = Displacement in Z direction.

x,y,z = Space coordinates.

α, β = Constants derived from eigen values.

θ = rotation angle.

δ = Displacement.

∆ = Incremental value of a variable.

ε = Normal strain.

φ = Frictional angle.

xiii

Page 20: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

γ = Bulk density of material or sharing strain

µ, λ = eigen values.

ν = Poisson ratio.

ρ = Density of material.

σ = Normal stress.

τ = Shearing stress.

xiv

Page 21: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

CHAPTER 1

INTRODUCTION

1.1 Background

As a result of rapid urban redevelopment in land scarce cities like Singapore, many

construction activities such as deep excavations for basement of buildings,

underground sewage treatment plants, and tunnels for utilities and transport are

inevitably built at close proximity to existing structures. This often requires engineers

to assess the adverse impact of such construction activities on nearby structures. A

fundamental understanding of soil structure interaction is essential to this assessment.

“What are the factors with adverse impact on the existing structures due to the current

construction activities?” “How do these factors undermine the safety allowance, or

even the safe use, of the existing structures?” The answers to these questions are just as

important as knowing how to minimize the adverse effects to the existing structures

induced by current construction activities.

Soil-structure interaction problems and problems related to the supported excavation

system, tunneling, and many other geotechnical construction activities involves

appropriate modeling of the mechanical behaviour of interface between soil and

supporting structures. Soil-structure interface is generally referred to as the contact

zone between soil and structure. The response of soil-structure systems is influenced

strongly by the characteristics of interface. Numerical models, constitutive models and

laboratory methods of replicating the mechanical behaviour of soil-structure interface

are vital to the robust assessment of the soil-structure interaction.

1

Page 22: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Numerical methods such as Finite Element Methods (FEM) are widely used in the

study of soil-structure interaction as a mathematical tool. Successful applications of

such tools for deriving meaningful results depend on the robustness of the numerical

models, the rigor of the numerical simulation of construction activities, the

comprehensiveness of constitutive model and the reliable representation of material

properties and parameters. Robust numerical models should have appropriate element

models, that are able to capture the salient mechanical behaviour of the structure, such

as shell element for modeling the flexural structure members in 3D analyses, slip

element for modeling the interface between soil and structures, and a number of other

element types for the modeling of the interaction between soil and solid inclusions

such as ground anchor, soil nail, earth reinforcement and geomembrane.

1.2 Objectives of the research

The main objective of this research work is to ascertain the soil-structure interaction

during the construction of supported excavation through numerical modeling. This

includes theoretical development and formulation of the FEM element model for

modeling anchor-soil interaction; the development of computer program to facilitate

the numerical analyses which involve soil-structure interactions; as well as

ascertaining the significance of the soil-structure interaction during supported

excavation through case studies. The program verification and validation are also

integral parts of the development in this research. The study hopes to provide technical

understanding to the design and construction of supported excavation system so as to

achieve safe designs and manageable ground movement control.

2

Page 23: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

1.3 Scope of the study

The research efforts were focused on the development of comprehensive and robust

FEM element models for ascertaining the anchor-soil interaction during the

construction of supported excavation, and rigorous simulation of construction activities

related to the excavation support system such as excavation, pre-loading of struts, or

pre-stressing of anchors/tieback, installation of struts/anchors, soldier piles, timber

laggings and so forth.

The software used in the study is built upon CRISP (Britto and Gunn, 1990), a

computer program developed initially by the Soil Mechanics Group at Cambridge

University. Refinements of the program were made with some re-development

including implementation of element types such as 3D interface element, which is

meant for modeling the interface between soil and supporting structure;

implementation of 16-noded thick shell element for modeling the flexural members of

the retaining structure such as diaphragm wall, sheet-pile wall, timber lagging tunnel,

concrete lining in 3D analysis. In particular, a special element, named

“anchor-interface element”, for modeling the interaction between soil and slender solid

inclusions such as anchor, soil nail, rock bolts and soil reinforcement, was formulated,

implemented and tested in this study.

The research work reported in this thesis covers the following:

(a) Literature review of past research publications in the field of soil-structure

interaction and deep excavation, with special emphasis on numerical analysis of

3

Page 24: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

excavation support system, is presented in Chapter 2. The literature review of

state-of-the-art developments of numerical modeling of the soil structure interface is

also included in this chapter.

(b) The theoretical formulation, program development and verification of 5-noded 3D

anchor-interface element are presented in Chapter 3 of this thesis. Closed form

solutions for some special cases of anchor-soil interaction were developed for the

purpose of verifying the computer program.

(c) The theoretical formulation, program development and verification of non-linear

algorithm for the proposed anchor-interface element are presented in Chapter 4. Closed

form solutions were also developed to verify the correctness of the proposed

algorithms and the computer program implementation.

(d) The theoretical formulation and the program code were further tested with reported

cases of in-situ pull out test data and full scale excavations supported with tie-backs

and ground anchor system. These case studies were conducted to investigate the ability

of the proposed element in the simulation of excavations supported with tie-back or

ground anchor system, and examine the performance of the proposed anchor-interface

element with different configurations and levels of complexity. As an integrated part

of the development work, comparative study with conventional ways of modeling

tie-back and ground anchors were made in these case studies to assess the superiority

of the proposed element over the conventional approach as well as to find out the

4

Page 25: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

limitations of the proposed element. This part of the work is included in Chapters 5, 6

and 7 respectively.

(e) The research work is finally summarized and proposals for future work in this

direction are outlined in Chapter 8.

The original work in this research lies in the theoretical formulation, computer

program development, and the closed form solutions of idealized cases of anchor soil

interaction for the purpose of program verification. In addition, the new findings from

the case studies are also original contributions to benefit the industry and academic

research.

5

Page 26: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

CHAPTER 2

LITERATURE REVIEW

2.1 Overview

Deep excavations are required in many civil engineering works such as basements of

buildings, road works, trenching for laying service cables and pipelines, underground

transport systems, underground water supply systems, drainage systems and waster

water systems. The main challenge for a geotechnical engineer is to achieve a safe,

economical and environmental friendly design with good serviceability so as to

minimize adverse impact on the surrounding.

The literature review presented in this chapter starts with a review of construction

aspects and design approaches as well as the numerical simulation of deep excavations.

It then zooms in to the areas related to the numerical modeling of interactions between

soil and solid inclusions, such as anchors, soil nails, reinforced earth and piles, during

excavations.

2.2 Construction aspects of deep excavation

The construction sequences of deep excavations in soft clay differ slightly depending

on the types of the vertical retaining structures and lateral supports. Commonly used

methods for supporting deep excavations in Singapore are:

(a) Vertical retaining walls

• Soldier pile with timber lagging

6

Page 27: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

• Shotcrete wall (normally used with soil nails)

• Sheet-pile wall

• Diaphragm wall

• Contiguous bored pile

• Secant pile wall

(b) Lateral supports

• Steel struts, or, occasionally, reinforced concrete struts, some times with

preload

• Tie-back or soil anchorage system

• Soil nails (often used with shotcrete wall)

• Floor slabs (in the case of top-down construction).

For excavations supported with diaphragm wall, sheet pile wall, contiguous bored pile

wall and secant pile wall, the sequence normally starts with the construction of the

vertical retaining structures like diaphragm wall or sheet-pile wall. The subsequent

construction activities are repeated sequences of removal of soil followed by the

installation of horizontal support system (waler and struts or tie-back anchorage) and

preloading until the maximum depth of excavation is reached.

For excavations supported with soldier pile and timber lagging or shotcrete with soil

nail systems, the construction starts with the installation of soldier pile. It is then

followed by repeated sequences of removing soil by lift, installing timer lagging or

shotcrete, installation of internal struts or tie backs or soil nails, applying preloading if

any (normally not for soil nails) until the maximum depth of excavation is reached.

7

Page 28: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

For deep excavations in poor ground conditions that are close to critical structures, jet

grouting (Yong, 1991, Hsieh et. al., 2003) or Deep Mixing Method (Tanaka 1994) are

used to improve the soft ground as a supplementary measure to control ground

movements.

2.3 Design considerations and design methods

The basic considerations for the design of deep excavation are safety, serviceability

and environmentally friendliness. The major issue in the consideration of safety is the

stability of the retaining system besides many other factors such as ground water

ingression, piping, soil floatation and excessive ground movement due to drawdown of

water table.

The stability considerations for a braced excavation design are:

• Avoiding global failure of retaining structure due to overturning or toe kick out;

• Ensuring that the lateral supports do not buckle nor be over stressed, and

• Limiting the base heave which may trigger deep surface sliding and thus jeopardize

the stability of the excavation-support system.

The main concerns in the consideration of serviceability are:

• Controlling the excavation induced wall deformation, especially the long term

component of wall deformation due to consolidation and creep of the surrounding

soil;

• Minimizing the excavation induced uneven settlement within the boundary of the

supper structure, and

• Avoiding adverse effects on nearby buildings and infrastructures to ensure the

normal use of the structures.

8

Page 29: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Various of ways of minimizing ground movement induced by excavation had been

reviewed by Chua (1985).

The commonly used design approaches in practice are empirical and semi-empirical

methods, including the traditional stability analysis methods from Rankine's earth

pressure theory and design figures and charts (Peck 1969). These designs are normally

ascertained with experimental methods and numerical simulations for major projects to

examine the adequacy of the design for a specific site conditions. This is mainly

because the design charts were developed based on the generic ground conditions

which may be different from the particular site. Recently, the use of numerical analysis

software in supporting the design is getting more indispensable as many government

authorities had set it as mandatory for their projects.

2.3.1 Empirical and semi-empirical approaches

The empirical and semi-empirical methods used in the design of retaining structures or

excavation support systems are represented by the theory based on the assumed

distribution of apparent earth pressure (Terzaghi, 1941; Peck, 1943; Tschebotarioff,

1973), the theory of limit equilibrium (Terzaghi, 1943; Bjerrum and Eide, 1956;

Gudehus, 1972) as well as empirical approaches based on field observations to address

the ground movement due to excavation (Peck, 1969; O’Rourke et. al. 1976).

Most of the existing empirical methods are aimed at ensuring the stability of

surrounding soil and the retaining structures based on limit equilibrium of wedge

theory (Terzaghi, 1941) and the apparent earth pressure theory (Terzaghi, 1943, 1954;

9

Page 30: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Peck, 1943; Peck, 1969). The soils behind the retaining structures are considered,

through the apparent earth pressure, either active or passive pressure rather than a part

of the excavation support system. The apparent pressure diagrams developed by Peck

(1969) and those developed by Tschebotarioff (1973) are some of the typical

approaches to estimate the maximum strut load during excavation. The apparent

pressure diagrams shown in Fig. 2.1 are empirical estimations derived from some field

observations and measurements. Deformation can not be determined directly by using

this empirical apparent earth pressure approach. The deformation control is implied in

the factor of safety (FS) as proposed by Terzaghi (1943).

FSC

HC H

B

u

u=

5 7

0 7

2

1

.

(2.1)

where Cu1 is the weighted average shear strength over the depth of excavation; Cu2 is

the average undrained shear strength from the base of excavation down to the depth of

0.7B below the excavation level; γ is the density of soil; H is the depth of the

excavation, and B is the width of the excavation.

Bjerrum and Eide (1956) studied the stability against base heave and proposed the use

of stability numbers. The factor of safety against base heave defined by Bjerrum and

Eide (1956) is

FS C NH q

u c=+γ

(2.2)

where q is the surcharge, Cu is the average undrained shear strength of soil within the

zone of influence (0.7B), and Nc is the bearing capacity factor of the ground which can

10

Page 31: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

be approximated as follows:

⎪⎪⎩

⎪⎪⎨

>β⎟⎠⎞

⎜⎝⎛ +

≤β⎟⎠⎞

⎜⎝⎛ +⎟

⎠⎞

⎜⎝⎛ +

=5.2

BH

LB2.015

5.2BH

BH2.01

LB2.015

Nc (2.3)

where β is a correction factor to incorporate the influence of the depth of underlying

hard stratum and is equal to 1 for a depth over 0.5B; L is the length of the excavation

area.

Empirical approaches based on field observations were also developed to address the

ground movement due to excavation (Peck, 1969; O’Rourke et. al. 1976). The

empirical diagram was developed by Peck (1969) for estimating ground settlement

adjacent to open cuts, as shown in Fig. 2.2, is frequently used as guidelines to estimate

the ground settlement around excavation.

Research efforts were also made by a number of researchers on the development of

semi-empirical methods for determining the struts load as well as the bending moment

in the vertical support (Lee et. al. 1984; Chua, 1985 ).

Many of the methods developed in the earlier days are still widely used in practice

because of their simplicity, but the limitations of such approaches are obvious. The

deformation profiles for a particular site with specific site geometry and supporting

system, the influence of construction sequence, the dissipation of excessive pore

pressure, the change of stresses in the surrounding soil as well as the effect on the

nearby structures, are not accounted for in these empirical approaches. After all, the

actual interaction forces between soil and the retaining structure may not follow the

11

Page 32: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

assumed earth pressure profile. There is no provision to the interaction between soil

and structure in those conventional methods.

Only with the development of analytical solutions for simple cases and the

approximate analytical methods based on the theory of subgrade reaction (Richart,

1959; Miyoshi, 1977; Mafei, et. al. 1977; Gudehus, et. al. 1985; Vallabhan and Dalogu,

1999) has the significance of soil-structure interaction been brought into the picture.

Obviously, the assumption of subgrade reaction has led to a sacrifice of modeling the

continuity of soil which is too important to ignore for a versatile model. This category

of analytical methods is normally weak in estimating ground movement.

Ground movement induced by excavation is becoming important because of the

presence of nearby critical structures such as Mass Rapid Transit (MRT) tunnels which

restricts the maximum absolute deformation to within 15mm and the differential

settlement not to exceed 1:2000 (3 mm per 6 m). For excavations around these

structures, more versatile design approaches supplemented with research, such as

experimental tests or numerical analysis rather than empirical approaches are required

to estimate ground movement.

2.3.2 Numerical analyses of supported excavation

In the category of numerical modeling, the applicability of 2D plane strain assumption

was examined by Tsui and Clough (1974), and it was pointed out that the 3D effect

would be significant when the stiffness of soldier pile, struts or tiebacks is much higher

than the stiffness of the retained soil.

The applicability of some FEM schemes for simulating excavation processes were

12

Page 33: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

examined by Ishihara (1970) and Clough and Mana (1976). Both papers concluded

that appropriate modeling of construction sequence is important to the reliable

prediction of the ground deformation.

Constitutive model of soil is another important factor affecting the numerical results. A

wide range of soil models had been used in past studies. Many of the earlier studies

were based on simple soil models like linear elastic model (Tsui and Clough, 1974),

bilinear elastic model (Hansen, 1980), nonlinear elastic hyperbolic model

(Balasubramaniam, et. al. 1976; Hansen, 1980) and elastic-perfectly-plastic model with

Von Mises criteria (Mana, 1978). Ever since the establishment of non-linear analytical

techniques (Zienkiewicz et. al. 1969), there has been an increasing use of the

elastoplastic analysis using various soil models. The following three main types of

soil models are widely used in geotechnical engineering and excavation analyses.

(a) Hyperbolic stress-strain curve such as those used by Clough, Duncan and their

colleagues as well as other researchers (Balasubramaniam, et. al. 1976; Hansen, 1980;

Wong and Broms, 1989; Ou and Chiou, 1993);

(b) Elastic-perfectly-plastic model of Mohr-Coulomb or Druck-Prager criteria (Brown

and Booker, 1985; Hata et. al. 1985; Yong et. al. 1989; Parnploy, 1990; Smith and Ho,

1992);

(c) The Cam-Clay family of models including the modified Cam-Clay (Simpson, 1972;

Britto and Kusakabe, 1984; Borja et. al. 1989; Lee, et. al. 1989; Hsieh et. al. 1990),

Schofield model (Schofield and Wroth, 1968), Critical State model and Cap model

(DiMaggio and Sandler 1971).

13

Page 34: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

In the present study, Elastic-perfectly-plastic model of Mohr-Coulomb type of yielding

criteria was used for its simplicity and convenience for comparison with existing

results.

2.3.3 The 3D numerical analysis of excavations

Most of the earlier works on the numerical analyses of excavations were based on the

2D plane strain analysis despite the factor that they are 3D problems by nature. Three-dimensional effects in excavation had been observed in many field

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

instrumented excavation. Simpson (1992) and Wroth suggested the use of axial

symmetric analysis to approximate the 3D effects in analyzing the deep excavation of

the British Library project which is roughly square in plane.

A case study involving 3D analysis of a deep excavation using the hyperbolic soil

model was reported by Ou and Chiou (1993). Lee et. al. (1997, 1998) presented 3D

coupled consolidation analyses of deep excavations in soft clay and highlighted the

significance of corner effects in 3D analyses. It was concluded that 3D analysis will

give smaller wall deflection than 2D analysis. Briaud and Lim (1999) analyzed a

tieback wall in sand using 3D FEM to model the in-plane bending of the timber

lagging and 3D reaction of soldier pile and tiebacks. Zhang et. al. (1999) carried out a

3D analysis of excavation supported soil nails. The soil arching behind the soldier pile

and timber lagging was studied by Vermeer et. al. (2001) using 3D Plaxis program. A

three-dimensional parametric study of the use of soil nails for stabilising tunnel faces

using ABAQUS was reported by Ng and Lee (2002). Three-dimensional pile-soil

interaction in soldier-piled excavation was studied by Hong et. al. (2003) using CRISP.

14

Page 35: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

In their study the deficiencies of the 2D “smeared” method in the modeling of soldier

pile were pointed out and the importance of using 3D analysis for such kind of

structure was highlighted. With the more sophisticated commercial software widely

available today, there is no technical obstacle for 3D analyses of excavation support

system.

In the present study, 3D analyses were adopted for the case study of a tie-back wall in

sand and an excavation in Singapore soft marine clay supported with internal struts

and ground anchors, as presented in Chapter 6 and Chapter 7. Emphases were given to

the modeling of soil-anchor interaction using the proposed element.

2.3.4 Drainage conditions and ground water draw down

In most of the numerical analysis for excavation in soft clay, undrained analysis is

considered to be suitable (Mana, 1978). This is based on the assumption that the

elapsed time during the excavation is not too long and the dissipation of the excessive

pore pressure is insignificant during the course of the excavation. This assumption is

reasonable enough for most of the small excavations in clayey soil. An undrained

analysis gives a lower bound of deformation prediction in most cases.

Drained analyses are needed to answer the questions associated with the long term

behaviour or the eventual settlement or deformation after the construction. This is

seldom the concern during the period of the excavation except for the case where an

aquifer is encountered. In most of the cases, a drained analysis is used as an upper

bound in deformation prediction (Lambe, 1970; Gudehus, 1972).

15

Page 36: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

On the other hand, the dissipation of the excessive pore water pressure depends not

only on the time elapsed but also on the permeability of the soil as well as the

existence of the drainage layers and boundaries. Consolidation analysis may be

required where the duration of the construction is long enough to cause significant

consolidation of ground soil. In fact, the dissipation of negative excessive pore water

pressure is one of the main reasons for a number of observed time-dependent

phenomena in field monitoring such as the development of observed sheet-pile wall

deflection and adjustment of strut forces (Lambe, 1970; DiBiagio and Roti, 1972;

Dysli and Fontana, 1982; Chua, 1985). The FE formulation of coupled consolidation

analysis using Biot theory was proposed by Sandhu and Wilson (1969) for elastic

medium and by Smith and Hobbs (1976), Small, et. al. (1976) amongst others for

elastoplastic materials. So far, consolidation analyses are applied more widely on the

analysis of embankment (Shoji and Matsumoto, 1976; Liu and Singh, 1977). Osaimi

and Clough (1979) studied a simple case of excavation using one dimensional

consolidation analysis. Hata et. al. (1985) analyzed a 25m deep braced excavation in

soft Yokohama clay supported with diaphragm wall using the elastoplastic constitutive

model with unsteady seepage flow based on the coupled algorithm proposed by

Sandhu and Wilson (1969). Coupled consolidation analysis of the time-dependent

behaviour of excavation support system in Singapore marine clay was also reported by

Yong et. al. (1989), Parnploy (1990), among others using elastic-perfectly-plastic

constitutive model, and by Lee et. al. (1989, 1993, 1997) using modified Cam-clay

models. In the present study, coupled consolidation analysis was adopted for the case

study of excavation in Singapore soft marine clay supported with internal struts and

ground anchors, as presented in Chapter 7. This is to demonstrate the fact that the

proposed element can also be used for coupled consolidation analysis.

16

Page 37: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

2.4 Experimental methods

As an efficient way to investigate the mechanical behaviour of soil and the excavation

support systems, experimental methods, such as the conventional soil tests and

centrifugal tests, play an important role in facilitating the design of excavation support

systems by providing more insight on the mechanical behaviour of soil and mechanism

of performance of the retaining systems. For example, the failure mechanism of

London clay was studied using centrifugal model test by Lyndon and Scholfield (1970);

The failure mode of a retaining wall in clay was investigated using centrifugal model

test by Bolton and Powrie (1987); Kimura, et. al. (1993) developed the in-flight

excavator for centrifugal modeling of excavation processes. Centrifugal tests were also

used to ascertain the interaction of pile and tunnel by Loganathan et. al. (2000),

piles-soil interaction during excavation by Leung et. al. (2000), and the failure

mechanism of soil nailed excavation (Tufenkjian and Vucetic, 2000), amongst many

others. Due to well known reasons, the centrifugal methods are not viable for a site

specific study. Numerical modeling using FEM, validated by centrifugal modeling if

possible, would be a best and more generic approach for studying soil-structure

interaction.

2.5 Soil-structure interface

Soil-structure interface is a weak spot in the system where a much stiffer and stronger

material, such as concrete or steel used for retaining structure such as diaphragm wall,

sheet-pile wall, soldier pile, tieback, soil nail or other solid inclusions meets with a

relatively softer material like soil. Pronounced relative displacements, both tangential

slippage and normal separation or compression between the inclusion and the soil, may

17

Page 38: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

take place within the limited space of the interface and therefore cause significant

uncertainty to the load transfer between the soil and the structure system. This is why

the numerical modeling of soil-anchor interface was chosen as the subject of this study.

A study on the effect of interface properties on retaining wall behaviour was

documented by Day and Potts (1998). A well-known relationship linking the load

transfer with the relative displacements of the interface was established by Goodman

(1968) through two stiffness coefficients, Kn and Ks in linear elastic domain.

Laboratory and field experiments were conducted by many researchers to find suitable

constitutive models for the interface in plastic or post-yield domain (Desai et. al. 1985;

Kishida and Uesugi, 1987; Yin et. al. 1995; Evgin and Fakharian, 1996).

Mohr-Coulomb criterion, with zero-effective tension limit, is widely used as one of

such models although many other constitutive models such as strain hardening model

(Fishman and Desai, 1987), Hierarchical single-surface model (Navayogarajah, et. al.

1992) and Hyperbolic model (Yin et. al. 1995) were also used. For simplicity, an

elastoplastic constitutive model of soil-anchor interface was adopted for the

development of the theoretical framework of the proposed anchor-soil interface

element. Other types of constitutive models can be easily adapted to the framework.

2.6 Numerical modeling of soil-structure interface

Numerical modeling of interface is one of the key tasks in the study on soil-structure

interaction. The contact zone between soil and other structures such as piles, footing

foundation, raft foundation, retaining structures like diaphragm wall, retaining wall,

sheet-pile wall or secant pile wall etc., tunnel lining, even embankment reinforced with

geomembrane, is generally referred to as soil-structure interface. The soil-structure

18

Page 39: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

interaction and the problems related to the mechanics of jointed rock involve interface

and its mechanical behaviour. The response of soil-structure systems, such as

shallow and deep foundations, lined tunnels, retaining walls, and reinforced earth, to

monotonic and cyclic loads, is influenced by the mechanical behaviour of interfaces.

Interface elements have been used widely in modeling various types of soil-structure

interaction problems, such as excavation support systems (Day and Potts, 1998),

anchor-soil (eg. Dicken and King, 1996), reinforced earth (eg. Mosaid and Lawrence,

1978; Cividini et. al. 1997), soil-nail (Smith, 1992; Zhang et. al. 1999; Tan et. al. 2000),

embankment and geomembrane (eg. Hird and Kwok, 1989); pile-soil interaction

(Cheung et. al. 1991), tunnel lining and soil (Bernat and Cambou, 1998), etc. Day and

Potts (1998) studied an excavation supported with sheet-pile and highlighted the

significance of modeling the interface between soil and sheet-piles. Soil nail was

modeled as a thin sheet and Interface between soil and nail was modeled using a thin

layer in the analyses conducted by Smith (1992), Smith and Su (1997). A 3D FEM

analysis of soil nail supported excavation was conducted by Zhang et. al. (1999) with

the soil-nail interface being modeled using bar elements. Herman and Zaynab (1978)

proposed a system to model reinforced soil with linker elements as shown in Fig. 2.3.

Tan et. al. (2000) investigated the failure mode and soil nail lateral interaction

mechanism using a linker element from commercial software, FLAC (Itasca, 1996), as

shown in Fig. 2.4. In their analyses, nail was modeled using 2D beam element which

can yield under tension or compression, and the interface between soil and nail was

modeled using spring-slide system which is also called Linker element in other

references.

Various types of interface element for FEM have been proposed in the past in order to

19

Page 40: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

model the discontinuous behavior at the interface. Some treated the interface by using

elements of zero thickness (Goodman et. al. 1968; Ghaboussi et. al. 1973; Wilson,

1977; Gens et. al. 1988) or by thin-layered elements of small thickness (Desai et. al.

1984; Schweiger and Hass, 1988). Other workers have suggested the connection of the

elements of soil and reinforcement to each other by discrete springs (Herrmann, et. al.

1978). There are also methods using the approach of constraint equations or penalty

functions to treat the interface as contact, stick/sliding element (Katona 1983).

Applications using isoparameter element with extreme aspect ratio to simulate the

interface behaviour is not uncommon as well (Griffths, 1985; Brown and Shie, 1990,

1991; Smith and Su, 1997; Wakai et. al. 1999).

Deficiencies and problems associated with the use of some of the interface elements,

such as the kinematical compatibility deficiency in the zero thickness slip element of

Goodman’s type (Kaliakin and Li, 1995), the ill-condition and numerical stability of

interface elements (Pande and Sharma, 1979; Day and Potts 1994; Ng et. al. 1998),

were reported and improvements to these problems have been proposed. However,

these problems have not discouraged the application of slip element in the analysis of

soil-structure interactions. The zero thickness slip element of Goodman’s type is still

widely used in the numerical studies (Ng et. al. 1998) as the reported deficiencies will

be only significant under extreme conditions.

The 3D slip element was absent in the existing version of CRISP (Britto and Gunn,

1990). To facilitate the numerical analysis for soil-structure interaction in 3D problems,

a 16-noded 3D slip element with zero thickness is formulated and implemented into

CRISP. Although modeling of soil-structure interface can be done using 20-noded 3D

brick element with small thickness (Smith, 1992; Bransby and Springman, 1996), the

20

Page 41: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

result of the analysis will be closely related with the thickness of the element (Ng et. al.

1998). There is no guideline on the selection of the thickness of the interface, therefore

it will be totally on personal judgment and make it a subjective issue difficult to justify,

thus compromising the reliability and the advocacy of the numerical result. The 3D

slip element used in the present study belongs to the category of zero thickness slip

element and is similar to the one reported by Gens et. al. (1988) but avoided the use of

mid-plane so as to make it more conveniently compatible with 20-noded brick element

existing in CRISP.

2.7 Numerical modeling of soil reinforcement / soil nail

Tie-backs, ground anchors, reinforced soil structures and soil nailing are widely used

in excavations as a replacement to internal struts so as to provide horizontal supports

for excavation bracing walls (Yong et. al. 1989; Briaud and Lim, 1999; Zhang et. al.

1999; Ganesh, 2002). The advantage of using this type of supporting system over that

of using internal struts is the cleared space within excavation uncluttered by braces

(Schnabel, 2002). It is a natural choice, and sometime the only choice, to use this type

of supporting system when large clear space inside the excavation pit is required for

the construction of other structures, such as the tunnel sections built in cut and cover

method (Parnploy, 1990). In some cases, it may result in a faster or less costly project

(Schnabel, 2002), especially when the dimensions of excavation pit are so large to

make the horizontal struts become too expensive, such as the excavation of the

underground MRT depot (Ganesh, 2002).

Conventional method of limiting equilibrium is still a basic design tool for tie-back,

ground anchor, soil nail and reinforced earth at present (Schlosser et. al. 1992, Thomas

21

Page 42: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

and Carlton, 2003), though there are many successful numerical analyses of the

problems in existence (eg. Mosaid and Lawrence, 1978; Cividini et. al. 1997; Tan et. al.

2000; Zhang et. al. 1999; Ng and Lee 2002). However, this method does not address

the issue of deformation of the reinforced soil which is actually the primary issue in a

soil-nail system. A full soil-nail interaction requires adequate provisions for modeling

the stress-strain relationship of soil, the reinforcing member as well as the soil-nail

interface. In fact, reinforcing members in these problems can be considered as solid

inclusions in soil.

Jewell and Pedley (1992) studied the soil-nail interaction and concluded that only a

small portion of shear strength of nail can be mobilized for normal soil nail diameter.

Byrne (1992) proposed an analytical approximation based on the assumption of

uniform stress distribution in soil, around and along the reinforcing member, which

allows for an estimation of soil nail interaction. Liang and Feng (2002) developed an

analytical model for anchor-soil interface which can be used to investigate the

anchor-soil interaction for single anchor. Like the analytical solutions for many other

problems, these analytical models can only be used in single anchor or nail in

homogeneous soil with no interaction with other structures such as soldier pile and

timber lagging. Therefore they are not able to assess the ground movement for

excavations supported with tieback or soil nails.

Herrmann and Zaynab (1978) proposed a linker element for modeling soil-nail

interaction using 2D FEM, as shown in Fig. 2.3. This type of element is now widely

used in many commercial software packages. Naylor (1978) proposed a strip slip

element which treats the soil and reinforcement interface in an equivalent stiffness

22

Page 43: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

manner. Other models based on equivalency can also be found, such as the multiphase

model proposed by Sudret and Buhan (2001). Truss elements bundled with joint

element were also used in the 2D analyses of similar type of problems (Ogisako et. al.

1988). Mauricio et. al.(1996) conducted parametric study of soil nailing system using

beam elements with interface.

FEM model of 2D analyses using an element similar to that proposed by Herrman and

Zaynab (1978), considering the slippage between soil and nail were also reported (Tan

et. al. 2000). This element is the same as the Linker element widely used in the

modeling of interaction between steel rebar and concrete, called “Bond-slip”, proposed

by Ngo and Scordelis (1967). Similar type of element was also found in commercial

software like Flexis and Plaxis as demonstrated in the work of Tan et. al. (2000) and

Vermeer et. al. (2001) respectively. The difference between this Linker element and the

proposed element formulated in Chapter 3 are highlighted in the later section of this

Chapter. There are a group of contact elements and a few special purpose elements in

ABAQUS but none of them has considered both stress equilibrium and deformation

compatibility within the element like the proposed element does. For example, the

pile-soil interaction element (type PIS24 for 2D and PSI36 for 3D), which was perhaps

specially designed for pipe soil interaction of buried pipes, uses linear or quadratic

shape functions for pile. Furthermore, the interaction force between pipe and soil are

evaluated by using the far field displacement in soil instead of soil displacement within

interface. In many case studies carried out using ABAQUS, bar/beam elements instead

of the pipe-soil interaction elements were generally used in 3D FEM for soil nail and

ground anchor types of soil reinforcements (Zhang et. al., 1999; Briaud and Lim, 1999;

Ng and Lee, 2002).

23

Page 44: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

A 3D analysis considering both the axial stiffness and flexural stiffness of nail by using

3D beam elements was reported in the work of Cividini et. al. (1997), but provision for

slippage was not accounted for due to the lack of appropriate element types. Analyses

of field experiment of reinforced soil using similar 3D model (without interface

element) was also reported by Akira et. al. (1988). The numerical model for rock bolt

proposed by Kawamoto et. al. (1994), as shown in Fig. 2.5, integrated the steel rebar

and grout annulus into one cylindrical element. The slippage in the interface between

grout and the wall of borehole was not formulated explicitly in the element. It can only

approximate the interface slippage by using a reduced stiffness for grout. Incidentally,

Briaud and Lim (1999) also adopted the similar kind of approximation by using a

reduced stiffness for the anchor bonded length in their analysis of a tieback wall. The

implication of this approximate approach lies in the selection of percentage of

reduction to the stiffness of grout which makes it a subjective parameter difficult to

justify. Briaud and Lim (1999) used a 40% reduction and it worked fine for the

particular case studied. There is no sign that this 40% is going to work for other site

with different ground conditions or configurations of anchor system. As such, the

development work described in Chapters 3 and 4 is desired for and is essential for a

full 3D analysis of problems like anchor, soil-nail interaction.

2.8 FEM model for modeling of bond-slip in reinforced concrete

An extensive literature review was conducted on publications related to the use of the

special element to model reinforcement in reinforced concrete (RC) members as well

as fibre reinforcement in composite materials. The conventional approaches in the

24

Page 45: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

modeling of steel or fibre reinforcement in RC or composite materials are:

(1) Equivalent stiffness method in which the reinforcement are smeared into the

structural member, such as the model proposed by Sprenger and Wagner (1999),

and

(2) Embedment method in which the reinforcement is considered separately as an

embedment in the structural member, such as the model proposed by Kwak and

Kim (2001).

An element called “Linker” element, as shown in Fig. 2.6, used to model the

interaction of reinforcement and concrete may share some concepts in common with

the proposed model for anchor-soil interface described in Chapter 3 and 4, but at the

same time is different in many other aspects. For example, both the proposed elements

and the “Linker” element use Ks and Kn to represent the interface stiffness in tangential

and normal direction to interface, but the formulation of the element stiffness matrix is

totally different. The Linker element simply integrates the Ks, Kn along the length to

arrive at the element stiffness. In the proposed element, the element stiffness is derived

from the internal equilibrium between solid inclusion and the interface without

prescribing any shape function for the displacement of the solid inclusion. The

bond-slip model for monotonic axial loads reported by Kwak and Kim (2001) has also

considered the internal equilibrium of the element like the proposed model, but the

attempt to model both concrete and steel reinforcement in the same element, which

leads to the use of steel ratio ρ=As/Ac and the assumption on the distributions of axial

displacement and stress across the section in concrete structural member, is a

significant limitation to the model and constitutes the major difference as compared to

the proposed Anchor-interface element.

25

Page 46: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

There are also models which simply combined the “bar element” and “bond link

element” together as one element (ASCE Report, 1982; Monti et. al. 1997; Monti and

Spacone, 2000). The internal equilibrium between bar and the bond-link is not

guaranteed although the displacement compatibility is satisfied if the same shape

function for bar and bond-link is used along the axial direction of the reinforcement.

This is due to the weak formulation of the element stiffness in FEM. Furthermore,

bundling a bar and a slip element may be applicable for 2D analysis of soil-structure

interaction such as sheet pile. It is obviously not applicable to 3D soil-anchor

interaction simply because the 2D slip does not match with 3D bar element in nodal

degree-of-freedom (DOF), nor does the conventional 3D slip element match with 3D

bar element. This is because the 3D slip element is formed between two surfaces

whereas the cross-sectional dimension diminishes in the 3D bar element, and thus

unable to represent the perimeter surface of the bar. There is no published literature on

the existence of a special 3D slip element which can wrap up around a 3D bar/beam

element to the best of the author’s knowledge.

