Finite Element
Finite Element
geotechnical engineering
Application
A catalogue record for this book is available from the British Library
All rights, including translation, reserved. Except as permitted by the Copyright, Designs
and Patents Act 1988, no part of this publication may be reproduced, stored in a retrieval
system or transmitted in any form or by any means, electronic, mechanical, photocopying or
otherwise, without the prior written permission of the Publishing Director, Thomas Telford
Publishing, Thomas Telford Ltd, 1 Heron Quay, London E14 4JD.
This book is published on the understanding that the author are solely responsible for the
statements made and opinions expressed in it and that its publication does not necessarily
imply that such statements andlor opinions are or reflect the views or opinions of the
publishers. While every effort has been made to ensure that the statements made and the
opinions expressed in this publication provide a safe and accurate guide, no liability or
responsibility can be accepted in this respect by the authors or publishers.
Contents
Preface
Authorship
xvi
Acknowledgements
xvi
1.
2.
Tunnels
2.1
Synopsis
2.2
Introduction
2.3
Tunnel construction
2.3.1
2.3.2
2.3.3
2.4
2.5
2.6
2.7
2.8
3.
Earth
3.1
3.2
3.3
3.4
Introduction
Open faced shield tunnelling
Tunnel Boring Machines (TBM), including slurry
shields and Earth Pressure Balance (EPB) tunnelling
The sprayed concrete lining (SCL) method
2.3.4
Ground response to tunnel construction
2.3.5
Simulation of the construction process
2.4.1
Introduction
2.4.2
Setting up the initial conditions
2.4.3
Important boundary conditions
2.4.4
Modelling tunnel excavation
2.4.5
Modelling the tunnel lining
Modelling time dependent behaviour
2.5.1
Introduction
Setting up the initial conditions
2.5.2
2.5.3
Hydraulic boundary conditions
2.5.4
Permeability models
A parametric study of the effect of permeable and
2.5.5
impermeable tunnel linings
Choice of soil model
2.6.1
Introduction
Results from a parametric study
2.6.2
Devices for improving the surface settlement
2.6.3
prediction
Interaction analysis
2.7.1
The influence of building stiffness on tunnel-induced
ground movements
The Treasury building - a case study
2.7.2
2.7.3
Twin tunnel interaction
Summary
retaining structures
Synopsis
Introduction
Types of retaining structure
3.3.1
Introduction
3.3.2
Gravity walls
ReinforcedJanchored earth wall
3.3.3
3.3.4
Embedded walls
General considerations
3.4.1
Introduction
3.4.2
Symmetry
3.4.3
Geometry of the finite element model
3.4.4
Support systems
Contents 1 iii
3.4.5
84
84
85
88
88
88
89
91
91
91
92
93
93
93
94
95
96
96
99
103
103
104
104
104
105
106
107
107
109
111
111
112
112
112
113
114
115
115
116
1 18
119
122
123
4.
Cut slopes
4.1
Synopsis
4.2
Introduction
4.3
'Non-softening ' analyses
4.3.1
Introduction
Cut slopes in stiff 'non-softening' clay
4.3.2
4.3.2.1 Introduction
4.3.2.2
Soil parameters
4.3.2.3
Finite element analyses
4.3.2.4
Results of analyses
Cut slopes in soft clay
4.3.3
4.3.3.1
Introduction
4.3.3.2 Soil parameters
4.3.3.3 Finite element analyses
4.3.3.4 Results of analyses
4.4
Progressive failure
4.5
'Softening' analyses
introduction
4.5.1
4.5.2
Choice of constitutive model
4.5.3
Implications for convergence
Cut slopes in London Clay
4.5.4
4.5.4.1
Introduction
4.5.4.2 Soil parameters
4.5.4.3 Finite element analyses
Results of a typical analysis
4.5.4.4
4.5.4.5 Effect of coefficient of earth
pressure at rest
Effect of surface boundary suction
4.5.4.6
4.5.4.7 Effect of slope geometry
4.5.4.8 Effect of surface cracking
4.5.4.9 Effect of subsequent changes to slope
geometry
4.5.4.10 Further discussion
Construction of cut slope under water
4.6
4.7
Summary
5.
Embankments
5.1
Synopsis
5.2
Introduction
Finite element analysis of rockfill dams
5.3
5.3.1
Introduction
5.3.2
Typical stress paths
5.3.3
Choice of constitutive models
5.3.3.1 Linear elastic analysis
Contents 1 v
5.4
5.5
5.5.8
5.6
6.
Summary
Shallow foundations
6.1
Synopsis
6.2
Introduction
6.3
Foundation types
6.3.1
Surface foundations
6.3.2
Shallow foundations
6.4
Choice of soil model
Finite element analysis of surface foundations
6.5
6.5.1
Introduction
6.5.2
Flexible foundations
6.5.3
Rigid foundations
6.5.4
Examples of vertical loading
6.5.4.1 Introduction
Strip footings on undrained clay
6.5.4.2
6.5.4.3 Effect of footing shape on the bearing
capacity of undrained clay
6.5.4.4 Strip footings on weightless drained soil
6.5.4.5 Strip footings on a drained soil
6.5.4.6 Circular footings on a weightless drained
soil
6.5.4.7 Circular footings on a drained soil
Undrained bearing capacity of non-homogeneous
6.5.5
clay
6.5.5.1 Introduction
6.5.5.2 Constitutive model
6.5.5.3 Geometry and boundary conditions
6.5.5.4 Failure mechanisms
Undrained bearing capacity of pre-loaded strip
6.5.6
foundations on clay
6.5.6.1 Introduction
6.5.6.2 Constitutive model
Contents / vii
6.6
6.7
7.
Deep
7.1
7.2
7.3
7.4
7.5
7.6
8.
Benchmarking
8.1
Synopsis
8.2 Definitions
8.3
Introduction
Causes of errors in computer calculations
8.4
8.5
Consequences of errors
8.5
Developers and users
8.6.1
Developers
Contents I ix
8.6.2
Users
Techniques used to check computer calculations
Benchmarking
8.8.1
General
8.8.2
Standard benchmarks
8.8.3
Non-standard benchmarks
8.9
The INTERCLAY I1 project
8.10 Examples of benchmark problems - Part I
8.10.1
General
Example l: Analyses of an ideal triaxial test
8.10.2
Example 2: Analysis of a thick cylinder
8.10.3
Example 3: Analyses of an advancing tunnel
8.10.4
heading
Example 4: Analysis of a shallow waste disposal
8.10.5
Example 5: Simplified analysis of a shallow waste
8.10.6
8.1 1 Examples of benchmark problems - Part II
(German Society for Geotechnics benchmarking exercise)
8.1 1.1
Background
8.1 1.2
Example 6: Construction of a tunnel
8.1 1.3
Example 7: Deep excavation
8.1 1.4
General comments
8.12 Summary
Appendix VIII.1 Specification for Example 1: Analyses of an
idealised triaxial test
VIII.1 .l Geometry
VIII. 1.2 Material properties and initial stress conditions
VIII. 1.3 Loading conditions
Appendix VIII.2 Specification for Example 2: Analysis of a thick
cylinder
VIII.2.1 Geometry
VIII.2.2 Material properties
VIII.2.3 Loading conditions
Appendix VIII.3 Specification for Example 3: Analysis of an
advancing tunnel heading
VIII.3.1 Geometry
VIII.3.2 Material properties
VIII.3.3 Loading conditions
Appendix VIIIA Specification for Example 4: Analysis of a shallow
waste disposal
VIII.4.1 Geometry
VI11.4.2 Material properties
VIII.4.3 Loading conditions
Appendix VIII.5 Specification for Example 5: Simplified analysis
of a shallow waste disposal
8.7
8.8
VIII.5. I
VIII.5.2
VIII.5.3
VII1.5.4
Appendix VIII.6
VIII.6.1
VIII.6.2
Appendix VIII.7
VI11.7. I
VIII.7.2
VIII.7.3
9.
Geometry
36 1
Material properties
36 1
Loading conditions
36 1
Additional boundary conditions
362
Specification for Example 6: Construction of a tunnel 362
Geometry
362
Material properties
362
Specification for Example 7: Deep excavation
362
Geometry
362
Material properties
362
Construction stages
363
396
List of symbols
410
Index
4 15
Preface
While the finite element method has been used in many fields of engineering
practice for over thirty years, it is only relatively recently that it has begun to be
widely used for analysing geotechnical problems. This is probably because there
are many complex issues which are specific to geotechnical engineering and which
have only been resolved relatively recently. Perhaps this explains why there are
few books which cover the application of the finite element method to geotechnical
engineering.
For over twenty years we, at Imperial College, have been working at the
leading edge of the application of the finite element method to the analysis of
practical geotechnical problems. Consequently, we have gained enormous
experience of this type of work and have shown that, when properly used, this
method can produce realistic results which are of value to practical engineering
problems. Because we have written all our own computer code, we also have an
in-depth understanding of the relevant theory.
Based on this experience we believe that, to perform useful geotechnical finite
element analysis, an engineer requires specialist knowledge in a range of subjects.
Firstly, a sound understanding of soil mechanics and finite element theory is
required. Secondly, an in-depth understanding and appreciation of the limitations
of the various constitutive models that are currently available is needed. Lastly,
users must be fully conversant with the manner in which the software they are
using works. Unfortunately, it is not easy for a geotechnical engineer to gain all
these skills, as it is vary rare for all of them to be part of a single undergraduate or
postgraduate degree course. It is perhaps, therefore, not surprising that many
engineers, who carry out such analyses andfor use the results from such analyses,
are not aware of the potential restrictions and pitfalls involved.
This problem was highlighted four years ago when we gave a four day course
on numerical analysis in geotechnicalengineering.Although the course was a great
success, attracting many participants from both industry and academia, it did
highlight the difficulty that engineers have in obtaining the necessary skills
required to perform good numerical analysis. In fact, it was the delegates on this
course who urged us, and provided the inspiration, to write this book.
The overall objective of the book is to provide the reader with an insight into
the use of the finite element method in geotechnical engineering. More specific
aims are:
element analysis;
To describe some of the more popular constitutive models currently available
and explore their strengths and weaknesses;
To provide sufficient information so that readers can assess and compare the
capabilities of available commercial software;
To provide sufficient information so that readers can make judgements as to the
credibility of numerical results that they may obtain, or review, in the future;
To show, by means of practical examples, the restrictions, pitfalls, advantages
and disadvantages of numerical analysis.
The book is primarily aimed at users of commercial finite element software both
in industry and in academia. However, it will also be of use to students in their
final years of an undergraduate course, or those on a postgraduate course in
geotechnical engineering. A prime objective has been to present the material in the
simplest possible way and in manner understandable to most engineers.
Consequently, we have refrained from using tensor notation and have presented all
theory in terms of conventional matrix algebra.
When we first considered writing this book, it became clear that we could not
cover all aspects of numerical analysis relevant to geotechnical engineering. We
reached this conclusion for two reasons. Firstly, the subject area is so vast that to
adequately cover it would take many volumes and, secondly, we did not have
experience with all the different aspects. Consequently, we decided only to include
material which we felt we had adequate experience of and that was useful to a
practising engineer. As a result we have concentrated on static behaviour and have
not considered dynamic effects. Even so, we soon found that the material we
wished to include would not sensibly fit into a single volume. The material has
therefore been divided into theory and application, each presented in a separate
volume.
Volume 1 concentrates on the theory behind the finite element method and on
the various constitutive models currently available. This is essential reading for any
user of a finite element package as it clearly outlines the assumptions and
limitations involved. Volume 2 concentrates on the application of the method to
real geotechnical problems, highlighting how the method can be applied, its
advantages and disadvantages, and some of the pitfalls. This is also essential
reading for a user of a software package and for any engineer who is
commissioning andlor reviewing the results of finite element analyses.
Volume 1 of this book consists of twelve chapters. Chapter I considers the
general requirements of any form of geotechnical analysis and provides a
framework for assessing the relevant merits of the different methods of analysis
currently used in geotechnical design. This enables the reader to gain an insight
into the potential advantage of numerical analysis over the more 'conventional'
approaches currently in use. The basic finite element theory for linear material
behaviour is described in Chapter 2. Emphasis is placed on highlighting the
Preface I xiii
Preface / xv
All the numerical examples presented in both this volume and Volume 1 of this
book have been obtained using the Authors' own computer code. This software is
not available commercially and therefore the results presented are unbiased. As
commercial software has not beenused, the reader must consider what implications
the results may have on the use of such software.
London
March 200 1
David M. Poas
Lidija Zdravkovic
Authorship
This volume has been edited and much of its content written by David Potts and
Lidija Zdravkovic. Several of the chapters involve contributions from colleagues
at Imperial College and the Geotechnical Consulting Group (GCG). In particular:
Dr Trevor Addenbrooke (Imperial College) wrote a large part of Chapter 2
(Tunnels);
Mr Kelvin Higgins (GCG) wrote Chapter 8 (Benchmarking) and contributed to
Chapter 3 (Earth retaining structures);
Dr NebojSa KovateviC (GCG) wrote large parts of Chapter 4 (Cut slopes) and
Chapter 5 (Embankments).
Acknowledgements
We would like to acknowledge our colleagues Professors John Burland and Peter
Vaughan and all the past and present research students at Imperial College, without
whose interest and involvement this book would not have been possible. In
particular we would like to acknowledge Dr Dennis Ganendra whose PhD work on
pile groups (supervised by David Potts) forms part of Chapter 7 (Deep
foundations).
David Potts
Lidija Zdravkovic
1.
1. l
Synopsis
1.2
Introduction
Since none of the currently available soil constitutive models can reproduce all of
the aspects of real soil behaviour, it is necessary to decide which soil features
govern the behaviour of a particular geotechnical problem (e.g. stiffness,
deformation, strength, dilation, anisotropy, etc.) and choose a constitutive model
that can best capture these features. Another factor that governs the choice of soil
models for finite element analysis is the availability of appropriate soil data from
which to derive the necessary model parameters. This often limits the use of
sophisticated soil models in practice, because their parameters cannot be readily
derived from standard laboratory or field tests.
The aim of this chapter is to give a brief description of standard and special
laboratory and field tests and to review the soil parameters that can be derived from
each of them. Due to restrictions on space it is not possible to give a detailed
account of how the parameters are deduced from the raw laboratory and/or field
data. For this information the reader is referred to the specialist texts on this
subject. It is recognised that a good site investigation should combine the strengths
of both laboratory and field testing.
1.3
Laboratory tests
1.3.1
Introduction
A laboratory investigation is a key feature of almost all geotechnical projects. The
activities in the laboratory can be divided into several groups:
- Soil profiling, which involves soil fabric studies, index tests (e.g. water content,
grading, Atterberg limits), chemical and organic content, mineralogy, etc.;
- Derivation of parameters for empirical design methods (e.g. CBR for road
pavements);
The pre-consolidation e
pressure is usually associated
Virgin consoi~dat~on
with the maximum previous
vertical effective stress that the
sample has ever been
he
subjected to. Hence the
original overconsolidation
ratio of the sample, OCR, can
be calculated by dividing(onl),
by the vertical effective stress
Figure 1.2: Typical results from an
existing in the field at the
oedometer test on clay soil
location the sample was taken
from. It should be noted that for struchred soils this approach may lead to an
overestimate of OCR, as the process of ageing has the effect of increasing the stress
associated with point 'b'.
To be able to use the oedometer data in general stress space, its is necessary to
express the gradients of the compression and swelling lines in terms of invariants.
By knowing the ratio of the radial to axial effective stress in one dimensional
compression (i.e. K,, = orf/o,'), it is possible to estimate the mean effective stress
p' (=(otI1+2K,,o,,')/3) and replot compression and swelling behaviour in terms of
specific volume, v (=l+e), and mean effective stress, p1(i.e. v-lnp' diagram). The
gradient of the virgin compression line is then usually denoted as A and is
calculated as:
The slope of the swelling line, K, in the same diagram is calculated in a similar
manner. These two parameters, A and K, are essential for critical state type models,
such as modified Cam clay, bounding surface plasticity and bubble models (see
Chapters 7 and 8 of Volume 1).
The difficulty is estimating the value of K, which is not measured in the
standard oedometer test. When the soil is on the virgin compression line it is
termed normally consolidated and it is often assumed that K,=K,N"=(l -sinpl) (i.e.
Jaky's formula), where p' is the angle of shearing resistance. If v' is known, then
K,, and hence p' can be estimated. As K T is approximately constant along the
virgin compression line, then if this line is straight in e-logo,' space it will also be
straight in v-lnp' space. In fact, it can be easily shown that CC=2.3R.However, on
a swelling line K,, is not constant but increases as the sample is unloaded. While
":,
varies, there is no
there are some empirical equations describing how K
universally accepted expression. Consequently it is difficult to calculate p' on a
swelling line. In addition, if K,OCvariesand the swelling line is straight in e-logo,,'
space, then it will not be straight in v-lnp' space and vice versa. Clearly the
determination of K from an oedometer test is therefore difficult and considerable
judgement is required.
The data from the oedometer test can also be used to calculate the constrained
modulus as:
If the soil is assumed to be isotropic elastic, this can be equated to the one
dimensional modulus to give:
= 0, - p j
0;= 0, - pj
(1.6)
- If o;, and ar are increased together such that the radial strain c,=O, the sample
will deform in one dimensional compression, similar to oedometer conditions.
The advantage over the oedometer test is that, by measuring pore water
pressure and radial stress, effective stresses can be calculated and the
coefficient of earth pressure at rest, K?, estimated as:
In a similar manner, if rr, and a, are reduced together such that zr=O, the
coefficient of earth pressure in overconsolidation, K?, can be estimated for
different OCRs. The results of such one dimensional compression and swelling
tests can be plotted in a e-loga,' diagram and parameters C, and C, estimated
as in Equation (1.1). The total volumetric strain, E,, in both cases equals the
axial strain, E,. It should be noted that these values can only be related to insitu conditions if the OCR is known. An alternative procedure for estimating
the value of K,, from undisturbed sample is discussed subsequently.
If the loading frame is locked such that it is not in contact with the top platen
of the sample and the sample is loaded only with an all-around cell pressure
(i.e. o,=or), it is deforming under isotropic compression. If the all-around cell
pressure is reduced, the sample is said to undergo isotropic swelling. The
results of such tests can be plotted in a v-lnp' diagram and used to obtain values
for the parameters A and K, see Equation (1.2).
After initial isotropic or anisotropic compression to a certain stress level,
samples can be subjected to either drained or undrained shearing. In this case
the total radial stress is held constant, while the axial stress is either increased
or decreased. Ifo;>o, the sample is undergoing triaxial compression and in this
case a,,=a, and a,=ae=q=a,. The parameter b (=(a,-a3)l(o,-a,)), which
accounts for the effect of the intermediateprincipal stress, is therefore zero. On
the other hand, if the sample is axially unloaded, such that au<ar,it undergoes
triaxial extension and in this case a,=q and o;=a,=a,=q, resulting in b=I .O. In
Volume 1 of this book the Lode's angle B was used to express the relative
magnitude of the intermediate principal stress. Using Equation (5.3) of Volume
1 it can be shown that 8=tan-,[(2b- 1)1J3] and consequently in triaxial
compression b=O and I%-30" and in triaxial extension b=l and e 3 0 " .
E.
\\\
E,
Ago
The ability to measure the radial strain to a very high resolution is essential for this
parameter to be estimated.
I-lowever, instead of using parameters E' and p, soil elasticity can be expressed
in terms of shear modulus G and bulk modulus K, see Section 5.5 of Volume 1.
Both can be calculated from triaxial compression or extension tests, although K
only from a drained test:
where:
JS
and
where:
where A , accounts for the change in pore water pressure due to the reduction in
deviatoric stress which occurs during sampling in the field. A value of A,=1/3
corresponds to an isotropic elastic material, whereas a stiff clay (e.g. London clay)
typically has a value of approximately 112; 0,' is the original vertical effective
stress in the sample in-situ and can be estimated knowing the depth of the sample,
the bulk unit weights of the overlying materials and the pore water pressure.
Accurate estimates of K,, from this approach require that the sample remains
undrained at all times. On setting up the sample in the triaxial cell care must be
exercised so that it does not come into contact with water. This can be difficult as
it is usual practice to saturate the porous stones at the top and bottom ofthe sample
and consequently special testing techniques are required, see Burland and
are installed in the sample it is also possible to obtain an estimate for the fifth
parameter.
While clearly an advantage over the conventional triaxial equipment, it is not
easy to obtain and set up the samples. Consequently true triaxial apparatus are
rarely used for commercial testing. Only a limited number of apparatus exist and
these are mainly located in universities.
'1
($lf
''
several reversals of loading may be
necessary. Residual strength can be p,.,
very important in some boundary value
problems involving high plasticity
clays, see Chapter 4. Because clay I-'
particles are elongated and, in their
natural condition, oriented in some
D~splacement
structure, after large deformations they
can become aligned with the direction Figure 1.13: Typical shear stressof shearing. such a situation could displacement diagram from direct
shear box test
occur, for example, during a landslide.
Low plasticity clays and granular materials have a less pronounced drop from
critical state to residual strength.
The direct shear box apparatus is also used for interface shear testing. In this
situation soil is sheared against some other material (e.g. concrete, steel, etc.) in
order to examine the angle of friction, 6, at the interface of the two materials. This
parameter is used, for example, in pile or retaining wall analysis, where it is
necessary to know the maximum value of the angle of friction between the pile or
wall material (concrete or steel) and soil (Jardine and Chow (1996), Day and Potts
(1998)). For such investigations the interface between the soil and the material
against which it is to be sheared is positioned at the split between the upper and
lower halves of the shear box.
I m"
'~
Top platen
tom
Bonom platen
u e
pressure
.l!
4:
2.8
Summary
Tunnels 1 73
5.
6.
7.
9.
alternative will alter the long term pore water pressure regime, for the same
hydraulic boundary conditions. This will alter the ground response during
consolidation and swelling.
The intermediate and long term behaviour is governed by many factors. In
particular, whether the tunnel acts as a drain or is impermeable, and whether the
initial pore water pressure profile is close to hydrostatic or not. It is important
to be aware of the dependencies, and so to view any prediction of intermediate
and long term behaviour with a critical eye.
It is important to select constitutive models capable of reproducing field
behaviour. For example, in a situation where pre-yield behaviour dominates the
ground response, it is essential to model the nonlinear elasticity at small strains.
Devices for improving settlement predictions can be developed. These
questions must be asked: What is the influence of this adjustment on the soil
behaviour? What are the knock on effects? For example, if one is adopting a
device to match a surface settlement profile, how does this alter any prediction
of sub-surface movement, the pore pressure response, or the lining stresses and
deformations.
The finite element method can be used to quickly assess the impact of different
influences on tunnelling-induced ground movements. Parametric studies can
prove extremely useful in the development of design charts and interaction
diagrams.
One of the great benefits of numerical analysis to the tunnel engineer is that an
analysis can incorporate adjacent influences. For example existing surface
structures, or existing tunnels. It is also possible to reproduce the effects of
compensation grouting to protect surface structures during tunnelling projects.
This chapter has demonstrated the power of the finite element method in this
respect.
Synopsis
This chapter discusses the analysis of earth retaining structures of various forms.
It draws heavily on issues that have to be considered when analysing real
structures. After a brief description of the main types of retaining structures in
current use, general considerations, such as choice of constitutive model,
construction method, ground water control and support systems, are discussed.
Attention is then focussed on the specific analysis of gravity, embedded and
reinforcedlanchored retaining walls.
3.2
Introduction
The purpose of an earth retaining structure is, generally, to withstand the lateral
forces exerted by a vertical or near vertical surface in natural ground or fill. The
structural system usually includes a wall, which may be supported by other
structural members such as props, floor slabs, ground anchors or reinforcing strips.
Alternatively, or additionally, the wall may be supported by ground at its base or
into which it penetrates. In most situations the soil provides both the activating and
resisting forces, with the wall and its structural support providing a transfer
mechanism.
The design engineer must assess the forces imposed on the wall and other
structural members, and the likely movements of both the structure and retained
material. Usually these have to be determined under working and ultimate load
conditions, see Chapter 1 of Volume 1 of this book. In addition, estimates of the
magnitude and extent of potential ground movements arising from construction of
the structure, both in the short term and in the long term as drainage within the
ground occurs, are required. This may be because of the effect construction may
have on existing, or planned, services or structures (buildings, tunnels, foundations,
etc.) in close proximity. Potential damage could occur which has to be assessed and
methods of construction considered which minimise these effects.
Design of retaining walls has traditionally been carried out using simplified
methods of analysis (e.g. limit equilibrium, stress fields) or empirical approaches.
Simplified methods have been developed for free standing gravity walls, embedded
cantilever walls, or embedded walls with a single prop or anchor. Some of these
are described in BS 8002 (1994) and Padfield and Mair (1984). Because of the
statically indeterminate nature of multiple propped (or anchored) walls, these have
often been dealt with using empirical approaches such as those suggested by Peck
(1 969). Simplified methods of analysis are also available for reinforcedlanchored
earth, for example see BS 8006 (1995).
Because all of these traditional design methods are based on simplified analysis
or empirical rules, they cannot, and do not, provide the engineer with all the
desired design information. In particular, they often only provide very limited
indications of soil movements and no information on the interaction with adjacent
structures.
The introduction of inexpensive, but sophisticated, computer hardware and
numerical software has resulted in considerable advances in the analysis and design
of retaining structures over the past ten years. Much progress has been made in
attempting to model the behaviour of retaining structures in service and to
investigate the mechanisms of soil-structure interaction. This chapter will review
some of these advances and discuss some of the important issues that must be
addressed when performing numerical analysis of earth retaining structures. It
begins by describing the different types of retaining structure and the general issues
which must be considered before starting an analysis. It then goes on to consider
the three main categories of retaining structures (i.e. gravity, embedded and
reinforcedlanchored earth walls) in more detail.
3.3
3.3.1
Introduction
The complexity and uncertainty involved in design and analysis increase with the
degree of soil-structure interaction and thus depend on the type of retaining
structure to be employed. It is therefore appropriate to categorise the types of
retaining structure on the basis of the soil structure interaction problems that arise
in design. In Figure 3.1 the main wall
types are shown in order of increasing
complexity o f soil-structure
interaction.
Gravity
Counterfort
Cantilever
Reverse cantilever
Soil nails
3.3.3
Reinforcedlanchored
earth walls
Embedded cantilever
Propped cantilever
Reinforced earth walls and walls
involving soil nails essentially act as
large gravity walls. However, their
internal stability relies on a complex
interaction between the soil and the
reinforcingelements (Or nails]. These
Multi-propped
walls generally have non-structural
level
membrane facing units which are
intended to prevent erosion, or are
purely aesthetic. The membrane has
to be designed to resist any bending
moments and forces that occur, but
this is not the primary method of earth
Multi-anchored
retention. The resistance to ground
movement is provided by soil nails,
Figure 3- 1: (...cant) Types of
anchors and ties of various forms (not
retaining structure
ground anchors) or strips of metal or
geo-fabric.
increases with the number of levels of props andlor anchors and hence the
structural redundancy.
3.4
General considerations
3.4.1 Introduction
Before starting any numerical analysis it is important to address a number of issues
to ensure that the most appropriate methods of modelling the soil and structure are
used. It is also important that the correct boundary conditions (e.g. displacements,
pore water pressures, loads, etc.) are applied in any analysis. This section briefly
outlines some of the more important issues. A number of these are discussed in
greater detail in subsequent sections.
3.4.2 Symmetry
In reality all geotechnical problems involving retaining structures are three
dimensional and, ideally, three dimensional analyses, fully representing the
structure's geometry, loading conditions and variations in ground conditions across
the site, should be undertaken. With current computer hardware this is not a
practical proposition for all, but a number of very limited and extremely simple
cases. To analyse any structure it is therefore necessary to make a number of
simplifying assumptions. Most commonly two dimensional plane strain or axisymmetric analyses are undertaken, see Chapter 1 of Volume 1 of this book.
For two dimensional and axi-symmetric
(E
analysis the assumption is frequently made
I
that there is an axis of symmetry about the
I
centre line of an excavation and that only a
'half section' needs to be modelled. In the
case of a three dimensional analysis two
planes of symmetry are often assumed and a
'quarter section' is considered. For example,
Figure 3.2 shows a plane strain excavation
supported by two parallel walls propped near
to the ground surface. If there is symmetry
I
about the vertical line passing through the
Of axicentre of the excavation, it is only necessary Figure 3.2:
symmetric
geometry
to analyse half the problem, either that half of
the problem to the right of the plane of symmetry or that half to the left. This
clearly reduces the size of the problem and the number of finite elements needed
to represent it. However, for such an analysis to be truly representative there must
be complete symmetry about the centre line of the excavation. This symmetry
includes geometry, construction sequence, soil properties and ground conditions.
In practice it is rarely the case that all of these have symmetry and therefore
analyses using a 'half section' usually imply further approximations.
[-:
Obviously, no analysis can model geometry in detail, or for that matter every
construction activity, but this example illustrates the care that must be taken when
making simplifying assumptions of any form.
In the first set of analyses the soil was represented by a linear elastic MohrCoulomb constitutive model in which the soil stiffness increased with depth. The
parameters used are shown on Figure 3.5. The procedure and boundary conditions
for all the analyses were the same. Only the dimensions of the mesh varied with
boundaries of differing depths and distances from the rear of the wall. This set of
analyses was then repeated modelling the soil with the same strength parameters,
but with the linear elasticity replaced by the small strain stiffness model described
in Section 5.7.5 of Volume 1. The parameters for the small strain stiffness model
are given in Table 3.1. In both sets of analyses the soil parameters are typical of the
same stiff clay and are therefore consistent with each other. All analyses simulated
undrained excavation to a depth of 9.3m, with a rigid prop at the top of the wall.
The surface settlement troughs
lwall
behind the wall predicted by analyses
with a lateral extent of lOOm (five
times the half width of the
excavation) and with depths of 36m,
52m and lOOm (1.8, 2.6 and 5 times
the half width of the excavation) are
presented in Figure 3.7. The results of
the analyses using linear elastic preyield behaviour of the soil shown in
- coordinate from the wall to the far side of the mesh (m)
Figure 3.7a indicate that the depth of
a) Linear elastic-plastic model
the mesh has a significant effect on
1""
the predictions. In contrast, the
l0
analyses with the pre-yield small
strain soil behaviour show a much
smaller dependency on the mesh Q
21 0
depth (Figure 3.7b).
The surface settlement troughs
behind the wall predicted by analyses g
with a depth of lOOm and with lateral m$ .'"
extents if loom, 180m and 340m (5,
,,
20
30
40
50
70
80
90
IW
9 and 17 times the half width of the
x - coordinate born the wall to the far side of the mesh (m)
excavation) are presented in Figure
b) Small strain stiffness - ~lasticmodel
3.8. These predictions indicate ;hat
for the analyses performed using the Figure 3.7: Effect of the depth of
small strain stiffness model, see the mesh on the surface settlement
trough
Figure 3.8b, there is little difference
between the surface settlement troughs predicted by the analyses using meshes
with 180m and 340m lateral extent. However, for the analyses employing linear
elastic behaviour pre-yield, see Figure 3.8a, but with stiffness increasing with
depth, the predicted surface settlement troughs differ in all three analyses. These
results therefore indicate that if the small strain stiffness model is being used to
represent elastic behaviour, the mesh with a lateral extent of 180m is sufficient.
However, if the elastic behaviour is linear then a much wider mesh must be used.
X
--
M)
M)
HZ
and not absolute movement and it is important to draw this distinction. This was
illustrated by the example given in section 3.4.2 above, the roof slab influenced the
differential movement of the walls relative to each other, but did not prevent the
sway of the structure.
Alternative ways to model the different types of support connection, shown in
Figure 3.1 1, and to model props and ground anchors, are discussed in Chapter 9.
Wall
Shear = 0
=0
Suppo~t
Support
Axial thrust + 0
Axial thrust + 0
SIMPLE
PIN-JOINTED
Moment + 0
Axial thrust + 0
FULL MOMENT
3.4.5.2 Soil
Regarding the properties of the soil, it is important to use realistic constitutive
models to represent the soil's behaviour. Appropriate methods of modelling soil
behaviour are discussed in detail in Chapters 5 to 8 of Volume 1, and obtaining the
relevant soil properties has been discussed in Chapter 1 of this book.
Two categories of soil can be identified, those soils which are used as backfill
material, and those that are found in-situ. While the latter category of soils is
present in all retaining structure problems, the former category is only relevant to
a limited set of structures. The significance of the behaviour of each category
depends on the type of retaining structure. For example, backfill will have a
dominant effect on gravity and reinforced earth walls, whereas in-situ soils will
dominate the behaviour of excavations supported by embedded and nailed walls.
Soils used for backfill purposes are usually free draining sands or gravels.
However, occasionally clays are used. These soils are often modelled using either
an elastic Mohr-Coulomb model, or a model of the Lade type (see Chapters 7 and
8 of Volume 1). If the former constitutive model is used, the elastic part of the
model should preferable benonlinear, with both bulk and shear stiffness dependent
on stress and strain level (see Chapter 5 of Volume 1).
In-situ soils should be modelled using the most appropriate constitutive model
that is available and that can be justified from the site investigation data. Ideally,
this model should account for both nonlinearity at small strains and soil plasticity.