The proposed anchor-interface element, however, is different from the “Linker”

element used in reinforced concrete analysis in the following aspects:

(1) No soil is included in the element unlike the bond-slip element proposed by

Kwak and Kim (2001) where both concrete and reinforcement are included.

(2) The displacement along the anchor in the element is explicitly derived from the

internal equilibrium within the interface around it. This is a major departure

from conventional approaches in modelling bond-slip in reinforced concrete

using bar and bond link elements.

(3) No prescribed shape function is required for anchor in the proposed

26

Page 47: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

anchor-interface element.

(4) The shape function on the soil side of the interface is prescribed so as to be

compatible with the shape functions of the elements modelling the soil around

the anchor.

(5) Both axial interaction and lateral interaction between anchor and soil can be

modeled using the proposed element whereas the “Linker” and other element

for modelling bond-slip focus on axial interaction between reinforcement and

concrete only.

To the best of the author’s knowledge, the proposed anchor soil interface element has

not been found in literature review. Therefore the work presented in Chapter 3 and 4 is

original development.

2.9 Interface properties, Ks and Kn

The physical meaning of Kn in soil-solid-inclusion interface may not be as

straightforward as that of Ks but it does not undermine the applicability of

mathematical modeling as presented in the proposed element. While taking Kn as a

phenomenological description to the response of solid inclusion subject to the lateral

load is a point of argument, the mathematical interpretation of Kn can be simply a

penalty function for solid inclusion-soil contact as far as the interface between solid

inclusion and soil is concerned. Many reported literature (Cheung et. al. 1991; Hird

and Kwok 1989) used Kn as high as 107~108 kPa which is generally large enough to

avoid significant overlapping of contacting faces compared to the elastic modulus of

soil in the order of 104 kPa. It should be pointed out that too high values for Kn might

cause numerical instability.

27

Page 48: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

2.10 Soil-pile interaction during supported excavations

Soldier piles are common in the supporting system for deep excavations (eg. O’Rouke,

et. al., 1976; Briaud and Lim, 1999; Vermeer et. al., 2001; Hong et. al. 2003). Recently,

the ground movement induced by construction activities is drawing more concern than

that to the deformation of supporting structures themselves during an excavation or

tunneling (Finno et. al. 1991; Loganathan and Poulos, 1998; Mroueh and Shahrour,

2002; Goh et. al. 2003). Therefore, there is a great need to address the issue of soil-pile

interaction during excavation or tunneling.

Analytical theory and practical approaches related to the design of piled foundations

are well established after comprehensive and continuing research efforts being made

by many researchers and practicing engineers over the decades (eg. Meyerhof, 1976;

Polous and Davis, 1980; Fleming et. al. 1992). Most of them are based on the elastic

solution of Mindlin (Polous and Davis, 1976), simple analytical solutions in closed

form (eg. Matlock and Reese, 1960; Randolph and Wroth, 1978), empirical load

transfer theory and numerical methods using subgrade reaction theory (eg.

D’Appolonia and Romualdi, 1963; Colye and Ressie, 1966), Boundary Element

Method (Butterfield and Banerjeer, 1971; Xu and Polous 2000), Finite Difference

Method (Polous 1994), Finite element Method (Balaam et. al. 1975; Ellison et. al.

1971) and even infinite layer method (Guo et. al. 1987; Ta and Small, 1997) or the

combinations of the afore mentioned methods.

28

Page 49: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Amongst the numerical methods mentioned above, FEM is thought to be the most

sophisticated and generic for its versatility and flexibility in modeling different

structure members with complex geometry and complicated ground conditions as well

as the ability for coupled consolidation analysis and many other advantages in

simulating construction processes vigorously.

The prediction of settlement of pile or pile groups is one of the main tasks in the

design and analysis of piled foundation. There are many methods of analysis available

directed to address the design issues of this type (and others). These methods can be

classified as:

(1) The empirical methods include those empirical formulae based on laboratory

tests or field observations such as those proposed by Terzaghi (1943) and

Meyerhof (1959) as well as the so-called “load-transfer methods” (eg. Coyle

and Reese,1966; Kiousis and Elansary, 1987).

(2) The analytical methods of closed form solutions for certain simple cases (eg.

Randolph and Wroth, 1978).

(3) The numerical methods include those based on subgrade reaction theory which

treats pile as a series of structural segments and soil as elastic continuum in half

space such as Polous and Davis (Polous and Davis, 1968, 1980) and Chow

(Chow, 1986); with some extended to incorporate non-linearity of soil as a

non-linear spring constant (Poulos and Davis, 1968), and the elastic continuum

methods.

(4) Elastic methods of analysis include the boundary element methods (BEM) and

the finite element methods. While the many researches had demonstrated the

versatility of FEM (Desai, 1974; Valiappan et. al. 1974; Ottaviani, 1975;

29

Page 50: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Pressley and Poulos, 1986; Smith and Wang, 1998), it is however considered

too expensive for routine analysis especially for 3D problems due to the lack of

an efficient way to model the soil-pile interaction. One of the important

considerations in modeling soil-pile interaction lies in the fact that pile cannot

be approximated with a line type of inclusions. Its transverse dimension has to

be modeled in the element.

In many pile problems, pile-soil interaction may have a significant influence on the

resulting pile and ground movement. For example, the importance of simulating

slippage between pile and soil in pile settlement analysis has been highlighted by

Trochanis et. al. (1991) and Cheung et. al. (1991), amongst others. By comparing the

load-settlement behaviour and axial load distribution along the pile computed from

two-dimensional axi-symmetric FE analyses with measured values from an

instrumented pile, Cheung et. al. (1991) concluded that interface elements and

non-linear soil behaviour are necessary to model the pile-soil interaction accurately.

Other instances of interface element usage in single-pile analyses include Ellison et. al.

(1971) and Desai et. al. (1974), and more recently, Rojas et. al. (1999) and Wakai

(1999). However, most of these interface element formulations are only applicable to

2D or axisymmetric analyses. Three-dimensional analyses of pile groups using

interface elements (e.g. Brown and Shie, 1990, 1991) remain relatively scarce and are

largely limited to relatively simple group configurations.

30

Page 51: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

CHAPTER 3

DEVELOPMENT OF ANCHOR-INTERFACE ELEMENT 3.1 Background

In this chapter, a new element for modelling anchor-soil interaction is constructed,

formulated and implemented in the CRISP. The verification of the model is conducted

in comparison with closed form solutions. The literature review regarding the issues in

anchor-soil interaction and the numerical models are presented in Chapter 2 and the

non-linear algorithm for the model as well as more case studies using the proposed

model are found in the following chapters.

3.2 The construction and the formulation of the anchor-interface element

This anchor-interface element is constructed based on the idea of wrapping a

cylindrical slip element around a beam element as shown in Fig. 3.1(a).

In the idealized 2D view, as shown in Fig. 3.1(b), Nodes 1 and 2 represent nodes on

anchor with 3 and 5 DOFs each for 2D and 3D respectively. Nodes 3, 4, 7 represent

soil side with 2 and 3 DOFs each for 2D and 3D respectively. The deformation of

nodes on soil side will follow the same shape function as that in 20-noded brick

element whereas the displacements of nodes on anchor are not prescribed and will be

determined through the internal equilibrium of the element.

In axial direction of the anchor, the internal equilibrium is governed by the equilibrium

between skin friction on anchor tendon and the axial force on the cross-section of the

31

Page 52: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

anchor. The lateral equilibrium is governed by the Euler beam theory taking the

interaction between anchor and soil as earth pressure acting on the anchor tendon.

3.2.1 The equilibrium in axial direction

Fig. 3.1(a) shows a beam element representing an anchor in contact with the

surrounding soil domain, which is modelled by 20-noded brick elements. A section

view of the slip element forming the anchor-soil interface is shown in Fig. 3.1(b).

Along the beam element, the rate of increase in axial stress can be related to the shear

stress τs which forms the skin friction via the equilibrium relation.

i.e. dxr

d sa τσ0

2= (3.1)

where d aσ is the axial stress increment in the anchor and sτ is the shear stress in

skin friction, and 0

2r

is the ratio of cross-sectional perimeter of anchor to its

cross-sectional area and r0 is its radius.

When the relative displacement between anchor and soil is sufficiently small, the shear

stress in skin friction τs can be proportionally related to the relative movement between

soil and anchor ∆u in a similar way as that used in Goodman’s interface element, i.e.

uK ss ∆=τ (3.2)

then udxrK

d sa ∆=

0

2σ (3.3)

Internal displacements along the soil side us are related to the displacement of nodes 3,

4 and 7 via the standard shape functions

32

Page 53: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

2)1(

3ξξ −

−=N ; 2

)1(4

ξξ +=N ; (3.4) 2

7 1 ξ−=N

where lx2

=ξ (3.5)

in which l is the length of the element; so that

us = N3u3+ N4u4+ N7u7 (3.6)

Let u be the axial displacement of the anchor, then

∆u=u-us=u-(N3u3+ N4u4+ N7u7) (3.7)

If the anchor is working in the linear elastic range, substitution of Equation (3.7) into

(3.3) leads to governing equation for the equilibrium of anchor-interface element in

axial direction.

)(22

77443300

2

2

uNuNuNrK

urK

dxudE ss ++−=− (3.8)

3.2.2 The elastic formulation of displacement of anchor in axial direction

The solution of the above homogeneous differential equation takes the form

)(21 xfeCeCu uxx ++= −µµ (3.9)

where fu(x) is the corresponding particular solution of the above homogeneous

differential equation, and

0

2ErK s=µ (3.10)

We note that

us = [ ] [ ] [ ] [ ] eu

eu AXAxxuNuNuN δδ ⋅⋅=⋅⋅=++ 2

774433 1)( (3.11)

where

33

Page 54: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

[ ]⎥⎥⎥

⎢⎢⎢

−−=

42200000

00001

2

2 lll

lA ; ;

⎪⎪⎪

⎪⎪⎪

⎪⎪⎪

⎪⎪⎪

=

7

4

3

2

1

uuuuu

euδ [ ] ⎣ ⎦21 xxX = (3.12)

Equation (3.8) can thus be written as

[ ] [ ] euAXu

dxud δµµ ⋅⋅−=− 222

2

(3.13)

Since us is quadratic in x, its second derivative is a constant, so that a particular

solution is

fu(x) = us + C0 (3.14)

Where C0 is a constant which can be evaluated by substituting Equation (3.14) into

Equation (3.13). This leads to

[ ] [ ] eulu AXxf δ⋅⋅=)( (3.15)

Where [ ]

⎥⎥⎥⎥⎥

⎢⎢⎢⎢⎢

−−

=42200

000

844001

22

22

2 ll

l

lAl

µµµ (3.16)

Coefficients C1 and C2 in Equation (3.9) can be determined using the boundary

conditions

12

22

uu

uu

lx

lx

=

=

=

−=

(3.17)

Substituting Equations (3.15) and (3.17) into Equation (3.9) leads to

[ ] [ ] euC

CS δΠ=

⎭⎬⎫

⎩⎨⎧

2

1 (3.18)

34

Page 55: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

where [ ]⎥⎥

⎢⎢

⎡=

22

22ll

ll

ee

eeSµµ

µµ

(3.19)

and [ ]⎥⎥⎥⎥

⎢⎢⎢⎢

−+−

+−−=Π

21]4)(1[

4)(0

2]4)(1[10

4)(

)(4

22

22

2 ll

ll

l µµ

µµ

µ

Thus, Equation (3.9) can be expressed as

[ ] euGu δ= (3.20)

where

[ ] [ ] [ ] [ ][ ]lxx AXSeeG +Π= −− 1µµ (3.21)

Substituting Equations (3.11) and (3.20) into Equation (3.7) yields the relative

anchor-soil displacement.

[ ] [ ] euAXGu δ)][( −=∆ (3.22)

3.2.3 Element stiffness for axial displacement

For any virtual nodal displacement vector Teu uuuuu *

7*4

*3

*2

*1

* ,,,,=δ , the relative

anchor-soil displacement is given by

[ ] euAXGu ** )][]([ δ−=∆ (3.23)

Considering only axial displacements, the increase in internal strain energy U due to

virtual displacement is given by

[ ] [ ] dxdxdu

dxdurEdxAXGKAXGr

dxrEdxurU

l

eu

ls

TTu

lls

∫∫

∫∫

+−−=

+∆=

*2

00*

20

**0

)][]([)][]([2

2

πδπδ

επετπ

(3.24)

35

Page 56: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

The virtual work done by external nodal forces is eTu RW *δ= (3.25)

where Re is the nodal forces on the element.

Application of the virtual work principle thus leads to

[ ] [ ] eeu

Ts

T

l

RxdGdxdG

dxdrEAXGKAXGr =+−−∫ δππ ])[(])[()][]([)][][(2 2

00 (3.26)

The element stiffness matrix [K]e can thus be re-written as

[ ] [ ] [ ] xdGdxdG

dxdrEAXGKAXGrK T

sT

l

e ])[(])[()][]([)][][(2 200 ππ +−−= ∫

(3.27)

Substitution of Equation (3.21) into Equation (3.27) leads to

[K]e ][][ III KK += (3.28)

where

⎥⎥⎥⎥⎥⎥

⎢⎢⎢⎢⎢⎢

+−+−−+−−−

−−−−

=

αχχαχαχχζχαχµχ

ζµχχµ

µ

µµπ

8)12(2)12(2221)1(21)cosh(

)cosh(1)cosh(1

1)cosh(

)sinh(2

][ 0

ll

symmetricll

lKr

K sI

(3.29)

[ ]

⎥⎥⎥⎥⎥⎥⎥⎥⎥

⎢⎢⎢⎢⎢⎢⎢⎢⎢

+−−

+−−−−

−−

=

α

αα

ααα

163

16

8344

311

8344

354

311

413043100

lEAK II (3.30)

and 2

2⎟⎟⎠

⎞⎜⎜⎝

⎛=

lµα ; ]1)[cosh( −= lµαχ ; )cosh()1(2 lµχαζ ++= (3.31)

36

Page 57: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

3.2.4 The equilibrium and anchor displacement in transverse direction

The normal force qn acting across the anchor-soil interface per unit anchor length may

also be proportionally related to the relative transverse displacement ∆v via a

Goodman’s spring constant for slip element, that is

qn = 2r0Kn ∆v (3.32a)

which allows the equivalent normal stress σn to be expressed as

σn = Kn ∆v (3.32b)

Letting the transverse displacement of the anchor be v and that of the soil be vs, then

relative displacement ∆v is given by

∆v=v-vs (3.33)

From Euler-Bernoulli beam theory,

nrdx

vdEI σ04

4

2−= (3.34)

Substituting Equations (3.32b) and (3.33), as well as 4 0

2EIKr n=λ

into Equation (3.34)

leads to

svv

dxvd 444

4

44 λλ =+ (3.35)

Since

vs = N3v3 + N4v4 + N7v7 (3.36)

which is quadratic in x, the general solution of this differential equation is

sxx vexCxCexCxCxv ++++= −λλ λλλλ )]cos()sin([)]cos()sin([)( 4321 (3.37)

where C1 to C4 are constants that can be determined from the boundary conditions

37

Page 58: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

⎪⎪⎪⎪⎪

⎪⎪⎪⎪⎪

=

=⎟⎠⎞

⎜⎝⎛−

=

=⎟⎠⎞

⎜⎝⎛

−=

=

2

2lx

2

1

2lx

1

dxdv

v2lv

dxdv

v2lv

θ

θ

(3.38)

where θ1 and θ2 stand for the angles of rotation at nodes 1 and 2 respectively.

Imposition of these boundary conditions leads to four linear equations in C1 to C4 as

follows:

[ ]

⎪⎪⎪⎪⎪⎪⎪

⎪⎪⎪⎪⎪⎪⎪

−=−−++

++−

−=+−−+

−++

−=−−+−

−=+++

)(1e)]sin(cosC)sin(cosC[

e)sin(cosC)sin(cosC

)(1e)]sin(cosC)sin(cosC[

e)]sin(cosC)sin(cosC[

vve]cosCsinC[e]cosCsinC[

vve]cosCsinC[e]cosCsinC[

3243

21

4143

21

324321

414321

θθλ

ββββ

ββββ

θθλ

ββββ

ββββ

ββββ

ββββ

β

β

β

β

ββ

ββ

(3.39)

where

2l

x

s3

2l

x

s4 dx

dv,

dxdv

−==

== θθ and 2lλβ = (3.40)

We note that, as in the case of axial displacement,

[ ] [ ] evs AXv δ⋅⋅= (3.41)

in which δve= [v1 v2 v3 v4 v7] T is the nodal transverse displacement vector, and

[X] and [A] are defined in Equation (3.12).

38

Page 59: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Therefore

[ ] [ ]

[ ] [ ] e

ve

vlx

ev

lx

lAdxXd

AdxXd

δδ

δ

θθ

⎥⎦

⎤⎢⎣

⎡−

−−=

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⋅⋅

⋅⋅

=⎭⎬⎫

⎩⎨⎧

=

−=

43100413001

)(

)(

2

2

4

3 (3.42)

Substituting Equation (3.42) into Equation (3.39) and solving for C1 to C4 allows

Equation (3.37) to be expressed as

(3.43) [ ] se

v vHv += θδ

or ∆v = [H]δvθe (3.44)

where [ ][ ] [ ]( )CFLH 1][ −= (3.45)

and [L] = )xcos(e)xsin(e)xcos(e)xsin(e xxxx λλλλ λλλλ −− (3.46)

(3.47) T7432121

ev vvvvv θθδ θ =

[ ]

⎥⎥⎥⎥⎥

⎢⎢⎢⎢⎢

−−++−+−−−+

−−=

−−

−−

−−

−−

)sin(cose)sin(cose)sin(cose)sin(cose)sin(cose)sin(cose)sin(cose)sin(cose

cosesinecosesinecosesinecosesine

F

ββββββββββββββββ

ββββββββ

ββββ

ββββ

ββββ

ββββ

(3.48)

and [ ]

⎥⎥⎥⎥⎥⎥

⎢⎢⎢⎢⎢⎢

−−

−−

=

l4

l1

l31000

l4

l3

l10100

00100100100001

C

λλλλ

λλλλ (3.49)

Now vs can be represented in terms of the enhanced transverse nodal

displacement-rotation vector δvθe by expanding δve to δvθe and [Al] to [Av].

Thus Equations (3.41) and (3.43) can be re-written as:

39

Page 60: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

[ ] [ ] evvs AXv θδ⋅⋅= (3.50)

and

[ ] [ ] [ ]( ) evvAXHv θδ⋅+= (3.51)

in which

[ ]

⎥⎥⎥⎥⎥

⎢⎢⎢⎢⎢

−−

=4220000

0ll0000

8l440000

l1A

22

22

2v

λλλ (3.52)

3.2.5 Element stiffness for flexural degrees of freedom

Given any virtual nodal transverse displacement-rotation vector

T*7

*4

*3

*2

*1

*2

*1

e*v vvvvv θθδ θ = (3.53)

the virtual displacement inside the element ∆v* is given by

[ ] e*v

* Hv θδ∆ =

Since axial and transverse displacements in a beam are independent of each other, we

can ignore the axial virtual displacement in the virtual work consideration for

transverse displacement. The increase of internal strain energy U due to virtual

transverse displacement can be expressed as

dx

dxvd

dxvdEIdx])H[K]H([r2

dxdx

vdEIdxvr2U

l2

2

2

*2e

vl

nT

0eT*

v

l2

2*

ln

*0

∫∫

∫∫

⎟⎟⎠

⎞⎜⎜⎝

⎛⎟⎟⎠

⎞⎜⎜⎝

⎛+=

⎟⎟⎠

⎞⎜⎜⎝

⎛+=

θθ δδ

εσ∆ (3.55)

The virtual work W done by the external force is given by

(3.56) eeT*v RW θδ=

40

Page 61: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

where Re is the transverse components of the nodal forces on the element.

The virtual work principle thus leads to

[ ] ])A][X[]H([dxd])A][X[]H([

dxdEI])H[K]H[(r2K v2

2T

v2

2

nT

l0v +++= ∫θ

(3.57)

Substituting Equation (3.45) into Equation (3.57) leads to

[ ] [ ] [ ] [ ] [ ]TIIIV

IIIV

IIV

IV

v KKKKK +++=θ (3.58)

where

[ ] [ ] [ ] [ ] [ ]JxdLLJEIJxdLLJKrKl

TT

l

TTnI

V ∫∫ += ][][][][2 0 (3.59)

[ ][ ] [ ] [ ] [ ][ ][ ][ ] ⎥

⎥⎥⎥

⎢⎢⎢⎢

−−−−

⎟⎠⎞

⎜⎝⎛=

=

×

×

×

××××

422021102110

00002

][][][][

41

41

41

141414444

2

2

2

2

ll

EI

AxdXdxdX

dxdAEIK v

l

TTvII

V

(3.60)

[ ] [ ] [ ]

[ ] [ ] [ ]

[ ] [ ]( )⎪⎪⎪⎪

⎪⎪⎪⎪

+=

⎭⎬⎫

⎩⎨⎧

⎟⎟⎠

⎞⎜⎜⎝

⎛+=

⎭⎬⎫

⎩⎨⎧

+=+

T

l

TTvv

TT

l

TTvv

TTIII

VIII

V

VVEI

JdxLXdxdAAxdX

dxdLJEI

HdxdX

dxdAAxdX

dxdH

dxdEIKK

][][][][][

][][][][][

"2

2

2

2"

2

2

2

2

2

2

2

2

(3.61)

in which [J] = [F]-1[C] (3.62)

[ ] [ ] [ ]22

2

2" L2L

dxdL λ== (3.63)

[ ] )xsin(e)xcos(e)xsin(e)xcos(eL xxxx2 λλλλ λλλλ −−−−= (3.64)

41

Page 62: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

and [ ] [ ]∫=l

vTT AxdX

dxdLJV ][][][ 2

2" (3.65)

Since , Equation (3.59) can be re-written as: n04 Kr2EI4 =λ

[ ] [ ] [ ][ ]JMJKrK TnI

V02= (3.66)

in which

[ ]⎥⎥⎥⎥

⎢⎢⎢⎢

=

)lsinh(0)lsin(00)lsinh(0)lsin(

)lsin(0)lsinh(00)lsin(0)lsinh(

1M

λλλλ

λλλλ

λ (3.67)

and

[ ] [ ] [ ]

[ ]

[ ][ ][ ][ ] ⎥

⎥⎥⎥

⎢⎢⎢⎢

−−−

−−

⎟⎠⎞

⎜⎝⎛=

=

×

×

×

×

22241

11141

22241

111412

2

2

22

20202020

22

][][2

aaaaaaaaaaaa

Jl

EI

AxdXdxdLJEIK

T

vl

TTIII

V

λ

λ

(3.68)

where

)

2lsin()

2lcosh()

2lcos()

2lsinh(a

)2lcos()

2lsinh()

2lsin()

2lcosh(a

2

1

λλλλ

λλλλ

−=

+= (3.69)

3.3 Global stiffness matrix--Coordinate system transformation

The transformation between local and global co-ordinates is same as that for a beam

element, e.g. Zienkiewicz and Taylor (1991). As illustrated in Fig.3.2, the local

co-ordinate axes can be related to the global co-ordinate axes via

42

Page 63: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

(3.70) →→→→

++= kljlilx zyx

(3.71) →→→→

++= knjniny zyx

(3.72) →→→→

++= kmjmimz zyx

where lx, ly and lz are the directional cosines between the local x-axis and the global

(X,Y,Z) axes, and nx, ny, nz and mx, my, mz are the corresponding directional cosine sets

for the local y- and z-axes.

The coordinate system transformation can thus be expressed as

(3.73) [ ]⎪⎭

⎪⎬

⎪⎩

⎪⎨

⎧=

⎪⎭

⎪⎬

⎪⎩

⎪⎨

⎥⎥⎥

⎢⎢⎢

=⎪⎭

⎪⎬

⎪⎩

⎪⎨

ZYX

TZYX

nnnmmmlll

zyx

zyx

zyx

zyx

1

The local nodal displacement vector u v wT are similarly related to the global

displacement vector U V WT by

(3.74) [ ]⎪⎭

⎪⎬

⎪⎩

⎪⎨

⎧=

⎪⎭

⎪⎬

⎪⎩

⎪⎨

WVU

Twvu

1

Owing to the requirements that Cartesian axes must be orthogonal (e.g. Jeffreys,

1969), [T1]T = [T1]-1 so that

(3.75) [ ]⎪⎭

⎪⎬

⎪⎩

⎪⎨

⎧=

⎪⎭

⎪⎬

⎪⎩

⎪⎨

wvu

TWVU

T1

The local angles of rotation θxy and θzx are related to the global angles of rotation Θxx,

Θxy and Θxz via the relationship

43

Page 64: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

(3.76) [ ]⎪⎭

⎪⎬

⎪⎩

⎪⎨

⎧=

⎪⎭

⎪⎬

⎪⎩

⎪⎨

⎥⎦

⎤⎢⎣

⎡ −−−=

⎭⎬⎫

⎩⎨⎧

xz

xy

xx

2

xz

xy

xx

zyx

zyx

xz

xy Tmmmnnn

ΘΘΘ

ΘΘΘ

θθ

Thus, if the local and global element displacement vectors δe and ∆e are given by

(3.77) T7xzxy111

e w......wvu θθδ =

(3.78) T7xzxyxx111

e W......WVU ΘΘΘ∆ =

then

[ ] eee ∆Τδ = (3.79)

where [Τ]e is the element transformation matrix and is given by

(3.80) [ ]

[ ][ ]

[ ][ ]

[ ][ ]

[ ]⎥⎥⎥⎥⎥⎥⎥⎥⎥

⎢⎢⎢⎢⎢⎢⎢⎢⎢

=

1

1

1

2

1

2

1

e

T0000000T0000000T0000000T0000000T0000000T0000000T

Τ

Hence the element stiffness matrix [Κ]e in global co-ordinates is given by

(3.81) [ ] [ ] [ ] [ ]eeeTe TKT=Κ

3.4 Body forces on anchor-interface element

The equivalent nodal forces on the anchor-interface element due to body forces, such

as self-weight, can also be obtained from the principle of virtual work.

Combining the axial and lateral displacements of the anchor in Equations (3.20) and

(3.43) leads to

44

Page 65: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

[ ]

[ ] [ ] [ ]

[ ] ee

v

eu

v

NAXH

Gvu

δδδ

θ

=⎭⎬⎫

⎩⎨⎧

⎥⎦

⎤⎢⎣

⎡⋅+

=⎭⎬⎫

⎩⎨⎧

00

(3.82)

Given a set of body forces f1 f2T and a set of arbitrary virtual nodal displacements

∆e in global coordinate, the corresponding internal virtual displacement in global

co-ordinates is given by

(3.83) [ ] [ ][ ] eeT TNTVU *

1*

*

∆=⎭⎬⎫

⎩⎨⎧

Therefore, the virtual work done by f1 f2T is given by

[ ] [ ] [ ]⎭⎬⎫

⎩⎨⎧

∆=⎭⎬⎫

⎩⎨⎧

⎭⎬⎫

⎩⎨⎧

= ∫∫−−

2

11

2

2

*2

22

1*

*

1 ff

TdxNTAAdxff

VU

W

l

l

TeTeT

l

l

T

(3.84)

The external work done by the equivalent nodal forces RGe in the global co-ordinates

is given by

(3.85) eG

eT RW *2 ∆=

Equating W1 and W2 leads to

[ ] [ ] [ ]⎭⎬⎫

⎩⎨⎧

= ∫−

2

11

2

2

ff

TdxNTAR

l

l

TeTeG (3.86)

45

Page 66: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

3.5 Stresses, Strain, Axial force, Bending moment and Shear force in an element

Once the displacements have been computed, shear stress and normal stress in the

interface can be determined from Equations (3.2) and (3.32) whereas axial force, P,

shear force, Q, and bending moment, M, in the anchor can be calculated as follows:

[ ] [ ] [ ][ ] eul

xx AXdxdSeeEA

dxduEAxP δµ µµ ⎟

⎠⎞

⎜⎝⎛ +Π−== −− 1)( (3.87)

[ ][ ] evJL

dxdEI

dxvdEIxQ θδλ 2

23

3

2)( == (3.88)

[ ][ ] [ ] [ ] evvAX

dxdJLEI

dxvdEIxM θδλ ⎟⎟

⎞⎜⎜⎝

⎛⋅+== 2

2

22

2

2

2)( (3.89)

in which

⎣ xXdxd 210][ = ⎦ (3.90)

⎣ 200][2

2

=Xdxd

⎦ (3.91)

[ ] )]sin()[cos()]sin()[cos(

)]sin()[cos()]sin()[cos(2

xxexxe

xxexxeLdxd

xx

xx

λλλλ

λλλλλ

λλ

λλ

−+

+−−=

−−

(3.92)

In many FEM programs, such as CRISP (Britto and Gunn, 1990), equilibrium is

checked after each loading increment. The nodal forces required to equilibrate the

internal stresses, shear forces and bending moments in the element can also be

determined through virtual work considerations. Within an element, the generalized

stresses and strains can be defined as

(3.93) Tn0s0 QMPr2r2 στπσ =

46

Page 67: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

[ ] eT

Bdx

vddx

vddxduvrur δπε =

⎭⎬⎫

⎩⎨⎧

∆∆= 3

3

2

2

00 22 (3.94)

where [B] is the generalized strain vs displacement matrix which can be easily

constructed from Equations (3.20), (3.22), (3.44) and (3.51).

The stress strain relationship or generalized Hook’s law for the element is

(3.95) [ ]

⎥⎥⎥⎥⎥⎥

⎢⎢⎢⎢⎢⎢

=

EIEI

EAK

K

Dn

s

00000000000000000000

The equivalent nodal force vector, in local and global co-ordinate systems, is then

related to the generalized stresses in the element via the relationships

[ ] [ ] [ ] dxDBdxBR

l

l

T

l

l

Te εσ ∫∫−−

==2

2

2

2

(3.96)

[ ] [ ] [ ] [ ] [ ] dxDBTdxBTR

l

l

TeT

l

l

TeTeG εσ ∫∫

−−

==2

2

2

2

(3.97)

3.6 Verification Studies

In this section, the formulation for the proposed anchor-interface element is validated

by comparison against the exact solution of two idealised examples. These two

examples deal with the axial and flexural responses of the proposed anchor-interface

element respectively.

47

Page 68: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

3.6.1 Anchor displacement in axial direction

In the first example, the axial displacement of a linear elastic anchor embedded in a

rigid medium is compared to the exact solution. Since the ground is rigid, ∆u = u,

and Equation (3.8) can be re-written as

02

02

2

=− urK

dxudE s (3.98)

The solution of the above homogeneous differential equation takes the form of

Equation (3.9) with fu(x) =0. The coefficients C1 and C2 in Equation (3.9) can be

determined using the boundary conditions that

0

lx

0x

PdxduEA

0u

=

=

=

=

(3.99)

where P0 is the axial load applied on the top of the anchor. With these conditions, the

axial displacement along the anchor u is given by

µµ

µEAP

lxu 0

)cosh()sinh(

= (3.100)

The axial displacement of the anchor at anchor top(x=l ) u1 is thus

µ

µEAP

lu 01 )tanh(= (3.101)

Fig. 3.3 shows the exact and finite element variation of axial displacement along an

anchor for E=1.0x104 kPa, 2πr0=1.0 m, A=0.0625 m2, Ks=1.0x102 kPa/m, l=10.0 m,

P0=10 kN. As can be seen, the agreement between the two solutions is very good.

This is not surprising since the stiffness matrix of the proposed element is formulated

based on the analytical distribution of anchor deformation rather than an assumed

shape function.

48

Page 69: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

The distribution of skin friction and axial force along anchor length can also be

determined easily through Equation (3.101).

00 )cosh()sinh(2 P

lxr s µ

µµτπ = (3.102)

As can be seen from Fig. 3.4, good agreement is observed between the skin friction

from FEM using the proposed element and that from Equation (3.102).

3.6.2 Anchor displacement in flexure

As shown in Fig. 3.5, the second example deals with a linear elastic beam resting on a

rigid ground, which is fixed at one end and subjected to a point load P at other (free)

end.

From Euler-Bernouli beam theory,

n04

4

r2dx

vdEI σ= (3.103)

Substituting Equation (3.32) into Equation (3.103), and letting ∆v = v leads to

02 04

4

=+ vKrdx

vdEI n (3.104)

The general solution takes the form:

)cos()sin()cos()sin( 4321 xeCxeCxeCxeCv xxxx λ+λ+λ+λ= λ−λ−λλ (3.105)

in which 4 0

2EIKr n=λ . The boundary conditions of this problem are

49

Page 70: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

⎪⎪⎪⎪

⎪⎪⎪⎪

−==

==

==

=

==

==

==

=

Pdx

vdEIQ

dxvd

EIM

dxdv

v

lxlx

lxlx

xx

x

||

0||

0||

0

3

3

2

2

00

0

θ

(3.106)

in which M and Q denote the bending moment and shear force, respectively.

Solving equation (3.105) in view of Equation (3.106) leads to

[ ] (3.107) l2l211 e)ltanh(e2)ltanh(

f2C λλ λλ

∆−− +−−=

[ ]l213 e)ltanh()ltanh(2

f2C λλλ

∆−++= (3.108)

)1(4 21

2 +∆

−= − le

fC λ (3.109)

)1(4 214

lefC λ−+∆

= (3.110)

where

( )( ) ( ) )1(1)tan(2)1(412 22222 −++++++=∆ −−−−− lllll eeleee λλλλλ ζλζζ (3.111)

)sin()cos()sin()cos(

llll

λλλλζ

−+

= (3.112)

and )lsin()lcos(Kr

ePfn0

l

1 λλλ λ

−=

(3.113)

Comparison was made with the finite element results based on r0 =0.25m,

I=0.000325521m4 and E=108kPa.

The FEM mesh for this analysis is shown in Fig. 3.6. As can be seen, the beam was

modeled by five anchor-interface elements resting on top of a layer of five very stiff

50

Page 71: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

soil elements constrained at the bottom and left side boundaries. Fig.3.7 shows the

analytical and finite element results for the lateral deflection of an anchor with

different stiffness ratios ⎟⎠⎞

⎜⎝⎛=

EI2Kr n04λ . Once again, very good agreement is obtained

between the analytical results and the finite element results using the proposed

elements. Furthermore, as is expected, as the stiffness ratio reduces, the deflected

shape of the beam approaches that of a cantilever.

3.7 Comparison of the proposed element model with existing bar element

Bar elements are widely used in the modelling of anchor type of inclusions in soil

medium with the trade off in omitting the relative displacement between soil and the

inclusions. To verify the theoretical framework adopted in this chapter and the

computer program code developed, an idealized case of anchor in stiff soil ground

grouted to its full length subject to a pull-out load is studied with the proposed element

model and the result compared with a mesh using the conventional 3-D bar elements.

In fact, if relatively large values for Ks and Kn are used in the analyses, the proposed

element should give the similar prediction of the anchor behaviour as compared to that

from using bar element. This was the main purpose of verification work in this section.

The 3-D mesh used in this verification work is as shown in Fig. 3.8. Typical deformed

mesh from analysis using the proposed anchor-interface element is as shown in Fig.

3.9 which clearly shows the relative displacement between anchor and the ground soil.

Similar deformed mesh for analysis using the conventional bar elements is shown in

Fig. 3.10 which shows no relative displacement between anchor and the ground soil.