The former being required to enable realistic predictions of displacements and the
latter to limit the magnitudes of the active and passive earth pressures.
In cases where a relatively large amount of field data has been collected
concerning movements around retained excavations, such as in London (e.g.
Burland and Kalra (1986)), empirical correlations have been derived for soil
stiffness. These are often obtained by back analysis. However, these relationships
are usually linear elastic and are based on observations of movements made around
deep excavations during construction and for only a very limited period after
completion. Consequently, such relationships are not necessarily relevant to other
ground conditions, the construction of other types of retaining structure or to long
term conditions.
3.4.6
Initial ground conditions
3.4.6.1 General
With the exception of embankments retained by reinforced earth wails or gravity
walls, there is usually some excavation associated with the construction of earth
retaining structures. Because the loading of the structure depends, in part, upon
stress changes associated with this process, it is important that these stresses are
estimated reasonably accurately.
Even if there is no excavation associated with construction of the structure, the
initial stresses in the ground are important. For a simple gravity structure
constructed on the ground surface the performance of the foundation, and
consequently the structure, will be influenced by these stresses.
It is important to know the state of stress in the ground if any form of nonlinear
constitutive model is being used to model soil behaviour. This includes linear
elastic perfectly plastic models.
3.4.6.2 'Greenfield' conditions
'Greenfield' conditions are the un-modified initial stresses that exist within the
ground, arising from the various geological processes that it has been subjected to.
These stresses have not been modified by construction or any other man made
activity. The earth retaining structure may have to resist these stresses either in
whole or in part (some stress may be partially relieved before the structure is
formed).
The geological processes that the soil has been subjected to will determine the
ratio of the horizontal to vertical effective stress (i.e. the value of K,,). This ratio
can have an enormous influence on the soil behaviour and consequently the
performance of any structure within it, as demonstrated by Potts and Fourie (1 986).
Consider an element of soil with an effective vertical stress of 400kPa. Points
A, B and C on Figure 3.16 show the position in effective stress space for K,, values
of 1.O, 1.5 and 2.0 respectively. The stress paths shown on this figure are those
associated with a reduction of effective vertical stress under one dimensional
conditions.
3.4.7
Construction method and programme
3.4.7.1 General
Consideration must be given to how the structure is to be constructed (temporary
as well as permanent works) and the time taken to construct it. Different sequences
of construction mean that soil may experience differing stress paths, yield may
occur in soil elements at differing times or be suppressed completely. As a
consequence additional forces, such as swelling pressures, might be imposed on
the structure. Further, if construction is slow some drainage may occur within the
ground, which might reduce long term effects. However, if significant drainage
does occur there may be some implications for the design of temporary works.
10.0
2.0 1.6
11.9
3
-800 -400
To make an assessment of
Figure 3.22: Bending moments and
time related movements,
analysis need to be time
displacements for the central wall
related and therefore an
analysis with fully coupled consolidation is generally required, see Chapter 10 of
Volume 1. For such analysis it is necessary to specify pore water pressure andlor
3.5
3.5.1
Gravity walls
Introduction
When the retained height is not large, typically less then 5m, soil movements are
not important and construction space is not limited, gravity type retaining walls are
often used. There is a variety of forms of gravity wall and some examples are
given in Figure 3.1. To maintain stability this category of retaining walls relies on
the weight of the wall (and any enclosed soil) and the shear resistance between the
base of the wall and underlying soil to resist the earth pressures generated by the
retained ground.
During design these walls
should be checked for overall
stability, sliding, overturning and
bearing capacity failure, see
Figure 3.24. As this type of
structure is not usually used where
b) Slldlng farlure
there are adjacent structures a)bL""OfOvemlisfabrli~
andlor services which are sensitive
to ground movements,
displacements are not usually
explicitly calculated. In many
d) Bearlng capaclty farlure
design codes it is often implicitly c ) O v e m g f a r l u r e
assumed that if there is a sufficient
margin o f safety against Figure 3.24: Limit states for gravity
instability, then movements will
walls
be acceptable. If movements are
of concern it may be better to use a different form of retaining structure, such as
an embedded wall, see Figure 3.1.
As there exists considerable experience with the design ofthis type of wall, and
as movements are not usually of concern, numerical analysis is rarely performed
for standard gravity wall types. However, there may be situations where
movements are of concern, or where the soil conditions are unusual or the structure
has dimensions outside current experience. In such cases numerical analysis can
be used. Such analyses can investigate all the potential instabilities shown in Figure
3.24, as well as provide an estimate of movements.
The rest of this section will consider some of the important issues concerned
with the numerical analysis of gravity retaining walls. Although reinforced earth
and nailed walls are gravity walls, they have additional complications associated
with internal stability. Consequently, this type of retaining structure is discussed
separately in Section 3.6.
3.5.2
3.5.3
Finite element analysis
The main issues involved with the numerical analysis, apart from selecting
appropriate constitutive models for the structure, foundation and backfill, are the
simulation of the construction procedure and the modelling of the interface
between the soil and structure.
As noted above, the stability of a gravity
wall is mainly derived from the shear stresses
mobilised along its base. This in turn is a
function of the vertical force on the base and
the cohesion and angle of interface friction
between the base and the underlying soil.
The vertical force is derived from the weight
of the wall (including any inclosed soil) and
the vertical shear force imposed on its back
by the retained soil. The latter force is
dependent on the cohesion and angle of
interface friction between the back of the
Figure 3.25: Position of
wall and the retained soil. Clearly, the .
interface elements in gravity
interface properties between the wall and the
wall analysis
soil, both foundation and retained, can have
a dominating effect and must be accurately modelled in any analysis. The most
appropriate way of doing this is to include interface elements within the mesh, see
Figure 3.25. However, there are drawbacks and limitations associated with the use
of such elements. These are discussed further in Chapter 9.
Once the construction procedure has been established, and in particular the
number of layers to be used in backfill construction, the finite element mesh can
be developed. Clearly, each layer of backfill must consist of a finite set of elements
(i.e. elements cannot span more than one construction layer). During the analysis
problems are often encountered due to inadmissible stresses arising in elements
that have just been constructed. This is usually a consequence of the elements
'hanging' on the wall during the construction process resulting in tensile stresses.
This problem, and some solutions, are discussed further in Chapter 9.
3.6
3.6.1
Following its first modern applications in France in the 1960's, reinforced earth has
become an increasingly popular alternative to gravity retaining walls. The method
has been shown to be significantly more economic in many situations, due to the
efficient use of material and its favourable distribution of load on soft foundations.
The mechanical properties of the soil are supplemented by the addition of
reinforcement with a tensile capacity, to allow the soil mass to become self
supporting.
Reinforced earth walls generally have
a facing membrane which is normally
non-structural and reasonably flexible, its
purpose being to prevent local
unravelling of soil. In their original
format reinforced walls were constructed
by backfilling in layers, see Figure 3.27.
A typical situation might be as shown in
Figure 3.28. A membrane unit is placed
on a purposely built foundation and is
strips
temporarily supported while a layer of
soil is backfilled behind it. When this
layer reaches mid-height of the Figure 3.27: Typical construction
membrane unit, a reinforcing strip is details of a reinforced earth wa//
l1\
Proposed
/ Geotextile sheet
the wall
11
1
2
A further mechanism induced within the reinforced mass is caused by the use
of reinforcement of insufficient stiffness. In this situation, if sufficient strains have
accumulated within the soil mass before full strength ofthe reinforcement has been
mobilised, a displacement induced failure can occur. This mechanism is
particularly important when considering the long term effects of plastic creep in
geotextile reinforcement. An example of this type of failure for a reinforced
embankment is given in Section 5.5.6.
Settlement
Rotation
Block sliding
on foundation
Block sliding
on reinforcement
Reinforcement
stiflhess failwe
Slip surface
Reinforcement
pull-out
Reinforcement
strength failure
Localised
unravelling
3.6.2
Lengthof
reinforcement $:-;-L-:-::$-:-;-:-;--~
F=inp
-----.
-----.--------;
-.--;
----7----7---...
-,-,
fi-> -.-:-:l
$-;..:...-.
*-C-;-r
,L,-.-,-
,---------
,-----,----..a
d--I
$.-.-;-r-n+-,-.-;-:--;-----;--------.
+b
J"-"-;tl:;-;-;-+-1--4
$;-;-l -,-,-,
I.-.-
Foundation
-.
-.-
r , - r -:-;-I-:+-,-.-;
~ - ~ - ~ 4 - : - : - 4-----;
-4awb-4
,A-'--,-+-:$-;-:-.-'--.-----.--.-------.
$:-L-L-I--+-,-+-J-:--2
----- J. X'--- - - --- - - -. - - - - - .
(.-.-:-:-::$-8-I-;-.--:--&$;-;-~-l-l-~~-:-:-,-
'--A
-----
where p' is the angle of shearing resistance of the backfill and ,u accounts for both
bearing and frictional effects. ,u can be found from shear box interface tests
performed on the reinforcement and can have values greater than unity.
If the reinforcement is steel strips then its pull out resistance is mobilised from
interface friction with the soil. However, as the reinforcement is not continuous in
the out of plane direction, it is not consistent with the plane strain assumption and
again an approximation has to be made.
One approach to the problem is to model the interface elements as having
equivalent frictional contact strength, by considering the relative areas a n d p values
of the soil to soil and soil to reinforcement interfaces. However this approach does
not provide logical results. For example, if the width of the strip increases the
proportion of soil to reinforcement surface area (p<l) in&-easesrelative to the soil
to soil surface area (p=l) and there is a corresponding drop in the effective contact
strength.
In addition, failure of the interface
elements will result in block sliding as
noted in Figure 3.33. This does not
accurately represent pull-out failure as it
disregards the soil to soil connectivity in
the discontinuities between the
reinforcement elements in the horizontaI
out of plane direction.
A possible solution to the problem is
to tie adjacent nodes of the soil elements
together across the reinforcement. For
example, the vertical displacement of
nodes such as E and G in Figure 3.35 are
Detail of a 3D slice
.,
,' ,
tied to nodes F and H respectively.
- - - - - - - - -f ' --.
Likewise the horizontal displacements of
II
:
'
;
these nodes are also tied. This approach
I
I
has the advantage of modelling the soil
I
continuity while still allowing failure
I
I
'
/
along the soil-reinforcement interface.
- - - - ; ,-~ ~
A better solution to the problem
would be to perform three dimensional
analyses. If the wall satisfies most of the Figure 3-37;30 modelling of a
reinforced earth wall
assumptions of plane strain, apart from
- - - - - - - - - - - - - - - - - - - - - - - - - - - - V - - - -
- - - - _ _ _ _ _ _ _ _ 7 _ _ _ _
I
lI
I
I
I
3.7
Embedded walls
3.7.1
Introduction
The most common use of concrete
embedded walls (such as diaphragm,
secant pile and contiguous pile walls,
see Figure 3.39) is to support the sides
of deep excavations, or tunnels which
are formed by various means (cut and
cover, top down, etc). While steel
sheet piles can be, and are, used for
secantpilewall
Conliguous pile wall
this purpose, concrete walls are also Diaphragmwall
used to form bridge abutments and
quay walls.
Figure 3.39: Types of embedded
Under some circumstances these
waN
walls may also be used to limit
ground movements within another structure such as an embankment or cutting, or
to control groundwater (slurry walls to cut off water flow into an excavation or
cutting). However, the following discussion is aimed solely at the first category.
3.7.2
Installation effects
3.7.2.1 General
This is an extremely complex issue which is generally ignored in most numerical
analyses. In most instances the walls are 'wished in place'. This is the term usually
used to indicate that installation effects have been ignored and implies the
assumption that there is no increase or reduction in stresses associated with
installation of the walls.
In reality, concrete piled walls are constructed as a series of individual bored
piles and diaphragm walls are constructed as a series of individual panels. During
construction the holes in which the piles or diaphragm wall panels are concreted
maybe supported temporarily (e.g. by bentonite slurry). This is obviously a very
three dimensional process and not easily approximated by a two dimensional
analysis. There is some stress redistribution around these individual elements, both
horizontally and vertically (Huder (1972), Davies and Henkel (1982)).
Likewise the installation of steel sheet piles will alter the stresses within the
adjacent ground.
3.7.2.3 Analysis
The issue is whether or not installation effects should be represented in any
analyses, or should it be assumed that measures will be taken during construction
to mitigate such influences. It should be remembered that in many situations
embedded walls are being used because movements are critical and therefore every
effort will be made to minimise movements during wall installation. As stress
changes within the soil only occur ifmovements are allowed, minimising the latter
will also limit the former.
A number of researchers have tried to analyse these effects (for example
Higgins (1983), Ng (1992), Gunn et al. (1993) and De Moor (1994)).
Unfortunately none ofthese approaches has been entirely satisfactory because they
have not realistically considered three dimensional effects, see St John et al.
More recently some results of three dimensional analyses involving wall
installation have been published (e.g. Schweiger and Freiseder (1994) and
Gourvenec (1998)), but these analyses have generally used very simple constitutive
models and very crude meshes. Such simplifications were necessary because, as
noted in Chapter l I of Volume 1, very large computer resources are needed to
perform full three dimensional analyses. It is therefore questionable how realistic
were the predictions of remote movements associated with wall installation. Since
the purpose of trying to model wall installation is normally to assess remote
movements, and because allowing soil stresses to relax is generally not
conservative as far as the design of structures is concerned, it is questionable how
valuable these studies were. To get a better understanding of the process more
sophistication is required.
One example where this issue had to be addressed was the design of a pin-joint
connection between an inclined road slab and a contiguous bored pile wall for an
underpass (Higgins et al. (1999)). The connection between the slab and the wall
was a pre-cast concrete hinge with a rounded stainless steel bearing, see Figure
The designers were extremely concerned about the ratio of the horizontal to
ertical thrust in this joint under long term conditions and they placed a limit on
his 'thrust ratio'. Assuming that the wall was 'wished in place' would allow no
tress relief during installation. For any given assumption concerning the
3.7.2.4 Comments
To model wall installation realistically a fully three dimensional approach is
required. Even then the sequence of pile or panel installation will need to be
known, in order to correctly account for the effects such as arching around the
individual excavations (piles or panels).
This is apparent from the Chater Station case history described above. If it had
been possible to analyse construction of the wall realistically (i.e. a fully three
dimensional analysis, a representative soil model and realistic boundary
conditions), then some indication of the problems might have been obtained.
Relatively simpler procedures, such as those currently proposed, are unlikely to
have been of much assistance.
Each case has to be considered on its merits. If it is necessary to account for
installation effects, care must be taken in any analysis to ensure that the results are
not simply a consequence of the applied boundary conditions, as discussed by St
John et al. (1995). Care must be taken to ensure that these effects are not overestimated, since this could have serious implicationsfor the design of the structure.
A cautious approach is required.
3.7.3
Modelling of walls
3.7.3.1 Element type
A major concern when modelling embedded walls is the choice of finite element
type to represent the wall itself. If the wall is a diaphragm, secant pile or
contiguous pile wall, see Figure 3.39, then it is relatively thick (usually greater than
0.8m) and solid elements are probably appropriate. However, if the wall is made
from steel sheet piles, then the use ofbeam (or shell) elements is more appropriate.
However, in either case solid or beam (or shell) elements can be used.
As noted in Chapter 3 of Volume 1, beam (or shell) elements provide an
efficient but approximate way of modelling structural members. They are
approximate in the sense that they reduce one dimension of the structural member
to zero. For example, if used to model an embedded retaining wall, the wall is
essentially modelledas havingzero thickness. At first sight this approximation may
seem reasonable and consequently the approximation is often overlooked in
practice. However, the approximation can have a significant effect on the analysis
of an embedded retaining wall.
As an example consider the embedded cantilever wall shown in Figure 3.13.
The wall is Im thick and 20m deep and is made of concrete. The properties of both
the soil and wall are given on Figure 3.13. To simplify matters the water table is
assumed to be at depth below the base of the wall.
Analyses have been
performed in which the wall
was modelled using a single
column of solid elements, with
the finite element mesh shown
in Figure 3.6. Two analyses
have been performed, one with 8
K,, = 2.0 and the other with K,,
= 0.5. The analyses have also
been repeated with a similar
mesh, but with Mindlin beam .20
-30 -25
-20
.l0
-5
0
-5W
0
1030
IJW
2030
elements used to model the
Wall b e n h g moment (kNmim)
m
,,,,,,,
wall. The predicted horizontal
wall displacements and Figure 3.4 1: Effect of wall element type
in a K, = 2.0 soil
associated bending moments
z,,
'"
-13
JM
for the analyses with K,) = 2.0 are given in Figure 3.41. These predictions
correspond to an excavation depth of 9.3m. Corresponding results from the
analyses with K,, = 0.5 are given in Figure 3.42.
For both K , values the use of beam elements to represent the wall results in
larger horizontal movements and bending moments. This is particularly so for the
analyses with K, = 0.5 were the maximum wall displacement and the maximum
bending moment are 33% and 20% respectively greater forthe analysis using beam
elements compared to the analysis using solid elements to represent the wall. The
differences are smaller for the analyses with K,, = 2.
"
sheet pile. This occurred for all three schemes. This clearly has implications on the
design of the wall.
End of conshuchon
10-
A BeUCommon
GmrgeOrem
m H o w of Commons
2Y
-R
10
15
10
Wall sllffness
(In El)
l5
End ofcoashuction
'
Wall S O ~ U (h
S EI)
Lnng term
Wall s h 5 w s (In@
In contrast, the lateral wall displacements increased as the stiffness of the wall
reduced. The magnitude of the increase in displacement depended on the scheme.
For the Bell Common tunnel, which had a single prop, the maximum displacement
increased by as much as a factor of three in the long term. However, as the number
of propping levels increased, the increase in displacement reduced. For the House
of Commons car park there was only a very small increase in the maximum wall
displacement as the stiffness of the wall reduced.
Hight and Higgins (1994) undertook a much simpler study of a hypothetical
scheme thought to be reasonably typical of deep excavations in central London.
When looking at wall stiffness, they only considered short term effects. Figure 3.48
shows a cross section of the problem they analysed. Their results are shown in
Figure 3.49. The wall was l m wide and in the standard case the Young's modulus,
E, of the section was 28GPa. Two other cases were also considered, one with E
AvT+ry
Unwcalbonsd
3.7.3.3 Interface
- IOE
behaviour
To model the interface
behaviour between the soil and
the wall, interface elements
can be placed around the Wall.
Horizontal wall
Day and Potts (1998) have
shown that the zero interface
elements described in Secti0n Figure 3.49: Predicted displacemen ts for
3.6 of Volume 1 work well in
different wall stiffness
this situation. They also show
that results are not too sensitive to the stiffness used for the interface elements. This
is good news, as interface stiffness is difficult to determine accurately. However,
they do show that the angle of dilation of the interface elements can have a
significant effect on soil movements.
3.7.3.4 Wall permeability
One important consideration which is often overlooked is the relative permeability
of the wall in relation to the surrounding ground. Obviously, some walls are likely
to be more permeable than others and measures can be taken to minimise seepage
through the wall.
An assessment of the effectiveness of the wall needs to be made and
appropriate boundary conditions applied in any analysis. The wall may be treated
as being completely impermeable, completely permeable, or even given some
finite permeability.
Solid elements representing the wall can easily model any of these conditions,
but if beam (or shell) elements are used more care must be taken. If elements with
zero thickness are used no distinction is made between nodes on either side of the
wall. The wall is therefore permeable. To make the wall impermeable a distinction
must be made between nodes on either side of the wall. This can be done by
introducing interface elements between the soil and the wall. However, a beam
element cannot be used to model a wall of finite permeability. This problem is
discussed in more detail in Chapter 9.
Support systems
3.7.4.1 Introduction
Section 3.4.4 discussed support systems in general tenns. Embedded walls can be
supported in a variety ofways. Different methods might be chosen to support walls
during construction to those methods adopted for the permanent condition. The
permanent support system may be used to support walls during construction, but
measures such as berms, anchors, grillages or props might be used which are
removed as the permanent system is installed. Generally, these temporary measures
are necessary to provide greater working space or more flexibility in the method
of working.
At the design stage uncertainties will exist over:
3.7.4
These analyses
used stiffness Figure 3.50: Cross-section of rhe Bell
values obtained from laboratory
Common tunnel
t e s t s p e r f o r m e d by t h e
manufacturer ofthe 'Kork' packing. However, field monitoring indicates that even
four years after construction little support is being provided by the roof slab. This
is due to the much higher compressibility of the packing in the field which has
significantly reduced the effective stiffness ofthe support to the wall. This example
denlonstrates the difficulties in determining appropriate support stiffness.
The simple study of a
typical d e e p basement
undertaken by Hight and
Higgins (1994), see Figure
3.48, considered the effect of
variations in prop stiffness on
wall and ground displacements
during construction. An initial
analyses was performed with a
prop stiffness K=50 MNImlm.
Horizontal wall
This value is typical of that
provided by a concrete slab.
The analysis was then repeated Figure 3.51: Predicted displacernents for
with prop stiffness values of
differing prop stiffness
0.2K, 5K and 10K. The results
are shown in Figure 3.5 1. It is apparent that there is a very nonlinear relationship
between stiffness of the props and the predicted displacements. Because the
displaced shape of the wall was modified, there was a consequential effect on the
bending moments in it.
It is apparent that the effective stiffness of the support system can have a
significant influence on the observed and predicted behaviour of walls. Care must
be taken not to over-estimate the effectiveness of this support. However, in some
instances struts are pre-stressed in order to minimise problems associated with
bedding effects, thereby maximising the effectiveness of the support. This prestressing can, and should, be modelled in any analysis.
Pm-jomted connection
FUII
3.7.4.5 Berms
Potts et al. (1993) described a study of the use of berms to provide temporary
support for retaining walls which considered variations in the geometry of the
berms. They related 'displacement efficiency' and 'normalised bending moment'
to the volume of the berm. There were a number of interesting results from this
study concerning the influence of the berm's geometry and the effectiveness of
berms. However, one of their conclusions was that:
"The horizontal resistance provided by the berm has a substantial effect on wall
behaviour. It is therefore not accurate to use an equivalent surcharge load to
model the effect of a berm. If such an approach is used, the effectiveness of the
berm is likely to be under estimated."
Berms are comparatively easy to model, although their stability, as drainage
occurs during construction, may need to be considered. There is also an issue
concerning the correct pore water pressure boundary conditions that should be
applied to their surface if they are left to stand for some time, see Section 3.4.7.4.
Problems can arise when berms are removed and a permanent support is
installed. The berm may be removed in sections and replaced with another form
of support, such as a slab cast in sections on the floor of the excavation. This
procedure is very difficult to model in a two dimensional plane strain analysis and
therefore approximate methods have to be considered. This may involve partial
berm excavation before installing the other form ofsupport or, considering a worst
case scenario, complete removal of the berm without another means of support.
+3
1I
Folkstone beds
I
I
Run 2
Run 7.1
Horizontal displacement(mm)
3.7.5
In stiff over consolidated clays, such as London Clay, long term effects can be
significant and can dominate the behavnour of a structure taking many years to
occur. London Clay is a relatively well understood soil (emphasis should be placed
on the word relatively), but even so there are few case records with measurements
of long term heave against which the predictive capability of constitutive models
can be assessed.
In any design possible long term effects need careful consideration and
measures taken to mitigate them or to cope with them. Ifconstruction is slow, some
drainage will occur during this period and this will tend to reduce post-construction
effects.
Apart from changes in earth 6
pressures that act on a wall as
drainage occurs and excess pore water
pressures dissipate, one of the most 8 4
Small stram shfhess-plastrc
significant long term effects is the 3
g
Oedometerdata
swelling of clays beneath deep 3
mples from 8-24m depth)
excavations. Figure 3.58 shows $ 2
Linear elashc-plast~c
measurements of swelling in
oedometer tests on London Clay,
together with two predictions. One of o
100
200
300
400
these predictions is based on a linear
Verttcal effectwe stress (&a)
elastic-plastic constitutive model. The
other has the same plastic behaviour
Figure 3, 58: PIedi.tjons and
but the elasticity is based on the small
measurements of S welling behaviour
strain model described in Section
of London Clay
5.7.5 of Volume l . It is apparent that
at low effective stress the test data indicate that large volumetric strains occur and
that large vertical displacements result. The predictions based on the linear elasticplastic constitutive behaviour are unable to reproduce this behaviour. However,
reasonable predictions are obtained from the small strain model.
To minimise the effects of these large volumetric strains and pore water
pressure changes, the swelling potential of the clays may need to be controlled by
drainage, or by other means. Alternatively, a void may be left beneath a basement
to allow heave to occur without imposing any stresses on the base slab of a
structure. Modelling ofthis is reasonably straightforward, but the depth ofthe void
needs to be known before the analysis is performed. Throughout the analysis the
heave is monitored at nodes along the soil surface beneath the base slab until the
displacement of the soil equals the depth of the void. Subsequently, dispiacements
at nodes along the soil surface become linked (i.e. tied) to the corresponding nodes
on the slab (this is not a zero displacement boundary condition, see Section 3.7.4
of Volume l), and forces transferred from the soil to the structure. Unfortunately,
this will not occur evenly across the base of an excavation due to the nonuniformity of stress conditions.
125
.;a loo
% 75
?
50
25
0
'
40
80
120
160
200
-gl*
140
120
60
40
20
10
From a drained test where only the horizontal effective stress is changing,
Poisson's ratios v,' (vertical straining due to horizontal stress) and',v (horizontal
straining due to horizontal stress) can be calculated as:
--'4%
E;,
E;,
--P&
E;
This compliance matrix is symmetric in the linear elastic region (i.e. elastic
plateau) of soil behaviour and in that case:
P;@ P;
---
E;
E;,
DP - direction of wave
propagation
DO - direction of wave
oscillation
BE - bender element
1 -3.1 1 Permeameters
The most common means of measuring
permeability in the laboratory is to use a
permeameter. However, because of the very
large range of permeability in soils, no
simple piece of equipment is suitable for all
cases. The types ofequipment most generally
used are:
-- - -----Waste
Water out'
1.4
In-situ tests
1.4.1 Introduction
In-situ soil testing offers complementary capabilities to laboratory testing by
providing information about current conditions in the ground (e.g. piezometric
levels, strength variation with depth, soil description,etc.). The principal advantage
of in-situ tests is that they assess soil behaviour in its natural condition, thus
avoiding sample disturbance. They can also be more economical than laboratory
tests and in some cases can provide more representative results through a higher
density of data collection.
However, the disadvantage of in-situ testing is that most geotechnical
parameters (e.g. strength, stiffness, etc.) are estimated from empirical correlations,
rather than from direct measurement. Also, in-situ tests load the ground in different
ways to conventional laboratory tests. Because soil properties are nonlinear,
anisotropic and rate dependent a wide range ofresults can be obtained from similar
tests in the same deposit. Therefore great care is required in test preparation.
In many practical situations the only information that exists about a particular
site comes from in-situ testing. This often restricts any finite element analysis to
the use of simple constitutive models.
This section describes the most common in-situ equipment and procedures for
soil investigation and the modelling parameters that can be obtained from them.
: Depth of
where a and b are material constants and g,,'is the effective overburden pressure.
He further suggested that it was convenient to normalise the blow count in the
above expression to a standard effective overburden pressure of 100 kPa (i.e.
l kg/cm2):
8
d
l00
01
NC coarse sands
'ES!
e
8 200
m NC coarse sands
Laboratow tests
.M
+a
Sand - stiffness
It is now widely accepted that soil stiffness is strain level dependent, i.e. it
gradually reduces as strains increase. To accurately predict movements of
structures in the ground (e.g. footings, retaining walls, tunnels etc.) it is necessary
to obtain a good estimate of the stiffness of the ground.
Vesid (1973) performed tests with a loaded foundation of a constant breadth in
sands of varying density. He found a unique relationship between settlement, p,
and the degree of loading, qlq,,, , expressed as a ratio of the design load, g, to the
ultimate load, g,,, . This suggested that qlq,,, could be used as an indirect measure
of shear strain and therefore linked to the average stiffness E' (expressed in terms
of Young's modulus).
For this purpose data from a wide range of strip footings, raft foundations and
plate tests on both normally and overconsolidated sands have been collected
(Stroud (1989)). They have been taken from case histories where SPT tests have
been carried out and a correlation, shown in Figure 1.28, has been derived between
the E'IN,, and qlq,,, values. Knowing the design loading and estimating the
ultimate bearing capacity, it is possible to use this graph to estimate the average
ground stiffness at working load on the basis of SPT blow counts.
Clays - strength
The undrained shear strength, S,,, of overconsolidated clays can be related to SPT
blow counts using the simple correlatio~lproposed by Stroud (1 974):
= jfi
(1.23)
N60
where the variation off, with plasticity index is shown in Figure 1.29.
Clay - stiffness
The drained stiffness, E', again expressed in terms of Young's modulus, of
overconsolidated clays in vertical loading can be estimated from case histories, in
a similar manner as for sands. Figure 1.30 shows such data plotted as E'/&, against
qlq,,,. Again, a trend of decreasing stiffness with increasing degree of loading (i.e.
increasing strain level) is evident.
consolidated
40
--35'
33'
31'
30
F w sand
skempton
CO- sand) (1986)
Well @add
sand
and gravel
D Sand
o F i saod
?
Average for overconsolidated
sands and gravels
Average for normally
consolidated sands
0.1
0.2
0.3
1.4.3
q,, and, if required, the frictional resistance on the side (sleeve) of the cone,f;, are
recorded at depth intervals. A typical CPT result is shown in Figure 1.31. The ratio
A/qCis termed the friction ratio.
Empirical correlations are available between these measured quantities and the
angle of shearing resistance and relative density for sands and undrained shear
strength for clays. These correlations should be used with care because they may
only apply to a particular set of conditions (e.g. penetrometer dimensions,
penetration rate, soil type).
Cone resistance, g, (MPa)
Local side friction,
L (~~lm')
Cone resistance,
9, m / m 2 )
Sand
The existing correlations of cone resistance with v', E and G have been derived
from tests performed in a calibration chamber.
Robertson and Campanella(1983) presented relationships between g, and peak
9' for normally consolidated quartz sands at increasing vertical effective stress, see
Figure 1.32. The angles of shearing resistance were obtained from triaxial tests
performed at confining stresses approximately equal to the horizontal effective
stress in the calibration chamber prior to cone penetration. Using the relationships
given in Figure 1.32 for overconsolidated sands will tend to slightly overestimate
9'-
Clays
The measurement of clay properties using the CPT is highly dependent on the rate
of cone penetration, because of the build up of excess pore water pressures. The
CPT in clays is principally used to estimate the undrained shear strength, S,,, and
the relationship with g, is expressed as:
qc = Nk
(1.24)
where: G,,,
is the total overburden stress;
Nk is the "one factor' analogous to the bearing capacity factor N,, see
Chapter 6.
Nk is established using empirical correlations, with shear strength measured
using other techniques. Lunne and Kleven (1981) found that for normally
consolidated clays N, was independentof plasticity index and suggested an average
value of 15, with the majority of data being between 11 and 19.
P - cavity pressure
Elastic expansion
The main parameters and variables for membrane expansion are shown in Figure
1.34. The radius of the cavity, when the cavity pressure P is equal to the initial
, , is marked as p,. When P>o,,,,, the cavity will
horizontal stress in the soil, a
deform, having a new radius p. The cavity strain, E,, is then given by:
The pressuremeter tangent shear modulus of the soil is then calculated as:
Estimates of a single (average) value of G,vcan be made from the initial portion
ofthe pressuremeter P-&,curve or from unload-reload loops.
Jardine (1992) compared triaxial G,v-~s
curves, where 6, is the shear strain
(=2/3(~,,-E,)), with pressuremeter Gp-c, curves and derived an empirical
relationship between E, and by comparing the strain values at identical shear
moduli levels:
With the above expression pressuremeter G,-&, curves can be transformed into
triaxial G,,-< curves, which are normally used in geotechnical design. For example,
such curves can be used to obtain parameters for the small strain models described
in Chapter 5 of Volume l .
When using the pressuremeter in sands it should be noted that the shear
modulus is likely to be stress level dependent. As the stress level in the sand
adjacent to the pressuremeter decreases with distance, the value determined from
the pressuremeter test represents an average value.
It is not possible to obtain estimates ofthe bulk modulus, K, or Poisson's ratio,
p, from pressuremeter tests.
The current cavity pressure, P, given by Equation 1.30 can therefore be reexpressed as:
Maintained load test, where load is applied in increments of about 115th the
design load and held at that level until the rate of settlement reduces to less than
0.004 mmlmin over a one hour period. The test continues until either the soil
fails in shear, or the bearing pressure reaches 2 to 3 times the design pressure.
The results are usually plotted as time vs. settlement and load vs. settlement
curves (see Figure 1.37).
Sand test
Using elastic theory the settlement of a rigid surface plate of diameter B, with a
uniform load q applied on a semi-infinite isotropic soil characterised by Young's
modulus E and Poisson's ratio p, is given by:
Terzaghi and Peck (1 948) observed that elastic theory tends to overestimate
settlements on sands and proposed that the settlementp, of a footing ofwith B was
better represented by the following correction:
where p, is the elastic settlement (i.e. Equation (1.33)) of a one foot plate (i.e.