From Fig. 3.11, the relative displacement, which is defined as displacement of anchor

51

Page 72: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

from analysis using the proposed anchor interface element over the displacement

calculated from using the bar element, approaches unity as the eigen value µ as defined

in Equation (3.10) increases. This indicates that as Ks increases, the displacement

calculated from using the proposed anchor-interface element converges to that from

using bar element as justified earlier on. On the other hand, when eigen value µ is

relatively small, the calculated displacement of the anchor point using proposed

element can be tremendously higher than that from using bar element. In another word,

when the soil anchor interface is weak, the ignorance of soil-anchor interaction can

lead to serious underestimation of anchor displacement.

3.8 Summary of the chapter

This chapter presents the theoretical development of the FEM formulation for an

anchor-interface element which is aimed at the modelling of anchor-soil interaction.

Verification of the proposed model within elastic framework through two closed form

solutions and a comparison with the existing element model, i.e. the bar element were

also presented. Further development on the non-linear algorithm for the proposed

element will follow in the next chapter.

52

Page 73: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

CHAPTER 4

NON-LINEAR ALGORITHMS FOR ANCHOR-INTERFACE ELEMENT

4.1 Background

In this chapter, non-linear approaches for soil-anchor interaction problems using the

proposed anchor-interface elements are formulated and verified. Formulations and

FEM code were developed for Incremental Tangential Stiffness (ITS) Method and a

non-linear algorithm called “Total Load Secant Iteration (TLSI) Method”. Closed form

elastic perfectly plastic solutions for both axially loaded anchor and laterally loaded

anchor with interface socketed in stiff ground were derived and used to verify the

correctness of the proposed non-linear approaches for the proposed anchor-interface

element. Numerical results from the proposed approaches were also compared with

published field measurement data and the numerical results using conventional FEM

element types to show the viability of the proposed anchor-interface element and the

non-linear approach for modelling soil-anchor interaction.

4.2 Proposed non-linear algorithms for anchor-interface element

4.2.1 Incremental Tangent Stiffness (ITS) Approach

Anchor-soil slippage often affects the axial response of anchor significantly (Briaud

and Lim, 1999, Zhang et al. 1999). Such slippage needs to be modelled using a

non-linear interface. The purpose of this anchor-interface element is to provide a

convenient way of approximating non-linear anchor-soil interaction, especially the

slippage of the anchor relative to the soil and the dependency of the anchor-soil

interface shear strength on the confining stress around the anchor. There are other

aspects of non-linear anchor behaviour such as anchor-soil separation and plastic soil

53

Page 74: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

flow around the anchor which can significantly affect the lateral response of anchors.

Explicit modelling of local soil flow and anchor-soil separation cannot be achieved

with such an element; solid anchor elements and interface elements will need to be

used for such purposes. Thus, the Goodman-type springs in this element allows such

non-linear lateral response to be modelled only approximately. Throughout this study,

the stress-strain response of the anchor is assumed to be linearly elastic and its flexural

response is assumed to follow the Euler beam theory.

In the linear elastic formulations formulated in the previous chapter, Ks and Kn are

constants. If interface behaviour is non-linear, Ks and Kn will not be constant within

an element, so that Equations (3.14) to (3.20) strictly no longer apply. Indeed, the

parameters µ and λ will not be constants so that Equation (3.37) also does not apply.

In order to simplify computations and retain the theoretical framework proposed here,

it is assumed herein that the spring constants Ks and Kn remain constant throughout the

element although they can change over successive load increments.

Various models have been used to represent slippage in soil-structure interaction. A

detailed study of the stress-strain relationships for soil-anchor interface is beyond the

scope of this topic. The main objective of this topic is to demonstrate the feasibility

of the non-linear algorithms proposed herein for use with the anchor-soil interface

element. As such, a simple elastic-perfectly-plastic relationship for anchor-soil

slippage and transverse displacement, as illustrated in Fig.4.1, will be adopted for the

analyses herein. In addition, interface behaviour is assumed to be undrained, so that

the limiting shear stress τf is independent of the normal stress acting transversely on

the anchor fixed length.

54

Page 75: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

The algorithm for modelling non-linear interface behaviour is as follows:

(1) The load increment to which the system is subjected is first computed.

(2) The incremental element stiffness matrix is computed according to the current

stress status, and therefore spring stiffness Kn and Ks, at the mid-point of the

interface. If the interface is in an elastic state, the incremental element stiffness

in the axial and lateral directions is calculated following Equations (3.29) and

(3.30). On the other hand, if the interface is in a plastic state, the incremental

interface stiffness vanishes and only the anchor stiffness remains. In this case,

the resulting element stiffness will be derived according to Equations (4.18)

and (4.35) (see below).

(3) Solution of the global equations gives incremental displacements, from which

cumulative displacements are calculated.

(4) Incremental stresses in each element are evaluated based on the incremental

displacements and then accumulated to give the total stresses;

(5) System equilibrium checking is performed and any out-of-balance load is

added to the load vector in the next increment.

(6) Steps (1) to (5) are repeated for the required number of the loading increments.

4.2.2 Total load secant iteration (TLSI) approach.

Although the incremental tangent stiffness approach is widely used in FE analyses e.g.

(Britto and Gunn, 1990, Zienkiewicz and Taylor, 1991), it is not without its limitations.

In particular, it cannot deal with problems involving strain softening material

behaviour. For such situations, a total load secant iteration approach may confer

better computational stability. Since anchor-soil interface behaviour may indeed

55

Page 76: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

follow a strain-softening trend, the algorithm for such an approach is also discussed

herein.

In this approach, the total load is first applied to the system assuming that the interface

is linearly elastic. The resulting stress τ at the mid-span of the anchor-interface

element is then determined and compared against τf to assess if yielding has occurred.

If yielding has not occurred, the interface remains in a linear elastic state and no

further iteration is needed. On the other hand, if yielding has occurred, an element

secant stiffness matrix is then computed for the interface based on elasto-plastic

behaviour and a new solution is obtained using the new secant stiffness matrix. Since

the new secant stiffness implies a more compliant interface, load will be shed onto

other components of the anchor-soil system, and the total load borne by the interface

consequently decreases, while the relative displacement across the interface increases,

with successive iterations. This iterative process is repeated until convergence is

achieved. In this study, convergence is assessed using the 2-norm of the relative

improvement in displacement Ri, defined as

i

iii u

uuR

−= +1 (4.1)

Iterations are terminated when Ri falls below a prescribed tolerance.

56

Page 77: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

4.3 Secant stiffness matrix for elastic perfectly-plastic anchor-interface element 4.3.1 Axial stiffness

Consideration of force equilibrium in the axial direction of the anchor leads to the

relation

pa r

dxd τσ = (4.2)

where aσ is the axial stress in the anchor, τ is the skin friction and rp is the ratio of

cross-sectional area of the anchor to its perimeter. When the relative displacement

between anchor and soil is large enough to fully mobilise the limiting skin friction τf

over the entire length of the anchor-interface element, as shown in Fig. 4.1, Equation

(4.2) can be expressed as

p

fa

rdxd τσ

= (4.3)

Since the axial stress-strain response of the anchor is linear elastic, Equation (4.3) can

be written as p

f

rdxudE

τ=2

2

(4.4)

where u is the axial compression and E is the Young’s modulus of the anchor.

The solution of Equation (4.4) takes the form of

212

2CxCx

Eru

p

f ++=τ

(4.5)

Substituting the nodal displacements of the anchor, i.e.

1

20

uu

uu

lx

x

=

=

=

= (4.6)

into Equation (4.5) leads to

57

Page 78: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

⎭⎬⎫

⎩⎨⎧

⎥⎦⎤

⎢⎣⎡ −+=

2

10 1)(

uu

lx

lxxuu (4.7)

where )(2

)(0 lxxEr

xup

f −=τ

(4.8)

and l is the length of the element.

Let be a vector of virtual nodal axial displacement

to which the anchor-interface element is subjected. The effect of this virtual nodal

displacement vector on the anchor-interface element is to cause a virtual displacement

u* and virtual axial strain ε

Tu uuuuu ******

74321=δ

* along the anchor, such that

⎪⎭

⎪⎬⎫

⎪⎩

⎪⎨⎧

⎥⎦⎤

⎢⎣⎡ −= *

**

2

1

uu

lx1

lxu (4.9)

[ ] *u

* 00011l1 δε −= (4.10)

We note that u0(x) does not appear in Equation (4.9) since it is independent of nodal

displacement and thus cannot be considered a virtual displacement. The virtual

relative anchor-soil axial displacement along the interface ∆u* is given by the

difference between anchor displacement and soil displacement, that is

[ ] *uep

7,4,3i

*i

** BuNuui

δ∆ =−= ∑=

(4.11)

where [ ] ⎥⎦⎤

⎢⎣⎡ −−−−= 743ep NNN

lx1

lxB (4.12)

The virtual strain energy in the element U is given by

[ ] dxdxdu

dxduEAdxBdxEAdxuU

ll

TTuf

llf ep ∫∫∫∫ +Ω=+∆Ω=

**** )( δτεετ (4.13)

where Ω and A are the cross-sectional perimeter and area of the anchor, respectively.

The virtual work done by external force is aT*

u RW δ= (4.14)

58

Page 79: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

where Ra is the nodal axial force vector on the element.

Applying the virtual work principle to the element leads to

[ ] [ ] aull p

f

l

Tepf Rdx

lEAdx

llx

ErAdxB =−

⎥⎥⎥⎥⎥⎥

⎢⎢⎢⎢⎢⎢

⎡−

+

⎪⎪⎪

⎪⎪⎪

⎪⎪⎪

⎪⎪⎪

⎥⎥⎥⎥⎥⎥

⎢⎢⎢⎢⎢⎢

⎡−

−+Ω ∫∫∫ 00011

0001

1

0001

1

1)2( 2 δτ

τ

(4.15) where T

u 74321uuuuu=δ is the nodal displacement vector of the

anchor-interface element.

Since , Equation (4.15) can be expressed as 0)2( =−∫ dxlxl

[ ][ ] [ ]

[ ] dxBRl

EA

l

Tfau

xx

xep∫Ω−=

⎥⎥⎥

⎢⎢⎢

⎡⎥⎦

⎤⎢⎣

⎡−

−τδ

3323

32

00

01111

(4.16)

Equation (4.16) can be written in the form

[ ] epauep RRK −=δ (4.17)

where

[ ] [ ][ ] [ ] ⎥

⎥⎥

⎢⎢⎢

⎡⎥⎦

⎤⎢⎣

⎡−

−=

3x32x3

3x2ep

00

01111

lEAK (4.18)

T

fep lR⎭⎬⎫

⎩⎨⎧ −−−

Ω=

31

34

3111

2τ (4.19)

For the incremental tangent stiffness method, [K]ep is also the element stiffness matrix

but τf=0 so that Rep vanishes, whilst Ra refers to the incremental nodal load.

59

Page 80: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

4.3.2 Flexural stiffness

Let σf be the maximum unbalanced earth pressure on the anchor that can be mobilised

by lateral displacement. Consider a beam element subjected to σf along its entire

length. The flexural equilibrium of this element can be written as

f4b

4

bdx

vdEI σ−= (4.20)

where vb is the lateral displacement of anchor, I the second moment of area of the

anchor cross-section in the direction of bending and b the projected width on which the

unbalanced earth pressure σf acts.

The general solution of the Equation (4.20) is

432

23

14f

b CxCxC21xC

61x

EI24b

v ++++−=σ

(4.21)

where C1, C2, C3, C4 are constants which can be determined by introducing the

boundary conditions.

⎪⎪⎪

⎪⎪⎪

=

=

=

=

=

=

20xb

2b

1lxb

1b

|dxdv

v)0(v

|dxdv

v)l(v

θ

θ (4.22)

This leads to

[ ] vepb Nxvv δ+= )(0 (4.23)

where (4.24) Tv vvvvv 7432211 θθδ =

22

f3f4f0 x

EI24lb

xEI12

lbx

EI24b

)x(vσσσ

−+−= (4.25)

[ ] [ ]0004321 BBBBN ep = (4.26)

60

Page 81: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

⎪⎪⎪⎪⎪

⎪⎪⎪⎪⎪

+−=

+⎟⎠⎞

⎜⎝⎛−⎟

⎠⎞

⎜⎝⎛=

−=

⎟⎠⎞

⎜⎝⎛+⎟

⎠⎞

⎜⎝⎛−=

xlx2

lxB

1lx3

lx2B

lx

lxB

lx3

lx2B

2

2

3

4

23

3

2

2

3

2

23

1

(4.27)

For any virtual nodal transverse displacement and rotation vector

T*7

*4

*3

*2

*2

*1

*1

T*v vvvvv θθδ = (4.28)

where v and θ denote transverse displacement and rotation, respectively, the virtual

displacement at any point within the element is given by *bv

[ ] *vep

* Nvb

δ= (4.29)

We note that v0(x) is independent of the nodal displacement vector and therefore does

not appear in . The relative anchor-soil transverse displacement is therefore *bv *

bv∆

[ ] *vep

7,4,3i

*i

** BvNvvibb

δ∆ =−= ∑=

(4.30)

where [ ] [ ]7434321 NNNBBBBB ep −−−= (4.31)

The virtual flexural strain ε* in the anchor is then given by

[ ] *vep2

2

2

*b

2* N

dxd

dxvd

δε == (4.32)

From the principle of virtual work, the increase of internal strain energy U due to

virtual displacement can be expressed as

dxdx

vddx

vdEIdxvbU

l

bb

lbf ∫∫ ⎟⎟

⎞⎜⎜⎝

⎛⎟⎟⎠

⎞⎜⎜⎝

⎛+∆= 2

2

2

*2*σ (4.33)

Equating the right-hand side of Equation (4.33) to the virtual work done by external

force vector Rf and incorporating Equations (4.29) and (4.30) leads to

61

Page 82: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

[ ] epV

fvepV RRK −=δ (4.34)

where

[ ] [ ] [ ]∫ ⎟⎟⎠

⎞⎜⎜⎝

⎛⎟⎟⎠

⎞⎜⎜⎝

⎛=

lep

Tep

V dxNdxdN

dxdEIK

ep 2

2

2

2

=

⎥⎥⎥⎥⎥⎥⎥⎥⎥

⎢⎢⎢⎢⎢⎢⎢⎢⎢

⎡ −−−

000000000l4000l612000l2l6l4000l612l612

lEI 2

22

3 (4.35)

[ ] Tf0

l

Tfep

V 422l6l66

lrdxBbR

ep−−−−== ∫

σσ (4.36)

4.4 Stresses, strains, internal forces and bending moment

Once the displacements have been determined through global equilibrium, the shear

stress and normal stress along the interface can be determined from Equations (3.2)

and (3.32b). The axial force P, shear force Q, and bending moment M, in the

elements where the interface has yielded can be calculated as follows:

⎟⎟⎠

⎞⎜⎜⎝

⎛−+−== )(1)2()( 21

0

uul

lxEr

EAdxduEAxP fτ

(4.37)

[ ] ⎟⎟⎠

⎞⎜⎜⎝

⎛+== e

vepBdxd

dxxvd

EIdx

vdEIxQ δ3

3

30

3

3

3 )()( (4.38)

[ ] ⎟⎟⎠

⎞⎜⎜⎝

⎛+== e

vepBdxd

dxxvd

EIdx

vdEIxM δ2

2

20

2

2

2 )()( (4.39)

The generalized stresses σ and strains ε can be defined as

62

Page 83: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

(4.40) Tns QMPστσ =

T

dxvd

dxvd

dxduvu

⎭⎬⎫

⎩⎨⎧

∆∆= 3

3

2

2

ε (4.41)

4.5 Validation with idealised problems

In this section, comparison is made between closed-form solution of idealised

anchor-soil interaction problems and that predicted by the above approaches.

Although closed-form solutions for piles loaded both axially and laterally are available,

relatively few considered the effect of the pile-soil interface. The closed-form solutions

presented below relate to highly idealized situations, and are meant only for the

verification of the proposed element and algorithms. They are not meant to replicate

any real problems.

4.5.1 Axially loaded anchor in rigid medium

4.5.1.1 Closed-form solution

In this problem, the ability of the proposed formulation to replicate the slippage of an

axially loaded circular anchor with cross-sectional radius r0 in rigid ground, as shown

in Fig.4.2, is examined.

Fig. 4.2 shows the anchor being loaded at the ground surface. When the load is

sufficiently high, plastic behaviour will be manifested on a segment of the anchor-soil

interface of length lp near the ground surface, whilst elastic interface behaviour will

prevail at the segment, of length le, deeper down where anchor-soil movement is

insufficient to cause permanent slippage. These regions have been designated plastic

63

Page 84: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

and elastic zones in Fig. 4.2. The axial behaviour of the anchor in the elastic zone is

governed by

02

02

2

=− urK

dxudE s (4.42)

The solution of the above homogeneous differential equation takes the form

xCxCu µµ coshsinh 21 += (4.43)

in which 0

2ErKs=µ (4.44)

The coefficients C1 and C2 can be determined using the boundary conditions that

⎪⎭

⎪⎬

=

=

=

=

elex

0x

PdxduEA

0u (4.45)

where Pe is the axial force at the section separating the elastic zone from the plastic

zone, hereafter termed as the critical section.

For equilibrium of the plastic portion of the anchor,

effpfe lrlrPlrPP τπτπτπ 000 222 +−=−= (4.46)

Incorporating Equations (4.45) and (4.46), into Equation (4.43) leads to

µτπ

µµ

EAlrP

lxu pf

e

02)cosh()sinh( −

= (4.47)

The axial displacement of the anchor at x=le is

µτπ

µµ

EAlrP

llu pf

e

ee

02)cosh()sinh( −

= (4.48)

At this point, the skin friction reaches its maximum value τf so that

fesuK τ= (4.49)

Combining Equations (4.48) and (4.49) leads to

64

Page 85: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

µ

τπµµτ

EA)ll(r2P

)lcosh()lsinh(

Kef0

e

e

s

f −−= (4.50)

which allows le to be calculated.

Following Equation (4.5), the axial displacement of the anchor in the plastic zone, i.e.

x>le takes the form

212

0

)()( ClxClxEr

u eef +−+−=

τ (4.51)

where the coefficients C1 and C2 can be determined using the boundary conditions

PdxdurE

uu

lx

elex

=

=

=

=

20π

(4.52)

This leads to the axial displacement for the plastic zone being given by

)2)(()(

02

0ee

f

s

fe lxllxErKrE

lxPu −−−−+

−=

ττπ

(4.53)

The axial displacement at anchor top is

2

02

0

)()(

ef

s

fet ll

ErKrEllP

u −−+−

=ττ

π (4.54)

while the skin friction τs and axial force P(x) along the anchor can be determined

through

⎪⎩

⎪⎨⎧

<−−

=

ef

esef

es

lxfor

lxforKEA

llrPlx

)(2

)cosh()sinh( 0

τµ

τπµµ

τ (4.55)

and

⎪⎩

⎪⎨⎧

≥−−

<−−=

ef

eefe

lxforxlrP

lxforllrPlx

xP )(2

)](2[)cosh()cosh(

)(0

0

τπ

τπµµ

(4.56)

65

Page 86: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

4.5.1.2 Comparison of FE and closed-form solutions

To compare the FE solution using the proposed element with the closed-form solution,

we consider a 10m-long fixed length of a ground anchor, with radius r0 of 0.5m, in a

very stiff soil, subjected to an axial load of 1000kN. The Young’s modulus of the

anchor fixed length, E is taken to be 10MPa whilst the anchor-soil interface is assumed

to be linearly elastic-perfectly plastic, with a reaction modulus Ks of 1MPa/m and a

limiting skin friction τf of 100kPa. Based on Equations (4.50) and (4.44), le=8.458m

and µ=0.6325/m.

The use of a constant τf may be adequate for cohesive soils or cases where loading is

largely undrained. In general, τf is probably more realistically described using a

Mohr-Coulomb type yield envelope, so that

τf = c + σc tan φ (4.57)

where c and φ represent the cohesion and angle of friction between anchor and

soil, σc represents the normal stress on the perimeter of the anchor.

Since the cross-section of the anchor-tendon is not represented explicitly in the

anchor-interface element, the variation of σc around the perimeter of the fixed length

of anchor cannot be represented. Instead, some perimeter-averaged definition of σc,

termed cσ hereafter, will need to be employed. Several definitions of cσ may be

envisaged and these may involve using the stresses at the integration points of the

elements surrounding the anchor, or the contribution of the individual soil element to

the nodal forces on the anchor-interface element. The additional issue of selecting an

appropriate definition of cσ will not be examined in this study. It is, however,

66

Page 87: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

important to note that cσ is different from σ; the latter being a measure of the

out-of-balance transverse stress on the anchor tendon rather than a perimeter-averaged

normal stress. Moreover, the local variation of stresses around the anchor tendon

cannot be represented by this anchor-element interface. This is consistent with the

limitation, noted earlier, that the element is unable to replicate the local soil flow and

anchor-soil separation around the anchor tendon.

The value of E adopted in this comparison is evidently too low to be representative of

realistic anchor materials. However, this does not detract from the objective of this

problem, which is to provide an idealized problem for the validation of the proposed

element. The anchor is discretized into 10 elements each 1m long. For the

implementation of the proposed element, the latter is assumed to remain in an elastic

state if the shear stress τ evaluated at its mid-span does not exceed τf . Once τ > τf ,

the interface of the entire element is considered to have yielded and the plastic

formulation applies. This invariably introduces some degree of approximation into

the analysis. Nonetheless, as will be shown later, this approximation does not appear

to significantly degrade the accuracy of the solution. In the results to be presented

below, the stresses are evaluated and plotted at 5 integration points in each element.

Fig. 4.3 compares the variation of axial displacement along the anchor obtained using

the closed-form solution as well as the FE analyses assuming elastic and elastoplastic

anchor-interface element. It is noted that the value of the axial displacement is

unrealistically large, however the results is intended to verify the correctness of the

FEM result using the proposed element through this hypothetic case. As can be seen,

the FE analysis using elastic anchor-interface element progressively underestimates the

67

Page 88: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

anchor displacement. Very good agreement was obtained between the elastoplastic

anchor-interface element and the closed-form solution. Furthermore, as shown in

Figures 4.4, 4.5 and 4.6, similarly good agreement was obtained between closed-form

solution and elastoplastic anchor-interface element for the variation of skin friction and

axial force along the anchor tendon. This underlines the theoretical viability of the

proposed formulation and indicates that the approximation of assessing onset of

plasticity using the mid-span stress state does not significantly degrade the accuracy of

the solution, provided sufficient numbers of elements are used.

4.5.2 Lateral loading on an anchor in a rigid medium

In practice, an anchor system is always designed to take axial load, but the uneven

lateral ground movement around the fixed length of the anchor may impose a lateral

load on the anchor. Indeed, the lateral reaction of soil nail is just as important as its

lateral reaction. In fact, the so-called “anchor-interface element” can be used in many

cases where a solid inclusion is present. In this problem, the ability of the proposed

approaches to replicate the response of a laterally loaded circular inclusion (such as

soil nail, soil reinforcement or the fixed length of an anchor) with interface in a very

stiff ground, as shown in Fig. 4.7, is examined. As noted by Trochanis et al. (1991) in

a case of pile, under lateral loading, pile-soil separation and generalized inelastic soil

deformation governs pile behaviour. Neither of these is modelled by the proposed

anchor-interface element. In any case, inelastic soil deformation is probably best

modelled by the soil elements around the inclusions. In this respect, the lateral spring

stiffness Kn would appear to be less relevant than Ks in 3D elastoplastic FE analysis.

However, in cases where local soil flow is not well-represented by the size of the solid

68

Page 89: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

elements used, the use of Kn may allow local anchor-soil interaction to be better

captured, albeit only in a global and phenomenological manner. It is for this reason

that Kn is included in the formulation. As far as the element formulation is concerned,

the parameter Kn provides the transverse constraint to make the inclusion (anchor for

this case) remain in the bearing medium (ground soil for this case). From mathematical

point of view, Kn is merely a penalty function to govern the contact/separation between

the solid inclusion (anchor) and the bearing medium (soil). However, the values of the

parameter must be carefully chosen as the extreme values may cause ill-condition in

numerical computation.

In the same vein, the problem of an anchor in a rigid ground may seem rather

far-fetched and unrealistic. However, it should be emphasised that the main objective

of this idealized problem is to assess if the prediction of the proposed elastoplastic

anchor-interface element conforms adequately to theoretical solutions. The assumption

of a rigid ground is only for the mathematical convenience in the development of the

closed form solution which is in turn for the benefit of verifying the correctness of the

algorithm and the formulation as well as the program code. The realism of the problem

is not really in question here. It is also plausible that local yielding of the soil may

occur around the anchor as a result of local stress concentration, even though the

surrounding soil is stiff (but not rigid). In such a situation, the size of the solid soil

element may preclude accurate modelling of this local stress concentration and soil

flow. In such circumstances, the use of elasto-plastic transverse interface “springs” in

the anchor-interface element may allow the load-displacement response of the solid

inclusions to be better captured.

69

Page 90: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

4.5.2.1 Closed-form solution

Fig. 4.7 shows a circular-sectioned anchor in rigid ground and fixed at anchor base,

being subjected to a lateral load P applied at its top end. As in the case of the previous

idealised example, the terms “elastic zone” and “plastic zone” refer to the segments of

the anchor along which the interface “spring” is in an elastic and plastic state,

respectively. In elastic zone, equilibrium of the anchor and interface leads to

02 04

4

=+ bnb vKr

dxvd

EI (4.58)

and the general solution takes the form

xxb exCxCexCxCxv λλ λλλλ −+++= )]cos()sin([)]cos()sin([)( 4321 (4.59)

The coefficients C1 to C4 can be determined from the boundary conditions

⎪⎪⎪

⎪⎪⎪

=

−=

=

=

=

=

=

elex2b

2

elex3b

3

0xb

b

M|dx

vdEI

Q|dx

vdEI

0|dxdv

0)0(v

(4.60)

Substituting Equation (4.60) into Equation (4.59) allows the coefficients C1 to C4 to be

evaluated, leading to

)]l(cos)l([coshKr)]xcosh()xsin()xsinh()x[cos(f)xsin()xsinh(f

)x(ve

2e

2n0

21b λλ

λλλλλλ+

−+= (4.61)

where

)lcos()lcosh(M2)]lsin()lcosh()lcos()l[sinh(Pf eee2

eeeee1 λλλλλλλλ ++= (4.62)

)]lsin()lcosh()lcos()l[sinh(M)lcos()lcosh(Pf eeeee2

eee2 λλλλλλλλ −+= (4.63)

Equilibrium of the anchor in the plastic zone leads to

effpfe lrlrPlrPP σσσ 000 222 +−=−= (4.64)

70

Page 91: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

)]()[(212 0

20 efepfpe llrPlllrPlM −−−=−= σσ (4.65)

From Equation (4.61), the lateral displacement of the anchor vb(le) at x=le is given by

)l(cosh)l(cosfMfP

Kr)l(v

e2

e2

beae

n0eb λλ

λλ++

= (4.66)

where

)lsin()lcos()lcosh()lsinh(f eeeea λλλλ −= (4.67)

)l(sin)l(cosh)l(cos)l(sinhf e2

e2

e2

e2

b λλλλ += (4.68)

At this section, the unbalanced earth pressure on anchor reaches its maximum value σf

thus

febn lvK σ=)( (4.69)

Substitution of Equation (4.69) into Equation (4.66) leads to

0)l(cosh)l(cos

fMfPr f

e2

e2

beae

0

=−++

σλλ

λλ (4.70)

whereupon le can be determined by substituting Equations (4.64) and (4.65) into

Equation (4.70).

The lateral displacement of the anchor in plastic zone (i.e. x>le) takes the form of

Equation (4.21). Continuity of beam deflection, slope, bending moment and shear

force at x=le allows the constants of integration to be determined using the boundary

conditions

71

Page 92: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

⎪⎪⎪⎪

⎪⎪⎪⎪

=

−=

=

=

=

=

=

=

elex2b

2

elex3b

3

elexb

n

flexb

M|dx

vdEI

P|dx

vdEI

|dxdv

K|v

θ

σ

(4.71)

where θe is the angle of rotation of the beam at x=le and can be determined by

differentiating Equation (4.61) as

)]l(cos)l([coshKr)lsinh()lsin(f)]lcos()lsinh()lsin()l[cosh(f

e2

e2

n0

ee2eeee1e λλ

λλλλλλλθ

+−+

= (4.72)

This leads to the lateral displacement of the anchor in plastic zone being given by

n

feee

ee

ee

fb K

lxlxEI

Mlx

EIP

lxEI

rxv

σθ

σ+−+−+−−−−= )()(

2)(

6)(

12)( 2340 (4.73)

The bending moment and shear force along the anchor can be determined using )lx(P)lx(rM)x(M ee

2ef0e −−−−= σ (4.74)

eef0 P)lx(r2)x(Q −−−= σ (4.75)

4.5.2.2 Comparison of FE and closed-form solution

To compare the proposed element with the closed-form solution, we consider a

10m-long fixed length of an anchor, with radius r0 of 0.25m, in a rigid ground,

subjected to an axial load of 100kN. The Young’s modulus of the anchor E is taken

to be 10GPa whilst the anchor-soil interface is assumed to be linearly elastic-perfectly

plastic, with a reaction modulus Kn of 10MPa/m and a limiting unbalanced stress σf of

100kPa. The anchor is discretized into 10 elements each 1m long. The FE

implementation is similar to that for the previous example.

As shown in Fig. 4.8 to 4.10, the deflection, bending moment and shear force predicted

72

Page 93: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

using the elasto-plastic anchor-interface element agrees well with the theoretical

solution whereas the elastic element shows significant discrepancy. The slight

discrepancy between the FE-predicted and theoretical bending moment near the anchor

top arises as a result of the approximation introduced by the use of the mid-span stress

as a parameter for determining onset of plasticity.

4.6 Case study for an instrumented pile subject to axial load

In this section, a comparison is made between the axial pile displacement predicted by

using the proposed algorithms and measured data from a 8.5 m long, 400mmx400mm

square-sectioned concrete pile installed in loose silt, reported by Cheung et. al. (1991).

Strictly speaking, the proposed element is not able to model the pile in 3-D as the nil

dimension of the element in cross-sectional view will make the soil-pile interface a

line rather than a cylinder in the actual case. This will lead to overestimation of the pile

settlement. As a special application in axi-symmetric condition, the proposed element

allows for the thickness of the interface to be represented so that the pile shaft can be

properly modelled as a cylindrical surface, but the end bearing effect of the pile base is

still not accounted for. Nevertheless, the error due to such approximation is not

significant for a slender pile as shown in the calibration study below.

4.6.1 Calibration of the element model and mesh accuracy

To calibrate the accuracy of the proposed element in modelling axially loaded pile,

investigate the limitations of the proposed element and examine the mesh

independency of the problem, an FEM mesh with the pile being modelled with

73

Page 94: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

isoparameter element and the interface being modelled using the conventional slip

element is adopted as a reference.

4.6.1.1 The effects of Ks

In this section, the proposed element is calibrated in a wide range of values for shear

stiffness coefficient of interface Ks against the conventional FEM model with

isoparametric element for pile and slip element for the interface between pile and soil.

For a case where shear stiffness coefficient of the interface Ks is relatively high, the

proposed model was also calibrated against the conventional FEM using bar element

as discussed in section 7 of Chapter 3.

The background of this numerical study was chosen as a floating pile of 8.5m long

subject to a axial load of 216kN installed in homogeneous ground of E=4000kPa,

ν=0.35 with a bulk unit weight of 20 kN/m3. Two meshes were used in the analyses,

with one using 8-node solid elements for modelling pile and slip element for interface

and with the other using the proposed element having a finite thickness to represent the

pile radius

The results from the analyses are tabulated in Table 4.1 for mesh using conventional

elements and Table 4.2 for mesh using the proposed elements. Fig. 4.11 shows the

comparison between the results from the two meshes. As Ks reduces, pile top

settlement increases whereas ground settlement near pile top decreases. This is due to

the factor that the lower the value of Ks, the less the skin friction will be transmitted

into surrounding soil at shallow depth. As shown in Fig. 4.11, the proposed element

74

Page 95: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

with a finite thickness (under axis-symmetric condition only) can model the pile-soil

interaction quite well when skin friction is dominant. The departure of the two curves

at the lower Ks values indicates the significant difference in capturing the pile base

reaction by two models. The proposed model is weak in this aspect compared to the

conventional model as a result of idealizing pile into a line without diameter and hence

unable to model the end bearing of pile base. This is a major limitation of the proposed

element.

Table 4.1 Results from mesh C1A3: 8-noded element + slip elements: _________________________________________________________________

Ks pile head Settlement (mm) 7500000 8.9536 750000 8.9934 75000 9.1579 7500 10.789 750 20.924 0.75 41.838

_________________________________________________________________

Table 4.2 Results from mesh C1B using the proposed elements _________________________________________________________________ Ks pile head Settlement (mm) 7500000 9.0522 750000 9.0768 75000 9.3136 7500 11.326 750 28.48 _________________________________________________________________

4.6.1.2 The mesh independency

The mesh dependency study carried out in this section is based on the actual ground

profile and the basic soil properties of an instrumented pile in silty clay as reported by

Cheung et. al. (1991) with the ground profile being tabulated in Table 4.3:

75

Page 96: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Table 4.3 Ground profile ____________________________ Soil strata: Properties ____________________________ 0~5.1 E=2500kPa ν=0.35 5.1~10 E=6000kPa ν=0.35 10~17 E=4000kPa ν=0.35 ____________________________

Elastic analyses were conducted for this mesh independency study. The pile was

subject to an axial load of 216kN. A relatively high value of Ks (=750000) was used in

the analysis to muffle the effects of Ks. The deformed mesh used as reference is shown

in Fig.4.12(a) and the deformed mesh with the pile and interface being modelled using

the proposed element is shown in Fig. 4.12(b) and the two being overdrawn in Fig.

4.13 to highlight the differences in the computed deformation.

The settlement of the pile shaft from both meshes is as shown in Fig. 4.13 and Table

4.4. A consistently larger settlement along pile shaft from the analysis using the

proposed element is observed as shown in Fig. 4.14. This is due to the absence of the

end bearing in the analysis using the proposed element as the two curves in Fig. 4.13

are almost parallel. The error between the two is within 5%, which is acceptable for

engineering applications.

Finer meshes are used to examine the mesh accuracy for both the conventional element

and the proposed element. The deformed meshes are overdrawn in Fig. 4.15 and the

pile settlement along the shaft is shown in Fig. 4.16 and Table 4.5. As can be seen, the

accuracy is improved to a level of 1% error and the deformed meshes are conforming

quite well, which indicates the proposed element is numerically stable and mesh

independent.

76

Page 97: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Table 4.5 Pile settlement along the shaft for fine meshes

Depth (m) 8-noded Proposed (Ks=7500000)

0.00 8.732 8.787 -0.85 8.623 8.731 -1.70 8.574 8.677 -2.55 8.522 8.627 -3.40 8.476 8.580 -4.25 8.432 8.536 -5.10 8.392 8.496 -5.95 8.357 8.461 -6.80 8.330 8.433 -7.65 8.311 8.414 -8.50 8.301 8.403

Table 4.4 Pile settlement along the shaft fro coarse meshes

Depth Settlement (mm)

(m) 8-noded Proposed element

0 7.006 7.302 -1.7 6.891 7.192 -3.4 6.793 7.094 -5.1 6.706 7.008 -6.8 6.641 6.943 -8.5 6.607 6.909

4.6.2 Non-linear analyses of an instrumented pile subject to axial load

Fig. 4.17 shows the mesh used for this study. The size and shape of this mesh is similar

to that used by Cheung et al.

The pile, as well as the pile-soil interface, is modelled by the proposed

anchor-interface elements vertically orientated and aligned along the centreline of the

mesh (see segment AB in Fig. 4.17). The properties of the pile material were not

reported by Cheung et al. For this study, the Young’s modulus Ep was assumed to be

that of concrete, that is 2x107kPa. The Ks value is assumed to be following a

hyperbolic relationship as Equation (4.76).

( 21 1 SR

PKK f

n

a

nws −⎟⎟

⎞⎜⎜⎝

⎛=

σγ ) (4.76)

whereφ

φσφσσ

σσσσ

σσσσ

sin1)sincos(2

)(;)()(

;)()( 3

3131

31

31

31

−+

=−−−

=−

−=

CSR f

fult

ff

77

Page 98: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

The soil was modelled using the modified Cam Clay model. Table 4.6 shows the

soil and interface parameters used in the analysis as well as those used by Cheung et al.

Cheung et al.’s values for soil parameters are largely adopted, but the angle of friction

of 18° was felt to be too low for a silty soil. Moreover, it is rather peculiar that

Cheung et al.’s adopted pile-soil friction angle is significantly higher than the soil

friction angle, when it is usually taken to be equal or slightly higher. In this study, the

adopted angle of friction of the soil is 28°, which is a typical angle of friction for loose

silt.

Fig. 4.18 shows the results from using the proposed pile-interface element and

computed and measured results reported by Cheung et al (1991). As can be seen,

analysis using the proposed non-linear approach for the pile-interface element agrees

reasonably well with the measured data.

Table 4.6 Soil properties Parameter Cheung et al.’s values Values used in this study

Soil: Cam Clay Model Compression index λ 0.181 0.181 Re-compression index κ 0.005 0.005 Friction angle φ (°) 18 28 Critical state void ratio under 1kPa confining pressure

0.9 0.9

Pile-Soil Interface Normal “spring” constant Kn

108 108

Shear “spring” constant Ks Hyperbolic Hyperbolic K0 (kPa) 7500 7500 N 0.437 0.437 Rf 0.86 0.86 Friction angle φ (°) 23.6 23.6 Cohesion (kPa) 1 1

78

Page 99: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

4.7 Summary of the chapter

Two algorithms for elastic-perfectly-plastic analysis of a pile-interface element are

formulated and verified in this Chapter. Comparison of the FE solution using the

proposed element with analytical solutions of idealized problems shows good

agreement, thus indicating that the algorithms are theoretically sound. The

limitations of the proposed element was also investigated with a finding that the

element is unable to capture the end bearing of pile base even when used in

axis-symmetric condition for axially loaded pile, but this does not prohibit the use of

the proposed element for long slender solid inclusions where the cross-sectional

dimension is not a major concern. Such examples may include ground anchor system,

soil nails where the main concern is over the overall effects of the anchors or nails to

the mechanical behaviour of the reinforced ground rather than a near field behaviour of

soil around the solid inclusions.

Comparison with Cheung et al.’s field data shows that, using Modified cam-clay model

for soil can result in a satisfactory prediction of pile settlement comparing to FE

analysis using conventional quadratic elements for pile and zero thickness slip element

for pile-soil interface which is also concluded by Cheung et. al (1991). On the other

hand, using the proposed element for modelling a floating pile may render significant

saving of system DOFs compared to the conventional way of using solid element such

as the case reported by Cheung et. al.(1991). In the axi-symmetric problem, one

element will save 10 DOFs with conventional 8-noded isoparametric element takes 16

DOFs compared to the proposed element of 6 DOFs for pile segment. This saving can

be even more significant in a full 3-D problem particularly for pile groups of large

79

Page 100: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

numbers of piles.

Of the two non-linear approaches, the ITS approach is probably better suited to

strain-hardening pile-interface behaviour, that is, where the tangent stiffness, Ks value

decreases with load but remains positive. The TLSI formulation presented above is

only strictly applicable to elastic-perfectly-plastic pile-interface behaviour. There is

also a possibility that pile-interface behaviour can follow a strain-softening trend, with

the shear strength decreasing with relative displacement after a strength peak is

reached. In current formulations, neither of the above approaches can model such

behaviour. The incremental formulation will present problems associated with

negative tangent stiffness, Ks values. In principle, the TLSI approach is probably

better suited to such strain-softening behaviour but some modification to the

formulation is likely to be needed.

80

Page 101: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

CHAPTER 5

CASE STUDY 1 – ANCHORED PULL-OUT TEST 5.1 Introduction

In this chapter, the performance of the proposed element in modelling single anchor is

examined. The loading and unloading behaviour of a single ground anchor recorded

during pull-out test was numerically simulated in this study. Both the bonded length

and the unbonded length of the anchor were modelled with the proposed element in the

numerical simulation. The computed displacement result of anchor was compared

against the in-situ pull out test data to investigate the mechanical behaviour of the

anchor-ground interface and assess the ability of the proposed element in modelling

the salient features of the observed behaviour of anchor during the test.

5.2 Background and site geological conditions

The ground anchorage system was designed to protect a slope which had suffered a

previous failure. The slope overlies the Bukit Timah Granite Formation. This

formation is formed as a result of the intrusion of a granite porphyry dyke into the

central and north-central part of the Singapore island, trending in a

northeast-south-westerly direction. The original rock mineralogy ranges from granite

through adamellite to granodiorite and several hybrid rocks are included within the

formation.

Although the rock is likely to be a competent foundation material, it has been

weathered in large areas. Owing to the variability of the weathering process, the

degree and depth of weathering as well as the weathering pattern vary from locality to

locality. At the site, the geological material can be divided into two quite distinct

81

Page 102: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

weathering grades. At and near the ground surface is a layer of highly weathered

residual yellowish to reddish coloured silty clay, occasionally with traces of sand in it,

which is derived from the chemical weathering of the granitic material. The thickness

of this layer of residual soil varies from less than 1m to more than 20m, depending

upon location. Within this layer, undrained shear strength cu often increases

progressively with depth, showing little or no abrupt increase. The SPT blow counts

often increase gradually from less than 10 nearer the surface to about 50 blow/300mm

at large depths. Test results reported for other sites (e.g. Dames & Moore, 1983)

indicate that, although it is not a conventional over-consolidated soil as such, the

residual granitic soil do behave in many ways like an over-consolidated soil with an

over-consolidation ratio (OCR) of about 3 to 4. For instance, the residual soil often

shows a peak strength and a Skempton’s Pore Parameter A ~ 0 under shearing. This

indicates that the soil is dilative upon yielding and that the peak strength is associated

with the onset of stress-induced dilation and the subsequent softening. Table 5.1 shows

the properties of this residual soil. Below this layer of residual soil lies almost

unweathered and intact granite, which has an unconfined compressive strength often

exceeding 150MPa.

Figures 5.1 and 5.2 show a plan view and details of the ground anchorage system,

which consisted of a grid of ground beams being held down onto the slope by a series

of anchors at the intersection points of the beams. Each anchor consisted of 5 strands

high strength of steel tendons of 15mm diameter (0.6 in) each. The cross-sectional area

and perimeter of each anchor was 912.1 mm2 and 239.4 mm respectively. Each anchor

was installed in a borehole of 200mm in diameter. The bonded length of the anchor

being modelled was socketed 3m into unweathered granitic rock formation and the

unbonded length of 12.5m lies in the residual soil.

82

Page 103: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Table 5.1 Summary of In-Situ Soil Properties from Soil Investigation Test Results

Parameter In-situ Soil Natural moisture content (%) 33 - 58

Bulk density (kg/m3) 1660 - 1900

Dry density (kg/m3) 1060 – 1430

• Liquid limit (%) 55 – 98

• Plastic limit (%) 30 - 42

Particle specific gravity 2.65 – 2.73

Undrained shear strength cu (kPa) 29 – 33

• Effective cohesion c’ (kPa) Most samples give ~0

• Effective frictional angle φ’ (°) 30 - 35

Each anchor was designed to work at 500kN load. In the in-situ load test, the anchor

was loaded to 120% of its designed working load (600kN) and then unloaded to 0

followed by another cycle of loading and unloading. The recorded load displacement

curve for the test is as shown in Fig. 5.3.

5.3 Finite element model

Since the pull out test was conducted on a single anchor, there is unlikely interaction

between the adjacent anchors. Furthermore, the bonded length of the anchor is rather

short (only 3m). For this reason, the influence zone around each anchor is likely to be

quite small; thus only one single anchor with a soil column of 3m radius around the

anchor was modelled as shown in Fig. 5.4(a). The assumed boundary conditions for

the mesh can be found in Fig. 5.4(b) with the side boundary being on roller support

and the bottom boundary fixed.

To verify the mesh dependency of the problem, a mesh with 10m in radial (x) direction,

as shown in Fig. 5.5, was used in the numerical study. In view of the limited extent of

83

Page 104: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

load transfer in the rock and soil, little disturbance to ground movement was expected

to be induced by the anchor load to the far field boundaries. Therefore radial

constraints were applied to all boundaries except for the ground surface as shown in

Fig. 5.5. No change in calculated anchor head displacement when the bottom

boundary was extended for another 3m.

Deformed meshes subject to 300kN pull-out load from analyses using the two meshes

for anchor are presented in Fig. 5.6, in which the vertical displacement of anchor head

can be discerned. No significant ground movement is observed. Node 52 and node 85

are anchor head nodes of the two meshes respectively. They were initially coincide

with node 41 and node 71 respectively and are pulled out after loading as shown in

Figs. 5.6(a) and (b) respectively. The calculated anchor displacements at 300kN load

from the two meshes are 4.1306mm and 4.1307mm respectively, indicating that no

significant difference for calculated anchor head displacement from the analyses

between the two meshes was evident. Therefore, the mesh used for the analysis as

shown in Fig. 5.4 was considered adequate. The ground anchor was assumed in-place

in the ground. The drilling, grouting and backfilling processes were not modelled in

this numerical study.

The elastic modulus for the rock was initially assumed to be 2.5GPa. It should be

noted that the strength and elastic modulus of granite is highly variable. This initial

assumed value only corresponds to the weathered state with low strength. As such, a

series of parametric studies was conducted and presented in section 5.4.2.2 in which

the effects of varying elastic modules will be further examined. In fact, the computed

anchor head displacement is not sensitive to the elastic modulus if the value is higher

than this initial assumed value of 2.5GPa. The other parameters studied are as follows:

84

Page 105: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Poisson ratio ν=0.25, c=200 kPa and φ=35°. Not much information on rock from this

particular site was available from site investigation. Parametric study on the influence

of the rock properties to the anchor head displacement versus load was carried out and

the results are presented in section 5.4.2.2 .

Following the recommendation by Dames and Moore (1983), the elastic modulus for

residual soil was taken as 2.0x104 kPa, c’=10 kPa and φ’=22°. Some of the soil

properties are tabulated in Table 5.1 and the properties used in this study are listed in

Table 5.2. Parametric study on the influence of the soil properties to the anchor head

displacement versus load was also carried out and the results are presented in section

5.4.2.2.

Table 5.2 Material Properties Used for the Analyses

No Material E (kPa) ν c ’ (kPa) φ’ (o) γ (kN/m3)

1 Soil 1.0x104 0.25 0 28 18 2 Rock 2.5x106 0.25 200 35 20

Anchor-soil Interface properties

E (kPa) Ks (kPa/m) Ksur (kPa/m) c’ (kPa) φ’ (o)

3 bonded length 2.1x107 1.5 x 104 1.8 x 104 200 20 4 unbonded length 2.1x108 5.0 x 103 6.0 x 103 1.0 10

The elastic moduli of each anchor tendon and grout are 2.1x108 kN/m2 and 2.5x107

kN/m2 being typical elastic moduli of steel and C30 concrete respectively. The

equivalent cross-section area and perimeter for anchor bonded length and anchor

unbonded length are 9.12x10-4m2, 0.239m, 0.0396m2 and 0.628m respectively. The

equivalent area of the bonded length is calculated as follows:

85

Page 106: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

c

sscb E

AEAA += (5.1)

where, Ab - the cross-section area of anchor bonded length;

Ac - the cross-section area of grout annulus;

As - the cross-section area of tendon strand;

Ec - the elastic modulus of grout;

Es - the elastic modulus of steel tendon.

It is assumed that the bonding between grout and tendons would be much stronger than

that between grout and the rock therefore the relative slippage would occur within the

interface between rock and grout.

A survey of existing literature shows that there is little or no information on the shear

stiffness Ks of the interface, both for the unbonded as well as the bonded length.

Furthermore, it is quite likely that this parameter is highly dependent upon the quality

of the construction and site conditions. For these reasons, a range of Ks values were

studied, and values which gave a good fit to the experimental data are highlighted.

The value of the cohesion c may similarly depend on the quality of construction and

site conditions; its range is also likely to be wide. For example, the bond of C30 grade

grout could easily reach to the order of 10MPa. On the other hand, where there is the

significant mixing between grout and soil, such as jet grouting, the strength of

resulting mix can be as low as 350kPa (Lee, 1999). For this reason, a range of values

between 200kPa and 10MPa was studied in the parametric study in section 5.4.2.3

below.

The initial stresses in the ground affects the numerical results especially in terms of the

86

Page 107: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

local yielding of rock and soil around anchor as well as the plastic behaviour of the

interface between anchor and the rock/soil around it. The initial site stress field for the

particular site is a complex one due to the complex site terrain and the geological

history. Little is known of the actual site stresses from site investigation. Furthermore,

drilling of boreholes for anchor disturbs the in-situ stress field.

As the main objectives of this numerical study are to investigate the soil/rock anchor

interaction as well as the mechanical behaviour of the soil/rock-anchor interface, the

ambient stresses are assumed to vary linearly along the anchor length, as shown in Fig.

5.7. This is because the site stresses are related to overburden which are generally

proportional to the depth. Since the anchor is straight and the slope of ground surface

is also uniform, the overburden should be varying linearly with the anchor length.

At the anchor head, a nominal confining stress of 10 kPa was assumed. At the lower

end of the unbonded length, the site stresses were assumed as σx=286.9 kPa, σy=248.6

kPa and these stresses were estimated based on the overburden depth of 17m (refer to

Fig. 5.2) and the horizontal earth pressure of Ko=0.75 (Dames and More, 1983).

Parametric study was conducted on several values of Ko to investigate the significance

of the site stresses to the calculated anchor load-displacement curve and the results are

presented in section 5.4.2.1.

5.4 Results and discussions

Table 5.3 shows the various cases examined and the combination of parameters. The

results of these cases are presented in Figures, 5.8 to 5.20.

87

Page 108: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Table 5.3 List of cases studied

No. ID Details / Objectives

1 POTM1,POTM2 Different meshes for mesh dependency study

2 P2DBAR Unbonded length modelled with bar element

3 P2D

Unbonded length modelled with the proposed

anchor-interface element

4 PostYield

To investigate the post yielding behaviour of anchor

using the proposed element.

5 P2DPR Erock=2.5E6, 2.5E4, 2.5E2 (kPa)

6 P2DPS Esoil=2.0E4, 2.0E3, 2.0E2 (kPa)

7 P2DPB

Parametric study on the bonded length properties : Ks, c,

φ, Ksres, Ksur

8 P2DUB

Parametric study on the unbonded length properties :

Ks, c, φ, Ksres, Ksur

9 P2DIS1~3

Parametric study on in-situ stress with Ko=0.5,0.75,1.0

respectively

As the anchor was socketed in rock formation, total stress analysis was adopted in this

study. Preliminary studies were carried out for the justification of mesh boundary

conditions as well as in-situ conditions as mentioned above. The so-called

back-analyses were then carried out to obtain the properties of the interfaces for

unbonded length and bonded length. Parametric studies were then conducted to

investigate the mechanical behaviour of anchor-rock interface as well as the sensitivity

of calculated anchor displacement to the material properties.

88

Page 109: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

5.4.1 Effects of different methods of modelling anchors

Fig. 5.8 shows the computed load displacement curves for a few cases in which the

unbonded length interface is considered to be perfectly smooth, this being the

conventional approach to simulating anchors (e.g. Briaud and Lim, 1999; Yoo, 2001).

As can be seen, where the cohesion of the bonded length interface is sufficiently high,

the load-displacement curves are perfectly elastic and exhibit no hysteresis, as is

expected. On the other hand, where the cohesion of the bonded length interface is too

low, yielding occurs along the interface and the load-displacement curves show a

plastic response, which is accompanied by a large permanent displacement. This

differs from the measured load-displacement curve in two aspects:

(1) The computed permanent displacement is usually much larger than the

measured values, as is evident by comparison in Fig.5.9.

(2) The slight curvature in what was supposed to be the elastic segment of the

measured loading and unloading curves is not reproduced in the computed

curves from analyses using friction free bar elements for anchor unbonded

length.

Figs. 5.9 and 5.10 show the computed load-displacement curves for a case in which the

unbonded length possesses a small amount of cohesion and friction. This is quite

plausible since a small amount of residual friction may be present in the unbonded

length, especially in this project where the unbonded length was not grouted with

grease. The source of the cohesion is less obvious. For this reason, the cohesion was

reduced to a minimum value of 1.0 kPa to replicate the load-displacement curve. As

can be seen from Fig. 5.10, this significantly improves the matching between the

89

Page 110: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

computed and measured results. As these figures show, the hysteresis loop and the

slightly non-linearity in the loading and unloading curves can now be explained in

terms of the interaction between the residual bond in the unbonded length and the

bonded length. The initial portion of the load-displacement curve, marked OA, in Fig.

5.11 of a typical observed hysteresis loop of anchor pull-out test, and as numerically

simulated in Fig. 5.12, reflects the combined interface stiffness of the fixed and

unbonded length. This occurs prior to the initiation of slippage at the unbonded length.

As the load is increased, a second stage appears (Point A and beyond as in Fig. 5.12) in

which the slope of the load-displacement curve is reduced. This can be explained by

the occurrence of slippage along the unbonded length, thereby reducing its tangent

stiffness to zero. The slope of the load-displacement curve at this stage would therefore

represent the stiffness of the bonded length alone. Point C reflects the yielding of the

anchor bonded length. Segment CD represents the unloading when reverse shear

develops over the unbonded length. From point D to E reflects the slippage along the

unbonded length due to shear yielding in a reversed direction.

Fig. 5.13 shows the distribution of the skin friction along the anchor during this stage

for one of the cases. As can be seen, although much of the load is carried by the

bonded length, owing to its higher stiffness, a proportion of the load is also distributed

to the unbonded length. Thus, during this stage, only part of the loading appears on the

bonded length. From the results of the parametric studies, a good match to the

measured load-displacement curve is obtained by setting the stiffness of the unbonded

length at about 30% of that of the bonded length. For this combination of relative

stiffness, the load distribution along the fixed and unbonded lengths is roughly 60-40

for this project. The relative low load transfer to the bonded length occurs because of

90

Page 111: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

the much longer unbonded length, relative to the bonded length. Since anchors

typically have much longer unbonded lengths than bonded lengths, this implies that

during the initial stages of the loading, a significant proportion of the load may not

appear over the bonded length. Thus, if the anchor is locked-off at a low-level preload,

much of the load may not appear over the bonded length. For example, about 80kN of

the load was taken by the unbonded length referring to the measured

load-displacement curve in Fig. 5.10. Thus if the anchor was locked-off after

preloading to 200kN, only 120kN would appear across the bonded length.

This may not be desirable since much of the unbonded length may lie within the

potential failure wedge, and a significant loss of load within this failure wedge would

mean that the net stabilizing force on the potential failure wedge is substantially

reduced. Furthermore, since the unbonded length is not grouted, its stiffness and load

carrying capacity may degrade in the long-term as moisture ingress into the void and

reduce the interface friction. For these reasons, it may be desirable to ensure that the

designed stabilizing force does appear across the bonded length. This can be easily

ensured during preloading, by projecting the second stage segment of the curve

backwards until it intercepts the load axis. The value of the intercept then represents

the load carried by the unbonded length.

In some of the load-displacement curve, a third stage occurs at even higher loading, as

indicated as segment BC in Fig. 5.11. This phenomenon was numerically simulated in

Fig.5.12 by using a set of deliberately selected parameters. This can be attributed to the

progressive yielding of the anchor-rock bond. This notion is supported by the

observation that, in such cases, the amount of irrecoverable displacement after full

91

Page 112: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

unloading is relatively large. As shown in Fig. 5.12, this can be replicated to some

extent in the analysis even though the interface is assumed to be

elastic-perfectly-plastic in behaviour. This is because yielding along the bonded length

occurs progressively along the length, rather than simultaneously along the entire

length. In other load-displacement curves, this portion is less evident, indicating less

degradation of the bonded length; thus the extent of degradation in segment BC may

be taken to be an indication of the quality of construction of that anchor. Since

premature yielding is an undesirable situation and may be indicative of defective bond

in part of the anchor, it is important to detect and assess its effect on the performance

of the anchor as a whole. This can be achieved partially by unloading and then

reloading the anchor and then checking to see if the degradation in anchor stiffness is

significant.

The sequence of events during unloading is similar in all the numerical simulations.

During initial unloading, residual friction along the unbonded length reverses

direction, causing a rapid reduction in loading with displacement. This occurs until

reverse slippage of the unbonded length occurs, which moderates the reduction in load

(e.g. segment DE of Fig. 5.12). The effect of this phenomenon is also reflected in the

redistribution of skin friction along the anchor length during unloading, as shown in

Fig. 5.13. The effects of interface drag on the unbonded length occur even at relatively

low stress level, as the unbonded length will reach its shear stress limits at low load

level. Unloading at low level of load will cause the same displacement reduction as

indicated in segment CD of Fig. 5.11 and thus the hysteresis. On the other hand, if the

unbonded length is modelled as a perfectly smooth segment, using a friction-free bar

element, this low stress level hysteresis will not be simulated. As Fig. 5.9 shows,

92

Page 113: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

hysteresis can be induced with a friction-free bar by allowing the bonded length to

yielding, but this hysteresis will initiate at much higher load level and is characterised

by an almost complete degradation of the anchor stiffness posterior to yielding.

5.4.2 Parametric study

5.4.2.1 The influence of in-situ stresses

Due to the complexity of the site terrain and ground conditions, the in-situ stresses

assumed in this study may not be accurate. In order to assess the effects of the

prescribed in-situ stress on the numerical results, a parametric study was conducted

with the different in-situ stresses derived from different values of Ko. The values of Ko

and the derived site stresses as transformed to the local coordinate system of the mesh

are tabulated in Table 5.4.

Table 5.4 In-situ stresses

Κo= 0.5 0.75 1 at the bottom of soil layer σx (kPa) 267.8 286.9 306 σy (kPa) 191.3 248.6 306 τxy (kPa) NA NA 0 at bottom of mesh σx (kPa) 396.3 424.6 452.9 σy(kPa) 283.1 368.0 452.9 Anchor head displacement versus load curves from this parametric study are presented

in Fig. 5.14. As can be seen from the figure, slightly larger anchor head displacement

was computed for lower values for Ko. This is expected because of the reduced

contribution from confining pressure to the ultimate shear strength of the anchor.

However, the influence of the in-situ stress to anchor head displacement is relatively

93

Page 114: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

insignificant compared to the influence arising from the method of modelling. This is

because :

(1) The bonded length is relatively short, being only 3m.

(2)The anchor was socketed in sound rock formation with no significant local

yielding, as the measured load-displacement curves shows, and

(3) The bond strength between anchor grout and rock along the bonded length is

largely contributed by the interface cohesion rather than the interface friction,

and is therefore not significantly affected by the in-situ stresses in the rock.

(4) Contribution of the unbonded length to the overall load capacity is

relatively minor.

However, this may not be so if the contribution from the unbonded length is significant,

for instance, in cases where the unbonded length is very long.

5.4.2.2 Influence of rock/soil properties

The elastic modulus and uniaxial strength of fresh granite can significantly exceed

10GPa and 10MPa respectively according to other literature (Galybin et. al. 1999).

Dames & Moore’s (1983) test results show that the unconfined compressive strength of

Bukit Timah granite in Singapore can range from as low as 500kPa to as high as

130MPa, depending upon the degree of weathering. In the initial analyses, relatively

low values of 2.5 GPa and 200 kPa for the Elastic modulus and cohesion respectively

were assumed in the study. These values are likely to lie at the low ends of their

respective ranges of variation. To assess the sensitivity of the results to variations in

these material parameters, a parametric study was carried out. Results of analyses

94

Page 115: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

using different values of Elastic modulus for rock are presented in Fig. 5.15. As can be

seen from the figure, no significant influence was observed for higher values of the

Elastic modulus. This is because the Elastic modulus of rock is sufficiently higher than

the Ks value of the interface so that the anchor head displacement is dominated by the

relative displacement of the anchor and the rock in the interface.

The effect of variation in the c’ and φ’ values of the bonded length is not sensitive to

the calculated anchor head displacement. In the practical ranges of c’=50 ~ 100 kPa

and φ’=30° ~ 35° for rock, the calculated load displacement curves are almost

identical.

The soil properties were well documented for the residual soil of the site (Dames and

Moor, 1983) and some of them were abstracted and tabulated in Table 5.1. The

parametric study to the sensitivity of soil properties on the calculated anchor head

displacement was conducted. This is partly because of the concern over the confidence

to the selected soil properties, especially the elastic modulus of soil, as the in-situ soil

properties may vary from spot to spot and it is impossible for the site investigation

program to cover all these possible variations.

Results from this sensitivity study on the influence of elastic modulus of soil are

presented in Fig. 5.16. As can be seen from the figure, lowering the elastic modulus of

soil leads to larger anchor head displacement, as expected. The sensitivity is more

pronounced when elastic modulus falls below 200 kPa which is very unlikely for the

soil at the particular site. Most importantly, the lower value of elastic modulus for soil

lead to insignificant friction along the anchor unbonded length. As shown in Fig. 5.16,

95

Page 116: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

for Esoil=200 kPa which is even lower than the Ks for the unbonded length (Ks =5000

kPa/m), the loading unloading curve almost replicate the results from analyses using

friction free bar element for anchor unbonded length. There is insignificant or no

hysteresis loop in the loading/unloading curve due to insignificant amount of friction

along anchor unbonded length which is even not enough to mobilize the slippage

within the interface between soil and the anchor. The relative displacement between

soil and anchor at the maximum load of 600kN is 10.16mm, though significant

heaving of 49.2mm is observed in soil around anchor, as are shown in the deformed

mesh, Fig. 5.17(a) and the displacement vector field, Fig. 5.17(b). In contrast, for the

case when elastic modulus of soil is 20000kPa, the relative displacement between soil

and anchor at the maximum load of 600kN is 46.6mm while the maximum heave

around anchor point is 9.6mm.

As discussed above, what can be concluded from this part of the parametric study is

that the elastic modulus of site soil is higher than 200 kPa, otherwise the initial

curvature of the loading curve would be hard to justify. The elastic modulus of 10000

kPa used in the fitting analysis is believed within reasonable range, though the

possibility of even high modulus cannot be totally excluded based on this parametric

study alone.

5.4.2.3 Influence of interface properties

The mechanical properties of the anchor soil/rock properties are the most sensitive

factors affecting the calculated load displacement curve. Unfortunately, little or no

information about these properties are available. Back-analysis of these properties by

96

Page 117: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

fitting the calculated curve with the measured one seems to be the natural choice. As

the back-analysis method does not guarantee the uniqueness of the results

automatically, necessary justifications and parametric studies are still indispensable to

the numerical analyses.

As shown in Fig. 5.18 and 5.19, the Ks value of the anchor bonded length affects the

slope of the loading/unloading curve as indicated as segment AB in Fig. 5.11 and the

Ks value of the anchor unbonded length affects the slope of the initial loading curve as

indicated as segment OA in Fig. 5.11. Similarly, the slopes of the segments BC, CD

and DE are dependant on the residual stiffness, Ks value for the anchor bonded length

after yielding, unloading/reloading stiffness Ksur for anchor unbonded length and

unloading/reloading stiffness Ksur for anchor bonded length respectively.

The ultimate shear strength of the anchor unbonded length, together with the

unloading/reloading stiffness of anchor rock/soil interface determine the dimension of

the hysteresis loop as shown in Fig. 5.20. By fitting the measured curve, it was found

that the unloading modulus of the anchor rock interface is 20% higher than the loading

ones for the particular anchor. The ultimate shear strength of the anchor rock/soil

interface is related to the cohesion, c’ and the frictional angle, φ’. It was found through

sensitivity study that the anchor head displacement is not sensitive for c’ larger than

500 kPa as the anchor interface will be working in elastic mode if the shear strength of

the interface is sufficiently high.

97

Page 118: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

5.5 Summary of the results and discussions

The numerical analyses and simulation as well as the parametric study based on the

proposed element model for anchor rock/soil interface presented above in this section

reasonably replicated the field measured anchor head displacement from a pull-out test

and explained some important observed phenomena such as hysteresis in the load

displacement curve at low load level. Comparison with the analysis result using

friction-free bar highlighted the advantages of the proposed element model over the

bar element in modelling the anchor rock/soil interface. The results presented above

also positively approved the proposed element model for the numerical analyses of

anchor soil/rock interactions.

98

Page 119: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

CHAPTER 6

CASE STUDY 2 -- AN ANCHORED RETAINING WALL IN SAND

6.1 Introduction

In this section, the performance of the proposed anchor element in a full-scale

retaining wall analysis was assessed by applying it to a case study of a full-scale

instrumented tieback wall in sand reported by Briaud and Lim (1999). The site

conditions and construction aspects have been reported in Briaud and Lim (1999), and

therefore, will only be summarized briefly herein. The excavation was generally 7.5m

deep and was supported with 9.15m long steel H-piles and timber lagging boards. Both

the steel H-piles and timber lagging boards were tied back into the retained soil by 2

levels of anchors. The alignment of the tieback wall is shown in Fig.6.1. A section

view of the wall with the two rows of anchor is shown in Fig. 6.2. The two rows of

high pressure grouted anchors are inclined at 30o to the horizontal. One is located at

1.8m, and the other is at 4.8m, below the top of the wall. Each of the anchors is 89mm

in diameter and has a total length of 12.35m of which 7.3m is bonded. Each steel

tendon consisted of a 25mm-diameter Dywidag bar. For the first level anchor, the

lock-off load is 182.35kN and the second level anchor lock-off load is 160kN. The

timber lagging boards have an elastic modulus of 1.365x106 kN/m2. The soldier piles

are HP6x24 I-sections and have an elastic modulus of 2.0x108 kN/m2.

The ground consisted of a 13m-thick layer of medium dense fine silty sand deposit

which overlies a bed of hard shale. The water table is 9.5m below the top surface of

the ground as shown in Fig. 6.3.

99

Page 120: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Ground investigation has revealed that the sand has a total unit weight of 18.5 kN/m3,

a standard penetration test blow count that increases from 10 blows per 0.3m at the

surface to 27 blows per 0.3m at the bottom of the soldier piles and borehole shear

friction angle of 32o with no cohesion. Cone penetration test returns a cone resistance

of 7 MPa at a depth of 9m. The soil properties from site investigation are tabulated in

Table 6.1.

Table 6.1 Soil properties

Depth Unit weight SPT, N CPT c' φ'

(m) (kN/m3) Blows/30cm kPa kPa o

Near ground surface 18.5 10 - 0 32

9 18.5 27 7000 0 32

According to the limited information from the Briaud and Lim (1999), the wall was

instrumented with vibrating wire strain gauges on the soldier piles to obtain bending

moment profiles, with inclinometer casings to obtain horizontal deflection profiles, and

with load cells at the anchor heads to monitor the anchor forces.

6.2 Finite Element Mesh

Both 2-D and 3-D FE analyses were conducted in this study. The 3-D mesh used in

present study is shown in Fig. 6.4, but the 2-D mesh is not shown as it is merely one

section of the 3-D mesh. The 3-D mesh comprises 2724 brick elements, 154 wedge

elements, 40 anchor/interface elements and a total of 13753 nodes. This 3-D mesh was

derived by extruding the 2-D mesh in the out-of-plane direction, except for the anchor

and soldier pile elements. The 2-D analyses were conducted to compare with 3-D

analyses so as to highlight the 3-D effects of the ground anchor system. In the 2-D

100

Page 121: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

analyses, the anchors and soldier piles are represented by their average properties

smeared in the out-of-plane direction.

The finite element domain consists of six components, namely the medium dense fine

silty sand; soldier piles (including waler), timber lagging, unbonded anchor, bonded

anchor and locking elements for simulating locking-off of the anchor heads. The soil,

soldier pile, waler and timber lagging were modelled using 20-noded isoparametric

linear strain, hexahedral elements in the 3-D analysis and by 8-noded isoparametric

linear strain quadrilateral elements in the 2-D analysis. The bonded and unbonded

lengths of the anchor as well as the locking elements were modelled using the

proposed anchor interface elements.

6.3 Boundary Conditions

Following Briaud and Lim (1999), the soil profile and the geometry of the tieback wall

on both sides of the excavation are assumed to be equivalent. This allows the finite

element method to be carried out for half of the excavation, by taking advantage of

symmetry. The thickness of the 3-D mesh in the in-plane direction is taken as the

thickness which represents the repeated pattern of elevation section.

The bottom of the mesh is chosen to coincide with the top of the old hard shale. The

distance from the bottom of the excavation to the old hard shale will hereafter be

denoted by Db and is equal to 5.5m. Briaud and Lim (1999) noted that, when using a

linear elastic soil model in the simulation, Db had an influence on the vertical

movement of the ground surface at top of the wall but comparatively very little

101

Page 122: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

influence on the horizontal movement of the wall face.

On the side of the retained soil, the side boundary is set at a stand-off of 39m from the

retaining wall. This is more than 5 times the depth of the excavation. Preliminary

analyses using different values of stand-off showed that, beyond this value, further

increases in the stand-off results in little changes to the computed results.

As shown in Fig. 6.5, nodes on the two vertical sides are constrained to move only

vertically. Nodes along the bottom boundary of the mesh are completely fixed in all

directions, to simulate good coupling between the rock surface and the soil.

6.4 Material properties and calculation parameters

Table 6.2 shows the material properties used in the numerical analysis for soil, soldier

piles, timber lagging and high-pressure grouted anchors. Soil properties used in the

original analysis by Briaud and Lim (1999) were obtained by matching with the in-situ

measured data from the instruments. In the present study, material properties for

soldier pile and timber lagging, cohesion and frictional angle for soil as well as the

at-rest earth pressure coefficient were adopted in accordance with those documented in

the literature (Briaud and Lim, 1999). A modified hyperbolic model for soil with the

minimum elastic modulus set at 300 kPa was used in the reported literature.

The soil model in present study is elastoplastic model of Mohr-Coulomb criterion, with a linear variation of elastic modulus

against depth. It was found that with elastic modulus at ground surface being set at 300 kPa and the rate of increase per meter

depth being set at 2200 kPa/m for soil in present study, the curve of elastic modulus versus depth in present study matches well

with that of the initial modulus of soil from Briaud and Lim’s data as shown in Fig. 6.6. The Poisson ratio of the soil was

102

Page 123: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

assumed to be a constant value of 0.3 which is a typical value for sand. The material properties used for the present study are also

listed in Table 6.2.