B=0.3m).
where:
1.4.6
Pumping tests
Because of the limited size of the samples that can be tested in the laboratory, they
may not provide a representative value of the in-situ permeability. For example,
the sample may not contain a thin seam of more porous material that may exist insitu. In addition, samples of granular soil have to be reconstituted in the laboratory.
For these reasons in-situ measurements of permeability are generally more reliable
than laboratory tests. However, they are usually considerably more expensive.
In the field the permeability of a stratum of soil is most commonly determined
by measuring the discharge from a well. A well (borehole) is sunk into the stratum,
and water is pumped from it at a constant rate. This pumping lowers the
piezometric level in the vicinity of the well, and the resulting hydraulic gradient
causes water to flow towards the well. The piezometric level may be determined
by sinking observation boreholes at various distances from the well, and by
measuring the height at which the water stands in each, once a steady state has
been established. From this data and using Darcy's law an estimate of the average
permeability can be determined.
While pumping tests are applicable to relatively permeable soils, they are not
necessarily appropriate for some clays. In these cases a borehole can be sunk into
the clay and sealed over its length in contact with overlying soils. If left water will
flow from the clay into the borehole. By measuring the rate of increase in the
height of the water level in the borehole and again using Darcy's law it is possible
to estimate the permeability. Alternatively, the borehole can be filled to the ground
surface with water and the rate of reduction in level noted. Coupled with Darcy's
law this allows an estimate of k to be made.
1.5
1.
2.
3.
4.
Summary
Both laboratory and field tests are available from which soil parameters can
be derived. Both types of testing have their strengths and weaknesses and
ideally most soil investigations should contain a combination of tests.
The aim of laboratory tests is to measure as accurately as possible and in a
controlled manner the response of a soil element to changes in stresses, strains
andlor pore water pressures.
Oedometer tests can be performed to provide information on
overconsolidation ratio, OCR, the compression index, C,, and the swelling
index, C,. With an appropriate assumption for K, it is also possible to estimate
the gradients of the virgin consolidation line,I, and swelling lines, K, in v-lnp'
space. It is also possible to obtain estimates of the constrained modulus, E,,
and the coefficient of permeability, k.
The conventional triaxial apparatus can be used to estimate K, values, strength
values c', p' or S,, and stiffness values E and p (or K and G). It can also be
used to estimate permeability. However, it is restricted to testing samples
under triaxial compression (b=O) or triaxial extension (b=l) conditions and
therefore cannot conveniently be used to investigate the effect of varying b or
for investigating anisotropic behaviour.
Tunnels
2.1
Synopsis
This chapter presents the application of numerical methods to the analysis of tunnel
construction, with emphasis on soft ground tunnelling. First, different methods of
tunnel excavation are presented, some of which have application in rock and soil.
The simulation of this excavation process is then discussed, with particular
reference to plane strain analysis, followed by the ways of modelling tunnel
linings. The chapter then focuses on the modelling of time dependent behaviour
and on constitutive models appropriate to the modelling of soft ground tunnelling.
The final section presents examples of analyses of tunnelling in stiff clay. These
examples include the analyses of case studies and parametric studies, covering
greenfield tunnelling, tunnelling beneath buildings, and tunnelling adjacent to
existing tunnels.
2.2
Introduction
In this chapter some ofthe numerical techniques presented in Volume 1 are applied
to the analysis of tunnel construction. Attention is restricted to bored tunnels.
Tunnels formed by the cut and cover constructiontechnique are covered in Chapter
3. The Authors' experience and expertise is in the area of tunnelling in soils and
soft rocks modelled as continuum materials, where structural features, such as
faults and joints, are not represented. This chapter therefore deals with soft ground
tunnelling.
The first question to be asked of anyone advocating the application of
numerical methods to tunnelling is surely '"hy use numerical methods to analyse
tunnel construction?" There are non-numerical ways of obtaining good predictions
of the likely ground response to tunnelling, and the likely loads in a tunnel lining.
These conventional design tools are arguably cheaper and quicker to use. But they
are characteristically uncoupled, i.e. the loads are determined by one technique
(usually an elastic solution), and movements by another (usually empirical), the
two not being linked together. Furthermore, the information gained from
conventional analysis is often limited. For example, empirical predictions are
limited to greenfield situations, where there is no existing surface or sub surface
structure to influence the pattern, or magnitude of ground displacement. In a real
tunnel, however, the different facets are clearly coupled and the problem is
Tunnels 1 39
This establishes a strong case for using numerical methods. It also focuses the
mind on the type of problem for which one might consider the use of numerical
methods. This chapter deals with all aspects of the above list: the construction
sequence and its simulation (excavation and lining); the choice of an appropriate
soil model (stress-strain-strength and permeability); the modelling of hydraulic
boundaries; and examples of analyses of tunnelling beneath buildings and
alongside existing tunnels.
2.3
Tunnel construction
2.3.1 Introduction
Historically soft ground tunnels were excavated by hand, using spades for cutting
clayey soils and picks for breaking up granular soils and weak rocks. Openings
would be supported by timbering and subsequently lined in brick. The first use of
a tunnelling shield was for construction ofBruneI's Thames Tunnel between 18251843 (Skempton and Chrimes (1 994)). The men excavating the face were protected
from falls by a large shield within which they worked. The advent of pneumatic
spades and picks clearly enabled more rapid hand excavation. Hand excavation is
still in use today on short lengths of tunnel, or where access for mechanical means
is restricted.
Civil engineering construction starts the 21" Century with many alternative
mechanical methods for creating underground openings. The engineer can design
the most remarkable tunnel excavations, cathedral like in proportion and complex
in shape. The decision as to which method to adopt is based on knowledge of the
geology and the size and shape of the opening required. This section briefly
introduces three of the most common tunnel excavation techniques, highlighting
some of the key features which have an implication for numerical modelling.
Temporary support is today provided by a number of means - shield support,
compressed air, slurry machine, earth pressure balance machine (EPB), or
shotcrete. Permanent support is also provided by various means, including wedgeblock precast concrete and cast iron, bolted cast iron lining, reinforced shotcrete,
or cast in-situ reinforced concrete. Section 2.4.5 of the chapter deals with
modelling the tunnel lining.
Tunnels / 41
For large openings using SCL it is always the case that the tunnel is created by
the method of advanced headings. This can involve excavation of the crown first,
leaving a temporary invert, or the use of left and right side drifts, or a combination
depending on the ground quality and the size of opening. In all cases the advanced
heading is fully lined by shotcrete before the following drift commences.
2.3.5
m{
percentage volume loss. During the driving of a tunnel there are various distinct
causes of loss of ground into the excavation. Figure 2.4 shows, as an example, a
conventional shield tunnelling machine geometry and highlights the sources of
short-term volume loss discussed here:
Face loss: Loss first occurs into the face of the excavation, causing ground
settlements to appear at the surface ahead of the tunnel. These losses can be
reduced through the use of a closed face TBM, compressed air, slurry shield or
EPB machine.
Shieldloss: Soil will relax radially towards the shield after the cutting bead has
passed, filling the bead width void if the deformations are large enough. Any
over-excavation will contribute to increase this radial loss. The shield drives
itself forward by pushing off the last ring of lining segments to be erected. If
the soil deformations are large enough to fill the bead width void, then the tail
to the shield provides support to the cut perimeter of the soil. Another source
of shield loss occurs if the shield
does not move forward with its
axis coincident with the axis of
see Figure 2.5. In such
the
a situation the size of the
excavated hole is larger than the
final lining. This behaviour is
often referred to as pitching and
Figure 2.5: Schematic view of
occurs due to bad steering, or
pitching
when tunnelling on a curved
alignment.
Post shield loss: If the lining is erected behind, rather than within the shield,
then there will be a length of unsupported soil between the back of the tail to
the shield and the erected liner, over which radial deformation can occur. Once
the lining is in place, whether constructed within or behind the tail to the shield,
ground will continue to squeeze into the excavation if a void exists behind the
lining. It is common practice therefore to grout up any voids between liners and
the exposed soil surface. Expanded concrete liners are theoretically expanded
against the soil leaving no void.
These three aspects form the major contribution to the short term volume loss.
If excavation is taking place in a low permeability clay soil, then rapid construction
progress may result in a no volume change, undrained ground response. Thus the
volume loss at the tunnel will be exactly reflected at the surface in the settlement
trough. This statement will only hold rigorously true if there are no other
boundaries which can deform in response to the volume of soil moving into the
new excavation, such as open excavations. In a clay soil, after the tunnel has been
constructed and the lining installed, ground displacements are likely to be seen to
continue at the surface. This is due to the consolidation movements of the clay as
pore water pressures equilibrate to the new steady state regime, dictated by the
Tunnels 1 43
presence of the tunnel. The tunnel lining may also deform as the earth pressures act
on it, causing ground displacements.
2.4
2.4.1
Introduction
Tunnel excavation is a three
,
dimensional engineering process.
Whilst recognising that three
dimensional analysis is becoming
possible in the work place, it is still
two dimensional modelling that
dominates. This is because there are
practical limits on cost and computer
resource which, when performing
analyses sufficiently sophisticated to
handle all the complexities outlined in
Section 2.2, restrict us to two
dimensional modelling. If multiple
shallow tunnels are to be analysed, or
if the ground surface response is key
Figure 2.6: Plane strain geometry
to the analysis, then a plane strain
representation of the transverse
section is required (e.g. to study
effects on structures, Figure 2.6). If a
single deep tunnel is to be
investigated, and surface effects are f i a m ~ ~ O ~1
~
not of prime interest, then an axially boundaries
symmetric approximation may be
appropriate and heading advance can
be studied, all be it within a simplified Figure 2.7: Axi-symmetric geometry
stress regime (Figure 2.7).
More recently three dimensional analyses have been carried out using the
Fourier series aided finite element method (FSAFEM), applied to an axially
symmetric geometry to analyse the advance of a single tunnel heading (Shin
(2000)). FSAFEM is covered in detail in Chapter 12 of Volume 1. Such an
approach allows the material properties, initial stresses and boundary conditions
to vary with 8 (Figure 2.8), so introducing the ground surface whilst maintaining
the 2 dimensional simplification for the finite element mesh. This still demands
substantial computer time and restricts the analyst to considering one tunnel, but
is a significant advance for research into tunnel heading behaviour.
When modelling in plane strain at least one assumption must be made, i.e.
something must be prescribed, rather than predicted. There are various accepted
assumptions which are addressed in Section 2.4.4 of this chapter. The reader might
2.4.2
Tunnels / 45
2.4.3
include the boundary displacement conditions, required to represent the far field
conditions or any symmetry of the problem; any surface traction; the excavation
of solid soil elements; the construction of structural shell elements; and the
hydraulic conditions at the far field boundaries, the soil strata interfaces (if an
interface is between consolidating and non-consolidating elements) and the tunnel
lining itself.
Loading and flow boundary conditions affect the right hand side of the global
equilibrium equations, whilst displacement and pore water pressure boundary
conditions affect the vectors of nodal displacements and pore pressures on the left
hand side of these equations, see Section 2.8 of Volume 1. Sufficient displacement
conditions must be prescribed in order to retain any rigid body modes, such as
rotations or translations of the complete mesh.
Analysis of tunnelling in drained granular materials will require careful
consideration of the hydraulic boundary condition, both during and after
excavation. In contrast, excavation in clay is usually rapid enough to be treated as
an undrained process, so the tunnel perimeter may remain impermeable until after
excavation is complete. Section 2.5.3 considers the hydraulic boundary condition
in more detail.
The void is placed around the final tunnel position and so locates the soil boundary
prior to excavation (Figure 2.9). This is achieved by resting the invert ofthe tunnel
on the underlying soil and prescribing the gap parameter at the crown. The gap
parameter is the vertical distance between the crown of the tunnel and the initial
position before tunnelling. The analysis proceeds by removing boundary tractions
at the perimeter of the opening and monitoring the resulting nodal displacements.
When the displacement of a node indicates that the void has been closed and the
soil is in contact with the predefined lining position, soilllining interaction is
activated at that node. The soil and the lining are actually treated as separate
bodies, related only by nodal forces (Rowe et al. (1978)).
5
l
t
Tunnels I 4 7
prescribes the volume loss that will result on completion of excavation, see Figure
2.12. This method is therefore applicable to predictive analyses of excavation in
soil types for which the expected volume loss can be contidently (and
conservatively) determined for the given tunnelling method. It is also invaluable
for worthwhile back analysis of excavations for which measurements of volume
loss have been made.
initial
E, = E ,
sti&ess reduction
and unloading
E,'
excavation, lining
installation and unloading
= p E,
vL= v,- v,
-jJ+([
Depending on the stiffness of the lining further volume loss can occur during the
latter process. It may therefore be necessary to install the lining at an increment
which has a smaller volume loss than that desired, so that after full excavation the
desired volume loss is achieved.
Tunnels / 49
>\
I 3
$:ients
Tunnels 1 51
M, and hoop force, N, developed at each joint between two lining segments can be
represented by the normal force, N, acting at an eccentricity, e, such that eN=M
(Figure 2.17b). As the induced bending moment increases during lining
deformation, the normal force translates across the joint, see Figure 2.17~.The two
limiting moment values, M, and Mi , above which the joint will open on the
intrados or extrados, are set by the model based on the defined position of the
neutral axis and the maximum compressive stress:
2.5
2.5.1 Introduction
It is often important to model the consolidation behaviour of a soil during and after
excavation of a tunnel. To reliably study the time variant behaviour of the ground
and the tunnel, coupled consolidation analysis must be carried out. For coupled
consolidation analyses the hydraulic boundary conditions must clearly be correct
and representative of the field, as must the permeability of the continuum. A
crucial, and often ill considered, hydraulic boundary condition is that controlling
flow around the newly constructed tunnel. This section discusses such
considerations.
2.5.2
Tunnels 1 53
pressure degrees of freedom at every node in amesh will significantly increase the
computer resource and time demand on that analysis. In a mixed situation the
engineer benefits therefore from being able to treat granular strata as free draining,
whilst only considering the consolidation behaviour of silty or clayey strata.
Any consolidating elements will require a model to define the permeability (see
Section 2.5.4). It is important to prescribe an initial pore water pressure distribution
which is in agreement with field measurements from the site. Reliable
measurements of permeability are less likely to have been made, but reasonable
judgement should be used when defining the permeability profile. The initial
stresses which prescribe the pore pressure profile at the outset of the analysis must
be consistent with the material permeability distribution. In the simple case of a
linear initial pore pressure profile, combined with a homogeneous permeability,
there is no problem (these will always be compatible). Modelling a nonlinear initial
pore pressure distribution, in combination with a more realistic inhomogeneous
permeability model, needs more care. Such a situation exists in London, and the
following example serves as a useful guide to approaching such situations, see also
Chapter 9.
-a
8
IT-J
,opbounduy
the actual boundary value
problem to be analysed. The
elements in the column are
given the stress-strain and
permeability constitutive
models that will be used in the
full analysis and the desired
steady state pore pressures at
the top and bottom are
prescribed as boundary
prrwnb; pm p-ur
conditions. Taking the analysis
Figure 2.20: Column method for
to the long term (through the
predicting
steady state pore water
use of sufficient time step
pressure
profile
increments) gives a prediction
of the steady state pore
pressure profile. This prediction can be compared with the data, and if the match
is not acceptable the parameters in the permeability model can be adjusted and the
analysis repeated. Using a unit width column for this exercise, rather than the full
mesh developed for the boundary value problem, is quicker. Once an acceptable
pore pressure profile has been obtained, this can be included in the full analysis in
the initial stresses, in the knowledge that it is compatible with the soil permeability.
The example analyses in Section 2.5.5 demonstrate the influence of an underdrained pore pressure profile on long tern ground movement predictions, through
comparison with a hydrostatic scenario.
,U
t o b m m ~
Tunnels 1 55
outward pressure is being applied to represent compressed air working, etc.). Soil
adjacent to a tunnel excavation might develop tensile pore pressures (suctions)
during excavation. This will arise if the reduction in pore water pressure due to
unloading is greater than the initial pore water pressure around the tunnel. In this
case, if a zero (or positive) pore pressure boundary condition is prescribed, the soil
will draw water across the tunnel boundary. This is clearly unrepresentative of
reality, as a new tunnel is not a source of water. A more sophisticated boundary
condition is therefore required.
The application of a precipitation boundary condition to a tunnel lining is
presented in Section 10.6.6 of Volume 1. When using such a condition, the
program monitors the pore water pressures at the boundary nodes. If, at any time,
at any node, the soil pore water pressure is tensile with respect to the boundary
pressure, then a no flow condition is maintained at that node to prevent flow from
the tunnel into the soil; if however it is compressive, then the boundary pressure
is prescribed and free flow at that boundary node is permitted. In this way, in the
intermediate term, the swelling soil is forced to take water from surrounding soil
and not across the boundary, but the long term steady state is a flow regime
towards the new tunnel. As the condition is applied node by node it is possible at
some stage in an analysis to have flow across the tunnel boundary (into the tunnel)
in places, whilst still forcing a no flow condition in other places.
..
. ..
, - m
pressure profiles
pressure profile was hydrostatic. Two permeability models were used and the
analyses taken to the long term. The first model was a linear, isotropic,
inhomogeneous permeability (reducing with depth). The second model was a
nonlinear model of the form presented below in Equation (2.2), in which
permeability varies with mean effective stress and is continually updated during
the analysis. The hydraulic boundaries in the long term are hydrostatic pressure at
the top of the London Clay (defined by the water table in the gravel) and zero pore
pressure at the permeable tunnel boundary. Figure 2.21 shows that the nonlinear
permeability model predicts a near hydrostatic profile down to within 10 m of the
crown, below which a rapid fall off of pore pressure is evident. The linear
permeability model however predicts a pore pressure profile significantly below
hydrostatic. If a back analysis is to be carried out, and piezometer readings have
shown the pore pressures to be near hydrostatic above the tunnel in spite ofthe fact
that the tunnel is acting as a drain, then clearly the nonlinear model is more
appropriate.
If the decision is made to use a nonlinear model for a tunnel analysis, then it is
noteworthy that the necessary data on which to base the parameters for such a
model are unlikely to have been obtained during a standard site investigation.
Hence the need to have reliable pore water pressure measurements which can be
matched, as outlined in Section 2.5.2. It is also noteworthy that a model in which
k depends on mean effective stress implicitly assumes something about the volume
change behaviour of the soil. Observations of seepage pressures in fills and in-situ
soils have shown that measured pore water pressures are generally very different
from those predicted using conventional theory, hence the development of more
sophisticated models for analysis (Vaughan (1989)). It is known that permeability
varies with void ratio, often by several orders of magnitude over the range of
possible void ratios for a given soil. Void ratio itself varies with effective stress,
and effective stress is dependent on pore pressures, so the problem is nonlinear. To
model London Clay a logarithmic law, relating permeability to mean effective
stress, has been adopted:
Tunnels / 57
----
'-
3
3
-_--__--:.
Tunnels / 59
25m. There is an increase in S,,,,, of 35mm for PH, but a decrease of 20mm for PU.
These movements are complete within 5 to 10 years. With an impermeable lining
the surface heaves by 45mm for IU, and by 40mm for IH. The drainage path
lengths in the consolidating clay are longer if the lining is impermeable, and this
is reflected in the increase in time taken for the movements to cease from within
10 years to 20 years.
During the consolidation period after construction (the intermediateperiod) the
predicted ground surface response depends on many factors: the equilibrium pore
pressure profile and the tunnel depth within this profile, the tunnel lining drainage
condition, and the tunnel diameter.
2.6
2.6.1
Introduction
The task of choosing an appropriate soil model for the analysis of tunnelling is a
specific, not a general one. Volume 1 deals with many ofthe available constitutive
models, and the particularities of them. This section aims, through example, to
show that the choice of soil model can be of great importance in tunnel modelling.
Some studies recently completed at Imperial College tested a number of pre-yield
constitutive models against each other through the comparison with field
measurements of surface and subsurface ground movements induced by tunnelling
(Addenbrooke (1996), Addenbrooke et al. (1997)). Results from these studies are
presented in Section2.6.2. When modelling tunnelling in an overconsolidatedclay,
adopting a valid soil model is not necessarily going to give good predictions of
ground movements when plane strain geometry is used. Section 2.6.3 discusses a
couple of devices for improving predictions of ground movement.
as undrained. The mesh for St. James's Park is presented in Figure 2.26 as an
example (only the deeper tunnel is considered here). Volume loss controlled
excavation was used, with a target of between 1.3% and 1.5% for Regents Park and
Green Park and 3.3% for St. James's Park (based on field measurement). The
initial stresses prescribed a pore pressure profile considered to be representative at
each site, and employed an initial stress ratio, K,, of 1.5 in the London Clay. In
each case linear isotropic, linear anisotropic, and nonlinear elastic constitutive
models were alternatively adopted for the pre-yield behaviour. All three cases had
stiffness increasing with depth. Plastic behaviour was modelled using the nonassociated Mohr-Coulomb model.
600Linear isotropic, linear anisotropic
and nonlinear elastic constitutive G,e,
- - . CAU compression test
models are introduced in Sections 5.5, -7
CAU extension test
5.6 and 5.7 of Volume 1 respectively.
The use of a nonlinear pre-yield
model recognises the nonlinear nature
of real soil behaviour in the elastic
region. One of the models which
reproduces nonlinear behaviour at
small strains is outlined in Section
I
I
0
5.7.5 of Volume l. Figure 2.27 shows
0 001
o 01
0.1
10
Shear straln (%)
the secant shear modulus (normalised
by mean effective stress p') decay
with shear strain, predicted by the Figure 2.27: Small strain Stiffness
curves from triaxial tests
model with appropriate parameters
for London Clay. This is compared
with laboratory data for undrained triaxial compression and extension (samples
consolidated anisotropically).
The surface settlement profiles presented in Figures 2.28a to c reveal a number
of interesting points. First, it is noteworthy how much more precise the settlement
monitoring was in the 1990s (Figure 2 . 2 8 ~ )compared with the 1970s (Figures
2.28a and b). A full description ofthe state-of-the-art instrumentation used and the
data obtained on the more recent project is presented in Nyren (1 998). Secondly,
linear elastic models, both isotropic and anisotropic, are wholly ineffective. In all
three cases the settlement profiles are too shallow and too wide when compared
with the field data, and the isotropic analyses predict profiles of the wrong shape.
Thirdly, the introduction of a more representative nonlinear elastic model
significantly improves the predictions, see also Addenbrooke et al. (1997). The
same improvement in prediction with the nonlinear elastic model can be
demonstrated for sub-surface ground movements.
,,,
Tunnels / 61
Distance from centre line (m)
symmetric and three dimensional
analyses have revealed that, as a
tunnel heading approaches and
passes a plane in space, the
effective stress ratio reduces to the
4
side and increases above and
below the tunnel. The introduction
a) Regent's Park, 34m deep tunnel
(data from Barratt and 51er (1976))
of a stress scenario such as this in
advance of the plane strain
excavation may be considered a
Distance from centre line (m)
logical assumption to represent
stress changes ahead of the tunnel
face. In this section the results of
PRE YIELD MODEL
analyses with a local zone of
--..-..-..- - - - anisotropic
isotropic linear elastic
linear elastic
reduced K,, (equal to 0.5) within
4
the London Clay with a much
$v1
b) Green Park, 29m deep tunnel
higher global K, of 1.5 are
(data fiom Attewell and Fanner (1974))
presented. This zone extended
vertically between the crown and
Distance from centre line (m)
the invert of the tunnel, and
horizontally a distance, a, of
approximately three timer the
radius ofexcavation (Figure 2.29). ,
The Jubilee Line tunnels were reanalysed with the reduced K, 3
approach. Figures 2.30a to C show
C)St. James's park, 30m deep tunnel
(data 6omNyren (1998))
the improvement in the predicted 20 *
surface settlement profiles.
Another device for obtaining Figure 2.28: Effect of a pre-yield model
improved surface settlement
on predicted settlement profile
profiles is to employ an
anisotropic nonlinear pre-yield
soil model. If the soil is cross
anisotropic then five parameters
must be defined: the vertical
Young's modulus, E,', the
horizontal Young's modulus, E,',
the Poisson's ratio for the
influence of increments of vertical
effective stress on horizontal
strain, p,,,',the Poisson's ratio for
the influence of increments of
horizontal effective stress on the
horizontal strain in the orthogonal Figure 2.29: Local zone of reduced K,
l!
g
stresses
maintain the
Tunnels / 6 3
condition in the London Clay. Figure 2.3 1 shows the predicted surface settlement
above the St. James's Park tunnel in comparison with the field data. The
introduction of an independent shear modulus based on field and laboratory data
gives little improvement over the isotropic model. A desirable improvement is
however achieved through the use of a very soft independent shear modulus.
The reason for using the term 'devices' for these adjustments which improve
the surface settlement profiles predicted by the finite element analysis, is that they
cannot at present be supported as representative of field conditions. Using a
reduced K,, on the premise that there have been stress changes ahead of the tunnel
is inconsistent with adopting a nonlinear small strain stiffness which assumes that
no straining has taken place ahead of the tunnel. The effect of reducing K,, to the
sides of the tunnel is to cause plastic zones to develop in these areas early on in the
incremental excavation procedure. This is in contrast to the high K,, analyses in
which plastic zones develop above and below the tunnel.
The effect of softening the anisotropic shear modulus in a nonlinear model is
to modify the pattern of ground movement close to the tunnel. It also increases the
horizontal component ofdisplacement across the whole finite element mesh. There
is no laboratory or field evidence for such a soft initial shear modulus G,,,,,,for
London Clay. It may be that softening the soil in this way represents softening of
the soil due to straining ahead of the tunnel face. A major drawback with a
nonlinear anisotropic model is that there is no information on how G,,,,varies with
strain level.
With both devices the percentage unloading to achieve the controlled volume
loss is reduced from the analyses in Section 2.6.2. This has an effect on the
subsequent soil and lining behaviour. If a lining is constructed early in an analysis,
then the growth of any zones of developing soil plasticity will be arrested, which
in a dilatant soil will reduce the magnitude of any pore water pressure reductions
around the tunnel. Also, the sooner the lining is constructed the greater the loads
transferred into the lining by the end of excavation.
Interaction analyses
2.7.1 The influence of building stiffness on tunnel-induced
ground movements
2.7
beam elements in plane strain). The results were synthesised into simple building
distortion assessment charts. These new charts allow the engineer to take account
of building stiffness when assessing the likely distortion and damage expected
from tunneliing-induced ground movements in an urban environment. Two new
soil/structure relative stiffness parameters were defined. The relative bending
stiffness, p* (with units m-'), and relative axial stiffness, a*:
P* =-
EI
E,, 'v4
a" =- E A
E,v H
where E, is a representative soil stiffness, being the secant stiffness at 0.1% axial
strain from a triaxial compression test of a sample taken from a depth 212, and H
is the building half width, Bl2.
Each analysis gave a building settlement profile and building horizontal
displacement profile. Examples for 60m wide buildings with zero eccentricity
above a 20m deep tunnel are shown in Figures 2.32 and 2.33 respectively. These
were interpreted in terms of building distortion parameters: a sagging deflection
ratio (DR,,,) combined with a compressive horizontal strain (E,,) , and a hogging
deflection ratio (DR,,,) combined with a tensile horizontal strain (c,,) for the
building. These parameters are often related to limiting tensile strain in a building
in order to predict the expected damage due to such a combination of distortions
(Burland and Wroth (1 974)).
centre line
E12 = 30m
centre line
' on
Figure 2.33: The effect of a
horizontal surface displacement
Tunnels / 65
In association with these analyses the greenfield situations were analysed for
comparison. Greenfield settlement and horizontal displacement are commonly used
for damage assessment, and it is relatively cheap and simple to acquire such
predictions empirically. For each building size and location (B and e respectively),
the greenfield ground surface movements were used to obtain greenfield values for
deflection ratio and horizontal strain. These represent the distortion parameters for
a completely flexible building. Potts and Addenbrooke (1997) therefore
investigated the degree of modification that buildings with different relative
stiffness (p' and a*) and relative position (elB) make to the greenfield values of
distortion. Modification factors, M, were defined which gave a quantitative
measure of the degree of modification for a given building. The sagging deflection
ratio for an analysis with a building was divided by the equivalent distortion
parameter from the greenfield analyses, giving the modification factor MDR,,,.
Similarly, the horizontal compressive strain for an analysis with a building was
divided by the equivalent distortion parameter from the greenfield analysis, giving
modification factor M,,. For those buildings which lay in part or in total in the
hogging region the modification factors for hogging deflection ratio, MDR,,, ,and
tensile horizontal strain, M,,, vvere obtained in the same way.
It was determined that in the likely range of true building stiffness the relative
bending stiffness controlled the degree of modification to the deflection ratio (i.e.
vertical settlement profile); and the relative axial stiffness controlled the degree of
modification to the horizontal strain (i.e. the horizontal displacement). Potts and
Addenbrooke (1 997) therefore plotted the modification factors for deflection ratio
in both sagging and hogging forms, MDR,,, and MDR,,,, againstp*for each elB. The
modification factors for horizontal strain (E,) in compression and tension, M",,, and
M,,, were plotted against a' for each elB. Empirical design curves were fitted
through the data and these are reproduced in Figure 2.34.
l
a) Deslgn curves to deterrmne deflect~onrat10
06
The new modified empirical damage assessment procedure for a given building
therefore requires:
- The relative bending and axial stiffness for the building in question. As a first
These modified parameters give the likely distortion to the building, taking account
of the building stiffness.
- a case study
The parametric study presented in Section 2.7.1 dealt with arbitrary buildings
above typical tunnels. This section presents the results from the finite element
analysis of one particularly sensitive building affected by construction of the
Jubilee Line Extension in London - the Treasury building at Westminster. The
building had been monitored during the project. Results are given here from those
analyses that were performed to model the tunnelling-induced response of the
building. To mitigate the effects of tunnel construction an extensive programme
of compensation grouting was implemented at the site and this too has been
analysed. Only a summarised description of the analyses and the results are
included here, further details are given in Standing- et al. (1998).
The Treasury is a massive stoneclad brick-masonry structure,
approximately 2 10 m long and l00 m
wide with four storeys above ground
and two basement levels. The
foundations consist of strips and pads
connected by an unreinforced
concrete slab founded in the Terrace
Gravels which overlie London Clay.
The top of the foundation is
approximately 6 m below ground
level. As Pa* of the Jubilee Line Figure 2.35: Plan of Treasury site
Extension project, two running
tunnels were excavated under one corner of the building (shown in plan on Figure
2.35). The westbound tunnel was the first to be excavated. Following this there was
a rest period before compensation grouting and excavation ofthe eastbound tunnel.
Compensation grouting was implemented after driving the westbound tunnel and
Tunnels 1 67
Tunnels / 69
Tunnels 1 71
1;
8; -,,
Figure 3.18 shows a cross section through the site of the new Jubilee Line
Extension (JLE) station at Westminster which was completed in 2000. Very close
to the station excavation are two underground tunnels and the Big Ben clock tower.
There was obvious concern over the potential movements of the clock tower.
Although this is a rather special case, it is riot an uncommon situation in a crowded
city such as London.
Quite often developments are proposed which are close to other buildings
andtor services. It is therefore necessary to make an assessment of potential
damage or distortion. The distortion of a tunnel may prevent an underground train
from running, or at least mean that speed restrictions have to be imposed, since
tolerances are generally quite tight. In some instances they may be no more than
a few millimetres. In addition, there may be machinery, such as an escalator, close
by which is very sensitive to movement. If distortions become too large certain
parts can fracture in a brittle manner and suddenly jam the escalator, potentially
causing loss of life.
There are a number of methods of assessing building damage (e.g. Boscardin
and Cording (1989), Burland (1995)), but to make an assessment it is necessary to
know how a structure will distort. Therefore in any analysis only soil models which
are capable of making reasonable predictions of the magnitude and pattern of
movement should be used. This normally rules out any constitutive model which
incorporates linear elasticity (including linear elastic perfectly plastic models), see
Section 3.4.5.
To make an assessment of the
A
movement of an adjacent structure,
this structure has to be modelled in an
appropriate way and in sufficient
detail. For example, if the object of
the analysis is to estimate the
potential distortion of an adjacent
segmental tunnel, it is pointless
modelling the tunnel as a stiff ring.
Equally, if the tunnel is modelled as
fully flexible then the predicted E'
(10storey)
distortions may be excessive and
YO*
unrealistic. In order to be able to
make reasonable predictions, special Figure 3.60: Plan of the site for the
techniques have been developed to
Waterloo International Terminal
allow rotation ofjoints between lining
segments once limiting conditions are exceeded, see Section 2.4.5. This may be a
limiting stress or some other criteria.
Hight et al. (1993) present a good case history of where accurate predictions
of tunnel distortions were necessary. Figure 3.60 shows in plan the site of the
Waterloo International Terminal in London. Beneath the basement are two
Bakerloo Line Underground tunnels and to the side of the excavation are two
R
PI-ad
Jubilee
had a deeper basement than Figure 3.61: Cross-section of the site for
the scheme that was
the Waterloo International Terminal
built, which made the base of
the excavation come to within less than 2m of the crown of one of the tunnels.
Plane strain finite element analyses were used to develop methods of construction
which limited the distortion of the tunnels and to prove to third parties that there
were unlikely to be any adverse effects. If movements had been excessive, at the
very least restrictions would have been imposed on trains running in the tunnels.