The geometrical/sectional properties of the anchors are calculated according to the

dimension of anchors and boreholes as given by Briaud and Lim (1999) and they are

listed in Table 6.3. It is worth pointing out that the stiffness of anchors was calculated

based on the geometric and material parameters of the anchors in the present study,

based on the rationale discussed below. This is to enable an evaluation of the

performance of the proposed element in a predictive, rather than back-analysis,

scenario. It should be mentioned that, in Briaud & Lim (1999), the measured stiffness

from preloading test was taken to be the stiffness of anchor unbonded length and a

reduced elastic modulus for grout, being 40% of that for concrete, was used for the

anchor bonded length to approximate the effects of anchor-soil interface.

The most important properties of the proposed model are Ks, c’, and φ’. These

interface properties were not given by Briaud & Lim (1999) as they modelled the

anchor bonded and unbonded lengths using beam and spring elements with no slippage

at the anchor-soil interfaces at the bonded and unbonded lengths. The parameters and

material properties for anchor interface element are listed in Table 6.4 and Table 6.5

for 3-D and 2-D analyses respectively.

Table 6. 2 Comparison of material properties used in the present study

Briaud and Lim (1999) Present study Soil (Hyperbolic model) Initial tangent modulus factor K= 300

Elastoplastic model of Mohr-Coulomb with E varies linearly with depth.

103

Page 124: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Initial tangent modulus exponent n= 0.85 Strength ratio Rf= 0.93 Friction angle φ’= 32° Cohesion c’ =0 Unloading-reloading modulus number Kur= 1,200 Bulk modulus number KB =272 Bulk modulus exponent nB =0.5 Unit weight γs =18.5 kN/m3

Earth pressure coefficient K0 =0.65

i.e. E=Eo+me*(y-yo) Eo=100 kPa me=2200 kPa/m yo =0 (y coordinates of ground level) Friction angle φ'= 32° Cohesion c’ =0 ν=0.3

Unit weight γs =18.5 kN/m3

At-rest earth pressure coefficient K0 =0.65 Anchor Tendon unbonded length =5.05 m Tendon bonded length =7.3 m Lock-off load—Row 1 =182.35 kN Lock-off load–Row 2 =160.0 kN Tendon stiffness—Row 1 = 19,846 kN/m Tendon stiffness—Row 2 = 19,479 kN/m Angle of inclination β =30° Diameter of the borehole =89mm Egrout=0.4Econcrete Econcrete =2.0 x 107 kN/m2

Properties of bonded/unbonded length are listed in Tables 6.3, 6.4 and 6.5. Lock-off load—Row 1 Fx= -64.72kN Fy=37.37kN (2-D) Fx= -78.96kN Fy=45.59kN (3-D) Lock-off load–Row 2 Fx= -56.79kN Fy=32.79kN (2-D) Fx= -69.28kN Fy=40.0kN (3-D) Egrout=2.0 x 107 kN/m2

Wall facing Wall height =7.5 m Thickness of wall facing =0.1 m Elastic modulus of wood board = 1.365 x 106 kN/m2

E = 1.365 x 106 kN/m2

Soldier pile Length of soldier pile = 9.15 m Embedment = 1.65 m Diameter of pipe pile = 0.25 m Thickness of pipe pile = 0.00896 m Horizontal spacing of piles =2.44 m Elastic modulus E =2.1 x 108 kN/m2

Flexural stiffness EI = 11,620 kN m2

Axial stiffness AE =1.47 x 106 kN

For 3-D analyses (0.25x0.25) I=3.255x10-4 (m4) E=11,620/I=3.57x107 (kPa) For 2-D analyses I=1.302x10-3 (m4) E=0.5*11,620/I/1.22=3.657x106 (kPa)

Table 6.3 The geometric/section properties of anchors

Material Diameter Area Perimeter E I

104

Page 125: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

(m) (m2) (m) (kPa) (m4)

Steel 0.025 0.000490

9 0.0785 2.1x108 4.0x10-8

Grout 0.089 0.0062211 0.2796 2.0x107 3.0x10-6

Composite (steel+ grout) 0.089 0.0062211 0.2796 3.5x107 2.0x10-6

Table 6.4 The material properties of anchors for 3-D analyses

Anchors Ks Ksur Kn c' φ’

(kPa/m) (kPa/m) (kPa/m) (kPa) (o)

Unbonded 1.0x107 2.0x104 1.0x108 1 10

Bonded 1.0x108 2.0x105 1.0x108 70 30

Table 6.5 The material properties of anchors for 2-D analyses

Anchors E Ks Ksur Kn c' φ’

(kPa) (kPa/m) (kPa/m) (kPa/m) (kPa) (o)

Unbonded 8.61x1074.1x106 8.2x103 4.1x107 1 10

Bonded 1.43x1074.1x107 8.2x104 4.1x107 70 30

The frictional angle φ’ for anchor-soil interface was estimated based on the frictional

angle of soil. A slightly lower value of 30o was adopted for this case to reflect the

relatively weakened interface between soil and anchor due to construction disturbance.

As for the unbonded length, nominal values of c’=1 kPa and φ’=10o which is the

typical value of the frictional angle between steel and polyester, were used in the

present study.

Since the bonded length was constructed by pressurized grouting, the cohesion of the

interface may vary over a wide range and can be significantly higher than the cohesion

of soil. Based on the perimeter of the grout annulus and the bonded length, the surface

105

Page 126: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

area of grout annulus is about 2m2. The recorded anchor preload is 182.35kN and

160kN for Row 1 and Row 2 anchors respectively. Using an interface friction angle of

30° and a confining stress of 92kPa, which is derived from the approximate depth of

the bonded length of the Row 1 anchors, the minimum cohesion of the anchor-soil

interface needed to prevent pullout during preloading is 36kPa. The actual value of

shear strength would be 2 to 3 times higher than 36 kPa taking into consideration of

Factor of Safety in the design. Schnabel (2002) recommended some values of bond

strength between anchor and soil as shown in Table 6.6. A value of 70kPa for the

interface cohesion of the bonded length falls within the recommendation as according

to the SPT number of the site of >8 blows per 300mm and it also compares well with

the minimum value of 70kPa.

Table 6.6 Interface properties (after Schnabel, 1982)

Soil type SPT (N) Adhesion (bond strength) between soil and anchor (kPa)

Silty clay 3~6 24~48 Sandy clay 3~6 35~48 Medium clay 4~8 35~60 Firm to stiff clay >8 48~72

As there is limited or no knowledge about the Ks between anchor and soil for the site,

the selection of Ks value for the present study was made with reference to the reported

properties for interface between sand and concrete pile used in Cheung’s study in a

soil-pile interaction problem (Cheung et. al., 1991). Thus a value of Ks=1.0x108

kPa/m was used for bonded length of anchor in present study. Considering the weaker

interaction between soil and anchor in unbonded length, the Ks value for the unbonded

length was taken as 10% of that in bonded length. According to the findings in section

106

Page 127: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

5.4.2.3 of Chapter 5 for pull-out test, this value only affects the initial slope of the load

displacement curve of the anchor. As the anchor in this case was stressed to a quite

high load level, the effects of the Ks value for the unbonded length on the performance

of the anchor would be insignificant and limited. The Ks value for unloading/reloading

was assumed to be twice that of the Ks value for loading mode. As this value affects

the hysteresis of the load displacement of the anchor, and the anchors in this case are

generally working in loading mode. Therefore, the effects of this parameter are also

insignificant. A nominal value of 100kPa/m was used as residual stiffness post-failure.

This is basically to avoid breaking down of numerical computation in case of total

yielding of bonded length.

As discussed in section 2.9 of Chapter 2, the stiffness, Kn, can be interpreted as a

penalty function to ensure the contact between soil and anchor in transverse direction.

A relatively high value of Kn =1.0x109 kPa/m was therefore used in the present study.

6.5 Modelling of construction sequence

The modelled construction sequence followed closely that described by Briaud and

Lim (1999) which consisted of the following steps:

1. Specify initial geostatic stress condition, and apply the gravity stresses to the soil.

The Ko condition is assumed in the analysis and the value for Ko is 0.65

according to Briaud and Lim (1999). The mesh at this stage is as shown in Fig.

6.7.

2. Install the solider piles. This involves the removal of soil elements and

replacement of those elements with elements for soldier pile. As shown in Figs.

107

Page 128: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

6.7 and 6.8, the soil elements are removed and replaced by the pile elements. The

elements in green colour represent the installed pile elements.

3. Excavate the first lift to 2.4m below ground level as shown in Fig. 6.9.

4. Install the timber lagging for the first lift and the first row of anchors. As same as

in the case of the soldier pile installation, the soil elements were removed and the

timber lagging were activated in this stage, Figs.6.10 and 6.11.

5. Stress the first row of anchor. In the actual construction, the installation of

anchors is normally carried out in three steps. The soil in the borehole is firstly

removed by drilling and the anchor tendons are then inserted into the borehole.

The borehole, with the anchor inside, is then grouted with high pressure. The

actual construction processes of borehole drilling, insertion of the anchors into

borehole and grouting were not modelled in the numerical studies. Instead, the

installation process was simulated by activating the elements for the unbonded

and bonded lengths, as well as the waler elements. The stressing of the anchor

was simulated by applying equal and opposite forces on the head of the tendon

element and the anchor attachment point of the soldier pile.

6. Lock-off the first row of anchor by inserting the locking element between the top

of the anchor and the anchor attachment point at the soldier pile, as shown in Fig

6.12 and 6.13. The locking elements are anchor elements with very high stiffness

and strength for interface, so as to minimise loss of preload during simulation of

locking-off.

7. Excavate the second lift (2.4m). The 3-D mesh at this stage is shown in Fig. 6.14.

8. Install the timber lagging for the second lift and the second row of anchors as

shown in Fig. 6.15.

108

Page 129: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

9. Stress the second row of anchors. The numerical simulation of pre-stressing is the

same as step 5 above and the mesh is shown in Fig. 6.16.

10. Activate the locking element replicating the effect of the locking-off of the

second row of anchors. The numerical treatment is similar to what was described

in step 6 and the mesh is shown in Fig. 6.16 with details shown in Fig. 6.18.

11. Excavate to the final formation level that is 2.7m below the last row of the

anchors as shown in Fig. 6.19.

12. Install the timber lagging for the final stage of excavation as shown in Fig.

6.20 as well.

6.6 Results and discussions

The main purpose of this case study is to examine the capability of the proposed

element in capturing the salient aspects of anchor-soil interaction which can

significantly affect the performance of the retaining system as a whole. The focus of

the analysis is on the following issues:

a) Calibration of the proposed element model for simulating anchor-soil

interaction.

b) The influence of the interface properties of anchor bonded length.

c) The performance of the 3-D anchor element and its 2-D approximation, the

latter being derived by smearing the anchor stiffness and forces by the anchors’

lateral spacing in the out-of-plane direction.

d) The effect of modelling anchor by different types of elements. In this study, the

types of elements investigated are the proposed element and the more

commonly used friction-free element which is totally decoupled from the soil.

109

Page 130: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

The cases investigated in the present study are listed in Table 6.7. The meshes used in

the present study are based on the configuration of section and file names are preceded

with letter “A”. For example, “A2D1” stands for a 2-D mesh using proposed elements

for anchors and “A2D2” stands for a 2-D mesh using bar elements for anchors. The

same goes for “A3D1” and “A3D2” mesh so on and so forth. As part of the calibration

exercises, analyses using the mesh configuration following Briaud and Lim’s work

were also carried out and the respective filenames are preceded with letter “B” such as

“B3D1” and “B3D2”

110

Page 131: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Table 6. 7 List of Cases investigated in present study

No ID Description

1 A3D1

Anchors were tied to waler beam. Grout modulus is not reduced.

Unbonded and Bonded length are modelled using Anchor-interface

element. (Referred to as the proposed mesh.)

2 A3D2

Anchor were tied to waler beam, Grout modulus is reduced by a

factor of 0.4. Unbonded and Bonded length were modelled using

Anchor-interface element.

3 A3D3 Analysis based on A3D1 with coordinates updating.

4 A3D1Ks

Sensitivity study, Ks =1.0x103 to 1.0x108 kPa/m for anchor bonded

length based on the proposed mesh.

5 A3D1C

Sensitivity study, c’=50 to 100kPa for anchor bonded length based

on the proposed mesh.

6 A2D1

Both bonded & unbonded length modelled using proposed element

in 2-D.

7 A3D2

Both bonded and Unbonded length are modelled using bar element

in 3-D based on the proposed mesh.

8 B3D1

Anchors were tied to soldier pile; Grout modulus was reduced by a

factor of 0.4; Unbonded and bonded lengths were modelled using

bar element.( Referred to as Briaud and Lim’s mesh.)

9 B3D2

Anchors were tied to waler beam; Grout modulus was reduced by a

factor of 0.4; Unbonded and bonded lengths were modelled using

bar element.

10 B3D3

Anchors were tied to waler beam; No reduction for grout modulus;

Unbonded and bonded lengths were modelled using bar element.

111

Page 132: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

6.6.1 Calibration of the FEM models

The calibration of the FEM models involves firstly conducting analyses using the mesh

configuration based on Briaud and Lim’s work (Briaud and Lim, 1999) to determine a

suitable set of soil properties for elastic-perfectly-plastic soil model. The results from

this analysis were then compared with analyses using the proposed mesh configuration.

The effects of variations in the assumptions made when selecting a mesh configuration

were investigated in the section.

Briaud and Lim’s mesh configuration was constructed based on the believing that the

best section (Fig. 6.21) would include one vertical pile at the centre of the section, one

stack of inclined anchors attached to the soldier pile and penetrating back into the soil,

and the soil mass. The width of the mesh was chosen to be equal to the pile spacing or

2.44 m in their study. Special moment restraints were required on the vertical edge

boundaries of the wall to maintain a right angle in plan view between the displaced

wall face and the sides of the simulated wall section; namely, a tensile force Tn and a

moment Mb were induced as shown in Fig.6.22.

In Briaud and Lim’s study, the general purpose code ABAQUS (ABAQUS 1992) was

used for all the numerical analyses. The soldier piles and the tendon bonded length of

the anchors were simulated with beam elements; these are one-dimensional (1-D)

elements that can resist axial loads and bending moments. The stiffness for the pile

elements was the EI and EA values of the soldier piles in their numerical analyses. The

tendon bonded length was treated as a composite steel/grout section to get the EI and

EA stiffness. A reduced grout modulus equal to 40% of the intact grout modulus was

112

Page 133: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

used to account for grout cracking. The steel tendon in the tendon unbonded length of

the anchor was simulated as a spring element which is a bar element that can only

resist axial load. The bar element was given a spring stiffness equal to the initial slope

of the load-displacement curve obtained in the anchor pullout tests. The timber lagging

facing material was simulated with shell elements. The soil was simulated with 3-D

eight-noded brick elements. The soil model was a modified Duncan-Chang hyperbolic

model.

In present study, CRISP—a software package specialized in geotechnical engineering

applications, was used with modifications and additions of the proposed element types

and algorithms. In this calibration exercise, same section of the mesh as used by

Briaud and Lim was adopted. Soil, Soldier pile and timber lagging board were

modelled using 3-D 20-noded isoparamtric elements and anchor bonded length was

modelled using 3-D 3-noded bar elements whereas anchor unbonded length was

modelled using a 3-D 2-noded bar element which is the same as spring elements used

in Briaud and Lim’s study. Fig. 6.23 shows the mesh based on the mesh configuration

used in Briuad and Lim’s study whereas the proposed mesh configuration used present

study were illustrated from Figs. 6.7 to 6.20.

Fig. 6.24 shows the measured as well as computed deflection profiles of the soldier

pile using the parameters set in Table 6.2 from the calibration analysis (B3D1). The

wall deflection predicted by the calibration analysis matches reasonably well with the

Briaud and Lim’s result. This indicates that the elastic-perfectly-plastic soil properties

used in the calibration analysis are reasonable approximations of the mechanical

behaviours of the soil as described by the Duncan-Chang hyperbolic model parameters

given in Briaud and Lim’s study.

113

Page 134: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

The discrepancy between the measured profile and those computed by Briaud and

Lim (1999) as well as the calibration analysis is primarily due to soil properties. The

soil properties in present study were in accordance with those obtained by Briaud and

Lim through a comprehensive back analysis study.

It is well known that the uniqueness of the result of back analysis is not guaranteed.

Sometimes unrealistic material properties may be obtained from back-analysis

exercises. Sufficient justification and verification of back analyzed soil properties with

the help of knowledge of the site soil and laboratory or in-situ test results are crucial to

back analysis results. Due to the lacking of first hand knowledge of the site soil,

further attempt of back analysis for a better matching with measured soldier pile

deflection profile was not conducted in the present study though it is not impossible.

Furthermore, soldier pile was modelled using 1-D beam element in Briaud and Lim’s

study. Hong, et. al., (2003) noted that the transverse dimension has a significant

influence on the computed displacement of the soil around the pile. The zero

transverse dimension 1-D beam element causes singularity in contact with soil at

transverse direction therefore overestimates near field deformation. In contrast, in

present study, 20-noded brick elements were used to model the soldier pile. This may

attribute to the difference between the two calculated soldier pile deflections.

Other simplifications assumptions and omission made in the numerical simulation of

construction processes such as the construction disturbances to the soil around

excavation, workmanship of the anchor installation and grouting and backfilling may

114

Page 135: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

also be sources of discrepancies in the numerical prediction. Therefore, it is often not

possible to replicate all the details of field measurement in numerical analyses (Hsieh

et. al. 2003).

The following sections are to ascertain some of the factors such as detailed modelling

of the anchor-waler-soldier pile system as well as the properties of the models which

may affect the numerical results.

6.6.2 Comparison with published results

As mentioned above, there is quite a significant discrepancy between the calculated

wall deflection profiles from present study and that from Briaud and Lim’s results as

shown in Fig. 6.24. Besides soil properties, the differences in the detailed numerical

modelling of soldier pile-waler-anchor-soil interaction may also attribute to the

numerical results.

According to the elevation view of the site under study, Fig. 6.1, the anchors were

installed in between the soldier piles. In Briaud and Lim’s study, a different section

configuration which is similar to the one-row anchored wall section as shown in Fig.

6.1 was used. To investigate the influence of the mesh configuration to the numerical

results, analysis using the proposed mesh configuration “B3D2” was carried out and

the soldier pile deflection profiles from this analysis is shown in Fig. 6.25. As can be

seen, with the modelling of the anchors being tied to waler instead of directly tied to

soldier pile, significantly larger pile deflection in the upper part of the deflection

profile is observed.

115

Page 136: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Fig. 6.26 shows the effects of different element type in modelling anchors. Analysis

named “B3D2” adopted the mesh configuration of tying the anchor point to waler

beam with the anchor unbonded length and bonded length being modelled using the

conventional bar elements. Analysis named “A3D2” modelled the anchor unbonded

and bonded length using the proposed anchor interface element with the rest of the

mesh details being the same as analysis “B3D2”. As shown in Fig. 6.26, the pile

deflection from analysis using proposed element type for anchor is consistently larger

than the analysis using bar elements. This is attributed to the relative displacement

between anchor and soil within the anchor-soil interface as well as the local yielding of

shear stress within anchor soil interface. The conventional bar element directly ties the

anchor to soil therefore unable to model the soil anchor interface thus results in smaller

soldier pile deflection. This means underestimation of system deformation when soil

anchor interaction is absent in the numerical modelling.

Most importantly for a “Class A” prediction where soil properties were obtained from

other sources such as lab tests and in-situ tests, the analysis using the proposed element

model gives a larger prediction of soldier pile deflection this may result in a safer

design as compared to the analysis using the bar element model. In other words, there

exists a risk of underestimating system deformation when bar elements are used to

model the anchor and hence unsafe design. More detailed investigation of influence of

different element types to numerical result will be discussed in section 6.6.5.

In Briaud and Lim’s study, parameters of the spring elements representing the

unbonded lengths of the anchors were back-deduced from preloading test measurement.

On the other hand, in analysis named “A3D1” ( and the analyses for parametric study),

116

Page 137: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

anchors are modelled using the proposed anchor-interface elements with parameters

which were, as far as is possible, predicted from the site and construction conditions,

with no factoring applied to match measured data. Fig. 6.27 shows the difference in

soldier pile deflection profiles due to the difference in stiffness of anchor unbonded

length and bonded length. Soldier pile deflection profile from proposed analysis

“A3D1” coincidentally matches with that from analysis “B3D2” which uses bar

elements with a reduced stiffness for modelling anchor. In this particular case, it means

the soil anchor interface can be approximated by using 0.4 reduction factor for elastic

modulus of grout concrete but there is little supporting evidence that this magic factor

is going to work for other cases with different ground conditions. On the other hand,

the proposed anchor-interface element uses an explicit approach in modelling this

anchor soil interaction in both the linear elastic domain as well as nonlinear

elastoplastic domain. Although many of the parameters for anchor-soil interface used

in the proposed model, Ks, Kn in particular, are uncertain due to lack of experimental

data, they are still material properties that can be determined experimentally such the

experimental method in work by Yin et. al. ( 1995) , Hu and Pu (2004), unlike the 40%

reduction factor for elastic modulus of grout concrete which is empirical in nature. The

40% reduction factor may work fine for the particular site condition but it does not

guarantee its effectiveness for other sites with different geological formation.

There exists a possibility that the curvature of the anchor may affect the tension and

skin frictions along anchor due to large deformation. The proposed anchor-interface

element is capable of capturing this by using a coordinates updating during analysis.

The result shown in Fig. 6.28 indicates that anchors in this case study remains straight

117

Page 138: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

throughout the course of excavation and therefore no observable difference from

analysis using coordinates updating as compared to the normal small strain analysis.

To address the uncertainty of interface parameters and gain confidence with the results

from the proposed model, sensitivity study of the parameters influencing the calculated

results is presented below.

6.6.3 Sensitivity study-- The influence of the interface properties of anchor

The calculated soldier pile deflection and ground movement are dependent on many

factors such as material properties, mesh refinement and element types simulating the

actual structure and ground conditions as well as the simulation of construction

activities. For the particular case, many factors were studied by Briaud and Lim (1999)

in their parametric studies. As this case study is focused on the performance appraisal

of the proposed element model, only those parameters which were related to the

anchor-soil interface were selected as the subjects of study.

In the present study, the Elastic modulus of soil is assumed to be increasing linearly

with the depth with a gradient of mE. Fig. 6.29 shows the effects of increasing mE.

With increased stiffness at depth, the soldier pile deflections are consistently reduced.

The analyses identified as “A3D1Ks” represent analyses with Ks=1.0x103 kPa/m,

1.0x108 kPa/m and 1.0x105 kPa/m, respectively with other parameters set as same as

“A3D1”. As shown in Fig. 6.30, the soldier pile deflection is not sensitive to the

changes of Ks within the estimated range of 1.0x108 to 1.0x105 kPa/m. This is because

118

Page 139: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

the plastic displacement due to local yielding of interface had been dominating the

deformation of anchors and the soldier pile, and this will be discussed later in section

6.6.4. Numerical instability was experienced for Ks=1.0x103 kPa/m which gives

extremely large displacement of anchor due to such a low value for Ks. No significant

difference in the soldier pile deflection profile when Ks for unbonded length varied

from 10% to 100% of Ks for bonded length as explained in section 6.4 earlier on.

Shear strength of the anchor-soil interface is another factor affecting the numerical

result. The lower bound of cohesion c’=50 kPa, and higher bound of cohesion c’=100

kPa were used to examine this effect. As can be seen from Fig. 6.31, a lower interface

shear strength gives rise to larger deflection of soldier pile on the upper part of the pile.

This indicates that the plastic deformation had occurred within the anchor-soil

interface especially for Row 1 and the evidence of the yielding of shear stress in

anchor-soil interface are provided below in section 6.6.4. Numerical instability was

experienced for c’<50 kPa because of the pull-out failure of the anchor at such low

value of cohesion.

6.6.4 The 3-D effects and 2-D equivalency

Despite the 3-D nature of anchored retaining structures, 2-D analyses are still

commonly used in practice due to the relative ease of 2-D modelling (Parnploy, 1992,

Tang, et. al. 2000, Yoo, 2001, Hsieh et. al. 2003) as well as limitation on computing

power and lack of appropriate 3-D software. In the present study, both 2-D and 3-D

analyses were conducted to ascertain the differences in performances between the

proposed anchor element, when implemented in 2-D and 3-D frameworks.

119

Page 140: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Comparisons between the results from 2-D analyses and 3-D analyses were made to

highlight the significances of the 3-D effects in the modelling of anchored retaining

wall system.

Soldier pile deflection

Results from 2-D and 3-D analyses using the proposed anchor element are presented in

Fig. 6.32. These analyses were based on the same construction sequence and same

material properties for soil and timber lagging. For the 2-D analyses, material

properties for soldier pile, anchor bonded length and anchor unbonded length were

derived by averaging (or “smearing”) the parameter of each pile or anchor over its

lateral spacing. The “smeared” properties are tabulated in Tables 6.2 and 6.4.

As shown in Fig. 6.33, the computed soldier pile deflection from 2-D analyses is

consistently larger than that from 3-D analyses. Furthermore, as shown in Fig. 6.33,

the calculated deflection from 2-D analysis is larger than the 3-D calculated deflection

of timber lagging around anchor points on the anchor plane which is mid-span between

the adjacent piles. It is, however, smaller than the deflection of mid-span of the timber

lagging at depth away from the anchor points. This can be attributed partly to the fact

that the stiffness of the retaining wall and anchor in the 2-D model is a “smeared”

stiffness of the actual soldier pile in the out-of-plane direction. In the 3-D model, the

retaining wall stiffness is non-uniform, being larger at the soldier pile and much

smaller at the timber lagging. The 3-D effect of soldier pile-waler and anchor system is

further complicated by the fact that the anchor points are located at mid-span of the

timber lagging in between the soldier piles. The vertical flexural rigidity of the timber

120

Page 141: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

lagging is too small to distribute the anchor load over the wall face. Thus, the anchor

force has to be transmitted firstly to the waler, on which the anchor blocks are mounted,

and then to the soldier pile. Flexural action along the soldier pile eventually distributes

the load in the vertical direction. Finally, lateral flexural action along the timber

lagging distributes the load in the horizontal direction over the timber lagging.

Fig. 6.33 appears to give an impression that the 2-D analysis gives a conservative

estimation of soldier pile deflection. Further investigation into the soldier pile

deflections at different construction stages showed that this is not always true. Fig.

6.34 compares the soldier pile deflection from 2-D analysis and that from 3-D analysis

at the end of 1st lift of excavation. At this construction stage, the retaining structure

was working under cantilever mode. The soldier pile deflection predicted by 2-D

analysis is larger at the top and slightly smaller at the depth 2m below ground surface.

The comparison of 2-D and 3-D deflection profiles indicated a more flexible cantilever

when soldier pile is modelled using 2-D analysis and a relatively stiffer cantilever for

3-D model. This is again attributable to the differential movement between the

mid-span and soldier pile which is captured by the 3-D but not the 2-D analysis.

Fig. 6.35 shows the comparison of soldier pile deflection from 2-D and that from 3-D

analysis after the pre-stressing of Row 2 anchor. At this construction stage, the reaction

force applied to anchor point during the pre-stressing of Row 2 anchor had

significantly pushed the soldier pile back into soil side. As can be seen, the calculated

soldier pile deflection around anchor point from 2-D analysis is significantly smaller

than that from 3-D analysis. This is attributable to the fact that the anchor load is not

applied directly at the soldier pile location; this being captured by the 3-D but not the

121

Page 142: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

2-D analysis. The difference is also enhanced by the fact that the stiffness of the soldier

pile using in the 2-D analysis is an averaging, and thus lower, stiffness, whereas that

used in the 3-D analysis is a localised soldier pile stiffness, which is much higher.

Ground Displacement Field

The general trends of ground movement from 2-D and 3-D analyses shared some

similarities, but the displacement around the anchor in the near field is quite different,

as can be seen from Figures 6.36 and 6.37, as well as Figures 6.38 and 6.39. These

figures show that the 3-D analysis captured more localized displacement around

anchor than the 2-D analysis. This is because of the difference in the ways the anchor

soil interaction is modelled in 2-D and 3-D models. In 3-D model, the load transmitted

from anchor to soil is represented with a line load along anchor whereas in 2-D model,

this is uniformly distributed on a sheet across the in-plane direction. There is therefore

a much higher concentration of load around the anchor in the 3-D analysis than the

2-D analysis. Therefore 3-D model gives larger soil displacement around anchor and

faster attenuation of the soil displacement. Comparison of the displacement fields and

deformed meshes at the end of excavation between the 2-D analysis and 3-D analysis

shows the same findings, as shown in Figures 6.40, 6.41, 6.42 and 6.43.

Stresses in soil

There are also evidences of significant anchor-ground interaction, which was modelled

differently by the 2-D and 3-D analyses. Figures 6.44 to 6.48, show principal

stress trajectories on certain horizontal planes at different level (represented by

122

Page 143: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

y-coordinate, yo) from 3-D analysis, and arching effect in the in-plane direction can be

discerned from these figures.

The arching effect in soil behind timber lagging is an important load carrying

mechanism in a soldier piles with timber lagging type of retaining structures (Vermeer

et. al. 2001, Hong et al. 2003). This is reflected in Fig. 6.44, which shows the principal

stress vectors on a horizontal plane 1.5m below ground level at the time when the

excavation was completed. As shown in the figure, the directions of principal stresses

within 3m distance from timber lagging were distorted due to the arching effect in soil.

Therefore the arching effect was estimated to occur within this 3m zone of soil. Fig.

6.45 shows the principal stresses on the plane which 3.0m below ground level and

1.2m below the anchor head of Row 1 anchor. As shown in this figure, that the stress

field is not significantly disturbed by the presence of anchor. This indicated that the

unbonded length of anchor does not have a significant influence to the stress field in

soil around anchor. On the other hand, as shown in Fig. 6.46 which is 4.5m below

ground level and 0.3m above the anchor head of Row 2 anchor, the presence of the

anchor bonded length of Row 1 anchor in the picture induces significant disturbance to

the stress field in the soil around the anchor as marked in the figures. This indicates a

pronounced load redistribution effect at this level in the soil around anchor bonded

length in the in-plane direction. This also an indication of the group effect of the

influence between the neighbouring anchors. Similarly, as can be seen from Fig. 6.47,

which is 6m below ground level, the unbonded length of Row 2 anchor in the picture

induces little disturbance to the stresses in the surrounding soil. Fig. 6.48, shows the

principal stress trajectories at a horizontal plane 7.6m below ground surface; the

bonded lengths of both Row1 and Row 2 anchors pass through this plane. Disturbance

123

Page 144: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

to the stress field is evident at 3 locations. The first location lies just behind the soldier

pile whilst the other two are around the bonded lengths of the two rows of anchors.

This underscores the complexity of the stress field arising from soldier pile-anchor-soil

interaction.

Skin friction along anchor length

Fig. 6.49 compares the skin friction along anchor computed by the 2-D and 3-D

analyses just after pre-stressing of Row 1 anchor. As can be seen, the skin friction

along the anchor length from 2-D analysis is consistently and significantly lower than

that from 3-D analysis. This underscores a deficiency of the 2-D equivalent stiffness

method of smearing what is essentially a 3-D problem; it may give nearly correct

displacement but correctness of stresses is not ensured. Where nonlinear deformation

takes place in the materials involved, even the displacement cannot be correctly

evaluated using the equivalent stiffness approach as the nonlinear behaviour of the

material is dependent upon the stresses in the material. This also explains the sharper

peak curve of skin friction distribution near anchor front in 2-D analysis which is

caused by the smaller extent of yielding arising from the lower interface shear stress.

As the 3-D analysis shows, the higher interface shear stresses cause the yielding at

anchor front, thereby resulting in stress re-distribution along the bonded length.

Fig. 6.50 shows the distribution of interface shear stress along Row 1 anchor at

different construction stages. As shown in the figure, the excavation of the 2nd lift

caused increase of skin shear along the bonded length of Row 1. Stressing of Row 2

anchor led to a small amount of unloading of the interface shear stress along Row 1

124

Page 145: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

bonded length. On the other hand, excavation of the final lift caused significant

decrease in interface shear stress near the front of the bonded length and an increase of

skin shear at the rear of the bonded length. This is caused by the movement of the soil

towards the excavation. As the depth of excavation increases, so the extent of the zone

of soil movement in the retained soil also extend rearward into the retained ground

away from the excavation. Towards the final lift of the excavation, the zone of soil

computed by the 3-D analysis reaches the front part of the bonded length. This inward

movement of soil at the vicinity of anchor front caused some reduction in the relative

displacement between anchor and the surrounding soil; thus causing unloading in this

part of the anchor interface. If this drop in skin shear is too much, it means that the

unbonded length of the anchor is not long enough. This is a useful point for checking

the design of anchored wall with FE analysis.

The respective figure for 2-D analysis, as shown in Fig. 6.51, registered similar

patterns of skin shear distribution at a few key construction stages but with the two

major discrepancies. The first one is the 2-D analysis showed that stressing of Row 2

anchor had a stronger influence to the unloading of the Row 1 anchor. The second one

is the skin shear distribution along anchor at the end of the excavation from 2-D

analysis showed larger reduction in skin friction around anchor front. This is

attributable to the fact that, in 2-D, the “smeared” anchor behaves effectively like a

membrane which has a larger range of influence. This may affect the decision on the

anchor unbonded length if used in design and may result in a more costly design. In

general, 2-D analysis will give smaller skin shear which if used for design may lead to

unsafe design or significant consumption of safety allowance.

125

Page 146: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

6.6.5 Influence of different models for anchor unbonded length

The performance of the proposed element is further investigated by comparing it with

other type elements in this case study. Analysis “A3D1” is a 3-D analysis using the

proposed anchor element whereas analysis “B3D3” is a 3-D analysis using bar

elements for the anchor. In “B3D3”, the element is assumed to be perfectly bonded to

the soil along the fixed length but shares no common node with the soil domain along

the unbonded length except for the anchor head and the rear end of the unbonded

length, where the bonded length commences. In other words, the bonded length is

completely coupled to the soil whereas the unbonded length is completely uncoupled

from the soil. Apart from the modelling of the anchor, all other aspects of the two

analyses are identical.

As shown in Fig. 6.52, the wall deflection from the analysis using bar element model is

significantly smaller than that from analysis using the proposed element. Further

details with respective to the differences in computed soldier pile deflection profiles

for a few key construction stages are as shown in Figs 6.53 to 6.56.

Fig. 6.53 shows the soldier pile deflection of analyses “A3D1” and “B3D3” after the

excavation of 1st lift. At this stage, the anchor has not been installed. Fig. 6.54 shows

the soldier pile deflection after stressing of the Row 1 anchor. As can be seen from

these figures, the deflection profiles from two analyses are virtually identical, thereby

indicating consistency of the two analyses up to this point of time. The two deflection

profiles started to drift away from each other after the locking of Row1 anchor. This

126

Page 147: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

clearly indicates that observed differences between the two deflection profiles were

due to the difference in element models modelling the anchor. Calculated soldier pile

deflection using the proposed element was consistently larger than that using bar

element. This difference persists as the simulated construction steps proceeded as

shown in Fig. 6.55 and 6.56 which are the deflection profiles at the end of excavation

of 2nd lift and stressing of Row 2 anchor respectively. This is attributed to the slippage

allowed at the anchor-soil interface in the proposed element, which cannot be

modelled by a perfectly bond bar element.

The displacement vector fields for the analysis using bar element is shown in Fig. 6.57.

Except for the difference in magnitude of displacement in soil, similar patterns of

displacement in soil are observed in analysis using bar element compared to that using

proposed element as shown in Fig. 5.42. Along the unbonded length, comparison of

Figs. 6.42 and 6.57 shows that there are slightly but discernible differences in the

displacement vector field of the soil. In Fig. 6.42, regions immediately adjacent to both

rows of anchors show a clear trend of upward displacement, caused by the pull of the

anchors on the surrounding soil. This is much less in evidence in Fig. 6.57. The

upward pull of the anchors in Fig. 6.42 also helps to mitigate the near surface

settlement. As a result, the surface and near-surface settlement of ground is smaller in

Fig. 6.42, compared to Fig. 6.57. The difference in surface settlement is also evident

from the comparison of the settlement profiles in Fig. 6.58. The significant differences

in the ground settlement curves within short distance behind the wall are attributed to

the difference in modelling shear along the unbonded length. Thus the interaction

between anchor and ground is not merely limited to their bonded length. The pull of

the unbonded length also has an influence on the ground deformation which can be

127

Page 148: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

manifested in the surface settlement profiles. Ignoring the interaction between anchor

and soil of the unbonded length in the case leads to larger computed ground settlement

in near field.

Principal stresses on a few selected horizontal planes from analysis using bar elements

are shown in Figs 6.59 to 63. As far as the locations of soil arching is concerned, no

significant discrepancy was found in the results from analysis using bar element

compared with those from analysis using proposed elements, as shown in Figs 6.44 to

6.48. This is because the only difference in the two analyses is the way the anchor

interface is modelled. The analysis using bar element, which is the conventional way

the anchors are modelled in numerical analysis, generally ignores the anchor-soil

interface. The analysis using the proposed element is superior to the conventional way

in the sense that it allows slippage to occur at the anchor-soil interface and allows for

plastic deformation to develop when yielding takes place within the bonded length.

6.7 Summary of the chapter

An excavation supported with soldier pile, timber lagging and tieback, was studied in

this chapter to investigate the performance of the proposed element type for

soil-anchor interaction. Reasonable matching with the published data for the analyses

using the proposed element type was achieved with similar material properties as those

in the published literature. It thus can be concluded that the proposed element

performed quite well in modelling soil-anchor interaction for this case study.

Differences between 3-D and 2-D analyses are also highlighted. Finally, the

differences between the proposed element for modelling anchors and another method

128

Page 149: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

using simple bar elements was explored. Differences were shown to be present not just

in the manner in which the bonded length interacts with the surrounding soil but,

perhaps more subtly, also in the manner in which the unbonded length interacts with

the surrounding ground. More importantly, for a “Class A” prediction during design

phase the analysis using the proposed element model gives a larger prediction of

soldier pile deflection this may render a safe design as compared to the analysis using

the bar element model. In other words, there exists a risk of underestimating system

deformation when bar elements are used to model the anchor and thus unsafe design.

129

Page 150: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

CHAPTER 7

CASE STUDY 3 – ANCHORED EXCAVATION IN SOFT CLAY

7.1 Introduction

This chapter focuses on the application of the proposed element model in analyses of

an excavation in soft ground. The excavation was supported by TradeArbed HZ775Z

heavy metal sheet-pile, internal struts and ground anchors. 2D analyses of the

excavation using conventional model for ground anchors had been conducted and

reported by Lee, et al. (1989), Yong, et al. (1990), and Parnploy, (1990). In the present

study, a series of 3D FE analyses are conducted to investigate the significance of some

issues affecting the numerical results such as the influence of detailed modeling of

ground terrain, unbalanced excavation and consolidation of clay layers as well as the

influence of constitutive models of soil and the material properties. The computed

sheet-pile wall deflection is compared with the field measurement to examine the

ability of the proposed anchor-interface element in modeling an excavation in soft

marine clay supported with strutted and anchored sheet-pile wall system.

7.2 Site conditions and site geology

As shown in Fig. 7.1, the selected site is a tunnel section of Central Expressway which

crosses under the Singapore river (Parnploy, 1990). The tunnel was constructed by

cut-and-cover method. To divert the river flow, the construction work was executed in

two stages. As shown in Fig. 7.1, the first stage involves the construction of a

strutted and anchored cofferdam of sheet-piles on the South bank of the river. The

130

Page 151: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

subject of this numerical study is the anchored section of the cofferdam. The layout of

the site plan of the 56m wide and 18m deep excavation is shown in Fig. 7.2. The

locations of five inclinometers, IM1 to IM5 are also shown in Fig. 7.2.

The site of the excavation overlies Kallang formation. This formation is typified by a

thin layer of fill or river deposit underlain by a thick layer of soft marine clay,

followed by thin layers of silty sand and sandy silt before hitting stiffer soil of the

Jurong formation (e.g. Dames & Moore 1983, Tan 1983). The latter consists of

sedimentary rocks which have been subjected to various degrees of weathering. The

grade with the highest degree of weathering is designated S4 and generally has SPT

ranging from 30 to 100. This is usually underlain by less weathered stiff to hard soil

which falls under the classification of S2 and often has SPT exceeding 100. In some

locations, a layer of bouldery clay, which consists of large boulders interspersed in a

stiff clay matrix, may be present between the S4 and S2 strata. This layer is classified

as an S3 material. Below the S2 soil lies the S1 material, which consists of largely

unweathered sedimentary rocks.

The supporting system of the excavation consists of a TradeArbed HZ775Z sheet-pile

wall driven into the firm bed rock by toe-pins, internal struts of steel pipes and

H-beams with ground anchors.

The section analyzed in the present study is designated X1 in Fig. 7.2. The subsoil

condition for sections X1 and X2 is shown in Fig. 7.3. The supporting system for these

sections comprises sheet-pile wall braced with 3 levels of internal struts at upper part

131

Page 152: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

of the excavation and 3 levels of ground anchor at lower part of the excavation. The

corresponding inclinometers near the two sections are IM5 and IM1 respectively.

At section X1, the soil profile below the river bed consists of a layer of fill material

followed by a thick layer of marine clay (M), a layer of medium stiff fluvial soil (F2)

and weathered sedimentary rock (S4 and S2) with decreasing degrees of weathering

with depth. Typical properties of the soil are summarized in Table 7.1.

Table 7.1 Typical subsoil properties (After Parnploy, 1990)

γ w LL PL SPT cu Soil Type kN/m3 % % % N kN/m2

Fill 18 - - - 1~4 - M 16 70~75 90 60 1~2 22~35 F2 18 - - - 2~10 28~40 S4 20 - - - 30~100 150* S2 20 - - - >100 500* * Estimated from pressuremeter tests.

The sheet-pile wall consists of a combination of double I-sections interlocked with

intermediate Z-Section to form a heavy composite section designated as HZ 775A. The

HZ 775A sections were driven into the weathered sedimentary rock to a depth varying

from 20 to 22 m depending on the degree of weathering. These depth roughly indicates

the interface of S4 layer which has a SPT value of between 30 to 100, and the S2 layer

which normally has an SPT value greater than 100. To enhance the stability of the

sheet-pile against the kick-out, toe-pins were installed at 2.065m intervals. Each toe

pin consists of a 5m steel H-pile HD260 (260m x 260m x 219kg/m) and inserted into a

prebored hole through the double I-section of HZ775A. The upper 2m of the steel

H-pile was then cement-grouted to integrate it with the TradeArbed HZ 775A leaving

the remaining 3-m socketed into the S2 layer.

132

Page 153: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

The first level of strut consists of double H-beam (2W 610 x 180 x 92 kg/m, Grade 50,

designated as 2W24x92 kg/m) braced together by short L 75 x 75 x 9mm links, laced

at 45o. At the 2nd and 3rd levels, the double H-beams (2W 610 X 325 X 172 kg/m,

Grade 50, designated as 2W24 X 174 kg/m) were used because of the expected higher

loads at the lower levels, and the two H-beams are braced together by 45o lacing. The

king posts (H 400 X 400 X 172 kg/m) with spacing of 4.12m centre to centre were

designed to support the struts and to provide the restraining force of strut in the

direction of buckling. Bracing in the lower part of the excavation consists of 3 levels of

ground anchors, A4, A5 and A6, inclined at an angle of 30o to the horizontal as shown

in Fig. 7.3. The 200mm-diameter ground anchors have lengths ranging from 20m to

30m, with bonded lengths of between 11m to 15m in the S2 material. The ground

anchors in all the 3 levels were installed at a horizontal spacing 2.065m. The design

properties of struts and ground anchors are summarized in Table 7.2 and 7.3

respectively.

The construction sequence follows that reported in Parnploy (1990), which was

regenerated from the recorded dates of strut installation and ground anchor preloading

with an assumed excavation rate of 2-3m depth per week in between. The simulated

construction sequence is summarized in Table 7.4. Before the commencement of the

excavation work, water inside the cofferdam was pumped out over a period of one

week. An unbalanced excavation scheme was adopted for the top 3 lifts of excavation

due to constraint from other construction activities. This involved excavating the soil

from one end of the excavation before the other end is excavated.

133

Page 154: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Table 7.2 Properties of Internal struts (After Parnploy, 1990)

Strut Level Section/Size Equivalent Stiffness Preload Remarks

S1 0 2W 610 x 180 x 92 kg/m 43080 kN/m/m 2W24-92 kg/m S2 -5 2W 610 X 325 X 172 kg/m 80000 kN/m/m 245.7 kN/m 2W24-174 kg/mS3 -8 2W 610 X 325 X 172 kg/m 80000 kN/m/m 245.7 kN/m 2W24-174 kg/m

Table 7.3 Properties of ground anchors (After Parnploy, 1990)

Nearest IM Stiffness Working Load Preload No. (kN/m) (kN) (kN)

A4 IM5 13250 1250 1200A5 IM1 14000 1500 1200A6 IM1 &IM5 14000 1500 1200

134

Page 155: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Table 7.4 Construction sequences used in the numerical simulation

No. Time StartsTime EndsLevel Level

X1 X2 Start IncEnd Inc Time Activity

1 0 10 0 1 3 10 Installation of sheet-pile wall and S1

2 10 17 -3 -6 4 10 7 Excavation of lift 1 3 17 77 -3 -6 11 15 60 Lull 4 77 78 Nil -5 16 16 1 Strut S2 Preload 5 78 79 Nil -5 17 17 1 Strut S2 Install 6 79 86 -9 -11 18 25 7 Excavation of lift 2 7 86 116 -9 -11 26 30 30 Lull 8 116 117 -8 -8 31 31 1 Struts preloading S3 9 118 119 -8 -8 32 32 1 Struts installation S3 10 119 122 -12 -13 33 40 3 Excavation of lift 3 11 122 132 -12 -13 41 45 10 Lull 12 132 133 -11 -11 46 46 1 Installation of Anchor A413 133 134 -11 -11 47 47 1 Preloading of Anchor A4 14 134 135 -11 -11 48 48 1 Lock-in Anchor A4 15 135 141 -15 -15 49 55 7 Excavation of lift 4 16 141 162 -15 -15 56 60 21 Lull 17 162 163 -14 -14 61 61 1 Installation of Anchor A518 163 164 -14 -14 62 62 1 Preloading of Anchor A5 19 164 165 -14 -14 63 63 1 Lock-in Anchor A5 20 165 171 -18 -18 63 70 7 Excavation of lift 5 21 171 204 -18 -18 71 75 33 Lull 22 204 205 -18 -18 76 76 1 Installation of Anchor A623 205 206 -18 -18 77 77 1 Preloading of Anchor A6 24 206 207 -18 -18 78 78 1 Lock-in Anchor A6 25 207 213 -20 -20 79 85 7 Excavation of lift 6 26 213 242 -20 -20 86 90 29 Lull Note : Unit for time in the table is Day.

135

Page 156: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

7.3 FE mesh

The section of the excavation in section X1 was adopted for the current numerical

study. A mesh of the entire excavation, together with the retained ground domain on

both sides, was used since one side of the excavation abuts against the river bank

whereas the other side is in the river itself, thereby creating an unsymmetrical

condition. Furthermore, unbalanced load was expected due to the unbalanced

excavation in the first 3 lifts of excavation.

To examine the influence of the stand-off distance, 2D meshes with different widths

were used for comparison. The mesh identified “AX1CA” as shown in Fig. 7.4

followed the stand-off distance used by Parnploy (1990). The total width of this mesh

is 158m wide. The other mesh, identified “AX1CB” as shown in Fig. 7.5, extends the

stand-off distance to 90m from the centerline of the excavation, which is roughly 3

times the width of excavation. For both meshes, the vertical height of the mesh

terminates at the rock head. The two meshes have the same number of elements. Each

mesh consists of 1318 linear strain quadrilateral elements, 99 linear strain triangle

elements, 3 bar elements for modeling struts and 66 numbers of anchor-interface

elements for modeling anchors.

The calculated sheet-pile wall deflection profiles at the final dredged level from

analyses using the two meshes are compared in Fig. 7.6. No significant difference

between the two results was observed. The error in maximum deflection is less than

1%. This indicates that the stand-off distances for the two meshes are sufficient.

136

Page 157: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

The mesh independency of the problem was also studied by comparing the results

from two meshes with different element size. Mesh “AX1CC” consists of 817 linear

strain quadrilateral elements, 40 linear strain triangle elements and 31 anchor-interface

elements. The typical element dimension for this mesh is about 2m within excavation

and the near field of excavation as shown in Fig. 7.7. Mesh “AX1CD” consists of

1885 linear strain quadrilateral elements, 68 linear strain triangle elements and 58

anchor-interface elements. The typical element dimension for this mesh is about 1m

within the excavation and the near field of excavation as shown in Fig. 7.8. The wall

deflection profiles from analyses using the two meshes are as shown in Fig. 7.9. As

can be seen, the relative error between the maximum deflections is within 4%,

therefore the size of elements for both mesh is considered accurate enough for the

particular problem. To strike a balance between accuracy and computation resources

needed, the 3-D mesh is generated by extruding the 2-D mesh AX1CB, which has

element sizes that are intermediate between those of meshes AX1CC and AX1CD.

All analyses in this case study were conducted in double precision.

Fig. 7.10 shows the 3-D mesh used in the present study. Since the horizontal spacing

of the struts and anchors are 4m and 2m respectively, symmetry consideration in the

out-of-plane direction allows a slice of 2m to represent the repeated pattern of the

sheet-pile, struts and anchor system. The 3-D mesh consists of 1322 linear strain

brick elements, 84 linear strain wedge elements and 66 3-D anchor-interface elements.

The 3-D mesh after the first 2 lifts of excavation and installation of strut 2 is shown in

Fig. 7.11. The two rows of anchors in out-of-plane direction can be seen in Fig. 7.12

which is a close-up view of the mesh at the end of the final excavation. All the vertical

faces in the 3-D mesh are not allowed to displace out of their respective planes. The

137

Page 158: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

bottom plane is completely fixed to simulate perfect coupling to the hard rock stratum.

Detailed fixity boundary conditions are as shown in Fig. 7.13.

The analyses which were conducted were fully-coupled Biot consolidation analyses,

which allow dissipation of excess pore pressure to be simulated concurrently with

construction processes. Fluid flow boundary conditions for consolidation analyses

were prescribed as zero excess pore water pressure on ground surface, excavation

surface as well as the two vertical surfaces at far field of the mesh as indicated in

Figures 7.14 to 7.17 for a few key construction stages.

7.4 Material properties for the FE analyses

The site geology belongs basically to a marine clay formation. The soil properties of

soils found in this formation have been documented by Dames and Moore (1983), Tan

(1983), Lee et. al. (1989), Yong et. al. (1990) and Parnploy (1990), amongst others.

Soil properties abstracted from studies conducted by Parnploy are tabulated in Table

7.5. In present study, similar soil properties as compared to the soil properties used by

Parnploy were used for the top 4 layers, i.e., top fill (F), marine clay (M), medium stiff

fluvial soil (F2) and weathered sedimentary rock (S4). As tabulated in Table 7.6, a

relative stiffer elastic modulus of 200 MPa for lightly weathered rock (S2) was used in

the present study. This is based mainly on the recommendation by Dames and Moore

(1983). The values for the elastic moduli of the site soil used in present study were

obtained by rounding Parnploy’s (1990) values. Comparison of Parnploy’s (1990)

with Dames and Moore’s (1983) shows that, apart the modulus of the S2 material, the

138

Page 159: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

other parameters are generally consistent. They were therefore adopted in this study

without any change.

The sheet-pile wall was modeled using solid elements with equivalent flexural

stiffness. This is unlikely to be strictly correct. A sheet pile is designed specifically for

flexure in the vertical plane. Owing to the presence of vertical joints and hinges, it is

unlikely to be equally stiff in the lateral, out-of-plane direction. Load distribution in

the lateral, out-of-plane direction is usually achieved using walers, which are not

explicitly modeled in this study. Given this, it is unlikely that the lateral flexural

stiffness will be the same as the vertical flexural stiffness. Thus, the use of isotropic

material properties for the wall is essentially an approximation. The equivalent

properties of sheet-pile elements are tabulated in Table 7.6 together with the properties

for rip-rap wall for river bank.

Struts are modeled using 2-noded bar elements. The material properties for struts are

derived from the section properties and tabulated in Table 7.7. As a slice of 2m thick is

used in the 3D mesh which is ½ of the spacing of the struts in in-plane direction, half

of the strut section values and preload were input to data file for analyses.

Table 7.5 Soil properties from site investigation report(After Parnploy,1990)

Level E ν c’ φ’ γ kx ky

Soil type (m) (kN/m2) (kN/m2) (Degree) (kN/m3) (m/sec) (m/sec)

Fill 0 5833 0.25 0 30 18 1 x 10-7 1 x 10-7

M -5 4333 0.3 0 22 16 6 x 10-8 2 x 10-8

F2 -18 5833 0.25 0 30 18 1 x 10-7 1 x 10-7

S4 -20 90000 0.25 30 30 20 2 x 10-8 2 x 10-8

S2 -22 150000 0.25 100 40 20 1 x 10-8 1 x 10-8

139

Page 160: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Table 7.6 Soil properties used in present study

Level E ν c’ φ’ γ kx ky

Soil type (m) (kN/m2) (kN/m2) (Degree) (kN/m3) (m/sec) (m/sec)

Fill 0 6000 0.25 0 30 18 1 x 10-7 1 x 10-7

M -5 4500 0.3 0 22 16 6 x 10-8 2 x 10-8

F2 -18 6000 0.25 0 30 18 1 x 10-7 1 x 10-7

S4 -20 90000 0.25 30 30 20 2 x 10-8 2 x 10-8

S2 -22 200000 0.25 100 40 20 1 x 10-8 1 x 10-8

Rip-rap 2.5 x 107 0.25 - - 24 2 x 10-6 2 x 10-6

Sheet-pile 1.18 x 108 0.25 - - 24 2 x 10-6 2 x 10-6

Ground anchors were modeled using the proposed anchor-interface element. The

material properties for ground anchor were derived from the design of the 7T32 anchor,

which consisted of 7nos. of 32-mm diameter tendons bundled into one 200-mm

diameter borehole. As the anchors were installed in fresh rock stratum, the interface

properties for anchor bonded length were assumed to be similar to the parameters for

the case study of the pull-out test discussed in section 5.4.2.3 of Chapter 5. The anchor

unbonded length in this case passed through soft marine clay. For this reason, lower

values were adopted for the interface properties of unbonded length as listed in Table

7.8. In reality, the bore hole of the anchor unbonded length is normally filled with

bentonite or spoils of the boring and there would be certain amount of friction along

the anchor unbonded length. This friction was modeled in the proposed element model

with a nominal value of 2 kPa for c’ and 10o for φ’ assuming the Mohr-coulomb type of

friction between soil and the anchor. Parameters and material properties pertaining to

the anchor interface elements were tabulated in Table 7.8. The horizontal spacing in in

the out-of-plane direction between adjacent anchors is 2m and two columns of anchors

in out-of-plane direction were modelled in the mesh. Based on symmetry consideration,

140

Page 161: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

each of the simulated anchors were assigned half of the section properties for each of

the actual anchor.

Table 7.7 Strut properties used in present study

2D (per meter run) 3D (½ strut in the 3D mesh)

Struts EA (kN/m) Pre-load (kN/m) EA (kN) Pre-load (kN)

S1 5.8 x 105 0 1.16 x 106 0

S2 1.08 x 106 245.7 2.16 x 106 491.5

S3 1.08 x 106 245.7 2.16 x 106 491.5

Table 7.8 Anchor/interface properties used in present study

Anchor

Length

A

(m2)

Perimeter

(m)

Ks

(kN/m)

Kn

(kN/m)

c’

(kN)

φ’

(ο)

I

(m4)

unbounded 0.0056 0.7 1.0x103 1.0x107 2 10 0.5x10-7

bounded 0.0314 0.628 1.0x105 1.0x107 300 30 0.8x10-4

Following Parnploy (1990), Ko is assumed to be 0.65 for marine clay, Ko=0.77 for the

F2 layer and Ko=0.8 for rest of the layers. These values are also consistent with the

recommendations in the report of Dames and Moore (1983). The profile of the

in-situ stresses is shown in Fig. 7.18.

7.5 Simulation of construction sequences and activities

The construction sequences were simulated following Table 7.4 and Fig. 7.19. The

simulated construction sequences are as follows:

(1) To equilibrate the in-situ stress field under the influence of river bank on the

left bank, the primary mesh was assumed to be as shown in Fig. 7.14 with

141

Page 162: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

water pressure being applied on the top surface. An analysis stage simulating

the overburden of the river bank was then introduced in the analyses. The mesh

after this introduction of river bank is shown in Fig. 7.20. A period of 50 years

was simulated in the consolidation analysis for this stage to allow for

consolidation in the marine clay layer due to the overburden to take place.

(2) The installation of sheet-pile wall and struts was simulated following the

conventional method by removal of soil elements and insertion of sheet-pile

wall elements and strut element. The pre-loading of struts was simulated with

point load. The mesh at this construction stage is as shown in Fig. 7.20.

(3) Dewatering of the cofferdam was simulated by applying the negated water

pressure within excavation area and the face of sheet-pile wall elements.

(4) Excavation was carried out in lifts according to the construction sequence as

listed in Table 7.4. Fluid flow boundary conditions within the excavation area

were changed accordingly in FEM mesh.

(5) Installation of ground anchor was simulated in the same way as that described

in Chapter 6. The anchor-interface elements were first introduced into the mesh,

the pre-stressing of anchor was then simulated with equal and opposite point

loads applied to the tip of anchor and the anchor attachment node on the

sheet-pile wall. The locking element was then introduced into mesh to tie the

anchor with the sheet-pile wall. The 3D mesh after installation of all anchors is

shown in Fig. 7.21.

One of the conventional methods for simulating ground anchor is to model the

anchor as a spring element (e.g. Briaud & Lim 1999). The 2-D analyses of the

anchored excavation using the above mentioned conventional method as reported

142

Page 163: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

by Parnploy (1990). This method was also used in the comparative study to

highlight the significance of modeling anchor with interface by using the proposed

element.

7.6 Calibration of FEM model

7.6.1 Deformation

A consolidation analysis with 3D mesh was conducted to calibrate the numerical

model used for this case study. Computed deflection profiles of sheet-pile wall at a few

construction stages were compared against the measured wall deflection from

inclinometer IM5 as shown in Fig. 7.22. As the inclinometer was installed only after

the installation of sheet-pile and before the pumping of water from within cofferdam,

the ground movement caused by the overburden of the river bank was not included in

the calculated displacement of sheet-pile wall. As Fig. 7.22 shows, the computed wall

deflection is generally in the same order of magnitude as the measured deflection

although some differences do exist. Given that no adjustment to the soil parameters

have been made, this level of agreement is not entirely unreasonable.

Typical displacement fields at different construction stages are shown in Fig.7.23 to 26.

As these figures show, there is clear evidence of a sway towards the left side of the

excavation due to the unsymmetric ground terrain and unbalanced excavation (Fig.

7.23 and 7.26). Significant basal heave is also computed especially near the wall. In

addition, horizontal movement of the basal soil was also computed near the center of

the excavation area where the two excavation levels meet, as shown in Figures 7.24

and 7.25.

143

Page 164: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

7.6.2 Excess pore water pressure

Figures 7.27 to 7.32 show contours of the excess pore water pressure distribution at

key construction stages. As can be seen, application of overburden of the river bank at

right bank led to high excess pore water pressure in marine clay layer beneath the right

side of river bank whereas the marine clay beneath the left side of river experienced

near zero excess pore water pressure change. A period of 50 years was simulated in the

analysis to allow for the consolidation of marine clay beneath right side river bank to

take place.

As expected, negative excess pore water pressure was generated within the excavation

area after dewatering of the excavation area as shown in Fig.7.28. As the depth of

excavation deepened and the marine clay layer was penetrated through, the negative

excess pore water pressure beneath the dredged line reduces in magnitude as shown in

Figs. 7.30 to 7.32 because of the higher permeability and stiffness of the soil at the

excavation base. Only small amount of negative excess pore water pressure exists

behind the wall in marine clay stratum which was generated as a result of the inwards

movement of the sheet-pile wall as shown in Fig. 7.31.

The analysis stopped at about 10 months since the commencement of the construction

work and excess pore water pressure within marine clay was almost totally dissipated

by the end of the analysis as shown in Fig. 7.32. Only small amount of excess pore

water pressure existed beneath excavation base but limited within rock stratum. This

144

Page 165: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

indicated the necessity of a consolidation analysis as an undrained assumption is

invalid in this case.

7.6.3 Ground anchors

The axial displacement of anchor head is as shown in Fig. 7.33. The curve labeled L1

represents Row 1 anchor installed at left bank of the excavation whereas R1 represents

Row 1 anchor installed at right bank of the excavation. The initial steep slope of the

curve is the result of prestress loading. The axial displacement of the R1 anchor starts

to pick up as the excavation near right bank deepened. As the excavation level in the

left back reached that of right bank from excavation of 4-th lift onwards, the axial

displacement of anchor heads of L1 caught up with and surpassed that of R1 as shown

in Fig. 7.33 due to a thicker lift of excavation at left panel. The picking up and

sudden drop in the later part of the curve are due to the excavation of following lifts

and the prestressing of next two rows of anchors. The back column of anchors

experienced almost the same trends of axial displacement as that in Fig. 7.34.

Not surprisingly, the Row 1 anchors installed at left bank had larger ultimate axial

displacement than those at right bank as shown in Fig. 7.33. This can be attributed to

the unbalanced excavation scheme. In the excavation of top 3 lifts, the depth of left

panel excavation was always shallower than that of right panel. This means less earth

pressure behind the sheet-pile wall at left bank would have been generated comparing

to the earth pressure behind the right bank wall. The strong presence of internal struts

had made the supporting system worked together as a frame. The relatively higher

145

Page 166: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

earth pressure behind right bank wall push the frame of supporting system to left bank

and thus restrained the inward movement of left bank wall. As the Row 1 anchors were

installed and followed by the excavation of 4-th lift, a thick layer of soil was removed

in the excavation of left panel at 4-th lift of excavation so as to level the depth of

excavation at both left and right panels. This resulted in the release of larger earth

pressure behind the left sheet-pile wall and significant part of this pressure was carried

by anchors at left bank with others being carried by the sheet-pile-wall-struts system.

Unlike the struts system which may transfer the load from one side of excavation to

the other, anchor system at either side of the wall carried load independently. Fig. 7.35

shows the distribution of skin friction along anchor tendon at some key construction

stages which is an indication of load taken by anchors. Fig. 7.36 compares the skin

firction along anchor towards the end of the analysis at left bank and right bank. The

elevated amount of skin friction along L1 anchor than that of R1 is attributed to the

deeper excavation depth at left bank in the excavation of lift 4, as explained above.

However the pattern of the distribution curve of skin friction along the anchor bonded

length differs from what was observed in Chapter 6. This is because the anchor bonded

length in this case penetrated through medium stiff layer of F2 silty sand and then the

stiff layer of weathered rock S4 and ended in very stiff rock layer of S2. Larger values

of skin friction was developed in the part of the anchor bonded length which is within

S2 layer while smaller values of skin friction were developed in the S4 and F2 layers,

respectively. On the other hand, the bonded length of anchors in Chapter 6 was

installed in one single layer.

146

Page 167: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

7.7 Proposed element vs bar element representation of anchor

As mentioned earlier, a common way of modeling the bonded and unbonded lengths of

anchors is by means of simple bar elements (Briaud & Lim 1999). In this section, the

computed results obtained by the two methods of modeling are compared, using this

case study. Following Briaud & Lim (1999), in the bar element model, the anchor

bonded length was modeled using bar elements with their nodes directly attached to

soil nodes. The unbonded length was modeled using a single 2-noded bar element

linking up the node representing anchor head with the node on the front end of the

anchor bonded length. This effectively allows the free length to move in a manner

which is completely decoupled from the ground.

The deflection profile at t=221 days from the commencement of the construction was

used for comparison as this is the stage at which anchors were already installed. As

shown in Fig. 7.37(a), the wall deflection profiles from the two models are virtually

the same immediately before the installation of anchors. As shown in Fig. 7.37(b),

after the installation of the anchors, some differences arises with the proposed model

showing prediction slightly less deflection at the top of the sheet pile wall and larger at

deeper depth where anchors were installed. The similar difference in deflection

profiles from two models are observed for deflection profiles at other construction

stages, as shown in Fig. 7.38. The deflection profiles of left bank at key construction

stages are as shown in Fig. 7.39. As can be seen from the figure, the computed wall

deflection of left bank from analysis using proposed element is consistently larger than

that using bar element. Taking deflection profiles of both right bank and left bank into

consideration, the computed wall movement from analysis using proposed element

147

Page 168: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

show significant inward movement to the excavation area at deeper depth where

anchors were installed. The upper part of the wall movement is dominant with

movement swaying to the right bank. Comparing this with the wall movement from

analysis using bar element, as can be seen from Figures 7.38 and 7.39, insignificant

swaying movement at upper part of the wall and relatively smaller inward movement

at deeper depth of the wall.

This can be explained by the interaction of the unbonded length of the anchor with soil,

which is modeled in the proposed model. This interaction gives rise to a drag on the

soil around anchor flow towards the excavation, thereby relieving some of the earth

pressure on the wall. The bar element model decouples the unbonded length from the

soil, thereby ignoring this interaction.

Figures 7.40 and 7.41 show the incremental displacement vector field around the

anchor in the two models. The displacements represented in the figures are computed

incremental displacement from the time of installation of A4 anchors. This is to filter

off the displacement caused by earlier excavation lifts so as to single out the effects of

anchor installation to the ground movement. As can be seen, the drag from the anchor

causes a perturbation to the displacement vector field predicted by the proposed

element, with the displacement showing a larger upward component and a smaller

horizontal component than that predicted by the bar element model.

The effect of the drag from anchor unbonded length is also reflected on the

incremental ground surface movement, as shown in Fig. 7.42. The stressing load

148

Page 169: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

pushes the sheet-pile wall into the soil thereby causing heave behind wall. The heave

as computed by analysis using proposed element is larger at near field and smaller in

far field, as compared to that from bar element. When this incremental ground

movement is combined with ground settlement incurred during the earlier excavation

stages, the computed ground settlement from using proposed element will be smaller at

near field and larger in far field compared to that from bar element. This is consistent

with the findings in Chapter 6.

7.8 Summary of the chapter

An excavation in soft Singapore marine clay supported with heavy sheet-pile, internal

struts and ground anchors was studied using the proposed anchor-interface element to

model the ground anchors. The 3–D consolidation analyses with rigorous simulation

construction activities were conducted with reasonably matching with the measured

sheet-pile wall deflections at different construction stages. Factors influencing

numerical results were investigated. Simplified 3-D analyses were conducted and the

results from them compared with the proposed analysis. It was found that sheet-pile

wall deflections from numerical analyses are greatly affected by the detailed modeling

of initial ground conditions such as ground terrain and the actual excavation scheme

such as unbalanced excavation. No significant difference between the analyses using

proposed element and the analyses using bar elements was observed mainly because

the anchors are bonded in rock layer and the bonding strength is high than the shear

stress in the interface. The performance of proposed element are better than the

conventional bar element in the modeling of ground anchor system only when local

yielding is mobilized in the interface. Modified cam-clay soil model was also used in

149

Page 170: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

the present study. It was found that modified cam-clay model can better describe the

mechanical behaviour of Singapore marine clay. This study also confirmed that the

proposed element works fine with this soil model.

150

Page 171: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

CHAPTER 8

CONCLUSIONS

8.1 Summary of findings

This study deals with the interaction between soil anchors and the surrounding ground.

Ground anchors are typically installed inside pre-drilled holes in the ground. Because

of this, some interaction between anchor and ground will invariably exist. Along the

bonded length, shear and normal stresses will be developed along the anchor-soil

interface. Along the unbonded length, shear stresses are likely to be much reduced but

they may not necessarily be reduced to zero since some amount of residual friction will

invariably exist between the anchor and the sleeve as well as between sleeve and

ground. In addition, normal stresses may also be present across the anchor-soil

interface along the unbonded length. These interaction stresses will give rise to a drag

on the soil when it tries to move across the line of the anchor. In numerical modeling

of anchors, the anchor-soil interaction is usually simplified. For instance, the unbonded

length of an anchor is often modeled as a bar connecting the anchor head at the

retaining wall and the front of the bonded length (e.g. Briaud & Lim 1999). In so doing,

the interaction between anchor and soil is effectively ignored.

In this study, a new finite element is developed to replicate soil-anchor interaction so

as to facilitate three-dimensional analyses of ground-anchored system. The proposed

element consists of a two-noded beam element wrapped on its outside by an interface.

The total number of nodes is five and the total number of degrees-of-freedom is 12 for

2D and 19 for 3D. The outside of the interface, which is fully coupled to the soil

domain, has three nodes whereas the beam has two nodes. By allowing relative shear

151

Page 172: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

and transverse displacement between the beam and the outside nodes, displacement

between anchor and soil can be modeled. The outside nodes have a quadratic shape

function for displacement; thus it is compatible with linear strain solid elements. There

are no prescribed shape functions for the two inside nodes along the beam. The

longitudinal deformation of the beam was derived from the equilibrium with skin

friction. The transverse deflection of the beam was represented using a function

derived from Euler beam theory. As such, flexural deflection of the anchor will be

represented in accordance with Euler beam theory. This eliminated the non-conformity

between the beam and the soil nodes which is arising from directly coupling a beam to

a solid element. The development work was conducted in two phases, the first

involving the development of a linear interface and the second involving its extension

to non-linear and plastic behaviours.

The proposed element was verified against idealized problems for which closed-form

solutions have been derived. These problems are :

(a) Elastic solution of anchor in rigid ground subject to axial load;

(b) Elastic solution of anchor in rigid ground subject to lateral load at anchor

head;

(c) Elasto-perfectly-plastic solution of anchor in rigid ground subject to axial

load;

(d) Elasto-perfectly-plastic solution of anchor in rigid ground subject to lateral

load at anchor head.

Comparison between the results computed by the proposed element and closed-form

solution shows very good agreement even though some of the meshes are fairly coarse.

152

Page 173: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

This indicates that the proposed element is theoretically correct and numerically stable

and has good convergence characteristics.

The limitations of the proposed element were also investigated through case study of

an instrumented pile subject to axial loading, which has been reported by Cheung et al.

(1991). Back-analysis of the instrumented pile settlement shows that end bearing

mechanism plays an important role in supporting this pile. The inability of the

proposed element to capture the end bearing of pile due to its one-dimensionality

means that the proposed element can only be applied to model relatively long slender

solid inclusions.