As an example of the results from this study, predicted deformations of the
station and running tunnels of the Bakerloo Line are shown in Figure 3.62. These
predictions were made before construction. However, during construction the
deformations of the tunnels were monitored, the recorded movements are also
shown on Figure 3.62. Overall, there is excellent agreement between the
predictions and the measurements.
There are a number of other examples, mostly unpublished, were such detailed
modelling has been necessary. These are truly soil-structure interaction analyses.
3.8
1.
2.
3.
4.
5.
6.
7.
8.
9.
Summary
The design ofretaining walls involves the assessment of forces imposed in the
wall and other structural members and the likely movements of both the
retaining wall and retained soil.
The design of retaining walls has traditionally been carried out using
simplified methods of analysis (limit equilibrium, stress fields) or empirical
approaches. However, such approaches cannot, and do not, provide all the
desired design information.
Numerical methods provide a viable: alternative and have the advantage that
they provide all of the desired design information in a consistent manner.
In reality, all geotechnical problems involving retaining structures are three
dimensional. However, at present, due to insufficient computing resources,
simplifications usually have to be made and two dimensional plane strain or
axi-symmetric analyses undertaken.
Care must be taken when assuming planes of symmetry. it is often tempting
to only analyse a 'half section'. This is only valid if the geometry, soil
conditions, construction sequence and loading conditions are all symmetric.
In practice this is rarely the case.
The lateral and vertical extent of a finite element mesh can have a significant
effect on predictions, if they are not large enough. Consequently, it is
important to insure that these boundaries are placed at a sufficient distance to
have a negligible influence. Unfortunately, it is not possible to give general
recommendations for the appropriate lateral and vertical extent of a mesh
because they will depend on the problem being analysed, the constitutive
models used for the soil and the facet of behaviour under investigation. In
practice it is therefore sensible to experiment with meshes of different size,
unless experience has already been gained of analysing a similar problem in
the past. It is always sensible to add some large elements to the sides and
bottom of a mesh. Relatively few of these elements are needed to expand the
mesh considerably and consequently they do not add a large overhead to the
analysis.
The methods of supporting walls have to be addressed (e.g. details of props,
ties, anchors, berms, etc.), as they can have a significant influence on
behaviour. In particular the type ofconnection between support and wall (e.g.
simple, pin-jointed or full moment) can have a large effect on displacements
and structural forces.
It is important to use realistic constitutive models to represent soil behaviour.
In this respect it is usually necessary to use models that can account for both
nonlinearity at small strains and soil plasticity.
The initial stresses within the ground before construction of the retaining wall
can have a significant influence on wall behaviour. The presence of adjacent
structures (e.g. existing tunnels, deep basements, etc.) will inevitably modify
the state of stress within the ground and these effects must be accounted for
in any analysis.
S=- J
Jf
where:
Jf is the value of J o n the failure surface at the current value ofp' and 8.
-current
(I11.2)
stress state
4.
Cut slopes
4.1
Synopsis
This chapter describes the application of the finite element method to the analysis
of 'man-made' cut slopes (slopes which are formed by geological processes are not
considered here). To perform any meaningful finite element analysis of a cutting,
it is necessary to model the process of soil excavation (see Section 3.7.10 of
Volume 1). This is true even when any other geotechnical structure (tunnel,
embankment, foundation, etc.) in the vicinity of a slope is analysed. In such cases,
to establish initial stresses in the ground, the modelling of slope formation by
excavation is essential.
Cut slopes can be formed in both granular and clayey soils. Granular soils,
which have a high permeability, behave in a drained manner, both during
excavation and subsequently. Consequently, they are much easier to analyse. An
exception is, however, when they are subjected to earthquake loading, when the
granular material may respond in an undrained manner. The superficial stability
andlor modelling of the various drainage measures are often the only problems
associated with slopes cut in granular soils. Clay slopes have proved to be more
elusive, and therefore most of this chapter is devoted to the problems encountered
with the finite element analyses of slopes cut in these materials.
introduction
An excavation in clays of low permeability is likely to be undrained. However, in
me normally or lightly overconsolidated soft clays, excavation may almost be
ained (Chandler (1984a)). These clays are often of moderate thickness and are
ely to have sandy or silty drainage layers which accelerate pore pressure
uilisation, despite their relatively low permeability.
Generally, the excavation process unloads the soil and pore water pressures
become depressed, both beneath and adjacent to the excavation. With time the soil
ells as the pore water pressures equilibrate to the long term steady state seepage
gime, imposed by the 'new' hydraulic boundary conditions on the excavation
. With swelling and equilibration, mean effective stresses reduce and the
state approaches failure. The stability of the slope therefore reduces with
e. A detailed discussion of this process is given by Bishop and Bjerrum (1960).
One of the fundamental distinctions made in limit equilibrium slope stability
analysis is that between short and long term conditions. It is usually assumed that
undrained conditions prevail in the short term. Given that it is not easy to predict
the pore water pressure changes immediately after slope excavation, stability
analysis in terms of total stresses (v, = 0" analysis) is often undertaken, in which
the undrained soil strength is assumed to be that existing prior to the formation of
the slope. in the long-term all excess pore water pressures generated during
excavation have dissipated, and the pore water pressures assumed to be in
equilibrium with the newly establishled hydraulic boundary conditions can be
determined. Consequently, an effective stress analysis can be undertaken.
Intermediate term situations cover the period between short and long term.
Unfortunately, the pore water pressure distribution during this transition period
cannot easily be calculated, and it is therefore difficult to carry out effective stress
stability analysis by the limit equilibrium method. However, finite element
analyses, which incorporate coupled consolidation/swelling, readily predict these
pore water pressures, and this is one reason (among others) why the finite element
method is particularly suited for this type of boundary value problem.
When analysing clay slopes, it is important to distinguish between the different
types of clay in which a slope is cut. The choice of constitutive model will largely
depend on the clay type. For example, stiff plastic clays are brittle and prone to
progressive failure. The mechanism of progressive failure is complex and
impossible to analyse without the incorporation ofstrain-softening in the analysis.
On the other hand, low plasticity tills do not usually soften to a significant extent,
and 'simpler' elastic-perfectly plastic constitutive models may well be employed
in analyses of slopes cut in this type of material. Soft normally to lightly overconsolidated clays usually undergo a substantial amount ofplastic straining during
the process of excavation. Then a constitutive model which incorporates plastic
behaviour pre-peak may be essential.
In this chapter examples of finite element analyses of slopes cut in the various
clay types are presented. Being more straightforward, the analyses of 'nonsoftening' clays are discussed first. Softening' analyses require a substantial
amount of computing effort, and they will be outlined later, after the mechanism
of progressive failure has been discussed in some detail.
4.3
'Non-softening' analyses
4.3.1
Introduction
The finite element method has only recently become a powerful alternative to the
limit equilibrium method of slope stability analysis. The recent papers by Griffiths
and Lane (1999) and Naylor (1999) outline the various techniques involved in the
analyses of slopes, and summarise many advantages of the finite element method
in conjunction with an elastic-perfectly plastic soil model of the Mohr-Coulomb
type.
Finite element analyses of this kind have been used here to shed further light
on the mechanisms taking place inside a slope since its formation. As it will be
Cut slopes 1 1 2 7
seen later (Section 4.5), these mechanisms can be extremely complex in cut slopes
formed in strain-softening materials. Thus 'non-softening' analyses provide a
useful reference point against which more complex slope behaviour can be judged.
Property
In-situ clay
Compacted f i l l
Discontinuity strength
c,'=OkPa, q,,'=20
no discontinuities
t
Bulk peak strength
cP1=7kPa,q,,'=20
Residual strength
(E;),
=0
- 5%
as peak intact
as in-situ clay
Permeability, k (mls)
varies
not relevant
o,50i
i.:
l_" - - - - - _ - - - _"LsY~!?!!u~~.:Io.?o..
- --- - - --
0.25~
P
Time since excavation, t (years)
Or-=
a)
3 ,,
10
5
g 0.50
Kpl.0
B
a
IS
20
25
30
35
40
45
50
K71.5
Ko=2.0
1.00
:;i
f 0.25
K,, .
c*
c*8
increasing with time, whereas in the thin rupture zone the soil is dilating and the
pore water pressures are reducing. At collapse the rate of decrease in pore water
pressure in the rupture zone is exactly matched by the increase due to swelling. The
generation of negative pore water pressure is controlled by the permeability of the
clay and the thickness of the rupture zone. This accounts for the gradual increase
in displacement, which is predicted with v = p' once a rupture surface has formed,
and for the higher average pore water pressure at collapse. To check the validity
of the latter statement, an attempt was made to restart these analyses but with the
angle of dilation, v, reduced to zero. In each case it was not possible to maintain
equilibrium of the solution and very large displacements occurred. The pore water
pressures generated by dilation essentially act as a partial brake on post rupture
movements. This effect is of great practical significance, since it helps to prevent
rapid post-collapse movement of slides in strain-softening clays (Vaughan (1 994)).
In reality, dilation would be restricted in the field to a much thinner rupture
zone than that predicted by the analyses, where the rupture surface is
approximately half an element thick (typically 0.5m for the current mesh).
Equilibration of pore water pressure would occur more rapidly in the field. Thus,
if it is assumed that dilation occurs post-peak and only on thin shear surfaces, as
occurs in laboratory tests (Sandroni (1977)), the analyses in which v=Oo are likely
to give a better representation of undrained effects in clay in which discontinuities
develop post-peak, unless an unrealistically large number of elements are used to
allow a thin rupture zone to develop. Consequently, as it will be seen later, for
analyses involving strain-softening, an angle of dilation v=OOwas adopted.
The angle of dilation also affects the position of the rupture surface predicted
and the stresses acting on it. The angles which the predicted rupture surfaces
(velocity characteristics) make with the plane on which the major principal stress
acts are &(45"+v/2). The shear stresses acting on these rupture surfaces are given
by:
c'cosp'cosv+ Q,', sinp'cosv
Zf =
1 - sinp'sin v
When v=pl, Equation (4. l ) reduces to the Coulomb failure criterion, namely:
r, = c' + Q; tan p'
(4.2)
If v#y~',the strength on the rupture surface from Equation (4.1) does not reduce to
the Coulomb criterion. For instance, if v=OO,Equation (4.1) reduces to:
(4.3)
For clays, the'differences are small, but they need to be taken into account
when comparisons are made between the results of finite element analyses and
limit equilibrium calculations in which the Coulomb equation is adopted (see Potts
et al. (1990)). If the rupture surface is kinematically constrained, then the
differences are likely to be mainly in the strength on the rupture surface. If the
rupture surface is not constrained, the differences are likely to be mainly in the
location of the rupture surface itself.
Cut slopes 1 1 31
v=cp'V=OAs noted previously, the rupture
f
surfaces predicted by the nonsoftening finite element analyses
depend on v. The two surfaces
predicted for v=@ and v=OOare shown
in Figure 4.3. Also shown is the
critical circular slip surface obtained
from a limit equilibrium calculation Figure 4.3: Comparison of predicted
using the Bishop rigorous method of
rupture surfaces from 'nonslices (Bishop (1955)). In the cut
softening' analyses
slope problem there is little kinematic
restraint on the position of the rupture surface, and it is therefore not surprising that
the rupture surfaces from the finite element analyses are dependent on v. It may be
noted that at the soil surface the major principal effective stress is vertical and the
rupture surfaces should be inclined at an angle of (45" - v12) to the vertical.
Inspection of Figure 4.3 shows this to be so.
From the limit equilibrium ,
analysis the average pore water
25
l*
pressure
ratio
on
the
slip
surface
was
g
v,,' = 0.233 at collapse. This is in $ 2 0
good agreement with the finite
element results (K*=0.235 for v=OO
Q
and r," = 0.25 for v=#). The average a 1s
L I ~ eqlhbnum
I ~
0 FEM - w ~ t hdllat~on
values of the shear and normal stress P
D FEM - no d~latlon
acting on the rupture surfaces at 3
E
30
35
40
45
50
failure are plotted in Figure 4.4 along P
Average normal effectrve stress, c' (Ha)
with Equations 4.2 and 4.3. As would
be expected, the results from the finite
Figure 4.4: Average shear stress at
element analyses with v=OOagree with
collapse predicted by 'nonEquation 4.3, whereas the analyses
softening' FE analyses and limit
with v=cpl and the limit equilibrium
equilibrium analysis
analysis agree with Equation 4.2.
In the softening analyses reported later, Equation 4.3 is used. Since there is
little apparent kinematic constraint on the rupture surfaces, and v=OOis assumed,
a small difference between the predicted rupture surfaces and those in the field is
likely.
slopes cut in these materials may achieve long-term pore water pressures either
during excavation or within a few months after excavation. The time required for
the pore water pressures to equilibrate to the steady-state condition is important
when designing a cut slope. Given that stability of a slope reduces with time, it
may be wise to either measure the pore water pressures during excavation, or
design slopes in soft clays on the basis of long-term conditions, even when the
slope is only temporary.
It has been appreciated for some time (Tavenas and Leroueil(1980)) that partial
drainage can play a significant role in short term problems involving embankments
on soft clays, where the measured pore water pressures during the early stages of
relatively rapid loading fell short of those expected for undrained conditions.
However, while Chandler (1984a) has argued that soft clays are often of moderate
thickness and are likely to have sandy or silty drainage layers, which accelerate
pore water pressure equalization despite their relative low permeability, Tavenas
and Leroueil (1980) concluded that these phenomena may be due to 'nonlinear
consolidation'. Namely, soft clays have high initial shear and bulkstiffnesses when
loaded (or, in the case of cuttings, unloaded) from their in-situ conditions.
Although these stiffnesses reduce rapidly as strains grow, the operational 'linear'
equivalent consolidation (swelling) coefficients would have been very high initially
as construction (excavation) proceeds, and when a significant factor of safety
operates in soft clays (further away from yield').
To investigate the possible effects of a high soil stiffness at small strains on the
time required for the pore water pressure equilibration, a series of finite element
analyses of a hypothetical slope cut in a soft clay was undertaken.
pressure (kPa)
A form of the modified Cam clay model (see -50 0 Pore50water100
150 200 250
\L'"'""""~'"'
Chapter 7 of Volume l), which had a MohrWeathered crust
Coulomb hexagon and a circle for the shapes
2
k
of the yield and plastic potential surface in
the deviatoric plane respectively, was used to
represent the soft clay (and its weathered
crust). The elastic behaviour was modelled in
two ways. In the first approach, the basic
critical state formulation, in which both bulk
and shear moduli are derived from the slope
of swelling lines, K, assuming a constant
value of Poisson's ratio, p, was utilised. In
the second approach the nonlinear elastic
behaviour proposed by Jardine et al. (1986)
Sand
(see Section 5.7.5 ofvolume 1) was used. As
Figure 4.5: Observed and
this model is able to capture the high soil
stiffness at small strains for stress states adopted pore water pressure
profile
inside the yield surface, it is more reakistic.
" " " I
33 ,
-a-Triaxial compression
-c Triaxial extension
I
W
Lab tests:
---Depth 5.5m
......,.,...Depth 8.5m
Depth 12.5~11
..
-100
100
200
-.
300
50
l0
.S-
20
'
40
FE prediction:
-+Loading
Critical State L i e
compression
- - - - Critical State Line
extension
-c Unloading
150
-100
-.
Lab tests:
--- Depth 5.5m
............Dep(h8,Sm
- .- .- Depth 12.5~11
Venical effective stress, a,' (kPa)
10
100
1000
The most critical issue when dealing with the modelling of soft clays is usually
their undrained shear strength, S,, . While the undrained shear strength is not an
input parameter to the modified Cam clay model, it can be calculated from the
input parameters and the initial stress conditions (see Section 7.9.3 of Volume 1).
The model parameters listed in Tables 4.2 and4.4, in combination with the profiles
of the overconsolidation ratio, OCR, (Figure 4.10) and the coefficient of earth
pressure at rest, K,, ,(Figure 4. l l) have generated the undrained strength profiles
given in Figure 4.12. It can be seen that predictions depend not only on the type of
the test modelled (triaxial compression or plane strain), but also on the elastic
component of the model used in the analysis. However, the trend observed in both
field (vane tests) and laboratory (triaxial compression tests) is reasonably captured.
Table 4.2: Soil properties assumed for soft clay (modified Cam clay)
Poisson's ratio, p
10 15 20 25
30 35 40
0.5
1.5
2.5
1o ' ~
0 10-'O
- o
-wiaxial compression
- - - plane s w m
Q
16t
Weath ed Clay
Stiff Clay
adopted
22
0 Scale 10m
1 ,,
excavation
Table 4.46: Coefficients and limits for nonlinear elastic secant bulk
modulus expression
c',
,,
8
2,
-A
-A
L
-A-OS
-B-07
4.4
Progressive failure
"";p
behaviour continues as the beam is loaded (i.e. displaced) further. Figure 4.2 1c
shows the situation when peak shear stress is mobilised everywhere along the
beam. With further displacement (see Figure 4.21d) the shear stress drops from
peak towards residual. Once enough displacement is applied to reach residual
conditions (see Figure 4.2 le), the stresses remain at this value with further loading
(see Figure 4.2 1f).
Peak
Oo
ti
p]
,
6
,
8
Distance
io@)
,
2
Dtstance
IO(~)
, DtsIanee
IO("')
, Dislance
IO(~)
Residual
L
, Distance "
to("')
0]
d)
c)
]
0
i " ~ - - - - - -
'0
e)
, Distance
10@)
O
0
Distance v
10(~)
--Residual
Distance
10("')
;. . .,..
2
.,6
Distance
IO(~)
Peak
Residual
10
"
'
"
Peak
Distance
Iso~
Peak
Residual
Distance
10(~)
Distance
IO("')
d)
c)
-- Peak
...............
- -
Residual
The data in Figure 4.24 is replotted in Figure 4.25 with the force F divided by
the contact area between the beam and surface to give a plot of average shear
resistance against displacement A. Also shown on this figure are two further
results. One is for the same compressible beam as above, but loaded at its centre
instead of at its end (see inset in Fjgure 4.25a). The other is for a similar
gm-
Displacement, A (mm)
Displacement, A (mm)
4.5
'Softeningpanalyses
4.5.1
Introduction
The delayed collapse of old British railway cutting slopes in stiff plastic clays, by
the formation of deep-seated slides, has been studied extensively (see e.g.
4.5.2
p\r
,C;
The model requires the specification of peak (q,,', c,') and residual (v,', c,')
strength, the angle of dilation, v, pre-peak stiffness, E, stiffness in unloading, E,,,
and the rate at which strength is lost with strain ((EJ), , (EJ'),).As the analysis
involves swelling, it is also necessary to specify a permeability, k.
The stiffness prior to failure is given by a simple elastic model in which
Poisson's ratio, p, is constant, and Young's modulus, E, varies with mean effective
stress,pl, but not with shear stress level. No distinction is made between unloading
and loading. Consolidation and swelling behaviour indicate approximately the
same modulus in stiffclays, and therefore the Young's modulus used in the present
analysis is based on the appropriate swelling modulus. The parameters used are
listed in Table 4.1.
4.5.3
Implications for convergence
Monitoring of convergence is difficult for analyses involving strain-softening. The
norms (see Section 9.6 of Volume 1) of the iterative nodal displacements, loads,
flows and pore water pressures were all kept less than one percent of the norms of
the associated incremental values. However, this procedure alone was not adequate
and, in addition, the residual stresses at all integration points were monitored and
kept small. For the early stages of an analysis, residual stresses were kept below 2
kPa. However, when collapse was approached more stringent conditions were
introduced and stresses were kept below 0.1 kPa. To achieve this level of accuracy,
small time steps (typically 0.0025 years) and a large number of iterations (typically
200) were required towards the end of an analysis. The time at collapse was
deduced by plotting horizontal mid-slope displacement against time. Once the
collapse time was exceeded, the solution became unstable, since the slope could
not then be in equilibrium. This was always confirmed by running the final
increment for a small time step of 0.0025 years and a large number of iterations
(typically 400). Instability was indicated when deformations increased according
to the number of iterations, without an improvement in convergence (KovaCevid
( 1 994)).
Brown London Clay (see e.g. Skempton (1977)), and that is why slopes in this
material have been chosen for this study.
100
3
6
'
.I-
Oo
100
200
300
(U,'+U,,)I~(Wa)
Peak strength
Residual strength
Measurements of residual strength are
summarised in Figure 4.28. Both ring
shear and field data are shown. The L
strength back-calculated from field slips
is slightly higher, probably due to the less
planar shear surfaces usually formed in +
the field. The field strength of cr1=2kPa
and qrl=l3" is adopted.
8
5
8
-9
StifJess
The clay in a slope swells with time, and
the strains and strain-softening are
Strongly influenced by the Strain energy
released by Swelling. Very little
experimental data is available for
100
200
Vertical effective stress, cr,' (kPa)
Permeability
The coefficient of permeability
Coefficientof permeability, k (tn/s)
was assumed to vary with depth as
shown in Figure 4.30, where field F:$'$
data for both the Upper Lias Clay
and the London Clay are also
shown. This permeability,
together with the swelling
modulus from Figure 4.29 for
K,,=1.5, gives a coefficient of
swelling c,,=2.7m21year at 5 m
depth. Walbancke (1976) quotes
field rates of swelling equivalent Figure 4.30: Variation of permeability
to c.v=3.2m2/~ear
for the Brown with depth - assumed in the analyses
London Clay.
and laboratory data
Rate of softening
The rate at which a clay strain-softens post-peak is difficult to establish. A
reasonable assumption was made based on precedent from other analyses.
Parametric studies, performed as part ofthe present investigation, have shown that
this property, when varied over a realistic range, does not have a major influence
on the results obtained (KovaEevid
(1994)). One or more shear surfaces
75 are likely to form as peak strength is
mobilised in real plastic clays, and .;:
loss of strength post-peak occurs as a d 50 consequence of sliding on these g
surfaces. Such thin discontinuities are
not reproduced in the finite element
0
analysis, in which the minimum
o
5
10
15
20
25
30
Shear strain, y (%)
thickness of a rupture surface is
approximately half the thickness of an
0
0.05
0.1
0.15
element. This must be taken into
Displacement across a 0.5m thick layer, A (m)
(112 the thickness, T, of a typical element)
account when specifying the rate at
which strength is lost post-peak. The
shear stress-displacement plot for a Figure 4.3 1: Assumed stress-strain
shear zone of thickness T in simple and stress-displacement relationship
shear is shown in Figure 4.3 1.
in simple shear
g
C-\
75
50
37 kPa
12 kPa
25
10
15
- Io
10
32s
Scale
20m
Cut slopes / 1 5 1
significance. The figures show that the outer part ofthe slip surface has developed
after 9 years, and that the complete slip has developed after 14.5 years.
Strain wfbiing starts when E:=5%.
It is complete when E,!=20%.
'
---
scale
Scale
I
h) 14.5 years after excavation just before collapse
0
;5m
25m
j
a) 9 years after excavation 0
2:m
Scale
0
25m
Contours of deviatoric plastic strain, EJ' (see Equation 4.5) are given in Figure
4.35. The 5% and 20% contours represent the start and end of strain softening. A
horizontal shear zone propagates from just below the toe of the slope, in a manner
similar to that observed in the field by Burland et al. (1977). The strength acting
on this base shear rapidly drops from peak to residual as the rupture zone develops.
,,
c*,
-9
2
'
/-
23
1
0.5
1.5
2.5
,,,
8'
0
l
Scale
25m
l
Scale
2qm
The predicted time-scale for the shallow slip to develop (i.e. 1.6 years) is
shorter than is usually observed in the field. However, the prediction involves a
monotonic swelling process with zero boundary pore pressure (equivalent to winter
conditions), with cracks already developed. The time scale in the field may be
controlled by the time taken for cracks to develop, and by the delay before the
slope is subjected to a long 'wet' winter.
""7
c*,2
,:
,:
ii
.g
"
-3
0.75
c*
;i
,,,
g , ,.S
0.3
"
O o 0 a0 p 0 s
20.2
gj
0;
o
a slope
Slope 23.0:1
. 5 : ~-predicted
-predicted
O$
[ O.'-do
g
50
Field
slope obse~ations
4.0: I -predicted
IW
150
200
%.o.l.0.2-*
Y
.$
C!
slope Z.O:I-predicted
o
A
sslop
l o p 2.21
3 . 0 : ~-predicted
predicted
Slope 4.0:L -predicted
3
*
50
Q
3
4
O0
50
100
I50
i ~0.5=0.41
8
E
k
"
F
4
0.3
o.2-
0.1 '-1
-0.1 -
a.2.
-O".0.4 L
after excavatron
steady state
This implies that, although the time to collapse would increase, collapse would still
occur (see Section 4.5.4.6). For collapse to be prevented, the surface suction would
have to be increased still further, so that the long-term value of r,' fell below the
'collapse' range indicated in Figure 4.53.
Figure 4.36 shows a sharp discontinuity in volumetric strain across the
horizontal rupture surface of 2-3%,equivalent to a change of water content of 12%. Such a discontinuity could be detected by site investigation. This would be a
potential way of examining the stability of old slopes. As can be seen from Figure
4.35a7the basal part of the rupture surface forms quite early on. Its formation does
not indicate that collapse will eventually occur, but its presence with continuing
movement on it (which could be
monitored) indicates decreasing
stability. The absence of movement
would indicate stability.
BIOWII London Clay
Figure 4.54 shows a section
through a slip on the M11 near
Loughton in Essex, in a slope 18
Water
Liquid
Plastic Remoulded
years old. Results from tests on
content l i t
limit quick undraiied
shear strngth
samples taken from just above and
W % WC%
W,%
S, kPa
below the rupture surface are given
SampleA 33.2
70
33
26
on the figure. There is nearly a 4%
sampleB 29,4
75
32
62
change in water content and a two
fold change in remoulded undrained
strength across the rupture surface. Figure 4.54: Slip on the east side of
the M I I near Loughton, Essex
This observation is consistentwith the
( 19931
numerical analysis.
4.6
Excavation of cut slopes underwater can be problematic. For example, consider the
situation shown in Figure 4.55a, where a trench is to be excavated in the sea bed.
At the beginning of the analysis, the initial stresses and pore water pressures in the
soil are in equilibrium with the sea level (the mean sea level is often assumed, but
its fluctuation can, in principle, be accounted for). Excavation is then simulated by
sequentially removing rows (or blocks) of elements. However, care must be taken
to properly account for the water pressures applied to the newly excavated soil
surface.
When excavating the first row (row 1) of elements shown in Figure 4.55b, the
process essentially also removes the water pressure acting on the original ground
surface between points A and H. It is therefore necessary to simultaneously (i.e.
over the same increments as excavation occurs) apply a boundary stress over the
newly excavated soil surface ABGH, to represent the water pressure acting on it.
This process must be repeated as each row of elements is removed.
Although the process described above is logical, it is cumbersome and often
r
This arises as many programs
difficult to achieve with some c o m p ~ e software.
4,7
1.
2.
Summary
c',
3.
general trend. Thus collapse will tend to occur in winter, when surface suction
is zero (see e.g. Chandler (1984b)).
The surface hydraulic boundary condition has a strong effect on stability. The
analyses show that an increase in surface suction from l0kPa to 20kPa goes
more than half way to stabilise the 3: 1 slope. Surface pore water pressures can
be reduced by increasing evapo-transpiration through the controlled use of
vegetation, or by surface drainage which reduces pore water pressures below
the depth of the drains.
The analyses show, beyond reasonable doubt, that the mechanism causing
deep-seated delayed slips in these slopes is progressive failure promoted by
swelling. They enable the importance of the controlling variables to be
established, and the effect of possible stabilising measures to be evaluated.
The new and sophisticated techniques of numerical analysis have an
encouraging ability to reproduce observed behaviour, provided that realistic
input data is used. No assumptions need then be made about behaviour
mechanisms, as these are established by the analysis.
No shallow slips were predicted in the analysis in which uniform soil
properties were assumed. However, shallow slides can be reproduced by
differential strains and progressive failure at the base of a superficial layer of
clay of higher permeability due to shrinkage cracks, although this is not the
only possible reason for such slides. This mechanism was inhibited by high
initial values of K,, , when early formation of a horizontal rupture surface
probably reduced the strain energy stored in the superficial 'cracked' layer.
Care should be exercised when widening amotonvay cutting by removing the
toe. The analyses show that widening greatly accelerates delayed failure,
although the amount of progressive failure remains about the same. The
increasing surface boundary suction produced by management of slope
vegetation, or surface drainage which reduces pore water pressures below the
depth of drains, may halt development of slips in the longer term.
Embankments
5.1
Synopsis
Many attempts have been made to predict the behaviour of various types of
embankments. The finite element method is a powerful tool for such predictions.
This chapter deals with applications of the finite element method in the analyses
of this type of structure. Examples of modelling are presented and discussed.
5.2
Introduction
Embankments 1 167
5.3
5.3.1 Introduction
According to the nature and position of the watertight element, there are two main
types of rockfill dam: (i) that containing a relatively impervious internal earth core,
either wide or narrow, central or inclined, and (ii) that with an upstream impervious
membrane, made either of concrete or asphalt. Usually rockfill and clayey fill are
involved in the former and just rockfill in the latter.
Being derived from rockfill only, dams with an upstream impervious membrane
are simpler to analyse and therefore an example of finite element analysis of this
type of dam will be presented first. This membrane is usually thin and flexible and
has little bearing on rockfill deformation during first reservoir impounding and
subsequent operation. First, some general features involved in embankment dam
modelling will be briefly discussed.
5.3.2 Typical s t r e s s p a t h s
A great variety of stress paths occurs in embankment dams (and their foundations).
They are often accompanied by a large rotation of principal stresses and may be
particularly complex during first reservoir impounding and subsequent operation.
It is difficult to reproduce all these real stress paths, even using sophisticated
laboratory equipment such as the hollow cylinder or true triaxial apparatus.
Nevertheless, it is common to
describe the behaviour of fills using
standard laboratory tests. This is
particularly SO for rockfills. Because
?for oned~menstonal
")::;Edopug
compresston
of their coarse nature, the tests have to
be performed on large samples and
200
4w
600
P'
are therefore expensive and
labourious. Consequently, their
number and types are limited.
Sometimes both large triaxial and
oedometer tests are performed for
rockfills, but quite often only results
from one of these two tests are
available. Therefore, it is necessary to
establish which of these is more
representative to simulate rockfill 4
behaviour in an embankment dam.
It can be argued that the
200
400
600
oedometer test, providing a constant
P'
stress ratio path, R = aI1/a,'
= lIK,,
can better capture fill behaviour, at Figure 5.1: Stress paths for three
least during the construction stage.
types of rockfill dam
Although the principal stress ratio, R,
(after Charles 119 76))
5.3.3
Choice of constitutive models
It has been repeatedly shown (see e.g. Duncan (1992)) that the most influential
factor in the finite element analysis of embankment dams is the modelling of the
stress-strain behaviour of the fill by an appropriate constitutive law. However, in
spite of the diversity of the stress-strain relationships being used, reasonable
agreement has usually been found when the results of finite element analyses
(typically movements) have been compared with field observations. This is not
surprising bearing in mind the fact that most of the analyses were done after the
field measurements had been made, resulting in after-the-event or so called 'Class
Cl' predictions (according to Lambe (1973)).
A review of different constitutive laws used in the numerical analysis of
embankment dams can be found in publications by Naylor (1991a) and Duncan
(1992). Both elastic and elasto-plastic formulations are available. In the following
the most representative models will be reviewed briefly, with emphasis on their
validity and parameter derivation. Some of them have already been described in
Chapters 5 and 7 of Volume 1 of this-book.
Embankments 1 l69
1 pPalD
(5.1)
where p,, is atmospheric pressure in the same units as G',C and D are
dimensionless parameters. Skinner (1 975) used the above relationship to derive the
closed form solution for the settlement profile on the centre line of an embankment
of height H constructed of rockfill of bulk unit weight y:
where S is the settlement of a marker placed at a height h above the foundation and
C , = ClpuD.Settlements in this case are treated as one dimensional and vertical
effective stress is presumed to be equal to the overburden pressure (pore pressure
is zero). The former results in an underestimate of settlements, the latter in an
overestimate. The two effects are likely to be relatively small and compensating.
By differentiating Equation (5. l), the expression for the tangential constrained
modulus, E,', can be obtained. The tangential Young's modulus, E,, can be derived
using the value of Poisson's ratio, p, estimated from the K,, value (i.e.
p=K,,/(I+K,,)) from elastic theory, where K, is related to the angle of shearing
where R, is the stress ratio at failure with a value always less than unity, p,, is
atmospheric pressure, M, is a modulus number, and n is an exponent determining
the rate of variation of the initial tangential modulus, E,, with confining stress, G,'.
The curved nature of the failure envelope for granular materials (ci=O) can be
accounted for by using the following logarithmic expression for p':
where q,,' is the angle of shearing resistance for a confining stress u,'=p,, and Aq'
is the reduction in p' for a tenfold increase in IS,'.
Instead of modelling the tangential Poisson's ratio, p,, Duncan et al. (1980)
preferred to model the volume change behaviour of soils by an exponentional
relationship between the bulk tangent modulus, K,, and confining pressure, c,':
(5.5)
K, = M , P , ( ~ 1 P,)"
where M, and m are dimensionless mode%parameters with a similar function as M,
and n in Equation (5.3).