The next case study involves some pull-out tests of single anchors. In all these pull-out

tests, the load-longitudinal displacement of each anchor was measured. Back-analyses

using simple bar elements to represent the bonded and unbonded lengths of the

anchors fail to replicate the hysteresis which were present in practically all of the

load-displacement curves. By using the proposed element and assigning an angle of

friction which is the typical to the anchor-sleeve contact, the observed hysteresis can

be reproduced. This indicates that the initial stiff portion of the load-displacement

curve may well be due to the friction along the unbonded length. Only when this is

overcome will all of the additional load be carried by the bonded length. If this is

indeed true, it may be prudent to make allowance for the fact that the bonded length is

not fully preloaded even though the anchor head may be. In order to preload the

bonded length to the prescribed level, a higher load may need to be applied to the

anchor head.

153

Page 174: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

The next case study involves an excavation supported with tieback systems includes

soldier pile and timber lagging, which has been reported by Briaud and Lim (1999). In

Briaud and Lim’s (1999) three-dimensional analysis, beam elements and spring

element were used to represent the bonded and unbonded lengths respectively. In order

to match observed and computed anchor behaviour during preload, Briaud and Lim

had to reduce the stiffness of the beam elements representing the bonded length by

40%. Back-analysis using the proposed element shows that, by considering the

materials constituting the interface and assigning interface parameters accordingly, the

observed wall deflection can also be replicated. This indicates that, by using the

proposed element, interface behaviour can be characterized in a heuristic fashion and

input parameters can be related to some real features of the interface, rather than by

arbitrary stiffness reduction. The study also shows that for a “Class A” prediction

during design phase the analysis using the proposed element model gives a larger

prediction of soldier pile deflection this may render a safe design as compared to the

analysis using the bar element model. In other words, there exists a risk of

underestimating system deformation when bar elements are used to model the anchor

and thus unsafe design.

The final case involves an excavation in soft Singapore marine clay supported with

internal struts and ground anchors. In this case study, the anchor was bonded into rock.

Back-analyses using bar element and the proposed element suggest that, in this case,

relative slippage between anchor bonded length and rock was insignificant. However,

the interaction between anchor unbonded length and the soft marine clay still gives rise

to noticeable differences in the predicted ground movement behind sheet-pile wall in a

near field, with the proposed element once again giving smaller ground settlement.

154

Page 175: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Thus, in the bonded length, significant differences in computed behaviour between the

proposed element and the bar element arise when the load level is high enough to

mobilize local yielding in the interface between anchors and surrounding soil. On the

other hand, if the load applied to anchor is so low so that the anchor soil interface is

within elastic domain, the differences in the performance of the proposed element and

that of the bar element may not be significant. More importantly, for all the three case

studies, modeling of the unbonded length with proposed element cause significant

localized differences in soil deformation in the retained soil domain compared to the

bar element. Since these three case studies cover three major soil types, namely stiff

residual soil, sandy soil and soft clayey soil, one may surmise that, in many instances

in the field, anchors may have discernible and, perhaps, even significant effect on the

deformation in the retained soil domain.

Besides the calibration of the proposed model and the demonstration of its applications

in the analyses of excavation support system, the significance of using 3D analyses in

modeling the interaction between soil and slender solid inclusions was highlighted

through comparison of the 3D model and 2D approximation. It was found that the 3D

modeling of anchor soil interface is highly recommended, especially when non-linear

behaviour of the interface is expected. This is because of the fact that the 2D

equivalent stiffness may give near correct answer for displacement, but the correctness

of the stresses in the anchor-soil interface are not guaranteed, therefore may lead to

incorrect modelling of the interface in non-linear domain.

155

Page 176: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

8.2 Recommendation for future research

The modeling of inclusions in soil often involves consideration of many issues. This

study models the interaction at a heuristic and rather basic level. There remain a host

of issues which need to be adequately addressed in future. Some issues which merit

future study are as follows:

(a) The proposed element models the inclusion using a one-dimensional

beam-interface element. In cross-section, this element has zero area. Thus, in

theory, the inclusion is a perfect stress concentrator, and stresses around this

element should theoretically reach infinity. In reality, this does not occur since

the inclusion always has a finite cross-sectional area. Neither is it reflected in

the finite element analyses because the solid elements used for the soil domain

cannot capture an infinite stress situation, owing to their finite sizes and finite

shape functions. In other words, the finite element size and shape function of

the finite element mesh impose a finite length scale onto the problem which is

totally unrelated to the reality of the problem. This limitation is not for the

proposed element only. The bar element, beam element, link element and other

line type of element for modeling solid inclusions also suffer from the same

deficiency. It is inherited from their geometrical nature of the element. The

implication of this will need to be investigated.

(b) Whereas the shear behaviour of the interface seems to have reasonably good

physical meaning, the meaning of the transverse behaviour of the interface is

less obvious. Since the interface is essentially an infinitesimally thin contact

surface between two media, the transverse stiffness and strength should

theoretically be infinite. In all the case studies considered, the shear behaviour

156

Page 177: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

of the interface is more dominant. However, if this proposed element is to be

extended to other types of inclusions, the transverse behaviour of the interface

may assume greater importance, and the philosophical issue of the meaning of

the transverse stiffness and strength will need to be resolved.

(c) Though issues like stress concentration around the proposed element are yet to

be addressed, as a reasonable approximation and simplification, the proposed

element, in its present form, can still be applied to many geotechnical

engineering problems where the global effects of the interaction between soil

and solid inclusions are of primary concern. Such applications include the

numerical analysis of soil retained with soil nail system, reinforced earth, and

analysis of tunneling using NATM. The proposed element can also be applied

to model many other solid inclusions such as the reinforcement in the

reinforced earth, geomembranes, diaphragm walls used as cut-off wall in

earthfill and rockfill dams.

(i) Soil nails grouted in a bored hole are essentially the same as anchors bonded

to the full length. The interaction between the relatively stiffer and stronger

nails and that of softer and weaker soil plays a significant role in affecting

the characteristics of the overall behaviour of the retained soil. For instance,

the ground movement induced by excavation supported with soil nail system

is greatly influenced by the interaction between soils and nails. The

proposed element can be applied to this case to model the soil nails installed

in the ground.

(ii) NATM had been widely used in tunneling within soft rock. Rock bolts are

often used in NATM together with shotcrete as a primary supporting system.

Numerical modeling of rock bolts in tunneling using NATM is traditionally

157

Page 178: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

done with bar element. As the proposed element is better than bar element in

modeling the non-linear interaction between rock and bolts, the application

of the proposed element to tunneling using NATM would be able to provide

improved accuracy in modeling rock bolt interaction during tunneling.

Although the proposed element was inspired and developed initially for modeling long

slender solid inclusions in soil, the element itself, as it stands, does not limit its

application only to geotechnical engineering. It can be used to benefit the users of the

FEM in other discipline. For example, the concepts of wrapping interface around a

solid element and establishing the element stiffness matrix through the internal

equilibrium between the interface and the solid element can be extended to the

modelling of prefabricated vertical drains for analysis of steady state fluid flow during

consolidation, and the modeling of cooling pipes installed in large concrete pours for

thermal analysis.

158

Page 179: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

REFERENCES

Akira Nagao & Teruki Kitamura, Jun Mizutani (1988). “Field experiment on

reinforced earth and its evaluation using FEM analysis” Int. Geotech.

Symp. On theory and practice of earth reinforcement. Fukuoka Japan. 5-6

Oct. 1988. pp329-334.

ASCE task committee on finite element analysis of reinforced concrete structures,

State-of-the-art reporton finite element analysis of reinforced concrete.

ASCE Special Publications, 1982.

Balaam, N.P., Poulos, H.G. & Booker, J.R. (1975) “Finite Element Analysis of Effects

of Installation on Pile- Load Settlement Behavior”. Geot. Engrg., Vol. 6,

No.1, pp33- 48.

Balasubramaniam, A.S., Brand, E.W., Sivandran, C. (1976). “Finite element analysis

of a full scale test excavation in soft Bankok clay.” Proc. of the 1976 Int.

Conf. on FEM in Eng., Adelaide, Australia. Bernat S. and Cambou, B. (1998) “Soil-structure Interaction in Shield Tunnelling in

Soft Soil” Computers and Geotechnics. Vol. 22, No. 3, pp221-242.

Bjerrum, L. and Eide, O. (1956). “Stability of strutted excavation in clay.”

Geotechnique, Vol.6, No. 1, pp32-47

Bolton, M.D. and Powrie, W. (1987). “The collapse of diaphragm walls retaining clay.”

Geotechnique Vol.37, No. 3, pp335-353.

Borja, I.R., Lee, S.R. and Seed, R.B. (1989). “Numerical simulation of excavation in

elastoplastic soils.” Int. J. for Num. & Anal. Methods in Geomech. Vol.13,

pp231-249. Bransby, M. F. and Springman, S. M. (1996). “3-D Finite Element Modelling of

Pile Groups Adjacent to Surcharge Loads”. Computers and Geotechnicas,

Vol.19, No.4, pp 301-324

Briaud, Jean-Louis and Lim, Yujin (1999) “Tieback Walls in Sand: Numerical

Simulation and Design Implications”. Journal of Geotechnical and

159

Page 180: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Geoenvironmental Engineering. ASCE, Vol. 125, No. 2, pp101-110

Britto, A.M. and Gunn, M.J. (1990). “CRISP90 User's and Programmer's Guide.”

Cambrige University.

Britto, A.M. and Kusakabe, O. (1984). “On the stability of supported excavation.” Can.

Goetech. J. Vol.21, pp338-348.

Brown P.T. & Booker J.R. (1985). “Finite Element Analysis of Excavation.” Computer

& Geotechnics. Vol.1 No.1. pp207-220

Brown, D.A. and Shie, C.F. (1990) “Numerical experiments into group effects on the

response of pile to lateral loading”. Computer and Geotechnics. Vol.10, No.

1, pp211-230.

Butterfield, R., and Banerjeer, P.K., (1971). “The elastic analysis of compressible piles

and pile groups” Geotechnique. Vol.21, No. 2, pp135-142

Byrne, R.J. (1992). “Soil nailing : a simplified kinematic analysis.” Proceeding of the

conference on Grouting, Soil improvement and Geosynthetics, New Orlean,

Louisana, Feb.25-28, 1992. Vol. 2, Geotechnical Special Publication No.

30, pp751-763

Cheung, Y.K., Lee, P.K.K and Zhao W.B. (1991) “Elastoplastic analysis of soil-pile

interaction.” Computer and Geotechnics. Vol.12, No. 2, pp115-132.

Chow, Y.K., (1986). “Analysis of vertically loaded pile groups” Int. Jnl. For Num.

And analytical Methods In Geomech., Vol.10, No.1, pp59-72

Christian, J.T. (1968). “Undrained stress distribution by numerical methods.” Journal

of soil mechanics and foundation division, ASCE, Vol.94, No.SM6, Nov.

pp.1333.

Chua, L.H. (1985). “Analysis and design of excavation support system on soft clay.”

Thesis submitted to the National University of Singapore for degree of

Master of Engineering. 152pp.

Cividini, A., Gioda, G. and Sterpi D. (1997). “An experimental and numerical study of

the behaviour of reinforced sand.” Computer methods and Advances in

160

Page 181: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Geomechanics, pp15-30, Yuan (ed.) Balkema, Rotterdam

Clough, G.W. and Mana, A.I. (1976). “Lessons learned in finite element analysis of

temporary excavations.” Proc. of the 2nd Int. Conf. on Num. Methods. in

Geomech. ASCE, Blacksburg, VA.

Coyle, H.M., Reese, L.C., (1966). “Load transfer for axially loaded piles in clay”, Jnl.

Of Soil Mech. And Found. Engrg. Div., ASCE. Vol.92, No.SM2, pp1-26.

D’Appolonia, E. & Romualdi, J. P. (1963) “ Load Transfer in End-Bearing Steel

H-Piles”. J. Of Soil Mech. And Foundation Engrg. Div., ASCE. Vol.89,

SM2, pp1- 25.

Dames and Moore (1983). “Detailed geotechnical study interpretation report”

sunmitted to Singapore Mass Rapid Transit System.

Day, R.A. and Potts D.M. (1994). “Zero thickness interface element-numerical

stability and application.” Int. J. Numer. Anal. Meth. Geomech. Vol. 18,

pp689-708.

Day, R.A. and Potts D.M. (1998). “The effect of interface properties on retaining wall

behaviour.” Int. J. Numer. Anal. Meth. Geomech. Vol. 22, No.12,

pp1021-1033

Desai, C.S. and Faruque, M.O. (1984). “Constitutive model for (geological) materials.”

J. Engng. Mech. Div., ASCE 110.

Desai, C.S., (1974).“Numerical design- analysis for piles in sand.” Jnl. Of Geot. Engrg.

Div. ASCE, Vol.100, No. GT6, pp613-635

Desai, C.S., Drumm, E.C. and Zaman, M.M. (1985). “Cyclic testing and modeling of

interfaces.” J. Geotech. Engng. ASCE, 111(6), pp793-815.

Desai, C.S., Johnson L.D. & Hargett C. M. (1974) “ Analysis of Pile Supported

Gravity Lock”. J. Geot. Engrg. Div., ASCE. Vol. 100, GT9, pp1009-

1029.

DiBiagio, E. and Roti, J.A. (1972). “Earth pressure measurement on a slurry trench

wall in soft clay.” Proc. of the 5-th Europpen conf. on Soil Mech. & Found.

Eng, Madrid, Vol. 1, pp437-484.

161

Page 182: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

DiMaggio, F.L. and Sandler, I.S. (1971). “Material model for granular soils.” J. of Eng.

Mech. Div. ASCE, Vol.97, No.3.

Dickin, E. A. and King, G.J.W. (1996). “Finite element analyses of vertical anchor wall

behavour in clay backfill”. Advances in Fnite Element Technology.

Pp301-307.

Dysli, M. and Fotana, A. (1982). “Deformations around the excavation in clayey soil.”

Proc of the Int. Symposium on Numerical Models in Geomechnics, Zurich,

Sept. pp634-642.

Ellison, R. D., D’ Appolonia, E., & Thiers, G.R. (1971) “Load-Deformation

Mechanism for Bored Piles”. J. of Soil Mech. And Foundation Engrg. Div.,

ASCE. Vol. 97, SM4, pp661- 678.

Evgin, E. and Fakharian, K. (1996). “Effect of stress paths on the behaviour of

sand-steel interfaces.” Can. Geotech. J., 33(6), pp853-865.

Finno, R.J., Lawrence, S.A., Allawh, N.F. and Harahap, I.S. (1991). “Analysis of

performance of pile groups adjacent to deep excavation.” J. of

Geotechnical Engrg., ASCE, Vol.117, No.6, pp934-955.

Fishman, K.L. and Desai, C.S. (1987). “A constitutive model for hardening behaviour

of rock joints.” in Desai et al. (eds), “Constitutive laws for Engineering

Materials: theory and Applications”, pp1043-1050.

Fleming,W.G.K., Weltman, A.J., Randolph, M.F. and Elson, W.K. (1992). “Piling

Engineering” 2nd edition. John Wiley and Sons.

Galybin, A.N., A.V. Dyskin, B.G. Tarasov And R.J. Jewell (1999). “An approach to

large-scale field stress determination.” Geotechnical and Geological

Engineering 17, 1999. pp267–289.

Ganesh, (2002). “Introduction to the excavation of MRT Depot at Upper Paya Leba

road.” Material prepared for the site visit organized by Institute of

Engineers Singapore (IES).

Gens, A., Carol, I., and Alonso, E.E. (1988). “An interface element formulation for

analysis of soil-reinforcement interaction.” Computers and Geotechnics.

162

Page 183: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Vol. 7, No.1, pp133-151.

Ghaboussi, J and Wilson, E.L. (1973). “Finite element for rock joints and interfaces”, J.

Soil Mech. Foundations Div., ASCE, 99(SM10), pp833-848.

Goh, A. T. C., Wong, K. S., Teh, C. I. and Wen D.Z. (2003) “Pile Response Adjacent to

Braced Excavation”. J. of Geotech. And Geoenv. Eng. ASCE, Vol.129,

No.4, pp. 383-386.

Goodman R.E, Taylor, R.L, and Brekke,T.L. (1968) “A model for the mechanics of

jointed rock”. J. Soil Mech. And Foundation Div., ASCE. Vol. 94, SM3,

pp637-659.

Griffiths, D. V. (1985). “Numerical modeling of interface using conventional finite

element”. Proc. Of 5-th Int. Conf. on Num. Meth. In Geomech. Nagoya.

Vol.2, pp837-844.

Gudehus, G., (1972). “Lower and upper bounds for stability of earth retaining

structures.” Proc. of 5-th European Conf. on Soil Mech. and Found. Engrg.

Madrid, 1972, Vol.1, pp21-28.

Gudehus, G., Sclegel, Th. and Schartz,W. (1985). “Subgrade moduli for shallow

foundation and piles.” The 5-th Int. Conf. on Num. Methods in

Geomechanics, Nagoya, Japan, pp731-745.

Guo, D.J., Tham, L.J. and Cheung, Y.K., (1987). “Infinite layer for the analysis of a

single pile” Computes and Geotechnics. Vol.3, No. 4, pp229-249

Hansen, L.A. (1980). “Prediction of the behaviour of braced excavation in anisotropic

clay.” ph.D thesis, Stanford University. pp439.

Hata, S., Yoshida, S., Ohta, H. and Honda, H. (1985). “A deep excavation in soft clay

performance of an anchored diaphragm wall.” Proc. of the 5-th Int. conf.

on Num. Methods in Geomechanics, Nagoya, April, pp725-730.

Herrmann, Leonard R. and Zaynab Al-Yassin (1978). “Numerical analysis of

reinforced soil system”. Pp428-457. Symp. On Earth Reinforcement.

Pittsburgh, Pennsylvania, April 27, 1978. Hird, C.C. and Kwok, C.M. (1989) “Finite Element studies of interface

behaviour in reinforced embankment on soft ground”. Computers

163

Page 184: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

and Geotechnics. Vol. 8, No. 2, pp111-131.

Hong SH., Lee FH., and Yong KY (2003) “Three dimensional pile-soil interaction in

soldier piled excavations”. Computers and Geotechnics, Vol.30, pp81-107.

Hsieh H.S., Kavazanjian, E.Jr., and Borja,R.I. (1990). “Double yield surface Cam-clay

plasticity model.” I: theroy. J. geotech. Eng., ASCE, Vol.116 No.9,

pp1381-1401.

Hsieh Hsii-Sheng, Wang, Chien-Chih, and Ou, Chang-Yu. (2003). “Use of jet grouting

to limit diaphragm wall displacement of deep excavation”. J. geotech. Eng.,

ASCE, Vol.129, No.2, pp146-157.

Hu Liming and Pu Jialiu (2004) “Testing and modeling of soil-structure interface” J.

geotech. And Geoenv. Eng., ASCE, Vol.130, No.8, pp851-860.

Ishihara, K. (1970). “Relationship between process of cutting and the uniqueness of

the solution.” Soils and Foundations. Vol.10,No.3, pp50-65.

Itasca Consulting Group Inc. (1996). FLAC Manual, Version 3.3 Itasca Consulting

Group Inc., USA.

Jeffreys, H. (1969). “Cartesian tensors”. Cambridge University Press.

Jewell, R.A. and Pedlay, M.J. (1992) “Analysis for soil reinforcement with bending

stiffness”. J. of Geotech. Eng. Vol.118, No.10, pp1505-1528.

Kaliakin, V.N., Li, J. (1995). “Insight into deficiencies associated with commonly used

zero-thickness interface elements.” Computers and Geotechnics, Vol.17,

No. 2, pp225-252.

Katona, M.G. (1983). “A simple contact friction interface element with applications to

buried culverts”. Int. J. Num. Anl. Meth. Geomech. Vol. 7, pp371-384.

Kawamoto, T., Kyoya, T and Aydan, O. (1994) “Numerical model for rock reinforced

by rockbolts”. Copmuter Meth. And Adv. In Geomech., Siriwardane &

Zaman, Eds. Blkema, Rotterdam. Pp33-45.

Kimura, T., Takemura, J., et, al (1993). “Stability of unsupported and supported cuts in

soft clay.” Proc. of 11-th SEAN Geotech. Conf. Singapore, pp61-70.

Kiousis, P.D. and Elansary, A.S., (1987). “Load settlement relation for axially loaded

164

Page 185: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

piles” Jnl. Of Geot. Engrg. Div. ASCE. Vol.113, No.GT6, pp655-661.

Kishida, H. and Uesugi, M. (1987). “Tests of interface between sand and steel in the

simple shear apparatus”, Geotechnique, 37(1), pp45-52.

Kwak, Hyo-Gyoung and Kim, Sun-Pil (2001). “Bond-slip behavior under monotonic

unaxial loads” Eng. Struct. Vol. 23 pp298-309.

Lambe, T.W. (1970). “Braced excavation.” Proc. of spec. conf. on lateral stresses in the

ground and the design of earth retaining structures, Ithaca, New York,

pp149-218.

Lambe, T.W., Wolfskill, L.A. and Wong, I.H. (1970). “Measured performance of

braced excavation.” J. of soil Mech. and Found. Eng., Div. ASCE, Vol.96 ,

No. SM3. pp817-836.

Lee, F.H., Liu K.X., Yong, K.Y. (1997). “Prediction of ground movement around deep

excavation”. Keynote paper,Geotechnical engineering development in Asia.

Proc. of 4-th Regional Conf. in Geotech. Eng. (GeoTROPIKA’97). Johor,

Malaysia, pp1-20.

Lee, F.H., Tan, T.S and Yong, K.Y. (1993). “Excavation in residual soils with high

permeability.” Proc. of the 11-th SEA Geotech. Conf. 4-8 May, Singapore.

pp739-744.

Lee, F.H., Yong, K.Y., Quan, K. C.N., Chee. K.T. (1998). “Effects of corners in strutted

excavation: Field monitoring and case histories”. J. of Geotech. And

Geoenv. Eng. ASCE, Vol.124, No.4, pp339-249.

Lee, F.H., Yong, K.Y., S.L. Lee, C.T. Toh (1989). “Finite element modelling of strutted

excavation.” Numerical Models in Geomechanics(NUMOG III), Niagra

Falls Canada, pp577-584.

Lee, S.L., Parnploy, U., Yong, K.Y. and Lee, F.H. (1989). “Time-dependent

deformation of anchored excavation in soft clay.” Proc. Symp. on

underground excavations in soils and rocks. Dec. Asian Int. of Tech.

Bankok, Thailand, pp377-384.

Lee, S.L., Yong, K.Y., Karunaratne, G.P. and Swaddiwudhipong, S. (1984). “Analysis

165

Page 186: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

of excavation support system in soft clay.” Proc. Sem. on Soil

improvement and construction techniques in soft ground, Singapore,

pp60-65.

Lee Yeong (1999) “A framework for design of jet grout piles in Singapore marine

clay” Thesis submitted to the National University of Singapore for partial

fullfillment of a degree of Master of Engineering.

Leung, C. F. , Chow, Y. K. and Shen, R. F. (2000). “Behaviour of pile subject to

excavation induced soil movement”. J. of Geotech. And Geoenv. Eng.

ASCE., Vol.126, No. 11, pp947-954.

Liang, Robert Y. and Feng, Y. (2002) “Development and application of anchor-soil

interface model”. Soils and Foundations. Vol.42, No.2, pp59-70.

Liu, H. and Singh K.J. (1977). “Numerical performance of some FE scheme for

analysis of seepage in porous elastic media.” Int. J. Ana. & Num. Meth.

Vol.1, pp177-194. Loganathan, N. and Poulos, H.G. (1998), “Analytical prediction for tunneling-induced

ground movement in clays”. J. of Geotech. And Geoenv. Engg. ASCE.

Vol.124, No.9, pp846-856.

Loganathan, N., Poulos, H.J., Stewart, D.P., (2000). “Centrifuge model testing of

tunneling-induced ground and pile deformations.” Geotechnique. Vol.50,

No. 3, pp283-294.

Lyndon, A. and Schofield, A.N. (1970). “Centrifuge model test of short term failure in

London clay.” Geotechnique. Vol. 20, No. 4, pp440-442.

Mafei, C., Andre, J. and CIFU, S. (1977). “Methods for calculating braced

excavations.” Proc. of Int. Symp. of soil structure interaction.” Roorkee,

India, pp85-92.

Mana, A.I. (1978). “Finite elment analysis of deep excavation behaviour in soft clay.”

PhD. Thesis, Stanford University, Stanford, California (1978).

Matlock, H. And Reese, L.C. (1960) “Generalized solutions for laterally loaded piles”.

166

Page 187: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

J. of Soil Mech. And Foundation Engrg. Div. ASCE. Vol.86, No.5,

pp63-91.

Matsui T and San. K.C. (1988). “Finite element stability analysis method for

reinforced slope cutting.” pp317-322. Int. Geotech. Symp. On theory and

practice of earth reinforcement. Fukuoka Japan. 5-6 Oct. 1988. Mauricio Ehrlich, Marcio S. S. Almeida and Alexandre M. Lima (1996). “Parametric

study of soil nailing system.” Proc. Earth reinforcement, pp747-752. Ochai,

Yasufuku & Omine (eds) Balkeman, Rotterdam.

Meyerhof G.G. (1976). “Bearing capacity and settlement of pile foundations”. J.

Geotechnical Engrg. Div. ASCE, Vol. 102 No. GT3, pp197-228.

Meyerhof, G.G. (1959) “ Compaction of Sands and Bearing Capacity of Piles”. J. of

Soil Mech. And Foundation Engrg. Div., ASCE. Vol.85, SM6, pp1- 29.

Miyoshi, M. (1977). “Mechanical behaviour of temporary braced walls.” Proc. of 9-th

Int. conf. on soil Mech. and found. Engrg. Tokyo, Vol.1, pp655-658.

Monti, G., Filippou, F. C., and Spacone, E. (1997). ‘‘Finite element for anchored bars

under cyclic load reversals.’’ J. Struct. Engrg., ASCE, Vol.123, No.5,

pp614–623.

Monti, Giorgio and Spacone, Enrico (2000) “Reinforced concrete fiber beam element

with bond-slip”, Journal of Structural Engineering, Vol.126, No. 6, pp

654-661.

Mosaid M. Al-Hussaini and Lawrence D. Johnson. (1978). “Numerical analysis of a

reinforced wall.” pp98-126. Symp. On Earth Reinforcement. Pittsburgh,

Pennsylvania, April 27, 1978.

Mroueh, H. and Shahrour, I. (2002). “Three-dimensional finite element analysis of

interaction between tunneling and pile foundations” . Int. J. Numer. Anal.

Meth. Geomech. Vol.26, pp217-230.

167

Page 188: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Navayogarajah, N., Desai, C.S. and P.D. Kiousis, (1992). “Hierarchical single-surface

model for static and cyclic behaviour of interfaces.” J. Engng. Mech.

ASCE, 118(5), pp990-1011.

Naylor, D.J. (1978). “A study of R.E. walls allowing strip slip.” pp618-643. Symp. On

Earth Reinforcement. Pittsburgh, Pennsylvania, April 27, 1978.

Ng, C.W.W., and Lee, G.T.K. (2002). “A three-dimensional parametric study of the use

of soil nails for stabilising tunnel faces.” Computers and Geotechnics.

Vol.29, pp673-697.

Ng, K.L.A. and Small, J.C. (1997). “Behavior of Joints and Interfaces Subjected to

Water Pressure.” Computers and Geotechnics, Vol. 20, Issue 1, pp71-93 Ng, P.C.F., Pyrah,I.C. & Anderson, W.F. (1998) “Asseement of three

interface elements and modification of the interface element in CRISP90.” Computers and Geotechnics. Vol. 21, No. 4, pp315-339.

Ngo D, Scordelis AC. (1967). “Finite element analysis of reinforced concrete beams.”

Journal of the American Concrete Institute. Vol.64, No.3, pp152-163.

O’Rouke, T.D, Cording, E.J. and Boscardin, M. (1976). “The ground movements

related to braced excavations and their influence on adjacent buildings”.

U.S. Department of Transportation, Report No. DOT-TST 76 T-23, 123 pp.

Ogisako, E., Ochiai, H. Hayashi, S. and Sakai, A. (1988). “FEM analysis of polymer

grid reinforced-soil retaining walls and its application to the design

method.” pp559-564. Int. Geotech. Symp. On theory and practice of earth

reinforcement. Fukuoka Japan. 5-6 Oct. 1988.

Osaimi A.E. and Clough G.W. (1979). “Pore pressure dissipation during excavation.”

Joural of the Geotechnical Engineering Divsion, ASCE. Vol.105, No. GT4.

pp481-498.

Ottaviani, M., (1975). “Three- dimensional finite element analysis of vertically loaded

pile groups.”, Geotechnique, Vol.25, No. 2, pp159-174

Ou, C.Y. and Chiou, D.C. (1993). “Three-Dimensional Finite Element Analysis of

Deep Excavation.” Proc. of the 11-th SEA Geotechnical Conference, 4-8

May, 1993, Singapore. pp769-774.

168

Page 189: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Pande, G.N., and Sharma, K.G. (1979), “on joint/ interface elements and associated

problems of ill-conditioning”. Int. J. Numer. And Analy. Methods in

Geomech. Vol.3, No.1, pp293-300.

Parnploy, U. (1990). “Time-dependent behaviour of excavation support system in soft

clay.” Thesis submitted to the National University of Singapore for partial

fullfillment of a degree of phD.

Peck, R.B. (1943). “Earth pressure measurements in open cuts of Chicago (Ill),

Sunway.” Trans. ASCE 108, pp1003-1036. Peck, R.B. (1969). “Deep excavation and tunnelling in soft ground.” Proc. of 7-th Int.

Conf. SMFE, Mexico, State-of-the-art volume, pp225-290.

Poulos, H.G. (1994). “An approximate numerical analysis of pile-raft interaction.” Int.

J. for Num. And Anal. Methods in Geomechs. Vol.20, No.1, pp57-72.

Poulos, H.G. and Davis, E.H. (1980). “Pile foundation analysis and design.” series in

Geotechnical engineering.

Poulos, H.G., Davis, E.H., (1968). “The settlement behaviour f a single axially loaded

incompressible piles and piers”, Geotechnique, Vol.18, No. 3, pp351-371

Pressley, J.S., Poulos, H.G., “Finite element analysis of mechanism of pile group

behaviour.” Int. Jnl. For Num. And analytical mtds. In Geomech. 10 No. 2

(1986) 213-221

Randolph, M.F., Wroth, C.P., (1978). “Analysis of the deformation of vertically loaded

piles”, Jnl. Of Geot. Engrg. Div. ASCE. Vol.104, No.GT12, pp1465-1488

Richart, F.E. Jr. (1959). “Analysis for sheet-pile retaining walls.” ASCE Transactions.

Vol.122, 1957 pp1113.

Rojas, E., Valle C. & Romo M. P. (1999) “Soil-pile interface model for axially loaded

single piles”. Soils and Foundations. Vol. 39, No.4, pp35-45.

Sandhu, R.S. and Wilson, E.L. (1969). “Finite element analysis of seepage in elastic

media.” J. of Eng Mech. Div. ASCE, Vol.95, EM3, pp641-652.

Schlosser, F., Unterreiner, P. and Plumelle C. (1992). “French research program

169

Page 190: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

CLOUTERE on soil nailing.” Proceeding of the conference on Grouting,

Soil improvement and Geosynthetics, New Orlean, Louisana, Feb.25-28,

1992. Vol. 2, Geotechnical Special Publication No. 30, pp739-750.

Schnabel, Henry and Schnabel, Henry W. (2002) “Tie-backs in foundation

engineering and construction.” 2nd edition. Balkema Publishers.

Schofield, A.N. and Wroth, C.P. (1968). “Critical state soil mechanics.” McGraw-Hill,

London.

Schweiger, H.F. and Hass, W. (1988). “Application of thin layer interface element to

geotechnical problems”. Proc. Of 6-th Int. Conf. on Num. Meth. In

Geomech. Innsbruk, Ed. Swoboda. Vol.2, pp907-912.

Shoji, M. and Matsumoto (1976). “Consolidation of embankment foundation.” Soil

and foundations, Vl.16, No.1, pp59-74.

Simpson, B. (1992). “Retaining structure: displacement and design.” Geotechnique Vol.

42. No.4, pp541-576.

Simpson, B. and Wroth, C.P. (1972). “Finite element computation for a model

retaining wall in sand.” Proc. of the 5-th European conf. on soil mechanics

and foundation engineering. Madrid, 1972, Vol. 1, pp85-95.

Small, J.R., Booker, J.R. and Davis, E.H. (1976). “Elasto-plastic consolidation of soil.

Int. J. of. solid & structures.” Vol. 12, pp431-448.

Smith IM. (1992). “Three dimensional analysis of reinforced and nailed soil.” In:

Pande GN, Pietruszczak S, editors. Numerical models in geomechanics.

Balkema, 1992. pp 829-838.

Smith, I.M. and Ho, D.K.H (1992). “Influence of construction technique on the

performance of a braced excavation in marine clay.” Int. J. for Num. and

Anal. Methods in Geomech. Vol. 16, pp845-867.

Smith, I.M. and Hobbs. R. (1976). Biot analysis of consolidation beneath

embankments. Geotechnique, Vol. 26, No.1, pp149-171.

Smith, I.M. and Su N. (1997) “Three-dimensional FE analysis of a nailed soil wall

curved in plan” Int. J. Numer. Anal. Meth. Geomech. Vol.21, No.9,

170

Page 191: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

pp583-597.

Smith, I.M. and Wang A. (1998) “Analysis of piled rafts.” Int. J. Numer. Anal. Meth.

Geomech. Vol.22, No.10, pp777-790.

Sprenger, W. and Wagner, W. (1999). “On the formulation of geometrically nonlinear

3D-Rebar-elements using the enhanced strain method.” Engineering

Structures. Vol. 21, No. 3, pp209-218.

Sudret, Bruno and Buhan, Patrick de (2001). “Multiphase model for

inclusion-reinforced geostructure application to rock-bolted tunnels and

piled raft foundations.” Int. J. Numer. Anal. Meth. Geimech. Vol.25,

pp155-182.

Ta, L.D. and Small, J. C. (1997) “An approximation for analysis of raft and piled raft

foundations”. Computers and Geotechnics. Vol.20, No.2, pp105-123.

Tan, S.A., Luo, S.Q. and Yong, K.Y. (2000). “Simplified models for nail-soil lateral

interaction.” Ground Improvement. Vol. 4 pp141-152.

Tan, S.L., (1983). “Geotechnical properties ant laboratory testing of soft soils in

Singapore.” Proc. of the Int. Semp. on Construction Problems in Soft Soils.

Nanyang Inst. of Tech. Singapore.

Tanaka, H. (1994). “Behaviour of a braced excavation in soft clay and the undrained

shear strength for passive earth pressure.” Soils and foundations. Vol. 34,

No.1, pp53-64.

Terzaghi, K. (1941). “General wedge theory of earth pressures.” Trans. ASCE 106.

pp68-97.

Terzaghi, K. (1943). “Theoretical soil mechanics.” Willey, New York, pp510.

Thomas C. Sheahan and Carlton L. Ho (2003) “Simplified Trial Wedge ethod for Soil

Nailed Wall Analysis”. J. of Geotech. Eng. ASCE. Vol. 129, No.2,

pp117-124.

Trochanis, A.M., Bielak, J., and Christiano, P. (1991) “Three-Dimensional nonlinear

study of piles”. J. of Geotech. Eng. ASCE. Vol. 117, No.3, pp429-447.