It can be noted that seven parameters are required to characterize the model. If
a stiffer response on unloading is to be modelled, the additional parameter M,, (an
increased value of M,) can be employed. All parameters can be evaluated using
data from conventional triaxial tests.
The hyperbolic model has been used extensively in the finite element analyses
of different geotechnical problems, particularly embankments. Consequently,
considerable experience with the model has been accumulated and this has enabled
Duncan et al. (1980) to compile parameter values for different fills taking into
account relative density (compaction), grading, particle shape and mineral
composition. This has been the main advantage of the model, especially when
applied to the analyses of rockfill dams, given the difficulties associated with the
laboratory testing of rockfills.
Embankments / 17 1
A nonlinear elastic model, similar in formulation to the above hyperbolic
model, has been used at Imperial College for numerical analyses of a number of
embankment dams (Wamza (1976), Dounias (1987), Kovaeevid (1994)). The
drained elastic parameters vary according to the following expressions:
E,, = H E,
(5.9)
where E and E,, are the drained Young's moduli on first loading and
unloading/reloading respectively, p is the Poisson's ratio, p, is atmospheric
pressure, and E,, , A, B, C, H, I, L are model parameters. The stress level, S,
represents the proportion of the shear strength mobilised at the current mean
effective stress, p', and varies from zero, when the state of stress is on the
hydrostatic axis, to unity, when it is on the failure envelope given by cohesion, c',
and angle of shearing resistance, p' (see Appendix 111.l).
Problems can be experienced with this type of model when deriving the model
parameters to satisfy both oedometer and triaxial test data. It is often not possible
to accurately model results from both types of test. Bearing in mind the above
discussion about typical stress paths, emphasis is often placed on fitting the
oedometer stress-strain curve. When results of analyses are compared against field
measurements, it emerges that the predicted settlements are in reasonable
agreement with the measured ones, but horizontal displacements are overpredicted
(see e.g. Potts et al. (1990), Kovatevid (1994)).
Embankments / l 7 3
Elasto-plastic models for sands usually separate the effects of consolidation and
shearing. During consolidation deformations are governed mainly by crushing and
yielding of interparticle contacts. During shearing to high stress ratios deformations
develop also due to sliding and rolling. To account for different plastic
deformations during consolidation and shearing, so-called 'double hardening'
models have been proposed. A double hardening model of the form proposed by
Lade (1977) has been described in Section 8.5 of Volume 1.
5.3.4
Embankments / l 75
~;~~~$;;t;;~''~ter
and sand waste material (Charles and Watts (1985)). These included investigations
of compressibility in Im diameter oedometer tests and strength (and
compressibility) in 0.23m diameter triaxial compression tests. The data from these
tests have been used to derive the parameters for the two constitutive models (see
Tables 5.1 and 5.2) to give the best fit.
Comparison of predictions from the two constitutive models with results from
triaxial and oedometer laboratory tests for the rockfill are shown in Figure 5.6.
Inspection of this figure indicates that the fit from the Lade's model is good for
both the triaxial and oedometer tests. The fit for the nonlinear elastic perfectly
plastic model is not so good but is still satisfactory, especially at low stress levels.
Lade's elasto-plastic
.-.,..
-.- ....-.
-.-,
'4: - - -.-?,--,.,-C-.-.-
.L.
A'"
C.*
----
.r
r r r r
C o a r , peravc
V 2SkPn -Modcl
%+Pm -Modcl
697kP.-Modcl
-697kPa-TcW
.-.%.-'
'
I
f-
'
,*-,
.-m
_.-'
_ ...
"
0 1 2 3 4 5 6 7 8 9 1 0
Axialstrain (%)
a) Thiaxial tests
b) Oedometer tests
Embankments 1 1 77
Table 5.7: Nonlinear elastic model parameters for Roadford rockfill
and sand waste fill
concrete membrane was modelled as a separate zone (see Figure 5.7b). Because of
the very high viscosity of bitumen, the asphaltic concrete of the membrane was
modelled as an undrained granular soil in terms of effective stresses, assuming the
properties of the sand waste fill and the drainage materials (K, = 100 K,,,, , see
Section 3.4 of Volume 1).
Table 5.2: Double hardening model parameters for Roadford rockfill
and sand waste fill
Embankments 1 179
drained conditions were
assumed in the fill material.
V4hile a static water pressure,
consistent with a water table,
was maintained at foundation
Initial and f d wat
level behind the grout curtain,
an increase in water pressure
in front of the grout curtain by ,jW
upstream ~ f g r oC~& t
due to impounding was
Figure 5.8: Simulation of reservoir loading
modelled.
during impounding
.,
"
>;.
\\
<:
,
': .
.X
-W
BI
-1
g,
-M)
I !
5.3.6
60
- - Double-hardeningmodel
-Observed deflection
Imvoundinp,level 14.5111above toe
Example: Analysis of
old puddle clay core
dams
Impounding level 25.0111above toe
5.3.6.1 in troduc tion
100
o
A large number of embankment dams
with central cores of puddle clay were
built in Great Britain in the 19'h
century. As the puddle clay was
60;
expensive, the cores were made as "
thin as possible. This made the cores
- Impounhng level 30.5111above toe
loo
vulnerable to cracking by hydraulic
fracture if they were subjected to full
Figure 5-1 1: Comparison
reservoir pressure on their upstream
boundary, where the seepage pressure predicted and t r ~ a s u r e d
may exceed the normal total stress by of as~halticconcrete membrane
an amount greater than the tensile
strength (which is usually negligible).
It has been deduced that hydraulic fracture of the core was the most probable
cause of the failure of the Dale Dyke dam at the end of first impounding (Binnie
(1978)). Dounias et al. (1996) have performed finite element analyses of the
construction and first impounding of this dam to see whether this mode of failure
could be recovered. Their findings will be briefly presented in the first part of this
section.
There is little information of how these old dams were built. In particular,
shoulder fill was often variable and placed in rather thick layers with little
compaction. This, in combination with extremely soft puddle cores, may give rise
to significant crest settlements as a result of operational cycles of drawdown and
re-filling (Tedd et al. (1994)). KovaEeviC et al. (1997) have analysed the
movements of four embankment sections to examine whether these measured
movements could be recovered, and their implications for long term stability.
Results for one section (Ramsden dam) will be briefly discussed.
Embankments / l 8 1
8# , \
8 I, 0
I,
$1
2
g
;::1
Embankments 1 183
Pressure (m/ml)
collapsed during a heavy storm,
and it is quite likely that the last
few metres were filled rapidly.
This scenario was investigated in
a further analysis where the last
7.5m was filled quickly, assuming
that the core behaved in an
undrained manner. The results are
presented in Figure 5.16. It can be
seen that minimum total stresses
at the upstream core boundary are
equal to the reservoir Pressure at a Figure 5.1 6: Minimum total stress near
between 7 and 12m
the upstream boundary of the core,
the crest. At a small distance
effect of undrained impounding
inside the boundary they are lower
(Dale D ykel
than the reservoir pressure almost
along the whole length of the core. It seems that the rapid undrained final
impounding may induce hydraulic fracture around a depth of 10m from the crest.
Once formed, it will propagate rapidly downstream, where the total stresses are
lower, as an undrained fracture.
predictions for conventional laboratory soil tests, on Figures 5.18a and b. The core
was simulated using a model based on modified Cam clay (see Section 7.9 of
Volume l ) and the fill by Lade's single haridening model (Lade and Kim (1988)).
This model is different from the Lade's double hardening model used in the
analysis of Roadford dam and briefly described in Chapter 8 of Volume1 . Plastic
yield before failure is still accounted for with this model, but this time utilising
only one yield surface. The results of a cyclic oedometer test on a typical fill from
a mudrock are shown on Figure 5 . 1 8 ~(Tedd et al. (1994)).
The load-unload properties adopted, which dominate predictions during
reservoir operation, were based on this test. The irrecoverable deformations during
operational cycles of drawdown and re-filling are due to the different elastic
stiffnesses adopted during unloading and reloading, not due to plastic straining.
The analyses were performed in terms of effective stress. The shoulder fill and
the foundation were assumed to be free-draining. Coupled consolidation/swelling
was adopted for the puddle clay core, with either a constant permeability or one
varying with the mean effective stress, p' (see Section 10.7.5 of Volume 1):
where k,, = 10v8m/s is the coefficient of permeability at zero mean effective stress
and a = 0.03 m2/kN is a material constant. The analysis using Equation 5.10 gave
a nearly perfect fit with the observations (see Figure 5.19). The degree of
consolidation in the clay core controlled by coupling of volumetric strain and
permeability enabled the time dependent response to load changes to be modelled
in real time. No separate approximate consolidation calculations were needed for
these analyses.
The purpose of the analyses was to examine
Wet boundary
behaviour during operation of the reservoir only.
Thus a simplified construction history was
simulated, involving construction in layers, full Predicted
consolidation of the core, impounding and the pressure
establishment of steady seepage, prior to the
impositions ofthe drawdown and re-impounding
cycles.
The observations and predictions of crest
movement are shown in Figure 5.20. Small
adjustments to the elastic properties shown in
Figure 5.18 were made to recover the difference
in movement observed between the core and the
shoulders. There is no doubt that better
agreement between prediction and observation
Figure 5-19: Seepage
could have been obtained by further adjustments
to the assumptions made. However, the reality is pressure through a narrow
core (lVughan (1994))
unknown and there is a limitation on what can be
achieved without a detailed site investigation.
Embankments 1 185
Predicted
Although the analyses cannot
displacements
predict the extent of the observed
irrecoverable deformation towards
downstream of the whole
embankment crest due to a
drawdown cycle, they show that the 4'
-16.5m
-16.51~1
large and irrecoverable deformations
observed can be reproduced by
a) 1988 drawdown
assuming
- soil parameters which are
consistent with the actual materials.
The average local factor of safety
,
(the ratio of strength mobilised to
b) lggg drawdown
strength available at the current stress
Movement scale (mm)
level) at the end of major drawdown
0
20
40
60
is quite high (about 1.8). Thus large
irrecoverable deformations occur
Stages: 0 - Reservoir full
despite a substantial factor of safety
1 - 1988 dmwdown (16.5m)
2 - Held at drawdo& levei
against overall slope stability, and
3 -Raised to top water level
such large movements do not
4 -Held at top water level
5 - 1989 drawdown (6.0m)
necessarily indicate incipient
6 - Raised to top water level
instability. In the analyses the
irrecoverability arises through the
Figure 5.20: Ramsden dam:
difference in elastic properties
Observed and predicted
assumed in unloading and reloading,
displacements during two
not through plastic strains. Thus
reservoir dra wdo wns
irrecoverability. is predicted
for the
.
second smaller drawdown. Large drawdowns do not eliminate irrecoverable
movements during subsequent smaller drawdowns.
I
5.4
5.4.1
Introduction
Finite element analysis of earth embankments is potentially more difficult than that
of rockfill dams, as an earth embankment is usually derived from clayey fill only
and may sit on a foundation of varying strength and permeability which will affect
its performance. The extreme condition of an embankment placed on a soft clay
will be presented in Section 5.5. Here, only embankments sitting on a relatively
'stronger' foundation will be considered. The first example is a typical road
embankment derived from a stiff clay on a stiff clay foundation, where behaviour
of the embankment is mostly governed by the undrained shear strength of the
foundation material due to changes in pore water pressure near the surface. The
second example is of a large embankment dam founded on a competent foundation
covered by a plastic clay deposit which was left in place below the embankment,
enabling the embankment to fail progressively during its construction.
5.4.2
Modelling of earthfill
During construction it is usually assumed that the earthfill is saturated and will
behave largely undrained. Because of simplicity, analyses of undrained saturated
fill in terms oftotal stresses is quite common (see e.g. Section 5.4.4). However, the
pore pressures could then only be estimated from the total stresses.
Analyses in terms of effective stresses are more tempting because they predict
undrained pore water pressures directly. Besides, in clay fills it is often necessary
to model initial suctions (negative tensile pore pressures). To do this, after the
element has been constructed, the mean effective stressp' can be reset to be equal
to an appropriate initial suction pjo, and the pore water pressure is reset to p, =
pfi,+p,', where p,' is the mean effective stress (equal to the mean total stress, p,)
after construction of the element. By essentially resetting the mean effective stress
p', it is possible to control the undrained shear strength, S,,,ofthe clay fill material.
It is worth noting that, by changing both mean effective stress and pore water
pressure, total stresses remain constant and equilibrium of forces is maintained.
In spite of their low permeability, real earthfills are partly drained at the end of
construction, at least at the outer boundaries of the dam. To model this time
dependent behaviour in terms of dissipation of construction generated excess pore
water pressures, 'coupled' approaches which consider the interaction between the
soil skeleton and the pore water are needed. The problem is solved on the basis of
equilibrium and the constitutive relationships ofthe soil skeleton together with the
water continuity equation (see Section 10.3 of Volume 1). This approach has been
used in some of the examples presented here (see Sections 5.3.6 and 5.4.3).
In the above approach it is assumed that (i) water completely fills all soil voids
(the so called 'coupled saturated' approach) and (ii) water is incompressible.
Nevertheless, these are reasonable assumptions for fine grained fills placed in a wet
compaction state, such as a typical British clay fill derived from a saturated borrow
pit.
In order to model the actual soil compressibility in unsaturated conditions, it is
necessary to use the 'coupled unsaturated' approaches (see, for example, Pagano
(1998)) which consider the actual multiphase character of earthfills (soil skeleton,
pore air, pore water). The problem is then solved on the basis of (i) equilibrium
equations, (ii) water continuity equation, Qiii) air continuity equation, (iv) soil
skeleton constitutive equations, and (v) water retention characteristic curve. By
assuming the air pressure to be equal to the atmospheric value, the air continuity
equation can be eliminated from the set of governing equations.
5.4.3
Example: Road embankments on London Clay
5.4.3.1 Introduction
Foundations for a typical road embankment may vary from an undrained clay to
a drained granular material. During embankment construction on a saturated low
permeability clay, such as stiff plastic London Clay, excess positive pore pressures
are generated. As these pore pressures dissipate, stability should then improve with
Embankments 1 l87
time. Undrained stability in the short-term may be of a concern, and clearly it will
depend on the undrained shear strength of the clay in the foundation. Regarding the
fill material, placement and compaction of clay fills may generate quite high pore
water suctions, particularly in more plastic clays. Thus in plastic clay fills of
modest height, typical of road works, long-term equilibrium pore pressures are
likely to be higher than those at the end of construction. The strength of the fill
may then decrease with time as the pore pressures equilibrate and there is a risk of
delayed slides.
Undrained shear strength
Probable avenge surface value
of stiff plastic clays depends
more on effective stress rather
than water content. There is a
seasonal variation of pore
water pressure in the surface
of a typical British clay
stratum as shown in Figure
5.21, and therefore the
undrained shear strength will
also vary seasonally. Figure 5.2 1: Pore water pressures near
Examples of short-ferm the surface of the ground: slopes in clay
in the UK (Walbancke ( 1 976),
behaviour of embankments
Vaughan ( 1994))
derived from London Clay and
constructed on London Clay foundations in either summer or winter weather
conditions will be now presented (Vaughan (1994)).
5.4.3.2
Material properties
Soil properties for the in-situ weathered London Clay were the same as those used
in the analyses of slopes cut in this material (see Chapter 4). The assumed
properties for the fill derived from London Clay are also listed in Table 4.1. The
higher value of cr=12kPa,
compared to ci=7kPa adopted S, max~al(kPa)
for the in-situ Brown London
Clay, reflects the absence of
seasonal
fissures in the fill. Also, a
reduced pre-peak stiffness and
an increased post-peak
straining to reduce strength
from peak to residual have
p'sm0'+c'cos~'
S.=
been assumed for the fill. The
l-sui+'/3
fill was simulated as a
Figure 5.22: Changes in undrained
saturated clay with an initial
suction as placed to give an
strength of stiff plastic clay due to
changes in pore water pressure near
undrained shear strength of
S,,=60 kPa, typical of a slightly
to the surface of the ground
wet fill. The permeability of plastic clay fill is low, and a constant value of the
coefficient of permeability of k=I .5xlO-'O m/s has been adopted to match, with the
stiffness assumed, typical field deduced values for the coefficient of
consolidation/swelling in the range of c, = 0.5-2.0 m2/year.
The predicted peak undrained strengths (the soil model described in Section 4.5
has been used in these predictions and subsequent analyses), as would be measured
in a triaxial compression test on a saturated sample obtained from the ground
without a change in mean effective stress, are shown in Figure 5.22. It can be seen
that there is a significant difference between the winter and summer strengths, the
former being much lower. Also shown on this figure are strength profiles based on
empirical relations proposed by Sandroni (1977). The implication of this with
respect to stability will be briefly discussed.
Embankments / 189
'YT
--m
Embankments 1 191
~~~~~
before failure
Embankments / 193
examined by performing analyses of
Redicled - A A
Predicted- PIB
the Carsington section with the same
- IM)
9
soil properties, but with different core
m
configurations (Dounias et al.
- 195 '3
(1989)). The geometries consisted of
- I90 5
a wide core, a narrow core and an
inclined core. These analyses are ,
l
I
l
I
!
185
summarised on Figures 5.32a to
'0
'O0
O
5.32d.
The collapse heights and the stress
Figure 5.3 1 :Observed and
concentration effects are virtually the
predicted pre-failure movements
same for the wide and inclined cores.
With the narrow core the embankment could not be built high enough to cause
failure through the foundation. However, Figure 5.32e shows that a stress
concentration was developing in the foundation near the core at the highest crest
height which could be reached (local instability near the crest due to applying
surcharge loading did not allow an eventual failure height to be determined).
790,,m~~disp~'ement7mm~
inelied 27.5m
wdc 26.5m
It seems that, given the same width, the influence of the core geometry (core
and boot, wide core, inclined core) is small, although the core and boot of the
original section give slightly worse stability than the other sections. While the
section with the narrow central core shows signs of eventual progressive failure,
the overall stability along the rupture surface through the core and the yellow clay
is much higher than for the other sections.
5.5
5.5.1
Introduction
Earth embankments are often built on soft clay foundations. Such soils usually
occur adjacent to river estuaries and, typically, the embankmentsare used for flood
potection or for carrying roads or railways. As the soft clays are weak,
embankments can usually only be built to a height of approximately 3m to 4m.
Embankments l l 9 5
However, higher embankments are often needed, for example for grade separated
junctions and flood protection, and consequently special construction measures are
required. Examples of such measures include the use of light weight embankment
fill, the provision of reinforcement in the bottom of the embankment andlor
stagged construction.
In the latter approach the embankment is constructed in stages. After each stage
a rest period is allowed for dissipation of excess pore water pressures in the soft
clay. Such dissipation is accompanied by consolidation and a gain in soil strength.
This in turn enables the embankment to be raised to a greater height in the next
stage of construction. Clearly, the amount of gain in soil strength during the rest
periods governs the height to which the embankment can be raised in the next
construction stage. Consequently, staged construction is often accompanied by the
provision of drainage in the soft clay to increase the amount of pore water pressure
dissipation during the rest periods.
Soft clays are often anisotropic, having both stiffness and strength dependent
on the orientation of the major principal stress. This complicates any analysis and
as sufficient data is not often available to define the magnitude of the anisotropy,
empirical factors are introduced into conventional design procedures (Hight et al.
(1 987)).
5.5.2
Typical soil conditions
Soft clays have not usually been subjected to
much erosion and therefore have not
O
experienced greater vertical stresses in the
past. Consequently, they are normally or very
2
lightly overconsolidated and in principle they
should therefore have an undrained strength
profile of the form shown in Figure 5.34. g
This distribution is based on a ground water
level 2m below the ground surface and an
a
overconsolidation ratio of 1.O. Note that pore
water suctions (hydrostatic) are assumed to
8
occur in the clay above the water table.
However, in most, but not all, locations
the surface layer of the clay has been
101
i
I
\ I
0
10
20
30
subjected to wetting and drying due to
Undrained strength, S, (Ha)
fluctuations in the ground water level. This
(in triaxial compression)
process results in the clay having a surface
crust of a higher strength than that indicated Figure 5.34:
of
in Figure 5.34. A typical strength profile for
strength with depth
a soft clay site is shown in Figure 5.35 (Mair for normally
clay
et al. (1992)). The surface crust extends
down to a depth of approximately 2m and below this level the strength is similar
to that indicated in Figure 5.34.
5.5.3
40
Choice of constitutive
Figure 5.35: Typical
model
distribution of undrained
strength with depth for
The most critical issue with embankments on
a soft clay
soft clay is usually the strength of the clay.
As the embankments are constructed
relatively quickly compared to the time
0
required for any significant consolidation, it
is the undrained strength of the clay that is
2
important. As noted above, this strength c m
vary significantly with depth and any crust
can have a marked effect on the behaviour of
24
an embankment.
-5
As much of the clay is lightly 3
overconsolidated and as critical state models
are reasonably good at reproducing the
behaviour of such soils, this type of model is
often employed to represent the behaviour of
the foundation. These models also have the
ability to predict changes in undrained
l0 o
2
4
6
strength with consolidation. It should be
OCR
noted that a Tresca model cannot reproduce
this behaviour. However, there are two Figure 5-36; Distribution of
potential difficulties when using conventional
OCR to give undrained
critical state models.
strength in Figure 5.35
The first difficulty is that the undrained
strength is not an input parameter to the model. It can, however, be calculated from
the input parameters and the initial stress conditions as described in Section 7.9.3
Embankments / l 9 7
of Volume 1. The required undrained strength profile can then be obtained by
adjusting the overconsolidation ratio (OCR).For example, to obtain the distribution
of undrained strength shown in Figure 5.35, using a form ofthe modified Cam clay
model which has a Mohr-Coulomb hexagon and a circle for the shapes of the yield
and plastic potential surfaces in the deviatoric plane respectively, and the
parameters given in Table 5.4, requires the distribution of OCR shown in Figure
5.36. Clearly, inputting such a distribution of initial OCR and its associated K,,
distribution may be problematic with some computer programs. This problem is
discussed further in Chapter 9.
Table 5.4: Properties for soft clay
It should also be noted that the undrained strength depends on the shapes
adopted for both the yield and plastic potential surfaces in the deviatoric plane. For
example, the undrained strengths shown in Figures 5.34 and 5.35 are appropriate
to triaxial compression conditions. For plane strain conditions the constitutive
model, using exactly the same input parameters, will produce smaller undrained
strengths. In this case the plane strain strengths will be 0.82 times the triaxial
compression strengths.
The second difficulty is that the majority of critical state models assume the soil
to be isotropic. For soft clays this is not necessarily correct and both strength and
stiffness may be anisotropic and depend on the orientation of the major principal
stress. If a clay does exhibit strong anisotropy, then either properties representing
some average strength and stiffness must be input into an isotropic model or,
preferably, a constitutive model such as MIT-E3 (see Section 8.6 of Volume l),
which can reproduce anisotropic behaviour, should be used. An example of the use
of such a model is given in Section 5.5.8.
As noted above, the embankment fill material is usually much stronger and
stiffer than the soft clay and is not therefore critical. However, if stability is to be
modelled in the analysis, the constitutive model used to represent fill behaviour
must simulate yield and ultimate strength. Consequently, a simple Mohr-Coulomb
model is often used. Sometimes the elastic properties are assumed to be stress level
dependent. More recently models of the Lade type (see Section 8.5 of Volume 1)
have been used.
5.5.4
Modelling soil reinforcement
As noted in Section 5.5.1,
reinforcement layers, in the form of
sheets of geotextile, are sometimes
added to the base of an embankment,
see Figure 5.37. The effect of such
reinforcement is discussed further in
Section 5.5.6. To model the
/
Embankments / l 9 9
soft clay in both analyses. To represent the strength profile shown in Figure 5.34
the OCR was assumed to remain constant with depth and have a value of 1.O. To
obtain the undrained strength profile shown in Figure 5.35 the distribution of OCR
with depth shown in Figure 5.36 was used. It should be noted that in this analysis
the K,, value also varied with depth in accordance with the equation given in Table
5.4. Soil suctions (hydrostatic) were assumed to exist in the clay above the water
table. Consequently, there were finite vertical effective stresses at the soil surface
and hence non zero undrained strengths, see Chapter 9.
In both analyses the embankment was unreinforced and assumed to be
constructed of a granular fill. The fill was modelled using a simple linear elastic
Mohr-Coulomb model with the properties given in Table 5.5. During the analyses
the embankment behaved in a drained manner (i.e. K, = 0), whereas the soft clay
behaved undrained (i.e. K, = 1600K,,,).
Table 5.5:Properties for embankment fill
ito,
1i
5.5.5.4 Results
The variation ofthe lateral movement
1.;.
of the toe of the embankment (point A I.;
....
on Figure 5.37) with height of the
embankment for both analyses is '
m
shown in Figure 5.39. The surface
crust has the effect of increasing the
embankment height at failure from
I
I
b) Wttb a surfacec w
2m to 3.5m.
Vectors of incremental
displacement during the last Figure 5.40: Vectors of incremental
increment of each analysis are
displacement just before failure
presented in Figure 5.40. These
vectors represent the magnitude and direction of the incremental displacements. It
is the relative magnitudes of the displacements and their directions that are of
interest and not their absolute values. In both analyses these vectors clearly identify
the failure surface. For the analysis with a surface crust the failure surface is
deeper and wider, passing through the centre line of the embankment.
,
b
S)
NO surface
5.5.6
Example: Effect of reinforcement
5.5.6, 1 Introduction
To examine the effect of reinforcement in the bottom of the embankment, a small
parametric study was performed on the example embankment geometry shown in
Figure 5.37. In this study both the stiffness and the strength of the reinforcing layer
were varied.
The reinforcement was assumed to be a layer of geotextile placed on the surface
of the soft clay prior to embankment construction. This was modelled by a layer
Embankments / 201
of membrane elements sandwiched
between two layers of interface
elements, see Figure 5.41. These
elements were added to the mesh
shown in Figure 5.38. Plane strain
finite element analyses were
performed.
! I
5.5.6.3 Results
As shown in the previous section, with no reinforcement the embankment can be
built to a height of 3.5m. When reinforcement is added, the embankment can be
constructed to a greater height. This can be seen in Figures 5.42 and 5.43 which
show the variation of the maximum embankment height with increase in
reinforcement strength and stiffness respectively.
Figure 5.42 shows results from a series of analyses in which the stiffness of the
reinforcement was 1 . 5l~O6 kN/mZ/m(a typical value for a geotextile) and its tensile
strength was varied. Increasing the tensile strength to 75 kN/m increased the height
of the embankment from 3.5m to 4m. Any further increase in the tensile strength
of the reinforcement produced an insignificant increase in the embankment height.
The results therefore indicate that for areinforcement stiffness of 1 . 51~O6 kN/mZ/m
the maximum height of embankment is 4m and, to achieve this, the reinforcement
must have a tensile strength greater than 75kN/m. If greater tensile strengths are
,,
,,
5.5.7
Example: Staged construction
5.5.7. 1 Introduction
As an example of an embankment on soft clay constructed in stages consider the
example shown in Figure 5.44. This road embankment is approximately 8m high
and was constructed in the west of the -United Kingdom. The foundation consists
Embankments 1 203
of ss thick deposit of soft clay
overlying sand. The depth of
the soft clay is not constant but
varies across the site. Because
it is not possible to construct
Figure 5.44: Example of a stage
an embankment 8m high on
constructed embankment
the soft clay in a single
- lift, a
staged construction approach is required.
At the design stage it was therefore necessary to decide on the number and
timing of the construction stages and the number and position of the drains that
must be provided in the foundation. As part of the design process, a series of plane
strain finite element analyses were carried out. One of these analysis will briefly
be described here.
For calculation purposes the sand below the soft clay was assumed to be rigid
and therefore the bottom of the soft clay provided a lower boundary to the finite
element mesh. As the depth of the soft clay varied and as the geometry of the
embankment was not symmetric (see Figure 5.44), the complete cross section had
to be modelled in the analysis.
5-
34
OCR
l4
OCR = 40
.
6 -
10
3,2 4 14
l6
1.24
20
30
19 ,
$ ,o
The finite element mesh is shown in Figure P
5.47. By today's standards this mesh is rather a 12
-Assumed m FE analys~s
14
coarse. However, the analyses were carried
out in 1983 when computer resources were
limited. The calculation process involved
four stages of construction, each separated by Figure 5.46: Distribution Of
strength prior to
a period of consolidation. The material
embankment
involved in each of the construction lifts is
indicated in Figure 5.48 and the main
calculation steps are summarised below:
Embankments / 205
5.5.7.4 Results
Lateral movements ofthe vertical line AA through the toe of the embankment (see
Figure 5.48) are presented in Figure 5.49. Settlements of the original ground
surface below the left hand side of the embankment are presented in Figure 5.50.
On both figures results are presented for calculation steps 'q' and 'r'.
From Figure 5.50 it can be seen that after construction (step 'q') the original
ground surface has settled nearly 2m. This increases to 2.5m in the long term after
dissipation of all excess pore water pressures (step 'r'). Such large settlements are
often associated with embankments on soft ground.
Kg
// 7
(i.e. steptern
'I,)
afier
applying 4th lift +
surcharge loading
(1.e step .q')
70
-50
- -0.5
0
o.o
-+
Immediately after
applying 4th lit? +
g
(i.e. step 'q')
In long term
(i.e. step '1')
5.5.8.2 Geometry
The geometry of this boundary value problem is presented in Figure 5.5 1. One side
of the embankment has a slope of 2:l (horizontal to vertical) inclination. The
Embankments / 207
stability of this slope is
enhanced by the presence of a
1.5 m high berm, so that any
k
237m
.,.,. 4 6 m
son Champlu slay
failure is forced onto the other - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - side of the embankment, where
the slope is steeper, with 1.5: 1
inclination. The predicted Figure 5.5 1: Geometry of the designed
embankment at St. Alban
height of the embankment,
based on conventional analysis, was 4.6 m.
The construction sequence was such that the first O.6m of embankment was
lifted in a day, followed by 0.3m per day until a height of 1.5m was reached. After
that two 0.3m layers were lifted per day, until failure occurred at about an
embankment height of 4 m, which was less than the predicted 4.6m.
5.5.8.5 Results
The finite element mesh for
the embankment analysis is
presented in Figure 5.53 and
the problem was analysed as
plane strain. The same
construction sequence and rate
were simulated as those
implemented on site, such that
failure occurred under
undrained conditions.
Figure 5.54 shows the
development of horizontal
displacement, U, of the
embankment toe with the
embankment height. This
horizontal displacement is
normalised by the maximum
horizontal displacement, U,, ,
predicted just before the slope
failed.
130m
a 2.0
$
*
%
.g
MCC
1.5
- 1.25
DSS
-2
.-
oo
1
The
Figure 5.54: Maximum embankment
using the
soil
height from different soil models
which simulated the
anisotropic soil strength as depicted in Figure 5.52, predicted failure at an
embankment height of 3.9m, which is very close to the 4m observed on site. For
Embankments / 209
the same embankment, but now analysed with the MCC model with an isotropic
undrained strength corresponding to the triaxial compression profile in Figure 5.52,
the failure height is grossly overestimated at 4.9m. Back analysis using the MCC
model to achieve a failure height of 3.9m with an isotropic strength resulted in the
necessary strength being 1.25 times the DSS strength. Analysis with the isotropic
strength being equal to the DSS profile, which is what some design practice
suggests, is largely conservative, predicting failure at only 3.3m embankment
height.
Table 5.8: MIT-E3 parameters for soft Champlain clay
A further set of analyses was performed, where the embankment geometry was
altered to have berms on both sides of the embankment, see Figure 5.55. Again
both anisotropic and isotropic analyses were performed and the results are
presented in Figure 5.56, as normalised horizontal displacement of the
embankment toe versus embankment height.
' soeA
with the
strength back-calculated from
the previous set of analyses
(i.e. 1.25 times the DSS
strength, see Figure 5.54).
This analysis predicted an
embankment height at failure
of 4.9m and therefore
overestimated that obtained by
the analysis with the MIT-E3
model. To obtain the same
failure height as the MIT-E3
analysis, the MCC analysis
needed a strength of only 1.15
times the DSS strength.
Several conclusions can be
drawn from this embankment
study:
p
-
2.0
1
B
g
1.0
.2
0.5
,
O
- An analysis which assumes isotropic soil behaviour and takes either the
triaxial compression strength profile or the DSS strength profile will either
overestimate or underestimate the embankment height, respectively;
Embankments 1 2 1 1
5.6
1.
2.
3.
4.
5.
6.
7.
8.
Summary
'Variable' elastic models were widely used in the past to predict behaviour of
embankment dams, primarily because of their simplicity and ability to model
nonlinear and stress dependent behaviour of fill materials.
A great variety of stress paths, accompanied by a rotation of principal stresses,
occurs in embankment dams during construction, and subsequent impounding
in particular. 'Variable' elastic constitutive models cannot account for stress
path dependency which is caused by nonelastic components of soil behaviour.
Compacted fills have a behaviour pattern similar to that of granular materials.
The double hardening elasto-plastic constitutive model of the Lade's type
accounts for pre-peak plasticity and is able to capture several important
characteristics of the stress-strain and strength behaviour of fill materials,
especially rockfills.
The limitations of computer modelling require relatively thick layers to be
used in the idealization of dam construction. However, significant errors can
occur if the layers are too thick, particularly when the effects of compaction
are to be modelled.
Comparison of the predictions with the observed behaviour during
construction of Roadford dam indicated that the predictions using the Lade's
double hardening model were in reasonable agreement with the field data.
The analysis using the nonlinear elastic model severely overpredicted lateral
movements. This prediction could be improved by adjusting the nonlinear
elastic model parameters, but then the laboratory test data could no longer be
recovered.
The analyses of old puddle clay core dams, which involved construction in
layers, impounding, development of seepage pressures, and cyclic
drawdowns, with soils behaving drained, undrained and partly consolidated,
demonstrated that the behaviour of these water retaining structures can be
fully modelled using the finite element method.
Cracking by hydraulic fracture ofpuddle clay in a cut-off trench seems almost
inevitable. Narrow puddle clay cores have stresses after construction and
during impounding which are close to those allowing hydraulic fracture. The
failure of Dale Dyke dam was recovered when rapid undrained final
impounding was modelled.
The analysis of Ramsden dam has shown that quite large and non-recoverable
movements can occur during operational cycles of drawdown and reimpounding, even though the factor of safety against an overall shear failure
Embankments 1 2 13
Shallow foundations
6. l
Synopsis
6.2
Introduction
Shallow foundations / 21 5
6.3
Foundation types
Shallow foundations may be broadly categorised into one of the following two
types.
6.3.1
Surface foundations
These are foundations which are placed
on, or very near to, the soil surface. They
can come in many shapes and sizes, but
are often categorized as either strip,
circular or rectangular footings, see
Figure 6.1.
/
Strip
6.3.2
Shallow foundations
Shallow foundations are similar to
surface foundations, except that they are
embedded below the soil surface, see
Figure 6.2. For economic reasons their
depth below the surface is usually not
great, hence the term shallow. Again,
they are often categorized as either strip,
circular or rectangular.
C~rcle
Rectan6le
6.4
Choice of soil model Figure 6.2: Shallow foundations
Conventional foundation design
considers bearing capacity and
deformations separately. Using these 3
approaches it is not possible to establish
Ultimate
the complete load-displacements curve. It
load
is only possible to provide estimates of
the ultimate foundation load and of the
initial gradient of the load-displacement
Displacement
curve, see Figure 6.3.
The calculation of bearing capacity is
Figure
Typical loadbased on stress field solutions andlor
limit analysis, combined with some
displacement curve
empirical correlations. The soil is
essentially assumed to behave as an elastic Tresca material, if undrained bearing
capacity is being considered, and as an elastic Mohr-Coulomb material, if drained
bearing capacity is under investigation. Such simple constitutive models can also
be used in numerical analysis and it can be shown, see below, that numerical
analysis can recover the analytical bearing capacity solutions where they are
available. However, while the conventional bearing capacity solutions are only
"3
available for simple foundations resting on one soil type, with either a constant
strength or a strength that varies linearly with depth, numerical analysis can
consider complex foundation geometries resting on inhomogeneous foundations.
Numerical analysis can also cope with complex loading combinations (e.g. vertical,
horizontal and moment loading).
Conventional bearing capacity solutions are restricted to either undrained soil
or fully drained soil. They do not therefore provide information on the variation of
bearing capacity with time. This may be important for foundations constructed on
clay soils. Using a coupled numerical approach, see Chapter 10 of Volume 1, it is
possible to simulate such time dependent behaviour, as long as a realistic
constitutive model is used to represent the soil. In this respect the use of a Tresca
model is not appropriate, as the strength of the soil remains constant. Although, in
principle, a Mohr-Coulomb model could beused, this is not advisable as unrealistic
predictions of bearing capacity can be obtained under undrained and partially
drained conditions, see Chapter 9. To obtain realistic predictions it is necessary to
use a strain hardening/softening model of at least the complexity of a simple
critical state model (e.g. modified Cam Clay).
Conventionally, foundation displacements are predicted using elastic theory.
Several approaches are available and they differ in the manner in which the
stiffness parameters are derived. Some assume the soil is linear elastic, others
obtain the stiffness from an oedometer test. While these approaches appear to give
reasonable predictions of average foundation settlement, they do not provide
accurate predictions of differential settlements, deformations under combined
loading, or movements in the soil adjacent to the foundation. This is perhaps not
surprising as the stiffness of the foundation is usually ignored, with the foundation
assumed to apply a uniform pressure to the soil. Numerical analyses have the
ability to deal with a range of constitutive models and can therefore produce more
realistic predictions.
In principle, a soil model that accurately simulates the behaviour of the soils
supporting the foundation under investigation should be used. If the foundation
consists of layered soils, then several soil models might be used. If bearing
capacity is of interest, then clearly the soil model should accurately simulate the
soil strength. If deformations are of concern, and in particular those adjacent to the
foundation, then it is advisable to use a constitutive model that can accurately
represent the nonlinear behaviour of the soil under small strains.
6.5
6.5.1
Introduction
If the foundations are subjected to vertical loading only, then a strip footing can be
idealised as plane strain, a circular footing as axi-symmetric, but a rectangular
footing remains a three dimensional problem. Consequently, while it is relatively
straight forward to analyse strip and circular footings under vertical loading, as
they can be reduced to two dimensions, considerably more computer resources are
Shallow foundations I 21 7
required to analyse rectangular footings
which involve a full three dimensional
analysis, although account can often be taken
of the two vertical planes of symmetry to
reduce the complexities of the analysis, see
Figure 6.4 (i.e. in this case only a quarter of
the full geometry needs to be considered).
If the loading is not vertical but inclined,
and the inclination is in the plane
I
r
/
Plane of
perpendicular to the length of the footing,
dJ
symmetry
plane strain analysis can still be performed
for a strip footing. Otherwise, and for both
v
the circular and rectangular footings, a three
Figure 6.4: Vertical planes of
dimensional analysis is required. For the
symmetry in 30 analysis of a
circular footing this analysis can be
rectangular footing
performed using either a full three
dimensional, or a Fourier series aided approach. A conventional three dimensional
analysis will be required for the rectangular footing.
In general, it is necessary to discretise both the soil and the foundation into
finite elements, see Figure 6.5. This enables the correct stiffness of the foundation
to be included in the analyses. Interface elements can also be positioned between
the underside of the foundation and the soil, so that the correct interface behaviour
can be modelled. However, as the foundation is often very stiff compared to the
soil, numerical problems can arise due to the large difference in stiffness between
adjacent elements on each side of the interface. Although these numerical
instabilities are not so much of a problem with modern computers, where it is
possible to use increased numerical precision (e.g. double or quadruple precision),
surface foundations are often analysed using one of two extreme assumptions (i.e.
a perfectly flexible or rigid foundation). Such assumptions are also used in
conventional soil mechanics practice. If one of these assumptions is made, then
there is no need to include the foundation itself in the analysis.
Foundation
Interface elements
6.5.2
Flexible foundations
If the footing is assumed to be flexible, then it is possible to simulate its behaviour
in a finite element analysis as shown in Figure 6.6. As the footing is flexible, it is
assumed that any loading is uniform and can be represented by a surface surcharge
pressure applied to the surface of the soil immediately below the position of the
footing. As noted in Chapter 3 of Volume I , such a uniform surcharge load can be
converted to give equivalent nodal forces, AFy . In fact, most computer programs
do this automatically for the user.
If the footing is smooth, Flexible foundation:
then the other boundary
Load control
Displacement control
rough
condition that must be applied ""00"
to the nodes under the footing
is that the horizontal nodal
NOT POSSIBLE
forces are zero (i.e. AF,=O).
Most computer programs
assume that AF,=O if no
b o u n d a r y condition is
Figure 6.6: Boundary condition options for
specified in the X direction for
flexible foo ting
a particular node.
Alternatively, if the footing is rough, then the horizontal displacement (Au) of
the nodes on the soil surface below the position of the footing must be restricted
and therefore set to zero (Au =0), see Figure 6.6.
It is not possible to apply displacements to the footing and perform the analysis
in displacement control. Consequently, care must be taken when approaching the
collapse load of the foundation, as the application of too much load will cause
unlimited displacements. This will probably manifest itself as a convergence
problem in the nonlinear solution algorithm. Some finite element programs have
facilities for automatically controlling the size of the incremental loads. This is
particularly useful for load control problems and can result in accurate estimates
of limit loads.
6.5.3
Rigid foundations
If the footing is rigid, analyses can be performed under either loador displacement
control. The various alternatives for the boundary conditions to be applied to the
soil surface below the position of the footing are given in Figure 6.7. As noted in
this figure, for load control it is necessary to tie the vertical displacements of the
nodes below the position of the footing, see Section 3.7.4 of Volume 1 . This will
ensure that all the nodes move vertically by the same amount. As the vertical
displacements are tied, the load can be applied as auniform pressure over the width
of the footing, or as a single point load at the node on the centre Iine of the footing.
In addition, as with flexible foundations, care must be taken when approaching the
collapse load of the foundation if the analysis is being performed under load
control.
Shallow foundations 1 21 9
Rigid footing:
Load conk01
smooth
Displacement conIr01
smooth
m
rough
AV a pli
where {R) is the vector of reactions in the coordinate directions, [B] is the strain
matrix (see Section 2.6 of Volume l), {c} is the vector of current total stresses,
{B,} is the vector of total stresses existing prior to displacing the footing, and the
volume integration is performed for all soil elements connected to the displaced
nodes. Most finite element software have facilities for obtaining such values.
6.5.4
Examples of vertical loading
6.5.4. 1 Introduction
As a simple example of the application of numerical analysis, the problem of a
vertically loaded surface footing resting on either an undrained clay or a drained
soil (clay or sand) is considered. To be consistent with conventional foundation
design, the undrained clay and drained soil are modelled using a linear elastic
Tresca and a linear elastic Mohr-Coulombmodel respectively. The results of these
analyses are compared with the conventional bearing capacity solutions.
Q,,,
(6.2)
where Q,,, is the maximum vertical
load applied to the footing, A its area Figure 6.8: Finite element mesh for
and NLthe bearing
In strip and circular footing analyses
Figure 6.9 the load is therefore
expressed in terms of the mobilised
P 6
bearing capacity factor NCmob
(= Ql(A 2"
S,,)), where Q is the load on the a?
2
footing. The displacement is $ 4
normalised by B, where B is the half $
width of the footing. Results from
2
three analyses are given. One of these
I
modelled a smooth footing and the 7
rl
other two a rough footing. The
difference between the two analyses
001
002
003
OM
00s
006
Normal~sedsettlement, &/B
modelling a rough footing is that in
one analysis a row of interface
Figure 6.9:Load-dis~lacement
elements was added between the soil
curves for strip footing
surface and the underside of the
footing, see Figure 6.10. In the other
Foundation width
analysis and in the analysis modelling .1
Interface elements1
a smooth footing, these interface
elements were not present. The
interface elements were given a shear
strength of IOOkPa and shear and
normal stiffness values of
Kf=K,,=105kNlm3 (note that these
stiffness values do not have a major
of mesh with
Figure 6. 10:
influence on the results). It can be
interface elements
seen that the three analyses give
slightly different load-displacement curves and slightly different N, values at
failure. These values are quoted on Figure 6.9.
Conventional bearing capacity theory indicates that, for a strip footing resting
on undrained clay with a constant strength, the bearing capacity factor N, should
be 5.1416 (= 2+x), for both smooth and rough footings. This solution can be
obtained from limit analysis (i.e. both upper (unsafe) and lower (safe) bound
solutions) and from a closed form plasticity solution (i.e. combination of stress and
=4
' Nc S u
97;
';
~
I1
i\Tmm
{. *
g
Oo
.g
2
O,
,,, ,,, ,
,
,,,
..............................................
Smoatb:
0.h
103
0.b5
0.k
0.h7
0.68
Q,,, = A S, NCstnP
S,,
(6'3)
curves for smooth rectangular
The shape factor S, depends on the
footings
LIB ratio of the footing and is usually
determined from the following fi
............................................................
2"
empirical relationship (Skempton g 6 1951):
. . . . . . . . . . . . . . . . . . . . . . . .
Rough:
S,
............ U B = Ill
- - -U
-B. U1
= 1+0.2 B / L
(6.37)
(6.02)
UB = 411 (5.72)
0.b7
0.08
1.3
SO;/
1.25
QInax
1 A = c'N,
--
+ q 1 N y+ y ' B N y
1
(6.5)
where B is the half width of
footing, q' is the magnitude of
surcharge pressure existing on
ground surface adjacent to
the
any
the
the
I
0
Nq =
For the situation where the angle of dilation equals the angle of shearing resistance
(i.e. v = p'), it is possible to show that a compatible displacement mechanism is
associated with these stress fields. It is also possible to extend the stress fields
throughout the soil mass, satisfying equilibrium and without violating the yield
condition. The solutions expressed by Equations (6.6) and (6.7) are therefore
theoretically exact, see Chapter 1 of Volume 1. The solutions can also be shown
to be applicable to both smooth and rough footings. If the angle of dilation, v, does
not equal the angle of shearing resistance, p', then the above solutions are only
approximate.
To assess the ability ofnumerical analysis to predict these results, finite element
analyses have been performed using the geometry and finite element mesh shown
in Figure 6.8. The Mohr-Coulomb model was used to model soil behaviour, see
Section 7.5 of Volume 1. The soil was assumed to be weightless (y = 0) and
cohesionless (c1= 0), with an angle of shearing resistance, p', of 25", a Young's
modulus, E', of IOOMPa and a Poisson's ratio, p, of 0.3. Four analyses were
performed with a surcharge, g', of 100kPa and one analysis with g' = IOkPa. The
four analyses with g' = 1OOkPa differed in that different combinations of the angle
of dilation and roughness of the footing were assumed. For the analysis with g' =
1OkPa the footing was smooth and the angle of dilation was zero. For the analyses
with rough footings interface elements were positioned between the footing and the
soil, see Figure 6.10. These were given the same values of p' and v as the soil, and
stiffness values K,=K,,= 10' kN/m3.
The predicted load-displacement curves are shown in Figure 6.2 1. The load is
expressed as the mobilised bearing capacity coefficient, N F b (=Q/(Aqr)). This
figure shows that there is little effect of the footing roughness, but that the angle
of dilation affects both the shape of the load-displacement curve and the ultimate
value of N,. The magnitude of the surcharge load, g', affects the amount of
displacement required to reach failure. However, it does not affect the ultimate
value of N,.
smooth, v = 25"
smooth, v = 0"
rough, v = 0'
. .
rough, v = 25'
q'>O) which has weight, due to the complexity ofthe governing hyperbolic partial
differential equations. Consequently, there is no theoretically exact expression for
N,. In fact, several alternative expressions can be found in the literature. One of the
most popular is that by Hansen (1970):
This equation, as with many of the alternatives, is often used for both rough and
smooth footings. However recent work by Bolton and Lau (1993) has shown that
footing roughness affects the magnitude ofNy.Their results are based on numerical
integration of the stress field equations and are therefore, again, approximate.
Results from four finite element B ,
analyses, using the mesh shown in 5 ,Figure 6.8 and with the properties
given above for the N, analyses, are
shown in Figure 6.23. Dry conditions
-smwth, v = 0"
were assumed and the soil had a bulk
------ smwth. v = 25"
unit weight of 18kN/m3and K,,=0.577
(i.e. 1 -sinvl). A different combination
of footing roughness and angle of soil
dilation was used in each
r71
.
... .
,... .
K, = 0.577, 4'=2s0
.::::-z::,:-
. . ... ., ..-,.... .. .
., . - .. .
..S,..
rough~tnp.V = W
?.
I_
El
C!
Plastic
smwth ship:
,$'=25'
v=@
K.
2.0
If a surcharge loading exists on the soil surface adjacent to the foundation and
the soil has weight, then both the N , and N, terms of Equation (6.5) are non-zero.
As noted above, this equation assumes that the effects of soil weight, y, and
surcharge, g', can be superimposed. To investigate this hypothesis the analysis
presented above for a rough footing on sand with an angle of dilation v = 25" was
repeated with a surcharge g' = 1OkPa. This analysis gave an ultimate load on the
footing Q,,, = 556 kN/m. Based on N, = 11.03 (see Table 6.1) and N, = 6.72 (see
Table 6.2), Equation (6.5) predicts an ultimate load of463 kN/m. Comparing these
values indicates that the superposition assumption implied in Equation (6.5) is
conservative. In this particular case by 17%.
One reason why the superposition assumption is not valid is that the failure
mechanisms associated with N, and N, differ. This can be seen by comparing the
failure mechanisms shown in Figures 6.22 and 6.24. The failure mechanism for N,
is deeper and wider than that for N,.
It should also be noted that the magnitudes of the footing settlements required
to mobilise full N, and N, values differ (see Figures 6.2 1 and 6.23). In addition, the
settlement required to mobilise the ultimate N, value depends on the magnitude of
the surcharge, g'. Conventional design procedures do not explicitly account for
this. However, in practice this could have significant implications, especially in
situations where the soil strength degrades (i.e. v' reduces) with straining. For
example, at ultimate load the average strength associated with the first mechanism
to form (i.e. N,) is likely to have decreased from its peak value, while that
associated with the second mechanism (i.e. N,) is unlikely to have reached its peak
value. In such a case Equation (6.5) may not be conservative (see Chapter 4 for a
detailed discussion on progressive failure).
rough, = 25'
footing the larger the ultimate N, $
value. In addition, and again in
contrast to the results for the strip
footing, the angle of dilation has only
a minor influence on the ultimate
value of N , .
One of the reasons for the
differences outlined above can be
seen in Figure 6.29, which shows
Normalised settlement, SIB
vectors of incremental displacement
at failure. These vectors indicate that
the failure mechanisms for the smooth Figure 6.28: Load-displacement
curves for a circular footing on
ciscnlar footings are shallower and of
weightless soil
a smaller lateral extent than those for
the rough circular footings. (effect of roughness and dilation)
Comparison of these failure
mechanisms with those for the strip footings shown in Figure 6.22 indicates that
even for the rough circular footings the failure mechanisms are smaller both
laterally and vertically than for the strip footings.
rough, v = 0"
rough, v = 25'
The ultimate values ofN, are listed in Table 6. l, where they are compared with
the equivalent values from the strip footing analyses. The values for the circular
footings are substantially larger than those for the equivalent strip footings. This
is in agreement with conventional design practice which, depending on the actual
design code or advice manual, suggests a shape factor of between 1.2 and 1.5 to
be applied to the N , value for a strip footing.
100kPa, v = 0"
9.92
18
1OOkPa, v = 25"
11.03
19.42
IOkPa, v = 0"
10
6.5.5.
Undrained bearing capacity of non-homogeneous clay
6.5.5.1 Introduction
In Sections 6.5.4.2and 6.5.4.3discussing the undrained bearing capacity of clay,
the strength was assumed to remain constant with depth. In many field situations
this is not a realistic assumption. For some simple linear and bi-linear strength
variations with depth, approximate analytical solutions are available (Davis and
Booker (1973)), however for more general situations such solutions do not exist.
where:
+2~:')[1+
D,:,
g(0) =
2(1+ 2 K
)':
BI]
(1
+ 2 K~')OCR[I + B']
]:
(6.9)
sin p'
sinOsinpl
cos@+
-,
OCR, K,
1'4).
Because the initial distribution of S,, with depth is linear for the stiff clay, it is
possible to obtain an estimate ofthe bearing capacity using results from Davis and
Booker (1973). This gives a value for the smooth footing of Q,,, lA=145 kPa. To
obtain this value interpolation and scaling from the figures supplied in the above
paper are necessary. Davis and Booker's figures are based on calculations which
involve some finite difference approximations and have been drawn to provide
conservative estimates for design purposes. The above result is therefore subject
to error. Nevertheless, this value of ultimate bearing capacity is within 5% of the
finite element result.
Similar results are also shown in Figures 6 . 3 6 ~and 6.36d, which are from
analyses of a 10m wide footing on soft clay. These results imply that the surface
crust, see Figure 6.34, does not have a major influence on the failure mechanism.
This is not so for the 2m wide footing on soft clay, as can be seen in Figures 6.36e
and 6.36f. In this case a deep Prandtl type mechanism is predicted for both the
smooth and rough footing. Here the failure mechanism, the extent of which
depends partly on the footing width, is controlled by the 2m deep surface crust
which has a strength reducing with depth. The failure mechanisms for both the
smooth and rough footings are forced into the weaker soil, as this provides a failure
mechanism involving the least resistance. As the failure mechanism is the same for
both smooth and rough footings, then the bearing capacity is also the same, which
is evident from the results shown on Figures 6.36e and 6.36f.
. . . . . . . . . .
. . . . . . . . . . . . . . 22.5m . . . . .
. . . . . . . . . . .
........................
. . . . . . . . . . .
.........................
. . . . . . . . .
.
. ... . .. . .. . .. . .. . .. . ..
............
.--a..
.........................
. . . . . . . . . . .
........................
. . . . . . . . . .
.........................
. . . . . . . . .
6.5.6
'0
60
--
40
80%
IW%pre-load
p-e-load
60% prc-load
-40%
pn-load
-o- 20% pre-load
zo
Oo
0.2
0.4
0.6
0.8
1.2
Displacement (m)
,,,
120
g ~w
1 so
--
60
40
20
O0
0.1
0.2
100%pm-load
80% PE-load
60% pm-load
40% pn-load
20% pn-load
0.3
Displacement (m)
0.4
0.5
,,
2 g ,,
4 ,,
+OCR 9
,,
$Tj 130
1;
. a
IIO
W
Im,
,, ,
7o
, ,,
Percentage of p~e-loading(%)
Figure 6-38,
(Jndrained bearing
capacity after pre-loading vs.
percentage of pre-load
::
::
3-
Parameter description
MIT-E3 value
strength profile to change, depending on the direction ofthe major principal stress,
as shown in Figure 6.42.
20
40
60
80
100
l20
I2O
loo
I
60
Figure
g,
HIV- 1.0
f#V= 0.3
d 30
__.-F
20
to
Q
lror ic M ~ Tmodel
____---------b l k__-model
HIV-0.1
______------
,,,, F-*m
.,.--.:----------------
l'
"
60
no
I4O
loo
M~
v* (W
Figure 6-46;
Finite element mesh
for analysis of circular footing
2c
g:
AH,,=O+A
sine+O...). The latter was of opposite sign but equal magnitude to that
set for the radial load, see Figure 12.2 and Appendix XII.2 of Volume I .
Again predictions using both sets of soil parameters were obtained. The results
are shown as an interaction diagram in Figure 6.47. The adverse effect of the
anisotropic behaviour is evident again. Comparisonwith Figure 6.45 indicates that
this effect is larger for the circular footing than for the strip footing.
6.6
6.6.1
Introduction
Most of the comments given above for surface foundations also apply to shallow
foundations. However, in this case it is necessary to account for the interface
between the soil and the sides of the foundation, see Figure 6.48. It is not usual to
model the construction of the foundation itself, but to simply assume it to be
'wished in place'.
6.6.2
- dcsc N:'~~S,
+ p(,
where d, is the depth factor andp, is the total overburden stress at foundation level,
see Figure 6.49. Exact theoretical solutions are not available to account for
foundation depth and therefore the magnitude of the depth factor is often based on
semi-empirical correlations. One of the most popular correlations currently in use
is that proposed by Skempton (1951) and shown in Figure 6.50. This correlation
l.4-
1
10
10
2
g
,
,
,
,
,
of approximately 0.5" and the present day inclination of the foundations is about
5.5". It can be seen from Figure 6.58 that significant inclination ofthe Tower only
began once the height exceeded the sixth cornice. If the inclination had been due
to much more compressible ground beneath one side than the other, it would have
developed much earlier. Therefore, another explanation for the rapid onset of
inclination is required and will be discussed later. It is the history of inclination
depicted in Figure 6.58 which was used to calibrate the numerical models
described later in this section.
For most ofthis century the inclination ofthe Tower has been increasing. These
changes in inclination are extremely small compared with those that occurred
during and immediately following construction. The rate of inclination of the
Tower in 1990 was about 6 seconds per annum. The cause of the continuing
movement is believed to be due to fluctuations ofthe water table in Horizon A. No
attempt has been made to model these small movements.
_ * _ * . . I . .
: : h
.__O_.a
( . , . . . _ . . . . I . . .
(Pc.$'
C,
c,
Glp,,'
eI
k
K,,
OCR
v
Given the above parameters, the value of the undrained strength S,, can be
determined from Equation (6.9). In Horizon A (see below) S,, turned out to be a
more reliable parameter than OCR and was therefore specified in its place.
The soil profile at Pisa was characterised in detail by the Polvani Commission
(Minister0 dei Lavori Pubblici 1971). Horizons A and B were divided into a
number of sub-layers, the descriptions of which are as follows:
Horizon A:
MG Top soil and made ground,
A1
Loose to very loose yellow sandy silt to clayey silt without
stratification,
Uniform grey sand with interbedded clay layers, broken fossils - Upper
A2
sand,
Horizon B:
Highly plastic grey clay with fossils,
B1
Medium plastic grey clay with fossils,
B2
Highly plastic grey clay with fossils,
B3
B4
Dark grey organic clay,
Blue grey to yellow silty clay with calcareous nodules,
B5
Grey, sometimes yellow, sand and silty sand - Intermediate sand,
B6
Medium to highly plastic clay, with fossils and thin sand layers in the
B7
upper part,
Grey clay with frequent thin sand lenses,
B8
Blue grey silty clay with yellow zones; calcareous nodules; some dark
B9
organic clay at centre,
B10 Grey clay with yellow zones; fossils in the lower part.
Laterally Horizon B is very uniform. However, there is much evidence to show that
in Horizon A, layer A I changes from predominantly silty sands and sandy silts
north of the Tower to clayey silts south of the Tower. There is also some evidence
from piezocones that the Upper sand layer A2 thins just south of the Tower. A
careful study of the detailed sample descriptions given in the Polvani Report
suggests that beneath the Tower, in Horizon A, there exists a lense of clayey silt
which thins from south to north.
It will become evident that it is the compressibility of the underlying soils
which has played the dominant role in the historical behaviour of the Tower (i.e.
leaning instability), so that emphasis is placed in this section on these properties.
Two major programmes of thin wall sampling and testing have been carried out,
Shallow foundations / 26 1
one in 1971 (at the instigation of the Polvani Commission) and another in 1986.
The results of these studies have been summarised by Calabresi et al. (1993) and
Lancellotta and Pepe (1990) respectively. Figures 6.66a, b and c show the
experimental values of C,', C, and OCR respectively, where C,' is the
compressibility of the reconstituted material - defined by Burland (1990) as the
intrinsic compressibility.The values of C,' were derived from the correlation with
the water content at the liquid limit, W,, established by Burland. Also shown on
Figure 6.66 are the sub-layers established by the Polvani Commission and
described above. It can be seen from Figure 6.66a that the values of C,' are
reasonably well defined for each sub-layer and the average values are shown by the
vertical lines. It is of interest to note that sub-layer B7 is made up of two distinct
soil types and should perhaps be considered as two separate sub-layers.
o 0,
c,'
0;s
1.:
cc
9.5
1.0 0I
OCR
f
It can be seen from Figure 6.66b that the experimental values of C,, determined
from high quality samples of the natural clay, show considerable scatter and it is
not easy to decide on suitable representative values for analysis. The use of the
intrinsic compressibility C', has been particularly useful in this respect (Burland
and Potts (1994)). It is well known that measured values of C, are particularly
sensitive to sample disturbance and the scatter in Figure 6.66b is a reflection of
this. Nash et al. (1992) presented the results of oedometer tests on a sensitive
marine clay from the Bothkennar test bed site and showed that the values of C,
obtained from high quality block samples were significantly higher than for other
sampling methods including thin wall sampling. Nash7sresults showed that the
measured values of C, for the best samples were between 1.9 and 2.3 times larger
than C,". In Figure 6.66b the vertical lines were obtained by multiplying the
average values of C,' by a factor depending on the plasticity of the material. For
the Pancone clay (B1 to B3) the factor was 2, for the Lower clay (B7 to B10) the
factor was 1.5 and for Horizon A and the Intermediate clay (B4 and B5) the factor
was unity. It is these values that have been used in the finite element analyses and
it can be seen that they tend to lie at, or a little beyond, the upper limits of the
experimental values, as is to be expected.
Figure 6 . 6 6 ~shows experimental values of OCR. The vertical lines are the
values used in the finite element analyses. Values were chosen near to the upper
limit of the experimental values, as the effect of sample disturbance is to reduce the
value somewhat.
Table 6.5 gives the values of all the soil parameters that were used in the
analyses. The mean values of q,' for each sub-layer were obtained from undrained
was taken as 0.1 for all the soil layers. The values of permeability
triaxial tests. 1~12
were derived from the oedometer tests, taking values in the upper quartile of the
range of results for each layer. The values of Glp,' were chosen on the basis of
experience with small strain testing of a wide variety of materials. The choice of
values was not important for modelling the history of inclination ofthe Tower, but
proved to be more crucial when predicting its response to small perturbations of
load in its present condition. The water table was assumed to be Im below ground
level and in hydrostatic equilibrium - a condition that must have existed at the time
of construction.
Table 6.5: Soil parameters used in the analysis
For the three dimensionalFourier series aided analysis the geometry is assumed
to be axi-symmetric. However, the spatial distribution of soil properties and
loading can be three dimensional. A mesh similar to that shown in Figure 6.67, but
only considering the geometry to the right of the centre line of the Tower, was
used. However, no tapered layer was incorporated to provide an 'imperfection'.
Instead, the soil properties were assumed to vary linearly beneath the Tower in
sub-layer A 1. To the north of the Tower the properties were as listed for layer A 1"
I, is a correction factor which takes account of the ratio between the second
moments of area of a rectangular and a circular foundation. For a rectangular
foundation of width 19.6m and the same area as the foundation of the Pisa Tower,
the value of I, for rotation about the centre is 1.266. For the three dimensional
analyses in which the circular shape of the foundation is modelled I, should
theoretically be unity. Equation (6.12) was incorporated into the analysis such that
any inclination of the foundations during or subsequent to construction
automatically resulted in the application of an appropriate overturning moment to
the foundations.
Analyses
All analyses involved coupled consolidation.The calibration analyses were carried
out in a series of time increments in which loads were applied to the foundation to
simulate the construction history of the Tower together with the rest periods, as
summarised Section 6.6.3.4. During a construction period it was assumed that the
load was applied at a uniform rate. The excavation ofthe catino was also simulated
in the plane strain analysis.
The only factor that was adjusted to calibrate the model was the factor I, in
Equation (6.12). For the first run, the value of l, was set equal to unity. At the end
of the run the final inclination of the Tower was found to be less than the present
value of 5.5". A number ofruns were carried out with successive adjustments being
made to the value of l, until good agreement was obtained between the actual and
predicted value of the final inclination. It was found that, for the plane strain
analysis with a value of I, = 1.27, the final calculated inclination of the Tower was
5.44". Any further increase in l, resulted in instability of the Tower. It is therefore
clear from this analysis that the Tower must have been very close to falling over.
"7
'3
,,
,,
,,,, ,,,,
1 *.y
that ageing significantly increases the stiffness of clays. For one dimensional
compression Leonards and Ramiah (1959) showed that the process of ageing
,
' as illustrated in Figure 6.72a. More
results in an increase in the yield stress a
generally a shift in the yield surface takes place as illustrated in Figure 6.72b. The
~ ' the intersections with the p' axis of the yield surfaces
parameters p,' a n d ~ , ,are
of a 'young' and the 'aged' clay respectively. The ratiopV1lp,,'is a measure of the
degree of ageing and is defined as the yield stress ratio. The effects of ageing can
be introduced into the finite element analysis by increasing the current value ofp,,'
for each integration point to give a prescribed value of the yield stress ratio.
In the laboratory, significant ageing effects have been observed over periods
of a few days. Leonards and Ramiah found that the yield stress ratio for a
reconstituted clay that was allowed to age for about 90 days was as high as 1.3. In
the present analysis the effect of introducingvarious values of yield stress ratio was
studied. The results presented here are for a value of 1.05, which was felt to be
conservative. It should be noted that ageing of the Pancone clay increases the
stability of the Tower foundations significantly.
The full line in Figure 6.73 shows
the predicted response of the Tower Ae(deg)
due the application of a counterweight 0,08
to the foundation masonry, at an
eccentricity of 6.4m to the north, after
allowing ageing of the Pancone and
Lower clays to give a yield stress
ratio of 1.05. No ageing of Horizon A
o
was assumed. The soil parameters are
e=6.4m
those given in Table 6.5. At the
design load of about 690t the
---- e = 7 . 8 m
inclination of the Tower is predicted
to reduce by about 27.5 seconds of Senlement(mm)
arc with a settlement of 2.4mm. More
importantly, the overturning moment
Figure 6.73: Predicted response o f
is reduced by about 14%. In Figure
the lower due to application o f
6.71 the broken line represents the
counterweight
predicted effective contact pressure
beneath the foundation after application of the counterweight.