Tschebotarioff, G.P. (1973). “Braced and anchored cuts. Foundations, Retaining and

171

Page 192: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

earth structures.” 2nd Ed., McGraw Hill, Chapter 11, pp415-457.

Tsui, Y. and Clough, G.W. (1974). “Performance of tie-back walls in clay.” J. of

Geotech. Eng. Div. ASCE. Vol.100, No.GT12, pp1259-1274.

Tufenkjian,Mark,R. and Vucetic Mladen (2000) “Dynamic failure mechanism of

soil-nailed excavation models in centrifuge”. J. of Geotech. And Geoenv.

Eng. ASCE., Vol. 126, No. 3, pp227-235.

Valiappan, S., Lee, I.K. and Boonlaulorh, P., (1974). “Settlement analysis of piles in

layered soil.” Proc. 7th Conf. of the Aust. Road Research Board, part 7,

pp144-153.

Vallabhan, C.V.G. and Daloglu, A.T. (1999). “Consistent FEM-Vlasov model for plate

on layered soil.” J. of Structural Engrg. Vol. 125, No.1, pp108-113.

Vermeer, Pieter A., Ankana Punlor, Nico Ruse (2001). “Arching effects behind a

soldier pile wall.” Computers and Geotechnics,Vol. 28, No. 7, pp379-396.

Wakai, Akihiko, Gode, Shingo, and Ugai, Keizo. (1999). “3-D Elasto-plastic finite

element analyses of pile foundations subject to lateral loading”. Soils and

Foundations. Vol.39, No.1, pp97-111.

Wilson, E. L. (1977). “Finite element for foundations, joints and fluids”. In Finite

element in Geomechanics, Ed. By Gudehus, Wiley Chichester. Pp315-350.

Wong, K.S. and Broms, B.B. (1989). “Lateral wall deflections of braced excavation in

clay”. J. Geotech. Eng. Div., ASCE, Vol.115, No.6, pp853-870.

Xu, KJ, and Polous, H.G. (2000) “General elastic analysis of piles and pile groups”. Int.

J. Numer. Anal. Meth. Geomech. Vol. 24, pp1109-1138

Yin Zong-Ze and Zhu Hong, Xu Guo-Hua (1995). “A study of deformation in the

interface between soil and concrete.” Computers and Geotechnics, Vol. 17,

No.1, pp75-92.

Yong, K.Y. (1991). “Grouting in substructure construction.” Proc. 9-th Asian Reg.

Conf. on Soil Mech. and Fdn. Engrg., Bankok, Thailand, Dec. pp1-9.

Yong, K.Y., Karunaratne, G.P. and Lee, S.L. (1990). “Recent development in in soft

clay engineering in Singapore.” Guest Lecture, Proc. Int Geotech.

172

Page 193: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Forum'90 on comparative geotech. engrg. Osaka, Japan, Oct. pp3-10.

Yong, K.Y., Lee, F.H., Parnploy, U. and Lee, S.L. (1989). “Elasto-plastic consolidation

analysis for strutted excavation in clay.” Computers and Geotechnics,Vol. 8,

pp311-328.

Yoo, Chungsik (2001). “Behaviour of braced and anchored walls in soils overlying

rock”. J. of Geotech. And Geoenv. Engg. Vol. 127, No. 3, pp225-233.

Zhang Mingju, Song Erxiang, and Chen Zhaoyuan (1999). “Ground movement

analysis of soil nailing construction by three-dimensional (3-D) Finite

element modeling (FEM).” Computers and Geotechnics. Vol.25, No.4,

pp191-204. Zienkiewicz, O.C. and Taylor, R.L. (1991) “The finite element method,

volume 1”. McGraw-Hill, 4th Edition, London.

Zienkiewicz, O.C., Valliapan and King, I.P. (1969). “Elasto-plastic solution of

engineering problems.” 'initial stress' finite element approach. Int. J. for

Num. method in Engrg. Vol. 1 No. 1, pp75-100.

173

Page 194: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Fig. 2.1 Apparent pressure diagrams for computing strut load (Peck, 1969)

174

Page 195: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

175

Not

es:

[1]

Zon

e I -

Wel

l- br

aced

exc

avat

ions

with

slur

ry w

all o

r sub

stan

tial b

erm

s lef

t per

man

ently

in p

lace

. [2

] Zon

e II

-Exc

avat

ions

with

tem

pora

ry b

erm

s and

raki

ng st

rut s

uppo

rt.

[3] Z

one

III -

Exca

vatio

ns w

ith g

roun

d lo

ss fr

om c

aiss

on c

onst

ruct

ion

or in

suffi

cien

t wal

l sup

port.

[4

] In

this

con

text

, the

term

“be

rm”

was

use

d fo

r a p

assi

ve so

il bu

ttres

s. Fi

g.2.

2 Em

piric

al c

hart

of se

ttlem

ent a

djac

ent t

o st

rutte

d ex

cava

tion

(O’R

ouke

et.

al.,

1976

)

Page 196: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Fig. 2.3 Element model for soil nail proposed by Herrman and Zaynab (1978)

Fig. 2.4 Element model for soil nail in FLAC (Itasca, 1996)

176

Page 197: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Fig. 2.5 Element model for rock anchor interface by Kawamoto et. al. (1994)

Fig. 2.6 Illustration of Linker element and the application in RC modeling. (Adapted from ASCE report, 1982)

177

Page 198: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

(a) 3-D view of anchor–soil interaction (b) Idealised 2-D view

Fig 3.1 The proposed anchor-interface element

(a) Local coordinate system (b) Local to global system

1

2

4

3

x 1 2

3 4

y

7

X

y

z

x

Z

Y

3 4 7

x

y

z

2 1

Fig. 3.2 Coordinate system transformation

178

Page 199: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Axial displacement along anchor length

01020304050

0 2 4 6 8 10

Distance from anchor end (m)

Dis

plac

emen

t (m

m)

TheoreticalFEM

Fig. 3.3. Variation of axial displacement along an anchor

Distribution of Skin friction along anchor length

-5

-4

-3

-2

-1

0

0 2 4 6 8 10

Distance from anchor end (m)

Skin

fric

tion

(kN

/m)

FEMTheoretical

Fig.3.4 Variation of skin friction along an anchor

179

Page 200: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

P y x

Fig. 3.5 Model for beam on elastic medium

Fig.3.6 Finite element mesh of elastic beam on stiff soil foundation

180

Page 201: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Anchor displacement under lateral loadA comparison between closed form solutions & FEM

-0.012-0.010-0.008-0.006-0.004-0.0020.000

0 0.2 0.4 0.6 0.8 1

Along anchor length (x/L)

Def

lect

ion

(mm

))

0.00192FEM0.192FEM1.92FEM9.6FEMCantilever

Fig. 3.7 Comparison of lateral deflection with different relative anchor stiffness ratio

Fig. 3.8 FEM mesh for comparison of the anchor-interface element with bar element

181

Page 202: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Fig. 3.9 Distorted FEM mesh from analysis using the anchor-interface elements.

Fig. 3.10 Distorted FEM mesh from analysis using 3-D bar elements.

182

Page 203: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

1

10

100

1000

0.1 1 10 100 1000

µ

Rel

ativ

e di

spla

cem

ent

Fig. 3.11 Relative displacement vs eigen value µ.

183

Page 204: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Kn

σf

σn

∆v

Ks

τf

τ

γ

Fig. 4.1 Typical model for stress-strain relationship of interface

x

1

y

Plastic zone: lpElastic zone: le

P

Skin Friction

Fig. 4.2 Anchor in rigid ground subjected to axial load.

184

Page 205: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Variation of axial displacement along anchor

00.050.1

0.150.2

0.250.3

0 2 4 6 8 10

Anchor fixed length (m)

Axi

al d

ispl

acem

ent (

m)

ElastoplasticElastic

Fig. 4.3 Variation of anchor settlement along anchor height from closed-form solutions.

Variation of axial displacement along anchorElastoplastic analysis

00.050.1

0.150.2

0.250.3

0 2 4 6 8 10

Distance from anchor point (m)

Axi

al d

ispl

acem

ent (

m)

FEM(Total Load Method)FEM(Incremental Method)Closed form solution

Fig. 4.4 Comparison of anchor displacement between proposed FEM and closed-form solutions.

185

Page 206: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Skin friction along anchor fixed length from Elastoplastic analysis

-120-100-80-60-40-20

00 1 2 3 4 5 6 7 8 9 10

Distance from anchor point (m)

Skin

fric

tion

(kPa

)

FEM(Total Load Method)

Closed form solution

FEM(Incremental Method)

Fig. 4.5 Comparison of skin friction between proposed FEM and closed-form solutions.

Variation of axial force along fixed length

-1200-1000

-800-600-400-200

00 1 2 3 4 5 6 7 8 9 10

Distance from anchor point (m)

Axi

al fo

rce

(kN

) FEM(Total Load Method)Closed form solution

FEM(Incremental Method)

Fig. 4.6 Comparison of axial force between proposed FEM and closed form solutions.

186

Page 207: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

x

y Earth pressure on

h σf

Elastic zone :le Plastic zone: lp

P

Fig. 4.7 Anchor in rigid ground subjected to lateral loading.

Variation of deflection along fixed length of anchor

-0.50

0.51

1.52

2.5

0 2 4 6 8 10

Distance from anchor point (m)

Def

lect

ion

(mm

) Closed form (Elastoplastic)Closed form (Elastic)FEM (Elastic)FEM(Total Load Method)FEM(Incremental Method)

Fig. 4.8 Variation of deflection along anchor fixed length.

187

Page 208: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Bending Moment diaghram from Elastoplastic analysis

-12-10

-8-6-4-2024

0 2 4 6 8 1

Distance from anchor point (m)

Ben

ding

Mom

ent (

kN.m

)

0

Closed form solution

FEM (Incremental Metod)

FEM (Total Load Method)

Fig. 4. 9 Variation of bending moment along fixed length of an anchor.

Shear Force Diagram from Elastoplastic analysis

-4-202468

1012

0 2 4 6 8 10

Distance from anchor point (m)

Shea

r Fo

rce

(kN

) FEM (Incremental Method)Closed Form SolutionFEM (Total Load Method)

Fig. 4.10 Variation of shear force along anchor fixed length

188

Page 209: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

0

5

10

15

20

25

30

100 1000 10000 100000 1000000 10000000Ks (kPa/m)

Pile

hea

d se

ttle

men

t (m

m) Conventional

Proposed

Fig. 4.11 Comparison of pile head settlement from conventional FEM (C1A3) and

proposed Element (C1B)

(a) 8-noded isoparametric elements (b) proposed elements

Fig. 4.12 Deformed mesh from FEA using different types of elements

189

Page 210: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Settlement along pile shaft

-9

-8

-7

-6

-5

-4

-3

-2

-1

00 1 2 3 4 5 6 7 8

Settlement (mm)

Dep

th (m

)

Proposed element8-noded isoparameter

Fig.4.13 Deformed meshes Fig. 4.14 Pile shaft settlement

Settlement along pile shaft

-9

-8

-7

-6

-5

-4

-3

-2

-1

00 1 2 3 4 5 6 7 8 9 10

Settlement (mm)

Dep

th (m

)

Proposed element8-noded isoparameter

Fig. 4.15 Deformed meshes Fig. 4.16 Pile shaft settlement

190

Page 211: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

(a) Original mesh (b) Deformed mesh

B

A A

B

Fig. 4.17 Finite element mesh used for comparison with result from Cheung et. al.

0

10

20

30

40

50

60

0 100 200 300 400 500 600

Load (kN)

Pile

Hea

d Se

ttle

men

t (m

m)

Present studymeasuredYK Cheung et al.

Fig. 4.18 The load settlement curve for an instrumented pile

191

Page 212: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Fig. 5.1 Site plan of the slope protection work (Courtesy, L & M)

192

Page 213: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Dep

th

Anchor under study

Fig 5.2 Section A-A of the slope protection work (Unit: m).

Measured Load Displacement Curve

0

10

20

30

40

50

60

0 100 200 300 400 500 600Load (kN)

Dis

plac

emen

t (m

m)

Fig. 5.3 Field measured load vs displacement curve from pull-out test

193

Page 214: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

(a) Mesh size (b)Boundary condition

Unb

onde

d le

ngth

=12.

5m

3 m

18.5

m

Bon

ded

leng

th=3

m

Fig. 5.4 Mesh for analyses of pull-out test (Mesh 1)

194

Page 215: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

10 m

Fig. 5.5 Comparative mesh (Mesh 2) for mesh dependency study

18.5

m

195

Page 216: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

(a) Mesh 1 (b) Mesh 2 Fig.5.6 Deformed meshes at load=300kN (Displacement x 100)

In-situ stress

02468

101214161820

0 200 400 600

σ (kPa)

y co

ordi

nate

in th

e m

esh

(m)

AxialRadial

Fig.5.7. In-situ stresses

196

Page 217: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

0

10

20

30

40

50

60

70

80

0 100 200 300 400 500 600

Load (kN)D

ispl

acem

ent (

mm

)

Yielding Non-yielding

Fig. 5.8 Calculated anchor head displacement vs load curve where unbonded length was modelled using friction free bar element.

0

10

20

30

40

50

60

70

80

0 100 200 300 400 500 600

Load (kN)

Dis

plac

emen

t (m

m)

MeasuredProposed element YieldingNon-yielding

Fig. 5.9 Comparison of calculated anchor head displacement vs load curves between analyses using the friction free bar element and those using proposed anchor-interface element for anchor unbonded length.

197

Page 218: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

0

10

20

30

40

50

60

70

0 100 200 300 400 500 600

Load (kN)D

ispl

acem

ent (

mm

)

NumericalMeasured

Fig. 5.10 Comparison with in-situ pull out test Bonded length: Ks =1.8x 104 kPa/m, Unloading Ks =1.5x 104 kPa/m, c=200kPa, φ=30°. Unbonded length: Ks =6.0x 103 kPa/m, Unloading Ks =5x 103 kPa/m, c=1.0kPa, φ=10°

Typical Load Displacement Curve

0

10

20

30

40

50

60

0 100 200 300 400 500 600Load (kN)

Dis

plac

emen

t (m

m)

A

E

D C

B

O

Fig.5.11 Typical Load displacement curve from pull-out test.

198

Page 219: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Load Displacement curve

0

5

10

15

20

25

30

35

40

45

0 100 200 300 400 500 600 700

Load (kN)A

ncho

r Hea

d D

ispl

acem

ent (

mm

)

Post-yielding

O A

E

B

D C

Fig. 5.12 Simulated hysteresis loop (Load to 610kN before unloading) Bonded length: Ks =5.0x 105 kPa/m, Unloading Ks =1.8x 106 kPa/m, c=85kPa, φ=20°. Unbonded length: Ks =5.0x 104 kPa/m, Unloading Ks =6x 104 kPa/m, c=10 kPa, φ=10°

Skin friction along anchor

-200

20406080

100120140

0.0 2.0 4.0 6.0 8.0 10.0 12.0 14.0

Length (m)

Skin

fric

tion

(kPa

)

Load to 200kNUnload to 200 kN

Fig. 5.13 Friction along anchor

199

Page 220: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

0

10

20

30

40

50

60

70

0 100 200 300 400 500 600

Load (kN)D

ispl

acem

ent (

mm

)

Ko=0.5Ko=1Measured

Fig. 5.14 Influence of in-situ stress

0

10

20

30

40

50

60

70

0 100 200 300 400 500 600

Load (kN)

Dis

plac

emen

t (m

m)

E=2.5E7E=2.5E5E=2.5E3

Fig. 5.15 Influence of elastic modulus of rock

200

Page 221: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

0

10

20

30

40

50

60

70

0 100 200 300 400 500 600

Load (kN)

Dis

plac

emen

t (m

m)

E=2.0E4E=2.0E3E=2.0E2

Fig. 5.16 Influence of elastic modulus of soil

(a)Deformed shape (b) Displacement vector Fig. 5.17 Deformation of soil at low elastic modulus

201

Page 222: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

0

50

100

150

200

250

300

0 100 200 300 400 500 600

Load (kN)

Dis

plac

emen

t (m

m)

Ks=1.0E5Ks=1.0E4Ks=1.0E3

Fig. 5.18 Influence of Ks for anchor bonded length

0

10

20

30

40

50

60

70

0 100 200 300 400 500 600

Load (kN)

Dis

plac

emen

t (m

m)

Ks=1.0E4Ks=5.0E3Ks=1.0E3

Fig. 5.19 Influence of Ks for anchor unbonded length

202

Page 223: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

0

10

20

30

40

50

60

70

0 100 200 300 400 500 600

Load (kN)

Dis

plac

emen

t (m

m)

c=15kPaMeasuredc=10kPa

Fig. 5.20 Effects of c’ in unbonded length. Bonded length: Ks =1.8x 104 kPa/m, Unloading Ks =1.5x 104 kPa/m, c’=200kPa, φ’=20°. Unbonded length: Ks =7.0x 103 kPa/m, Unloading Ks =5x 103 kPa/m, c’=1 and15kPa, φ’=10°

203

Page 224: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Analysis area

Fig 6.1 Elevation view of the Texas A & M University tieback wall

Fig 6.2 Section view of the Texas A & M University tieback wall (after Briaud & Lim, 1999)

204

Page 225: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Fig 6 3

205

Soil profile

Page 226: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Fig 6.4 Initial mesh

We = 10m Be = 39m

H = 7. 5m

Db = 5. 5m

Fig 6.5 Boundary conditions

206

Page 227: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Variation of E modulus with depth

-14

-12

-10

-8

-6

-4

-2

00 10000 20000 30000

E (kPa)D

epth

(m)

Briaud & Lim (1999)Present study

Fig. 6.6 Variation of elastic modulus of soil with depth

207

Page 228: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Fig. 6.7 Installation of steel H soldier pile

Fig. 6.8 Zoomed-in view of soldier pile elements

Fig. 6.9 Excavation of 1st lift

208

Page 229: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Fig. 6.10 Installation of timber lagging for the 1st lift

Timber lagging elements

Fig. 6.11 Zoomed-in view of timber lagging and soldier pile elements

209

Page 230: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Tie-back elements

Fig. 6.12 Installation and pre-stressing of anchor row 1

Locked-in elements

Fig. 6.13 Zoomed-in view of the locked-in elements

Fig. 6.14 Excavation of 2nd lift

210

Page 231: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Fig. 6.16

Fig. 6.17

Fig. 6.15 Installation of timber lagging for the 2nd lift

Installation and pre-stressing of anchor row 2 anchor elements

Activate the locked-in elements

211

Page 232: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Fig. 6.18 Zoomed-in view of the locked-in elements

Fig. 6.19 Excavation of the 3rd lift

212

Page 233: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Fig. 6.20 Installation of timber lagging for the 3rd lift

Fig. 6.21 Analysis area used by Briaud and Lim (1999)

213

Page 234: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Fig. 6.22 Mesh configuration and boundary conditions in Briaud and Lim’s study

214

Page 235: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Fig. 6.23 Mesh configurations used in calibration study (B3D1) based on the study of Briaud and Lim (1999)

215

Page 236: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

-10

-9

-8

-7

-6

-5

-4

-3

-2

-1

0-0.05 -0.03 -0.01

Deflection (m)

Dep

th (m

)

Briaud & Lim (1999)

Measured

B3D1

-10

-9

-8

-7

-6

-5

-4

-3

-2

-1

0-0.05 -0.03 -0.01

Deflection (m)

Dep

th (m

)

Briaud & Lim (1999)MeasuredB3D2B3D1

Fig. 6.24 Pile deflection from calibration analysis Fig. 6.25 Effects of mesh configurations

216

Page 237: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

-10

-9

-8

-7

-6

-5

-4

-3

-2

-1

0-0.06 -0.04 -0.02 0

Deflection (m)D

epth

(m)

Briaud & Lim (1999)MeasuredB3D2A3D2

-10

-9

-8

-7

-6

-5

-4

-3

-2

-1

0-0.06 -0.04 -0.02 0

Deflection (m)

Dep

th (m

)

MeasuredB3D1B3D2A3D1A3D2

Fig. 6.26 Effects of element types for anchors Fig. 6.27 Summary of pile deflections from calibration analyses

217

Page 238: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

-10

-9

-8

-7

-6

-5

-4

-3

-2

-1

0-0.06 -0.04 -0.02 0

Deflection (m)

Dep

th (m

)

Briaud & Lim (1999)

Measured

A3D1

A3D3

-10

-9

-8

-7

-6

-5

-4

-3

-2

-1

0-0.06 -0.04 -0.02 0

Deflection (m)

Dep

th (m

)Briaud & Lim (1999)MeasuredmE=2200mE=2500mE=3000

Fig. 6.28 Effects of coordinates updating on numerical results Fig. 6.29 Effects of soil stiffness on pile deflections

218

Page 239: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

-10

-9

-8

-7

-6

-5

-4

-3

-2

-1

0-0.06 -0.04 -0.02 0

Deflection (m)D

epth

(m)

Briaud & Lim (1999)

Measured

Ks=1E8

Ks=1E5

-10

-9

-8

-7

-6

-5

-4

-3

-2

-1

0-0.06 -0.05 -0.04 -0.03 -0.02 -0.01 0

Deflection (m)

Dep

th (m

)

Briaud & Lim (1999)MeasuredC=50kPaC=100kPa

Fig. 6.30 Parametric study on the influence of Ks Fig. 6.31 Influence of interface cohesion

219

Page 240: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

-10

-9

-8

-7

-6

-5

-4

-3

-2

-1

0-0.04 -0.03 -0.02 -0.01 0

Deflection (m)

Dep

th (m

)

Briaud & Lim (1999)Measured3D2D

-10

-9

-8

-7

-6

-5

-4

-3

-2

-1

0-0.04 -0.03 -0.02 -0.01 0

Deflection (m)

Dep

th (m

)Measured

3D(Pile plane)

2D

3D(Anchor plane)

Fig. 6.32 Soldier pile deflection from 2D and 3D analyses Fig. 6.33 Soldier pile and wall deflection from 2Dand3D analyses

220

Page 241: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

-10

-9

-8

-7

-6

-5

-4

-3

-2

-1

0-25 -20 -15 -10 -5 0

Soldier pile deflection (mm)

Dep

th (m

)

2D

3D

-10

-9

-8

-7

-6

-5

-4

-3

-2

-1

0-0.03 -0.02 -0.01 0

Deflection (m)

Dep

th (m

)3D

2D

Fig. 6.34 Deflection profiles for 2D and 3D analyses Fig. 6.35 Deflection profiles for 2D and 3D analyses

at end of 1st lift excavation at end of stressing Row 2 anchor

221

Page 242: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Fig. 6.36 Displacement vector fields at the end of stressing the Row 1 anchor from 2D analysis using proposed element model.(Displacement x 50)

Fig. 6.37 Displacement vector fields at the end of stressing the Row 1 anchor from 3D analysis using proposed element model. (Displacement x 50)

222

Page 243: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Fig. 6.38 Deformed mesh at the end of stressing the Row 1 anchor from 2D analysis using proposed element model.(Displacement x 50)

Fig. 6. Deformed mesh at the end of stressing the Row 1 anchor from 3D analysis using proposed element model. (Displacement x 50)

223

Page 244: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Fig. 6.40 Displacement vector fields at the end of final excavation from 2D analysis using proposed element model (Displacement x 50).

Fig. 6.41 Deformed mesh at the end of final excavation from 2D analysis using proposed element model (Displacement x 50).

224

Page 245: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Fig. 6.42 Displacement vector fields at the end of final excavation from 3D analysis using proposed element model (Displacement x 50).

Fig. 6.43 Deformed mesh at the end of final excavation from 3D analysis using proposed element model (Displacement x 50).

225

Page 246: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Arching

Sold

ier p

ile

Tim

ber l

aggi

ng

Fig. 6.44 Plan view of principal stresses at end of excavation on plane Yo=-1.5

Arching

Sold

ier p

ile

Tim

ber l

aggi

ng

Fig. 6.45 Plan view of principal stresses at end of

226

Segment of anchor

excavation on plane Yo=-3

Page 247: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Fig. 6.46 Plan view of principal stresses at end of excavation on plane Yo=-4.5

Fig. 6.47 Plan view of principal stresses at end of excavation on plane Yo=-6

Fig. 6.48 Plan view of principal stresses at end of excavation on plane Yo=-7.6

Distribution of skin shear along Row1 anchor at end of stressing

-20

0

20

40

60

80

100

120

0 2 4 6 8 10

Anchor length (m)

Skin

She

ar st

ress

(kPa

)

12

3D2D

Fig. 6.49 The distribution of skin shear along Row 1 at end of stressing Row 1 anchor.

227

Page 248: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Distribution of skin shear along Row 1 anchor ( 3D)

0

20

40

60

80

100

120

140

160

0 2 4 6 8 10

Anchor length (m)

Skin

She

ar st

ress

(kPa

)

12

End of stressing Row 1End of 2nd lift excavationEnd of stressing Row 2End of excavation

Fig. 6.50 The distribution of skin shear along Row 1 at different construction stages

(3-D).

Distribution of skin shear along Row 1 anchor ( 2D)

0

20

40

60

80

100

120

140

0 2 4 6 8 10Anchor length (m)

Skin

She

ar st

ress

(kPa

)

12

End of stressing Row 1End of 2nd lift excavation

End of stressing Row 2End of excavation

Fig. 6.51 The distribution of skin shear along Row 1 at different construction stages

(2-D).

228

Page 249: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

-10

-9

-8

-7

-6

-5

-4

-3

-2

-1

0-0.06 -0.04 -0.02 0

Deflection (m)

Dep

th (m

)

Proposed ElementBar ElementBriaud & Lim (1999)Measured

Fig. 6.52 Deflection profiles from analyses using different model in 3D analyses

229

Page 250: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

-10

-9

-8

-7

-6

-5

-4

-3

-2

-1

0-0.04 -0.03 -0.02 -0.01 0

Deflection (m)

Dep

th (m

)

Proposed element

Bar element

-10

-9

-8

-7

-6

-5

-4

-3

-2

-1

0-0.04 -0.03 -0.02 -0.01 0

Deflection (m)

Dep

th (m

)

Proposed element

Bar element

Fig. 6.53 Soldier pile deflection at the end of 1st lift excavation Fig. 6.54 Soldier pile deflection at the end of stressing Row 1 anchor

230

Page 251: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

-10

-9

-8

-7

-6

-5

-4

-3

-2

-1

0-0.04 -0.03 -0.02 -0.01 0

Deflection (m)

Dep

th (m

)

Proposed element

Bar element

-10

-9

-8

-7

-6

-5

-4

-3

-2

-1

0-0.04 -0.03 -0.02 -0.01 0

Deflection (m)

Dep

th (m

)

Proposed element

Bar element

Fig. 6.55 Soldier pile deflection at the end of 2nd lift excavation Fig. 6.56 Soldier pile deflection at the end of stressing Row 2 anchor

231

Page 252: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Fig. 6.57 Displacement vector fields at the end of final excavation from 3D analysis using bar element model (Displacement x 50).

Ground surface settlement

-4

-3

-2

-1

00 10 20 30

Distance from wall (m)

Set

tlem

ent (

mm

)

40

Bar elementProposed element

Fig. 6.58 Ground settlement

232

Page 253: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Arching Arching

Sold

ier p

ile

Tim

ber l

aggi

ng

Fig. 6.59 Plan view of principal stresses from analysis using bar element on yo=-1.5

Arching

Anchor unbonded length

Sold

ier p

ile

Tim

ber l

aggi

ng

Fig. 6.60 Plan view of principal stresses from analysis using bar element on yo=-3.0

233

Page 254: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Arching

Sold

ier p

ile

Tim

ber l

aggi

ng

Segment of anchor Fig. 6.61 Plan view of principal stresses from analysis using bar element on yo=-4.5

Segment of anchor Fig. 6.62 Plan view of principal stresses from analysis using bar element on yo=-6.0

Segment of anchor

Fig. 6.63 Plan view of principal stresses from analysis using bar element on yo=-7.6

234

Page 255: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Fig. 7.1 Location map of the site (After Parnploy, 1990)

235

Page 256: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Fig. 7.2 Site plan (After Parnploy, 1990)

236

Page 257: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Fig.

7.3

Gro

und

prof

ile (A

fter P

arnp

loy,

199

0)

237

Page 258: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

57m

32m

158m

Fig. 7.4 The 2D mesh at initial stage (AX1CA)

57m

Fig. 7.5 The comparative mesh (AX1CB) for determining stand off distance

32m

180m

238

Page 259: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

-30

-25

-20

-15

-10

-5

0-150 -100 -50 0

Wall deflection (mm)

Dep

th(m

)

AX1CBAX1CA

Fig. 7.6 Comparison on wall defection for different stand-off distances

239

Page 260: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Fig. 7.7 The comparative mesh (AX1CC) for mesh dependency study

Fig. 7.8 Mesh AX1CD for mesh dependency study

Wall deflectionat the end of excavation

-25

-20

-15

-10

-5

00 10 20 30 40 50 60

Deflection (mm)

Dep

th (m

)

MESH2

MESH1

Fig. 7.9 Wall defection at the end of excavation

240

Page 261: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Fig. 7.10 3D Mesh for present study

Fig. 7.11 Uneven excavation of 3D mesh

241

Page 262: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Fig. 7.12 The 3-D mesh at the end of the final excavation

Fig. 7.13 The 3-D mesh at the end of the final excavation with fixed boundary conditions shown

242

Page 263: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Fig. 7.14 Fluid flow boundary condition at initial stage

Fig. 7.15 Fluid flow boundary after excavation of 1st lift

243

Page 264: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Fig. 7.16 Fluid flow boundary after 2nd lift excavation

Fig. 7.17 Fluid flow boundary after at the last stage of construction

244

Page 265: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

In-situ stress field

-30

-25

-20

-15

-10

-5

00 100 200 300 400

Stress (kPa)

Lev

el

Horizontal site stress

Pore water pressure

Vertical site stress

Construction sequenses

-25

-20

-15

-10

-5

0

0 50 100 150

Time (Days)D

redg

ed L

evel

(m)

200 250

Left panel (Section X2)Right Panel (Section X1)Installation of struts and anchors

Fig. 7.18 The In-situ stress profile assumed in present study Fig. 7.19 Simulated construction sequences

245

Page 266: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Fig. 7.20 The 3D mesh after introducing the overburden of river bank

Fig. 7.21 The 3D mesh after installation of anchors

246

Page 267: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Wall deflection (mm)

-25

-20

-15

-10

-5

0020406080100

Dep

th (m

)

T=12 DaysFEM

Wall deflection (mm)

-25

-20

-15

-10

-5

0020406080100

Dep

th (m

)

T=74 Days

FEM

(a) (b)

Wall deflection (mm)

-25

-20

-15

-10

-5

0020406080100

Dep

th (m

)

T=129 DaysFEM

Wall deflection (mm)

-25

-20

-15

-10

-5

0020406080100

Dep

th (m

)

T=211 DaysFEM

(c) (d) Fig. 7.22 Wall deflection profile from analysis using proposed element

247

Page 268: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Fig. 7.23 Displacement vector field (Scaled by 20 times) and excess pore water pressure contours after the pumping of water within cofferdam

Fig. 7.24 Displacement vector field (Scaled by 20 times) and excess pore water pressure contours after the excavation of top two lifts.

248

Page 269: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Fig. 7.25 Displacement (Scaled by 20 times) after excavation of top 3 lifts

Fig. 7.26 Displacement (Scaled up by 20 times) at end of construction

249

Page 270: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Fig. 7.27 The excess pore water pressure contours after introducing the overburden of river bank

Fig. 7.28 The excess pore water pressure contours after pumping water within cofferdam

250

Page 271: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Fig. 7.29 The excess pore water pressure contours after excavation of top two lifts

Fig. 7.30 The excess pore water pressure contours after excavation of top three lifts

251

Page 272: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Fig. 7.31 The excess pore water pressure contours after excavation of top 5-th lifts

Fig. 7.32 The excess pore water pressure contours at the end of analysis

252

Page 273: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Front anchors

0

0.002

0.004

0.006

0.008

0.01

0.012

150 170 190 210 230 250 270 290Time from the commencement of construction (Day)

Anc

hor

head

disp

lace

men

t (m

)

R1(Node 827)L1(Node 758)

Fig. 7.33 Development of anchor head deformation (Front column anchors)

Back anchors

0

0.002

0.004

0.006

0.008

0.01

0.012

150 170 190 210 230 250 270 290Time from the commencement of construction (Day)

Anc

hor

head

disp

lace

men

t (m

)

R1B(Node 3957)L1B(Node 3888)

Fig. 7.34 Development of anchor head deformation (Back column anchors)

253

Page 274: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Anchor R1 (Right side Row 1)

-200

204060

80100120140160

0 5 10 15 20 25

Anchor Length (m)

Skin

Fri

ctio

n (k

N/m

2 )

After stressing A4After stressing A5End of excavation

Fig. 7.35 Skin friction along anchor tendons at key construction stages

Skin friction along anchor tendon at the end of analysis

-200

20406080

100120140160

0 5 10 15 20 25

Anchor Length (m)

Skin

Fri

ctio

n (k

N/m

2 )

L1 AnchorR1 Anchor

Fig. 7.36 Comparison of skin friction along anchor tendons for different element models

254

Page 275: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

255

Wall deflection (mm)

-25

-20

-15

-10

-5

0020406080100

Dep

th (m

)

T=129 DaysBar elementProposed element

Wa

6080100

ll deflection

-25

-20

-15

-10

-5

002040

Dep

th (m

)

T=211 DaysBar elementProposed element

(a) Before pre-stressing of anchor (b) After pre-stressing of Row 1 anchors Fig. 7.37 Wall deflection profiles from analysis using different element types

Wall deflection (mm)

-25

-20

-15

-10

-5

0020406080100

Dep

th (m

)

Bar elementProposed element

(a) After excavation of lift 5 (b) At the end of the analysis Fig. 7.38 Comparison of wall deflection profiles at key construction stages (right bank)

Wall def6080100

lection (mm)

-25

-20

-15

-10

-5

002040

Dep

th (m

)

Bar elementProposed element

Page 276: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Wall deflection

-25

-20

-15

-10

-5

0-80 -60 -40 -20 0 20

Deflection (mm)

Dep

th (m

)

Stressing of A4Bar elementProposed element

Wall deflection (mm)

-25

-20

-15

-10

-5

0-60 -40 -20 0 20

Dep

th (m

)

Stressing of A5Bar elementProposed element

(a) (b)

Wall deflection (mm)

-25

-20

-15

-10

-5

0-60 -40 -20 0 20 40

Dep

th (m

)

Excavation of lift5Bar elementProposed element

Wall deflection (mm)

-25

-20

-15

-10

-5

0-60 -40 -20 0 20 40

Dep

th (m

)

End of analysisBar elementProposed element

(c) (d) Fig. 7.39 Comparison of wall deflection profiles at key construction stages (Left bank)

256

Page 277: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Fig. 7.40 Incremental displacement field by analysis using bar element (Displacement x 100 at t=211 days)

Fig. 7.41 Incremental displacement field by analysis using proposed element (Displacement x 100 at t=211 days)

257

Page 278: NUMERICAL MODELLING OF ANCHOR-SOIL INTERACTION

Ground movement behind right bank after installation of anchors

-5

0

5

10

15

20

25

30

35

5 15 25 35 45 55 65

Distance behind wall(m)

Hea

ve (m

m)

Bar elementStressing A4Stressing A5Stressing A6Proposed elementStressing A4Stressing A5Stressing A6

Fig. 7.42 Ground movement

258