Also shown in Figure 6.73 are the effects of increasing the load above the
planned level and the results are of considerable interest. It can be seen from the
full lines that, as the load is increased, the rate of increase of inclination to the
north reduces, becoming zero at about 1400t. With hrther increase in load the
movement reverses and the Tower begins to move towards the south. The
settlement rate also begins to increase once the load exceeds 1400t. Figure 6.73
also shows the results of increasing the eccentricity of the counterweight. At an
eccentricity of 9.4m (dotted lines) the Tower continues to rotate northwards as the
counterweight is increased. At an intermediate eccentricity of 7.8m (broken lines)
a curious response is obtained in which the Tower first moves northwards, then it
,,
-+@
Critical line
Simple studies carried out on one-g models on sand at Imperial College pointed to
the existence of a Critical line. Soil extraction from any location north of this
Critical line gave rise to a reduction in inclination, whereas extraction from south
of the line gave rise to an increase in inclination. The first objective of the
numerical analysis was to check whether the concept of a Critical line was valid.
Figure 6.78 shows the finite element mesh in the vicinity of the Tower.
Elements numbered 1,2,3,4 and 5 are shown extending southwards from beneath
the north edge of the foundations. Five analyses were carried out in which each of
the elements was individually excavated to give full cavity closure, and the
a)
b)
c)
d)
e)
-g
g
12
l
16
20
6.6.3.15 Comments
This section describes the development and calibration of two finite element
computer models of the Pisa Tower and underlying ground using the modified
Cam clay model with coupled consol-idation.One ofthe models is plane strain and
seconds has been achieved and the observed behaviour has been consistent with
that predicted by the analyses. It is doubtful that this very sensitive operation on
a Tower that is on the point of leaning instability would have been undertaken
without the positive results of the numerical analysis.
6.7
1.
2.
3.
4.
5.
6.
7.
8.
Summary
Shallow foundations can have different shapes in plane. If they are long in one
dimension, they are classified as strip foundations and can be analysed
assuming plane strain conditions. If they are circular, they can be analysed
either assuming axi-symmetric conditions, if the loading is vertical, or using
the Fourier series aided finite element method for general loading. For other
shapes a full 3D analysis is required.
When modelling surface foundations both the soil and the foundation should,
in general, be discretised into finite elements. However, if the loading is
vertical and the footing is assumed to be either very flexible or very stiff
compared to the soil, further approximations can be made in which it is no
longer necessary to include the foundation in the mesh.
The bearing capacity for a square surface footing is smaller than that of a
circular footing.
Finite element analyses enable theoretical shape factors to be determined for
surface foundations. These differ depending on the roughness of the footing,
but in general agree with the empirical formulae used in most design manuals.
For drained soils there are three coefficients in the general bearing capacity
equation, namely N, , N , and N, . For strip footings exact theoretical
expressions are available for N, and N, . Only approximate solutions (e.g.
stress fields) exist for N , . It has been shown that finite element analyses can
recover the theoretical correct values of N, , but more importantly such
analysis provide considerable insight in to the values of N , . They have also
indicated some surprising differences between the behaviour of strip and
circular footings.
No design guidance is available for the undrained bearing capacity of preloaded strip foundations on clay. It has been shown how finite element
analyses can be used to tackle this problem. It was concluded that for the
majority of real situations it is unlikely that pre-loading will give rise to a
substantial improvement in undrained bearing capacity.
Results of finite element analyses using the sophisticated MIT-E3 have been
presented to show how the effects of observed anisotropic soil behaviour can
be reproduced in finite element analyses. The effect of this anisotropic
behaviour on the bearing capacity of both strip and circular surface
foundations has been quantified by finite element analysis.
For shallow footings founded below the ground surface it is important to
correctly model the interface between the sides of the footing and the soil.
Interface elements are useful for this purpose, but must be given a zero tensile
stress capacity, as well as the appropriate shear strength.
Deep foundations
7.1
Synopsis
7.2
Introduction
In the previous chapter surface and shallow foundations have been considered.
While such foundations are usually adequate where the soils below the foundations
are both strong and stiff andlor the foundation loading is light, they may not be
appropriate in weak soils or in situations where the imposed loads are high. In such
situations it is common to extend the foundations deeper into the soil. This has two
beneficial effects, firstly soils usually become stronger and stiffer with depth such
that bearing capacity increases, and secondly shear and lateral stresses can be
mobilised on the sides of the foundation which provide additional resistance to
support the applied loads.
Another reason for increasing the depth of a foundation is where pull-out forces
have to be resisted. For example, the foundations for some offshore structures often
have to resist upwards forces. This can be resisted by the shear and lateral stresses
mobilised on the sides of a deep foundation.
The most common form of deep foundation is the pile. This can either be
constructed first and then driven into the ground, or a hole can be bored and the
pile cast in-situ. Driven piles can be constructed from steel, wood or reinforced
concrete and come in many different cross-sectional shapes (e.g. circular,
rectangular, H-shaped, etc.). Bored piles on the other hand are usually constructed
from reinforced concrete. Sometimes an undersized hole is bored, a pile is then
driven and grouted afterwards. There are clearly many other combinations of
installation that are possible.
Piles are rarely used in isolation but are installed in groups. They are usually
7.3
7.3.1
Single piles
Introduction
Although single piles are rarely installed in isolation, it is useful to consider their
analysis, as this is similar to, but less complex than, the full analysis of a pile
group. As will be discussed subsequently, the latter analyses are currently beyond
the capacity of present computing resources for all but some very simple cases.
If the pile is circular in cross section, installed vertically in the ground, which
has a horizontal lithology and horizontal water table, then the geometry to be
analysed is axi-symmetric. If only vertical loading is applied to the pile then an axisymmetric analysis is appropriate. However, if either inclined andlor moment
loading is applied to the pile then a three dimensional analysis is required. As the
geometry is axi-symmetric this can be performed with the computationally
efficient FSAFEM. If any ofthe above conditions are not satisfied, for example the
pile is not circular, or is not installed verticalfy in the ground, or the soil lithology
is not horizontal, then a full 3D analysis using three dimensional finite elements
must be used even if axial loads are applied to the pile.
For the remainder of this section the geometry will be assumed to be axisymmetric. Consequently, for vertical loading axi-symmetric analysis, and for
lateral loading three dimensional analysis using the FSAFEM, will be considered.
7.3.2
Vertical loading
One of the important issues when analysing a pile subject to vertical loading is the
modelling of the interface between the pile and the soil adjacent to the pile shaft.
As noted in Section 6.6.2, problems can arise if the elements representing the soil
are too large and no interface elements are used. To overcome the problem it is
important to either use thin solid elements in the soil adjacent to the pile shaft
andlor use interface elements. However, for the pile problem care must be
exercised when assigning material properties to the interface elements.
interface. This problem has been discussed in Section 6.6.2. The ultimate base
capacity suggests a bearing capacity factor Nc=9.
Three further analyses were then performed, all of which used the finite
element mesh shown in Figure 7.4, but with zero thickness interface elements
positioned adjacent to the pile shaft and its base. The same soil properties as above
were used and the interface elements were assigned an undrained strength
S,,=100kN/m2 (i.e. the same as the soil) and equal shear, K, , and normal, K,, ,
stiffnesses. The only difference between the three analyses was the values assigned
to K, and K,, ,these being 103kN/m3, 105kN/m3 and 107 kN/m3 for each analysis
respectively. As noted in Chapter 9 the choice of stiffness values for interface
elements is problematic, especially because they have different units (kN/m3) to
those of the soil (kN/m2).
The resulting total pile head load8ow
displacement curves for the analyses
with interface stiffnesses of 103kN/m3
and 105kN/m3are compared with each X
other and with that obtained from the
4000
analysis with no interface elements in .B
Figure 7.6. The analyses with an 3t: 7.000
interface stiffness of 105kN/m3andno
interface elements agree quite well,
oo 0.02 0.04 0.06 008 o 1 0.12 0.14
the former producing a slightly less
Displacement (m)
stiff response and having a slightly
smaller limit load. Closer inspection
,=igure7.6: Effect of interface
of the results indicates that the element stiffnesson pile behaviour
ultimate shaft capacity is now
6279kN, which is near to the value obtained from the simple calculation in which
the undrained strength is multiplied by the shaft area (i.e. 6283kN). As the
integration points for the interface elements are located on the pile-soil interface,
the predicted shaft capacity is likely to be more accurate than that obtained from
the analysis with no interface elements.
The results from the analysis with an interface stiffness of 107kN/m3 are not
shown on Figure 7.6 for clarity. During the early stages of loading the predicted
load-displacement curve is indistinguishable from that of the analysis without
interface elements. The ultimate load, however, is slightly smaller than that
obtained from the latter analysis, but is in agreement with that from the analysis
with an interface stiffness of 105kN/m3.
The analysis with the softer interface stiffness of 103 kN/m3 predicts a much
softer load-displacement curve, see Figure 7.6. It also does not appear to reach the
same ultimate load. The reason for this can be seen in Figure 7.7, which shows the
separate mobilisation of shaft and base resistance for this analysis. While the shaft
resistance is fully mobilised at a displacement of 0.105m, the base resistance is
extremely small and requires considerably more displacement before it becomes
fully mobilised.
g ,,
d
5
g
g
g
Displacement (m)
Displacement (m)
The failure of the analysis with v=25O to predict a limiting pile load is a
consequence of the kinematically constrained nature of the problem, combined
with the continued plastic dilation predicted by the Mohr-Coulomb model. The
problem is not confined to this model but will occur with any constitutive model
that predicts finite plastic dilation indefinitely, without reaching a critical state
condition. This potential pitfall is discussed in greater detail in Chapter 9.
The use of interface elements
6000
between the soil and the pile can, in
some respects, suppress the above ,
5000
shortcomings of the constitutive g 4000
model used to represent the soil. For
o,,
example, the drained analyses have
.B 20,
been repeated with zero thickness
1000
interface elements adjacent to the pile.
These elements were given a stiffness
0
0.01 0.02 0.03 0.04 0.05 0.06 0.07
K,,=K,=lOS kN/m3, an angle of
Displacement (m)
shearing resistance 9'=2S0 and an
angle of dilation v=OO. The loadFigure 7.10: Drained loading of a
displacement curves are shown in
single pile with interface elements
Figure 7.10. As with the undrained
along its shaft
analysis discussed above, the
interface elements dominate the behaviour and the results of the analyses are much
closer. Figures 7.1 1a and 7.1 1b show the mobilisation of the shaft and base
resistance for the two analyses. There are differences in the base resistance. If the
pile is subject to further displacement a limiting base resistance is predicted by the
analysis with v=OOin the soil, while no such limiting value is predicted for the
analysis with v=2So in the soil. There are also small differences between the
mobilised shaft resistance from the two analyses. This arises due to the different
normal stresses mobilised along the pile shaft, especially in the vicinity of the pile
base. It can therefore be concluded that while the behaviour, and therefore
properties of, the interface elements dominate the mobilisation of pile shaft
resistance, they have only a small influence on the base resistance.
l-
Displacement (m)
Displacement (m)
7.3.3
Lateral loading
This results in an incremental lateral displacement of the top of the pile of Ad. The
remaining boundary conditions consisted of restricting all three components of
displacement (i.e. U,, U , and U,) along
4000
the base of the mesh and the radial
displacement (U,) on the right hand
IOW
side of the mesh. Unlike the vertical
analyses described above, no $
displacements were restricted on the 4 2000
left hand side of the mesh which
looo
represents the axis of symmetry of the 8
geometry. Under lateral loading this
o
0.02
0.04
0.06
0.08
0.1
boundarv must be free to move. The
Displacement (m)
analyses were performed using
parallel symmetry and ten harmonics,
Figure 7.12: Undrained lateral
see Chapter 12 of Volume 1.
loading
of a single pile with and
Initially the soil was assumed to
without
interface
elements along
be undrained with E,=1O5 kN/m2,
its
shaft;
constant
S, with depth
,u=0.49andS,,=lOOkN/mZ.Theresults
.g
Horizontal displacement
of pile (m)
g
E
500
400
3oo
'3 '0
100
.,
g 10
-
.$
4
PI
20
0
7.4
7.4.1
introduction
Displacement(m)
100
200
300
400
I;-
-6
'
<
y ; :F
This new method is divided into two parts, the analysis of a pile group
(Ganendra and Potts (2002a)) and the design of a pile group (Ganendra and Potts
(2002b)). The proposed method of analysis of a pile group is able to give
predictions of pile group rotation, lateral deflection and vertical settlement for a
particular applied load, which may consist of any combination of axial load, lateral
load and turning moment. This method is essentially an extension ofthat described
by Jardine and Potts (1988), (1992) which was for vertically loaded pile groups
only.
The proposed method of design for a pile group provides 'design charts' for
any particular pile group such that predictions of pile group rotation, lateral
deflection and vertical settlement can be obtained for any combination of axial
loads, lateral loads and turning moments. Note, however, that the design method
does not give any recommendations regarding the number, size or configuration
of piles in a pile group.
As this approach is based on nonlinear finite element analysis of a single pile
and as it appears to provide many advantages over current design approaches, it
will be described in subsequent sections of this chapter. The application of the
proposed analysis and design methods to the foundations of the Magnus oil
platform will then be discussed.
The manner in which the loads applied to one of the piles in a group affects
the displacement of the other piles.
The manner in which the loads applied to the group are distributed between
the piles in the group.
The proposed method of analysis uses superposition techniques for the first
assumption and a rigid pile cap criterion for the second assumption, both of which
are described below.
7.4.3
Superposition
The displacements of any pile are affected by the loads imposed on the
surrounding piles. This effect is accounted for in the proposed method of analysis
- Simple superposition;
- Nonlinear superposition;
Virtual pile.
Of these the simple superposition has been used in this chapter and it is described
below. The other techniques are presented in Ganendra (1993).
where r, and 8, are the coordinates expressed in terms of the local cylindrical
coordinate system for the thpile, which has its axis about the centre of the pile, see
Figure 7.18. The function C,(r,,8,) is obtained from a 3D analysis (i.e. using the
FSAFEM) of an isolated pile.
The simple superposition method assumes that the additional displacement U/
of the jth
pile, induced by the load F, on the thpile, is equal to C,(r:, B:). r,' and 8,'
are the coordinates of the/lh pile expressed in terms of the local coordinate system
for the ithpile. Thus the displacements U,;/of the jth
pile in a pile group can be
obtained using simple superposition, by adding the displacement contributions
from all the piles in the pile group:
where:
U/ = C,(r,',
jthpile;
(r,', 0,') = (0,O) are the coordinates for the centre of the jlh pile;
U,! is the additional displacements induced at the jThpile due to the
loads imposed on the other piles in the group.
The vectors of displacements U,,' and u j are expressed in the pile group global
Cartesian coordinate directions X, y and z (where z is the depth from ground
surface), see Figure 7.18. The origin for the x and y coordinates is located at the
plan view centroid of the pile group and shall be referred to as the group datum.
Thus the ground surface displacements (obtained from a FSAFEM analyses)
associated with the loads on each pile have to be resolved from the local axisymmetric coordinate system of the pile into the global pile group Cartesian
coordinate system.
Only pile head and ground surface displacements are considered using this
superposition technique. Methods of extending these superposition techniques to
predict pile displacements with depth are described subsequently. An important
feature of this method is that the ultimate load of a pile is unaffected by the loads
on the other piles, however the displacements at which this load is mobilised will
be different.
This approach is a generalisation of the method used by Jardine and Potts
(1988) and is schematically illustrated in Figure 7.19.
Load
Displacement
The ratio between the superposition pile head displacements and the
isolated pile head displacements is used to factor the isolated pile
displacements (along the length of the pile). The stresses in the pile
associated with these factored displacementswould no longer be consistent
with the loads applied to the pile.
The difference between the superposition pile head displacements and the
isolated pile head displacementsare added to the isolated pile displacements
(along the length of the pile). The resulting pile stress distribution is the
same as the isolated pile stress distribution and thus is consistent with the
7.4.4
The manner in which the loads are distributed between the piles in a pile group
depends on the nature of their pile cap connectivity and the stiffness ofthe pile cap.
In the Ganendra and Potts approach a rigid pile cap criterion is used, which
assumes that the pile cap is fully rigid and there is full connectivity between piles
and pile cap at ground level. In this case the pile head displacement and rotations
of any pile can be evaluated from their location within the pile group and the pile
group displacements and rotations. These pile group displacements can be
expressed as U,, U, and U:, which are the displacement of the pile group datum
in the X, y and z (where z is depth) global coordinate directions respectively. The
pile group datum and the origin for the global coordinate axes are taken to be the
plan view centroid of the pile group at ground level, see Figure 7.18. The group
rotations can be expressed as 0, and B,, which are the rotations of the group about
the X and y axes respectively. As a simplification, torsions applied to the pile group
and accordingly rotations about the z axis are not considered. U,, U,, U,, Or and
O, are schematically illustrated in Figure 7.20.
Due to the rigid pile cap criterion, the displacements U,', U,,' and U,' (in the X, y
and z directions respectively) of any thpile in a pile group can be related to the pile
group displacements via the equations:
where X' and y' are the x and y coordinates respectively of the thpile.
Due to the assumption of full connectivity between pile and pile cap, the pile
head rotations 0; and 0; (about the X and y axes respectively) of the any thpile
in a pile group should equal the global group rotations:
Note that the global group displacements and rotations (U,, U,, U:, 0, and 0,) can
be evaluated from the pile head displacements, rotations and locations (U,', U,,',U,',
O,'? Q,,' X' and y') from any one pile in the pile group.
The loads that are applied to a pile group can be expressed as forces F,, F, and
F, and turning moments M, and M,. F,, F, and F, are the forces in the X, y and z
directions respectively, and M, and M, are the turning moments about the X and y
axes respectively. They are schematically illustrated in Figure 7.2 1. These loads
are distributed among the piles in the group in such a way that the rigid pile cap
criterion is satisfied, i.e. the same values of global group displacements and
rotations (U,, U,, U,, Oxand 0,) are evaluated from the pile head displacements
and rotations (U:, U,', U:, 0:' and 0)'; of each pile in the group. An iterative
pra'cedure can be used to evaluate these individual pile loads. The steps in the
procedure are:
Obtain an initial trial division ofthese applied loads between the piles in the
group;
Evaluate, using a superposition technique, the displacements of each
individual pile due to the loads applied to it and the loads applied to its
surrounding piles;
These pile displacements are used to check if the rigid pile cap criterion is
satisfied:
- If it is satisfied (to within a small tolerance) this trial division of loads
is assumed correct and the resulting solution for the global group
displacements is accepted;
- If the rigid pile cap criterion is not satisfied, a new trial division of loads
is obtained and the procedure repeated until the criterion is satisfied.
M;p, the push-pull moment, which is the moment resulting from the
difference in pile axial loads in the x direction;
M?, the pile head bending moment, which is the sum of the bending
moments about they axis which are applied to the pile cap from each pile
in the pile group.
My"P=M
KY""
K,PP
and M; = M y
K~+K?
K ~ P+P~
yhb
(7.6)
2.
3.
4.
5.
6.
7.
8. Check that the values of O, obtained from checks 3 and 6 are the same. If
not, the ratio between M,PP and Myhbis adjusted accordingly.
9. Check that the values of Oxobtained from checks 4 and 7 are the same. If
not, the ratio between M,PP and Mxhbis adjusted accordingly.
If any one of these conditions is not satisfied to within a prescribed tolerance, the
appropriate corrections are made and a new set of pile head displacements is
evaluated and all nine checks repeated. If all nine conditions are satisfied
concurrently, the solutions for U,, U,, U,, Oxand O, are accepted.
The above procedure seems very cumbersome. However, the trial division of
loads described in Section 7.4.4.1 should give solutions that satisfy (to within a
reasonable tolerance) the majority of the checks described. For most cases it is
anticipated that only checks 8 and 9 may not be fulfilled by the trial division and
that relatively few iterations would be required to obtain a solution that satisfies all
nine conditions. The solutions obtained from this proposed method ofanalysis only
make two basic assumptions:
loads applied to one of the piles in a group affect the displacements of the
other piles;
The rigid pile cap criterion can be used to determine the manner in which
the loads applied to the group are distributedbetween the piles in the group.
7.4.5
Pile group design
The design of an offshore platform is a complex interdisciplinaryprocess of which
foundation design is a component. The method of analysis of a pile group
described in Sections 7.4.2 to 7.4.4 is of limited use in such a procedure, since it
is an iterative process for obtaining solutions for a particular set of applied loads
and it requires that 3D finite element analyses (i.e. using the FSAFEM) of each pile
be undertaken for each iteration. Predictions of this type are more suitable for back
analyses than for design. Thus to provide a more flexible tool for the design of an
offshore structure a 'design chart' procedure was devised which gives predictions
of pile group displacements and rotations for any set of design loadings (without
recourse to additional 3D finite element analysis). The manner in which these
design charts are obtained and how they should be used is described below. The
assumptions that are made in addition to those made for the analysis of a pile group
are stated. The procedure used for obtaining these design charts extends the linear
elastic procedure described by Randolph and Poulos (1982) to incorporate
nonlinear soil behaviour.
to the pile group, F,, F,, F,, M, and M, (also defined in Section 7.4.4), as vectors
Uand Frespectively. These vectors are related via a pile group stiffness matrix K:
In the linear elastic formulation described by Randolph and Poulos (1982), the
components ofthe matrix [Kj are constants. Thus if all the components of [Kj were
known, displacement U can be calculated for any applied load F by premultiplying Fwith the inverse of [K]. Evaluating these values of U for an applied
load F is more complex for nonlinear soil behaviour. The components of [W are
no longer constant and vary as a hnction of the applied loads and displacements.
Let the loads required to cause displacements U,, U, U,,,, UOxand U,, be F,, F,
F,, ,F,,, and F,,,, respectively. The assumption of superposition of loads means that
the applied load F, required to produce group displacement U, is equal to the sum
of the of the loads F,, .Fm,F,,, , F,, and F(,:
The components of the load vector F, are given by the reaction forces that have
to be applied to the pile group in order that it displaces a distance U.vertically,
while all other components of displacement and rotation are fixed (at zero). The
same principle can be used to relate the other load vectors, F,, F,, ,F(,,and F,, ,
with their associated displacements, U,, U,, B, and B, respectively. The manner in
which these load vectors can be obtained from single pile FSAFEM analyses is
described below.
- (Ml$')pp, the push-pull component, which is the moment resulting from the
The value of M(,,? for a particular Oy value is found by performing the above two
calculations and then summing the values of (M@Y)PPand (M,$)hb obtained.
The components of the load vector F,, are evaluated in a similar manner. Note
that these values of (Ml?;)PP and (M,,?')hb can be used to give the estimates of K,PP
and K,hb required in Section 7.4.4.1 :
(M&IPP and ~ , h =
~ v p p,b(M;)hb
@Y
@Y
(7.1 1)
The various terms in [W and the method in which each term is derived is
depicted in Equation (7.12):
where:
K:
KZ,;
K,,,"Y
K,,?
K,,"
K,,;
K,,"'"
K,,""
K,,?
Using [K] the displacement U can be calculated for any applied load F.
However, the nonlinear nature of soil response means that [Kl is not a constant but
varies with U. An iterative technique is described below which ensures that the
value of U due to an applied load F is evaluated from a [K] which is consistent
with U.
The procedures described above allow the variation of F,, F,, F,, F,, and F,,
with respect to their associated scalar displacements U : , U, , U , , O, and O,
respectively, to be evaluated. These can be used to relate the vectors of secant
stiffness K,, K,, K,, , K,, and K,, with their associated scalar displacements U:,
U,, U,, Ox and O, respectively. These relationships can be expressed as design
charts which can be used to obtain the pile group displacements, U, due to any
applied pile group load, 8;: A simple iterative procedure for obtaining U is as
follows:
A trial value of pile group displacement, U', is used to evaluate from the
design charts the initial trial values of K,, K,, K,, ,K,,, and K,,;
These vectors are combined to obtain a trial value of pile group stiffness
EK!;
A trial solution for the pile group displacement, U""',is obtained based on
this value of [K] and the: applied loads F;
The validity of this solution is evaluated:
- If F = U""'(to within a specified tolerance) this solution is accepted;
- Otherwise this value of U"'is used as the new value trial displacement,
U',and the procedure repeated.
7.4.6
Magnus
7.4.6.1 Introduction
The design and analysis procedures
Deckload 330 MN
for pile groups described above were
applied to the foundations of the
Magnus platform. Magnus is located
in the UK sector of the North Sea and
was installed in 186m of water in
1982. The structure is shown in
Figure 7.27. When the structure was NO
being designed, in 1979, it was to be
the deepest and most northern
structure in the North Sea. This
resulted in particularly severe
environmentaldesign criteria. This, in
with the very high deck Figure 7.27: Schematic presentation
of the Magnus structure
loads, results in the structure being
one of the heaviest steel structures
installed off-shore (at float out it weighed 340,000 kN). To verify the existing
foundation design methods the foundations for one ofthe legs (A4 in Figure 7.27)
were instrumented. Jardine and Potts ( 1 992) studied the vertical response of this
foundation using an axi-symmetric finite element analysis. The predictions
obtained were in good agreement with the field measurements.
In this section the work carried out by Jardine and Potts (1992) is extended to
consider the Magnus foundation response to general loading (axial and lateral
forces and turning moments). Design charts ofthe type described in Section 7.4.5.3
are derived. These design charts and the method of analysis described in Section
7.4.4 are used to analyse the performance ofthe foundations when subjected to the
loads measured from the largest wave of the stom on the 22ndJanuary 1984.
DESCRIPTlON
GEOTECHNICAL.
a
.!m
10
20
1
'I1
80
t
U
Elastic behaviour is governed using the small strain stiffness model described
in Section 5.7.5 of Volume 1. Minimum values of E,and E, are specified such that
at strains below these values the shear and bulk moduli respectively are assumed
to be constant. Similarly, maximum values of E, and E,, are specified. The same
small strain parameters were used for all the stratum and are shown in Table 7 . 2 .
+ O g -3O a P
0-= Gap
where:
0 is
r
lnro
The overconsolidation ratios, OCR,specified for the soil were such that the
same undrained shear strength (in triaxial compression) profile with depth, as
shown in Figure 7.29, was obtained from the soil at all radial locations. The
resulting OCR variation with depth adjacent to the pile and in the far field is plotted
in Figure 7.30. To obtain the required undrained shear strength profile, a semi-log
variation, similar to that given by Equation (7.13), was used for the OCR of the soil
less than twenty radii away from the pile. The details of how these soil parameters
and the initial stress regime were derived are described fully in Jardine (1985).
OCR
Table 7.3: Specified effective stress regime at pile face and far-field
B.
5
9
'5
Deep foundations / 31 1
Q
Q
Deep foundations / 3 1 3
rotation, 0,. This is presented in Figure 7.39. Due to the linear nature of its two
constituent parts this response is also strongly linear, as can be seen from the
variation of secant stiffness, K@my,shown in Figure 7.39. K,,? is assigned a
constant value of 1 159GNm for design purposes.
evaluated above
The values of secant stiffnesses K,', K,,' KGy, Kwx and KhnJY
can be combined to give the pile group stiffhess matrix
[a:
pile 3&8 A
pile 2&9
-0.2
- coordinate (m)
- coordinate (m)
Horsnell et al. (1992) evaluated an average pile group stiffness matrix on the
basis of the measured response of the Magnus platform:
This conventional group stiffness matrix is significantly softer than the stiffness
matrix inferred from field measurements. The axial stiffness, K;, is over 4.5 times
softer than the measured axial stiffness and the other terms in the stiffness matrix
are approximately twice as soft as the corresponding measured terms.
The predicted displacements due to this imposed load obtained from the new
design method, Equation (7.14), and from the conventional design stiffness matrix,
Equation (7.16), are presented in Table 7.4. The new design method required four
iterations to obtain the solution value of U,. Also shown on Table 7.4 are the
measured displacements of the pile group for this wave (Horsnell et al. (1992)).
These measurements are based on the readings from accelerometers at the base of
Deep foundations / 31 5
the leg. The new design predictions are in good agreement with the measured
displacements, underpredicting lateral displacements and rotations and
overpredicting settlements. The conventional design predictions are significantly
larger than the measured displacements,with settlements over four and a halftimes
larger and lateral displacements and rotations approximately twice the size of
measured lateral displacements and rotations.
Table 7.4: Predicted and measured pile group displacements
displacements
9
B
- coordinate (m)
-Measured
displacemen&
-10
- coordinate (m)
The method of analysis described in Section 7.4.4 was also used to give
predictions of pile group displacements and rotations. The initial trial division of
loads gave predictions that satisfied all but one of the rigid pile cap conditions
given in Section 7.4.4.3, the values of 69, from pile head bending and push-pull
loading were not equal. Five iterations were required to obtain a solution that
satisfied all the conditions. Note that additional FSAFEM analyses were required
for each iteration. The predictions for lateral displacement of each pile from the
method of analysis are plotted against X coordinate in Figure 7.42. Also presented
are the measured displacementsand the predicted displacementsfrom both the new
and conventional design methods. There is good agreement between the new
design and new analysis predictions. This implies that the assumption of
superposition of loads described in Section 7.4.5.2 is reasonable.
Figure 7.42 also shows the good agreement of these predictions with the
measured displacements, and the large overprediction in displacements obtained
from the conventional method. The validity of the assumption of superposition of
loads and the accuracy of the predictions from the new methods of design and
analysis are also illustrated in Figure 7.43, which is a plot of vertical settlement
against X coordinate for the same imposed load.
Kenley and Sharp (1992) provide distributions for bending moment with depth
for a pile in the pile group subjected to the above wave loading. Distributions
obtained from both conventional design methods and measurements were
presented. The measured bending moments were derived from strain gauges at
different levels in different piles to give a typical bending moment distribution
within a pile. These bending moments are shown in Figure 7.44 with the predicted
bending moments from the new method of analysis. The new method of analysis
gives a better agreement with the measured bending moments than the
conventional method. In particular, the new method does not overpredict the depth
of the point at which the maximum bending moment occurs by as much as the
conventional method.
Bending moment ( m m )
Figure 7.44: Distribution of
bending moment with depth
Displacement (mm)
Figure 7.45: Distribution of
lateral displacement with depth
Deep foundations 1 31 7
7.5
Bucket foundation
7.5.1
Introduction
Suction caissons (or 'bucket' foundations)
are becoming extensively used in the
offshore industry as deep water anchors for
floating structures or foundations for oil
platforms. The caissons are hollow
cylindrical structures which have a top cap
and a relatively thin wall, the so called skirt.
One example of floating tension leg oil
platform on bucket foundations is the Snorre
platform, illustrated in Figure 7.46, which
has a cluster of three buckets connected to
each of its four foundation legs
(Christophersen et al. (1992), Jonsrud and
Finnesand (1992)).
Installing a bucket foundation involves
initial penetration into the sea-bedunder selfweight. The pressure in the water trapped
inside the bucket, between the soil surface
and the top cap, is then lowed by pumping, to Figure 7.46: Schematic view
cause a positive differential water pressure of Snorre platform foundation
template
across the top of the bucket, thus forcing the
bucket further into the soil until its final
position is reached.
In the case of the Snorre structure, dead weights were added to ensure that
compressive loading was acting on the bucket foundations. However, oil
exploration is moving more into progressively deeper waters, and there is
7.5.2
Geometry
The reference bucket foundation
geometry adopted for the research
study is shown in Figure 7.47 and is
based on the geometry of the bucket
foundations installed at the Snorre
platform. A single concrete cylinder is
analysed, with a.skirt 0.4m thick and
a top cap 1.Om thick. The diameter of
the cylinder, D, is 17.0m, while the
skirt penetrates the sea-bed to a depth Figure 7-47:Idealised geometry for
finite element analysis
L=12.0m, giving a D1L ratio of 1.4.
The depth of water is also based on that at Snorre and is assumed to be 3 10m.
Undrained uplift loading is applied at the centre ofthe top cap, at an inclination
/3 to the vertical, see Figure 7.47. Analyses were performed varying the inclination
p from 0 to 90'. The scenario considered is typical of that for a tension leg
platform (TLP). Other types of floating structure may apply their loads in different
ways.
I
7.5.3
Deep foundations / 31 9
the bucket and in the soil. This implies that any realistic numerical analysis of this
problem must be three dimensional. Only in the special case of vertical loading
(i.e. p = 0") can the problem be assumed to be axi-symmetric.
The FSAFEM method was D,2=8,5m
used in this study for the analyses
ofbucket foundations subjected to
inclined loading, adopting the
finite element mesh presented in
Figure 7.48. As indicated in this
figure, prescribed displacements
were imposed on both the bottom
and right hand side mesh
boundary. No such displacement
restrictions are imposed on the left
hand side boundary, because it
represents only the geometrical Figure 7.48: Finite element mesh for
axis of symmetry. The parallel
bucket foundation analysis
symmetry option was adopted and
ten harmonics were used to represent the variation of all quantities in the
circumferential direction. A small parametric study, varying the number of
harmonics, was performed to ensure that this was adequate. For a vertically loaded
foundation (i.e. P=O0) a conventional axi-symmetric finite element analysis was
performed, which further reduced the computational effort required.
The bucket foundation was modelled as concrete material, assuming linear
elastic behaviour (with Young's modulus E = 30* 106kN/m2 and Poisson's ratio
p=0.15). This implies that the concrete has sufficient strength and reinforcement
to sustain the applied loads. The installation of the foundation was not modelled
in the analyses, the 'bucket' was assumed to be 'wished in place'.
For the conventional axi-symmetric analyses, when P=O0,the foundation was
loaded to failure by applying increments of vertical uplift displacement, v, at the
centre of the top cap. The reaction force at this location represents the applied load.
For the foundation subjected to inclined loading, when P+O0,the analyses had to
be performed under load control, with increments of vertical and horizontal force
being applied as line loads (in the circumferential direction) at the node
representing the top right hand corner of th: top cap, point A in Figure 7.48. As
with the analyses described in Section 6.5.7.4 for circular footings, for the vertical
load (V) only the zero harmonic coefficient was specified as non-zero. To obtain
the horizontal load, both radial (H,) and circumferential (H,,) line loads were
specified. For the radial load only the first cosine harmonic was specified as nonzero, whereas for the circumferential load only the first sine harmonic coefficient
was set. The latter was of opposite sign but equal in magnitude to that set for the
radial load. The magnitudes of these increments were kept in the proportions
required to give the desired loading inclinations.As failure was approached, it was
necessary to reduce the size of the force increments to obtain an accurate estimate
of the ultimate loads. The resultant load from these boundary conditions acts at the
centre of the top cap as desired.
The finite element mesh shown in Figure 7.48 was used for both axi-symmetric
and FSAFEM analyses. It consisted of 350 eight noded isoparametric elements.
Undrained conditions in the soil were simulated by assigning a high bulk stiffness
(=1000K,,,,, where K,,, is the effective bulk stiffness of the soil skeleton) to the
pore water (see Section 3.4 of Volume l).
7.5.4
Modelling of the interface between top cap and soil
To model the interface between soil and the top cap, zero thickness interface
elements have been used (see section 3.6 of Volume 1). At the outset of this
research programme, three different conditions at the interface between the soil and
the top cap were identified. These are shown diagrammatically in Figure 7.49. In
Case 1 the bucket foundation has been perfectly installed such that there is no
water layer between the soil surface and the underside ofthe top cap. It is assumed
that the bucket foundation is water tight. Case 2 represents the other extreme where
a water layer exists between the soil surface and the underside of the top cap and
the pressure in the water remains equal to that at the sea bed outside the foundation
(i.e. there is a hole in the top cap). Case 3 models the intermediate situation in
which a water layer exists, but the foundation is water tight.
All three situations were
hole
modelled in a conventional
axi-symmetric finite element
analyses in which the
foundation (D= 17m,DIL= 1.4)
was subject to a vertical pullout load. The isotropic~soil
Case 1: NOinterface
Case 2: Interface elements
properties described in the
with no tension capacity
next section were used for
these analyses. Case 1 was
modelled by omitting the top
cap interface elements and
Case 3: Interface elements
thereby simulating perfect
with tension capacity
contact between the soil and
the underside of the top cap.
Case 2 was mdelled by Figure 7.49: Possible scenarios for the
letting the interface ekments
interface between the top cap and
open freely under any vertical
undersoil
loading and Case 3 was
modelled by giving the interface elements a very low shear stiffness, but a high
normal stiffness. In addition, if the water pressure was to reduce to an absolute
value of - IOOkPa, a facility was provided to reduce the normal stiffness to a very
small value, thus simulating cavitation. For all analyses full wall adhesion was
assumed between the soil and the skirts of the bucket foundation.
with depth (see Figure 7.5 l), giving Su/a,,'=0.33, which is typical of soft clays in
triaxial compression (Hight et al. (1987)).
Table 7.5: Modified Cam clay parameters for soft clay
In each of these studies the inclination, p, of the pull-out force, T, was varied
between 0" and 90".
7.5.5.3 Results
Results from a typical analysis are shown in Figure 7.52. This analysis is for a
,4
bucket foundation with D=17.0m
and L=12.0m (i.e. DIL =1.4), 8
12
subjected to a pull-out force, T,
10
inclined at 70" to the vertical (i.e. 3
P=70). The vertical component, V,
of this force is plotted against the 8 6
vertical displacement and the
4
horizontal component, H, of this .g 2
force is plotted against the X
'0
0.2
0.4
0.6
0.8
1
1.2
1.4
horizontal displacement of the
Horizontal and vertical displacement (mm)
foundation. Figure 7.52 shows that
clear limit values are predicted for
Figure 7.52: Typical load-displacement
both the horizontal and vertical
curves for inclined loading
load components at failure.
!g
. . . . . . . . . , . . . . . . . .. . . . . . . . .
: . . . . ._. .. , .
. . .. .
. / I .
l
I
The results plotted in Figure 7.54 are replotted in a normalised form in Figure
7.55. The data from a particular foundation geometry and for a particular load
inclination has been normalised by dividing the ultimate horizontal force, H,,,, and
the ultimate vertical force, V,, , by the maximum vertical force, V, . V, is the
maximum vertical load found from the analysis involving purely vertical loading
for that particular foundation geometry. The values of H,,/V,, and V,,,/V,,, found
for all the geometries and pull-out inclinations indicate a unique relationship which
can be represented by an ellipse of the following form:
All data fit this ellipse with a maximum error of only (1.5 to 3.0)%.
Clearly, Equation (7.17) cannot hold for large DIL ratios because in the limit,
when the embedment depth L approaches zero, the ultimate horizontal force also
becomes zero (i.e. the water interface between the underside of the top cap and the
soil cannot sustain shear stress). Further analyses have been performed to
determine the limiting value of DIL over which Equation (7.17) does not apply.
These analyses indicate that Equation (7.17) is valid for DIL < 4.
e D/L=17/12
Fitted ellipses
90"
E
A
DIL=17/17
DIL=17/8
D/L=17/24
DIL=8.5/8.5
- The finite element predictions for a particular geometry can be fitted with
The results are also plotted in normalised form in Figure 7.57, together with the
ellipse from Equation (7.17).They agree with this curve with a maximum error of
6%.
Fitted ellipses
D/L=I7/12
D/L=17/24
7.5.6
Anisotropic study
7.5.6.1 Introduction
To model the anisotropic behaviour of the soil the MIT-E3 model has been used.
Selection of appropriate parameters for a soft sea bed clay have been discussed in
some detail in Section 6.5.7. The parameters are listed in Table 6.4 and these have
been adopted for the analyses presented here.
The undrained triaxial compression strength profile for this soft clay is shown
in Figure 6.42. It varies linearly with depth, starting from a finite value at the soil
surface, giving S,,/cv'=0.36. This triaxial compression profile differs slightly from
that used in the previous isotropic analyses, but is a consequence oftrying to model
the anisotropy experienced in the Hollow Cylinder Apparatus (see Section 6.4.1).
The difference has no bearing on the conclusions that come.
7.5.6.2 Results
In order to quantify the effect of soil anisotropy on the pull-out capacity of bucket
foundations two sets of analyses were performed, assuming the foundation to have
a reference geometry (DIL=17/12) and full skirt adhesion:
Analyses where the bucket foundation was loaded with a tensile force
whose inclination, p, varied from 0" to 90" and in which the soil was
assumed to be anisotropic, i.e. to possess the variation of undrained strength
presented in Figures 6.4 1 and 6.42.
Analyses where the bucket foundation was loaded in the same way, but the
soil was assumed to be isotropic. The soil properties for the MIT-E3 model
were chosen such that, for any inclination a of the major principal stress,
the undrained strength was the same as that obtained from the anisotropic
soil when sheared with a=OO, see Section 6.5.7.3. This means that the
effective stress paths for any a match the effective stress path for test MO
in Figure 6.4 1 .
Again, an interaction diagram of ultimate vertical and horizontal forces for each
particular inclination of loading is produced, as shown in Figure 7.58. The upper
envelope represents the results from the isotropic analyses, while the lower
envelope is for the anisotropic analyses. Similar to the previous analyses, both
isotropic and anisotropic numerical predictions can be fitted with a simple elliptical
curve, with a maximum error of k (0.6 to 1.2)%. However, the main conclusion
from these analyses is that taking account of soil anisotropy reduces the pull-out
capacity of the bucket foundation. For this particular soil the anisotropic analyses
lead to capacities that are about 22% lower than those for isotropic cases. Clearly,
the reduction of pull-out capacity due to anisotropic soil behaviour will depend on
the amount of strength anisotropy that each particular soil possesses.
The resulting normalised interaction diagram is presented in Figure 7.59,
together with the ellipse from Equation (7.17). The analytical data points conform
to the curve with a maximum error of 6%.
7.5.7
Suction anchors
7.5.7. 1 Introduction
In the previous sections the behaviour of bucket foundations loaded at the centre
of the top cap has been considered. Such a loading situation is typical of that of a
tension leg platform, but may not mobilise the maximum possible capacity of the
anchor. Recent results from large scale field test have shown that by lowering the
attachment point of the loading tether from the top to half way down the skirt, the
load capacity increases (Anderson (1998)). To illustrate this effect the results of
some numerical analysis in which the position of the loading tether is varied will
now be described.
7.5.7.2 Geometry
The bucket is now assumed to be made of steel and have a diameter of D=3m and
a skirt length of L=9m, see Figure 7.60. The loading tether is assumed to be
attached to a point some distance l down the skirt.
As the attachment point of the loading tether is positioned on one side of the
bucket and at some distance down the skirt, care must be exercised when
specifying the loading conditions. In the present analysis a single line load was
applied at the node on the outside of the bucket at the required depth down the
skirt. In addition a local axis (see Section 3.7.2 of Volume 1) was specified at this
node so that one of the axis aligned in the direction of loading (i.e. P to the
vertical). The line load was then oriented in this direction, but its magnitude varied
in the circumferential direction as shown in Figure 7.62, which shows the line load
distribution for an increment ofthe analysis. The load has an intensity of IOOkNlm
over a 40" sector of the skirt and is zero over the remaining part. This gives a
resultant load in the direction of the tether of lOOkN per increment.
This loading condition was specified using the fitted method described in
Appendix XII.7 of Volume 1. In this approach a series a line load intensities and
the values of B at which they act are input.
7.5.7.3 Results
Analysis have been performed with
the attachment point at four positions,
namely at the top of the skirt, one
third of the way down the skirt, half
way down the skirt and at the bottom
of the skirt. For each attachment point
analyses have been performed with
the direction of loading P equal to 0"
(i.e. vertical), 45", 70" and 90" (i.e.
horizontal), see Figure 7.60.
Angle of anchor inclination to the vertical c)
The ultimate capacities predicted
by these analyses are summarised in Figure 7.63: ultimate anchor pullFigure 7.63. Interestingly
the out force for different anchoring
analysis with vertical loading predict
points
approximatelythe same load capacity,
independent of the depth of attachment of the tether. However, this is not so if the
loading is inclined, where the analyses indicate that the greatest capacity is
predicted for the analysis with the attachment point located half way down the
skirt. The analyses are therefore in agreement with the field observations.
7.6
1.
2.
3.
4.
Summary
When analysing a single pile subject to axial loading either thin solid elements
or special interface elements should be placed adjacent to the pile shaft. If
solid elements are used and they are not sufficiently thin, the analysis will
over estimate the pile shaft capacity.
If interface elements are positioned adjacent to the pile shaft care should be
taken when selecting their normal and shear stiffness. These values, if not
sufficiently large, can dominate pile behaviour. However, if the values are too
large numerical ill conditioning can occur.
Soil dilation can have a dominant effect on pile behaviour and consequently
care must be exercised when selecting an appropriate constitutive model and
its parameters.
When analysing axially loaded piles using an effective stress constitutive
model it is unwise to use a model which predicts finite plastic dilation
5.
6.
indefinitely, without reaching a critical state condition (e.g. the MohrCoulomb model with v>Oo). Such analyses will not predict an ultimate pile
capacity.
When analysing a single pile subject to lateral loading the possibility of a
crack forming down the back of the pile (i.e. gapping) should be considered.
If this is likely to occur interface elements should be installed along the pile
soil interface, which cannot sustain tensile normal stresses. Wether or not
gapping is likely to occur depends on the soil strength and in particular its
distribution with depth.
A new method of design and analysis for pile groups is proposed. This
method is thought to be an improvement on the existing methods because:
the loads applied to one of the piles in a group affects the displacement of
the other piles;
The rigid pile cap criterion can be used to determine the manner in which
the loads applied to the group are distributed between the piles in the
group.
The design method produces design charts from which the pile group
displacements and rotations can be evaluated without recourse to additional
FSAFEM analysis. The basic assumption made by this method, in addition to
the two method of analysis assumptions, is that the loads required to cause
displacements U can be obtained by summing the loads required to cause each
component of U (superposition of loads).
9. The proposed analysis and design methods for a pile group were applied to
the foundations for the Magnus platform. The pile group stiffness obtained
from the new method of design was in good agreement with the measured
foundation stiffness.
10. The methods of design and analysis were used to predict the pile group
displacements and rotations for the loads imposed by the largest wave of a
storm on the 22ndJanuary 1984. The predictions obtained were shown to be
in good agreement with each other and with the measured displacements and
rotations. The predictions from the new methods are shown to be much more
accurate than the predictions from conventional methods.
8.
8. Benchmarking
8.1
Synopsis
This chapter discusses the benchmarking of finite elements programs, but the
comments made are equally applicable to the testing of other computer software.
The aim ofthis chapter is to demonstratethat problems can arise when undertaking
finite element analysis, which, unless they are detected, can have serious
implications.
A number of geotechnically orientated benchmark problems are presented and
discussed, together with the results of a benchmarking exercise carried out to
assess the reliability of a number of finite element programs.
8.2
Definitions
Benchmarking / 333
software development process to ensure compliance with software
requirements.
VerlJication: Establishing the truth or correctness by examination or
demonstration. The process of determining whether or not the products of a
given phase of the software development cycle fulfil the requirements
established during the previous phase.
When applied to the testing of computer software, and its operation, the terms
'validation' and 'verification' are often confused, since to most people they have
a very similar meaning. Although in some technical literature a distinction is made
between the two, in this chapter they are taken to mean the process of testing
computer software to evaluate its performance, assessing its accuracy and correct
operation by an individual and on specified hardware.
8.3
Introduction
Computers have for many years been available to engineers to aid design. In recent
years engineers have had the possibility to routinely use advanced numerical
methods (finite element analysis, boundary element analysis and finite difference
techniques) in design. The possibility to do so is largely due to the increased
availability of such programs, combined with the rapid pace of hardware
development. It is apparent that the use of these methods is becoming more
widespread, as more engineers appreciate the benefit of using such techniques to
examine complex problems and the sensitivity of these problems to various
influences.
When undertaking any calculation, an engineer has a responsibility to check
that the calculation is appropriate and numerically correct. A check should
therefore be made to ensure that the tools used are working correctly and that the
results obtained are reliable. It is important to remember that computers and
computer programs do not have a mind of their own, they simply execute a series
of pre-defined instructions. The question is, do these instructions correctly follow
the desired method of analysis and produce the intended result, andtor does the
user understand what the result is.
There is a number of reasons why errors may occur in calculations performed
using computer programs. These include errors or bugs in the software through
mistakes in coding or interpretation of the method of analysis, the use of the
software, or, in exceptional circumstances, there may be hardware problems. A
technique for checking the computer program and the use of it is known as
benchmarking. Essentially, it is a technique which forms part of the validation or
verification process.
This chapter starts by discussing, in general terms, the issue of verification and
validation of computer software and then goes on to discuss the benchmarking of
finite element programs, with particular reference to geotechnical engineering.
Most of the principles and techniques are equally applicable to the finite element
analysis of other problems (structural, thermal, hydraulic, etc.) and to the use of
computer programs in general.
8.4
The computer programs and hardware need to be checked before they are used,
in order to ensure that none of these problems exist. As the three examples of
failures resulting from errors demonstrate (see Section SS), the consequences of
Benchmarking / 335
an error being made could be severe and could have extremely serious
implications. It is obviously important that any of the potential problems with the
computer program, or the use of it, are identified and dealt with.
8.5
Consequences of errors
the project (NCE (1 99 l)). At the time concerns were raised about the safety of
other turbine blades produced by the same manufacturer.
Sleipner offshore platform (1993). In August 1993 the concrete gravity base of
the Sleipner A platform sank during ballasting trials. An explosion was heard
and 15 minutes later the final 10m ofthie structure disappeared below the water
surface and the structure was lost. Twenty two people were on or around the
structure at the time and fortunately all were evacuated safely. The cost of the
structure was approximately $1 80m, but the economic loss was about $700m.
Problems arose when one of the buoyancy cells imploded under load. In the
design of the structure extensive use was made of finite element analysis, but
for whatever reason, problems with the use of the software or problems
associated with interpretation of the results, a critical design condition was not
identified.
In all three cases the consequences for the Client and the designers were
serious, although potentially in two ofthe cases matters could have been far worse.
At the time all three cases received unwanted publicity and were an unpleasant
reminder to all engineers that computer programs and calculations need to be
checked. Had adequate checks on the software been undertaken, before they were
used for design, errors in the computer code, or problems with the use of it, could
have been detected and the failures avoided.
8.6
8.6.1 Developers
Traditionally, developers of software have tended to be either organisations or
academics who program specific analytical andlor numerical techniques. The
software they produce may be made available to users commercially by selling or
leasing it.
Some software developers expend enormous resources on checking their
programs and ensuring that they are robust. Limitations may even be placed on the
use of the program, to prevent misuse, by introducing default xalues for certain
control features. They may submit their software for independent assessment or
type approval (fitness for purpose), which is undertaken by certain organisations
(e.g. Lloyds Register). This is a check on the 'product' rather than the procedures
used for development. Other developers cany out relatively little testing beyond
fairly basic checks on the code and its operation. They tend to adopt the principle
of 'user beware'. Such effort, or lack of it, is usually reflected in the cost.
Obviously, cheaper programs are likely to be less robust than more expensive
programs, which are likely to have been subjected to more testing. A potential user
should take this into account when considering the purchase of any program.
Although the developer may be responsible for the code, and should carry out
sufficient checks to ensure that the program is working correctly and that the
relevant theory has been correctly applied, the developer cannot be responsible for
the way the program is used, or the interpretation of the user on the way it is used.
Benchmarking 1 337
Further, the developer cannot trace all the bugs in a program, which may consist
of many thousands of lines of code, and some of these bugs may only become
apparent if the program is used in a certain way. Therefore, the user has to be
aware that such problems may exist.
In the past the user was rarely given the opportunity to modify code. If changes
had to be made this was usually done by the developer. However, in recent times
the situation has become more complicated. Some computer programs that are sold
commercially allow the introduction of 'User Defined Subroutines'(UDS) to allow
special features to be added, or new constitutive models to be introduced into the
program. Under these circumstances the traditional user or purchaser assumes the
role of the developer. Not only must the person writing these features ensure that
the subroutine is working correctly, but the interface with the rest of the program
needs to be checked. The subroutine may function perfectly adequately by itself,
but it may cause errors in other parts of the program which were otherwise
functioning correctly. This requires a very extensive checking procedure,
particularly with the more complex codes. The UDS will have to include all the
error control routines, which further complicates the issue. There are obviously still
further complications if there is limited access to the source code. Essentially, if
changes are made the program has to be re-validated (consider the example of the
wind turbine that failed).
8.6.2 Users
Users should expect that developers have carried out sufficient checks to ensure
that the program is working correctly, but this cannot, and should not, be taken for
granted. A number of issues arise: does the user or engineer have a responsibility
to carry out sufficient validation of any computer applications used in design; if so,
what procedures have to be adopted; is the implication of this that if the engineer
does not adopt such procedures, then there has been a breach of professional
responsibility?
Some people are of the opinion that users should be given greater access to
computer source codes. This would allow users to better understand how the
program was operating and might, under certain circumstances, allow them to
customise it. This is obviously desirable when testing the program by
benchmarking or other means because, if problems occur, it might be possible to
identify the source of any errors.
There is, however, a number of drawbacks. In the case of finite element
programs, and other programs based on advanced numerical methods, there are
very few people capable of interpreting the code and even fewer with sufficient
knowledge to successfully customise it. Even after a relatively minor modification
a sufficiently rigorous testing programme needs to be initiated to ensure that no
unforeseen errors have been introduced, or interface problems have developed with
other sections of the code.
Does the fact that the user has access to the source code, and has had the
opportunity to inspect it, shift the onus of responsibility for checking the program
further away from the developer onto the user? Is there even more need for the user
to comprehensively test or benchmark the program? If problems did arise, even in
an unmodified code, would the developer be able to use the fact that the user could
have identified these problems if it had been correctly tested, or benchmarked, to
minimise or even evade any responsibility?
The professional responsibility of an engineer to adequately check computer
programs has long been recognised by the Institution of Civil Engineers. The
following question was for sometime included in the list of essay questions
forming part of their professional examination (Q.24, The place of the engineer in
the community and management topics; questions applicable to Section B of the
Essay for the 1981 April and October Professional Examination):
"Discuss the impact of increasing computer applications on the civil engineer's
design responsibilities and the procedures that must be adopted to ensure the
proper use of such applications."
In the early 1980's there were numerous articles in New Civil Engineer, the
Institution of Civil Engineers' weekly journal, discussing these issues. Much was
made of a comparative study undertaken by the Construction Industry Research
and Information Association (CIRIA). The exercise compared the results obtained
from seven different programs that were based on code of practice for the design
of reinforced concrete beams. No two programs gave the same answers.
In 1990 the UK's National Economic Development Office warned the users of
computer hardware and software that, under current UK legislation, they were very
unlikely to be able to sue suppliers or developers if faults were found in hardware
or software. Commenting on this statement, the Association of Consulting
Engineers reminded engineers that, irrespective of any legal redress, they had a
responsibility to ensure that calculations were accurate, whether or not they were
done using a computer.
Identifying that there was potentially a problem with the use of computer
programs, in 1994 the Association of Geoenvironmental and Geotechnical
specialists (AGS (1994)) published a guide for the validation and use of
geotechnical software. This document was not aimed specifically at users of
advanced numerical methods, but it did draw on the work of NAFEMS (National
Agency for Finite Element Methods and Standards) for setting out strategies for
testing software. When published, the guide stimulated a number of debates as to
how these issues should be dealt with. A number of suggestions were put forward,
including formation of an industry sponsored testing organisation for software.
However, to date this has not materialised.
In recent times many engineers work in an environment where quality
assurance (QA) procedures are applied. Although at times these procedures may
appear to be unnecessarily bureaucratic, they lay down management principles
which are measured against a standard, most commonly I S 0 9000. There is a
requirement to ensure that the 'tools' being used to produce the product (i.e.
design) are of an adequate quality and that checks are made to ensure that this
Benchmarking 1 339
quality is maintained. Obviously, the results of computer analyses could have a
major impact on a design, or a product, and therefore sufficient checks need to be
made on the program and the results obtained to ensure their adequacy.
8.7
There are several techniques that may be used by the developer or user to check
programs. These include the following approaches:
- A check carried out on the code by reading it, to check if there are no
inadvertent errors.
technique or an earlier version of the program, if that version has been tested
and found to be correct.
Checks made by running the same problem with the same computer program
can different hardware, to ensure there are no interface problems between
different software packages and the hardware or operating system.
$.8
Benchmarking
General
8.8.1
There may be a number of reasons for wanting to undertake a benchmarking
exercise. These include:
These benchmarks are actually quite demanding in the way the problem is
formulated. The specification for the problem is quite detailed and very rigid. Not
only are the essential features, such as geometry, loading, boundary conditions and
material properties defined, but so are the element types and the dimensions of
each element.
Benchmarking 1 341
8.8.3
Non-standard benchmarks
For the reasons NAFEMS explained, the geotechnical engineer cannot rely on
those problems produced by the developer to properly check the program.
However, these may be useful when first installing the program, as they allow any
obvious problems with the hardware and operating system to be identified at an
early stage. The user of a geotechnical software package therefore has a serious
problem, since standard geotechnical benchmarks do not exist.
Users therefore need to develop sufficient and exacting benchmarks to
adequately test the software. Problems need to be formulated which have a known
solution that can be compared with the results of the program being tested.
8.9
There were three stages to the project. The first stage involved the solution of
hypothetical problems with a defined geometry (not element type or size as for the
NAFEMS standard benchmarks), loading conditions and material properties. Stage
2 involved the back-analysis of two laboratory tests to predict results for a third
experiment. The third stage involved the analysis of in-situ tests at the Mol
(Belgium) underground facility. In the context of benchmarking only the first stage
of the project is of interest.
The results of the analyses undertaken during Stage 1 are discussed in Section
8.10 of this chapter, in which the Stage 1 INTERCLAY I1 problems are used as
examples of benchmarking problems. However, when assessing the results of the
project as a whole Jefferies and Knowles (1994) commented as follows:
"The project highlights the need to recognise, more formally, the opportunity for
human error and to devise a means of controlling it. The experience fvom this
exercise is that in most cases 'errors' were not discovered until the results were
compared with the experimental results and those of other participants. Errors
were primarily due to misinterpretations of the benchmark specijkation as a result
ofeither incomplete definition of the requiredparameters or misinterpretation of
their meaning".
Despite these comments some participants did discover that there were
fundamental problems with their software, or that it was not ideally suited to the
analysis of some problems. The exercise proved to be a beneficial experience for
all concerned and highlighted the need for caution when using computer programs
of any form.
- Part I
Benchmarking 1 343
(1993)), although specifications for these problems are provided in Appendices
VIII. l to VITI.5. The specifications are less detailed than the NAFEMS standard
benchmarks, because there are no restrictions on the mesh details or the solution
technique. However, they are no less demanding.
Example 1
Example 2
Example 3
Example 4
Example 5
Results
Analytical solutions for ideal (no end
effects) drained and undrained triaxial
tests on modified Cam clay are given in
Chapter 9 of Volume 1. Figure 8.1 shows
the results for the drained triaxial test,
using the modified Newton-Ralphson .F
technique with a substepping stress point 4
algorithm for solving the nonlinear finite
element equations, while Figure 8.2 8
shows similar results for the same test
performed using the tangent stiffness
approach. In each figure results from 3
finite element analyses are given. These
differ in the size of the increment of axial
strain applied, i.e. OS%, 1.0% and 2.0%.
Comment
There is an analytical solution available
which can be used for comparison with
-Analytical solution
"
Axial strain (%)
Benchmarking / 345
Cornrnent
Like Example 1, this is an ideal benchmark because there is an analytical solution
available which can be used to compare with the results.
The results obtained from the INTERCLAY I1 exercise all showed apparently
good agreement between all participants and the analytical solution. All of the
participants were aware of the analytical solution when undertaking their analyses
and could have made adjustments to the increments of loading to obtain
sufficiently accurate results. However, given that all of the participants were very
experienced in the use of their own software, it was reasonable to assume that they
were able to use these programs well enough to obtain accurate results. The same
may not be true if an inexperienced user or a user of some less robust software was
to analyse the problem.
0.01
0.00
0.05
0.10
0.15
0.20
0.25
0.0
0.0
2.0
4.0
6.0
8.0
10.0
t
Not to scale
TOU~
stress analysis
Range of values
;at 2.5m oRsct
10
l5
20
25
The
On this work
Figure 8.7: Radial convergence with
(EUR 15285 (1993)) has a
progression of excavation
series of plots which show
(total stress analysis)
stresses at three specified offsets
from the tunnel at different stages in the analyses. Figure 8.10 summarises some
of these results for the radial stress at the end of excavation.
Figure 8.1 11 shows the distribution of the radial total stress on the tunnel's
periphery (2.5m offset) along the first 12m of the tunnel (from the start of the
tunnel up to section A-A, refer to Figure &.6),when the tunnel has reached point
B. Both this and the previous figure relate to the total stress analysis.
Benchmarking 1 347
0 l0
Cornment
This is not an ideal benchmark.
Unlike the previous two examples,
there is no closed form solution 8 006
available. Unlike the standard
OM
NAFEMS problems, no
restrictions were placed on the
oo2
geometry of the finite element
ow
mesh.
0
5
10
IS
20
25
There was significantly more
Length of excavahon (m)
scatter in the results obtained from
the analysis of this problem by the
INTERCLAY I1 participants. The Figure 8.8: Radial convergence with
progression of excavation
plots of displacement (Figures 8.7
(effective stress analysis)
and Figure 8.8) show more scatter
close to the tunnel than remote
030
from it. For the total stress 3 ozo
approach there is a difference of
almost 20mm between
0.10
participant's results for a point on
,
the edge of the tunnel, 2 . 5 1 ~ g
offset. This is probably a
-O.'O
lbo results hfferen
consequence of an ill-conditioned 2
problem, because of the boundary 2
-0 30
condition applied to the sides of
0
S
10
IS
20
25
Lengtb ofexcavation (m)
the excavation to represent the
tunnel lining (zero radial
Figure 8.9: Displacement of B vs.
displacement), rather than the
length of excavation
accuracy of individual codes. It is
important therefore, when
formulating benchmarks, that care Q
a Mmmum range
is taken to ensure that the
A Maximum range ~gnonng
one analysis
boundary conditions are
reasonable and realistic.
.g l0 However, close to the tunnel
all the participants obtained some
A
4 5oscillations in radial stresses, as 2
illustrated by Figure 8.1 1. There
a
a
was considerable scatter, not only 3 (I
I
I
between participants, but some of
o
2s
5o
75
10
Radial distance fiom tunnel centre lrne (1.e rad~aloffset)
the predictions showed
considerable fluctuation along the
Figure 8.10: Range of radial stress
length of the tunnel for the 2.5m
variation at different offsets
offset (along the boundary of the
-5
1
2
3
.g
. /
,,
Benchmarking 1 349
Results
Again the INTERCLAY 11 report gives detailed results (EUR 15285 (1993)).
Figures 8.14 to 8.18 show vertical displacements along the underside of the base
slab, at each of the five stages in the analyses (see Appendix VIII.4). Figure 8.19
compares d,isplacementsof point B (refer to Figure 8.12), which is on the centre
line at the base of the sand capping layer.
25m
1-
25m
1 1-.
25m
(Not to scale)
material' '
material
B d'
0.3m
Cornment
This is the most demanding of the five examples and, like Example 3, there is no
analytical solution available and therefore there is no direct check on the results.
With this problem some of the deficiencies with some of the codes used by
INTERCLAY I1 participants became apparent. There were also some problems
with interpretation of the specification.
With the exception oftwo participants, similar results for the displacernents of
the underside of the base slab were obtained for the first stage of the analysis
(Figure 8.14). For stages 2,3 and 4 there was some scatter in the results, although
2?:
0.30
0.30
0.20
'2
020
O.I0
O.W
c.
-Bg
3;
4.10
O'rO
0.w
4.10
f 4.20
"20
.-
4.30
25
75
50
>
4.30
25
50
I
0
29
M
HorizontalLstancc from c c n a line (m)
75
75
Benchmarking 1 351
.g ,,,
,ezm,
Results
Figure 8.21 shows the displacement
of a point B at base slab level at
different stages in the analysis, for
case 5a. Similarly Figure 8.22 shows
the vertical total stress beneath the
base slab.
CI
OW
005
.a
3
.,,
1
p
g
015
3
Slage
Comment
at point B
Figure 8.21 shows that, with the
exception of one of the codes, there _.I,
was quite good agreement between $
those who chose to study this
loo
problem. The agreement might have
been even better, but at least one of f 2w
the participants was using fully 3
8 3w
coupled consolidation to model an .;d
undrained event. The time steps used
400
may not have been sufficiently short
,
and some drainage may have occurred
'
2
Stage
3
4
during this stage, resulting in some
settlement. However, the final Figure 8.22: Vertical total stress at
displacement obtained from this
point B
Benchmarking / 353
analysis was very close to that obtained from the other analyses. Again, with the
exception of one participant, the predicted total stresses beneath the base slab were
very close to each other (Figure 8.22).
Careful examination of the results of the analyses of this example (5a) and the
additional analysis (5b) of the same problem, using alternative methodologies,
allows conclusions to be drawn as to which analyses of Example 4 were the most
reliable (the INTERCLAY I1 report provides these comparisons, but the figures
presented in this chapter are not sufficient to do so).
This is another useful example of a benchmark problem, the specification is
very clear and the problem is simple, but it involves the use of a number of
different features of a program.
8.11
8. l 1 . l Background
Schweiger (1998) (see also Carter et al. (2000)) describes a benchmarking exercise
sponsored by a working group of the German Society for Geotechnics. The aim of
the exercise was to develop a series of problems which could serve the following
functions:
The aims of this project were therefore broadly similar to the aims of the
INTERCLAY I1 project. This study aimed to use actual practical problems,
simplified in such a way that the solution could be obtained with reasonable
computation effort.
Two examples are given by Schweiger (1998). For the first example (Example
6) construction of a tunnel was modelled, and for the second example (Example
7) the formation of a deep excavation was modelled. Both are problems that are
commonly analysed by geotechnical engineers.
These problems were analysed by a number of participants and the results
compared. It should be noted that for both these problem no closed form solution
exists and so some reliance has to be placed on a consensus being obtained (this
was also true for the more complex INTERCLAY I1 problems).
;;?
Results
Figure 8.24 shows the predicted surface
settlements for one step tunnel
l
excavation, from 10 groups that
undertook the exercise, while Figure 8.25
-I
summarises the predicted bending
moments and forces in the tunnel lining. Figure 8-23:Geometry for tunnel
{J
excavation
Comment
Although half of the predictions gave
,
maximum surface settlements between
52mm to 55mm, overall there were B "
I
significant differences in the predicted 1
settlements. Two groups used a different
,
method of tunnel construction to that m
specified, because of restrictions of their
codes. It was difficult to match the
S
10
l5
20
different approaches. This probably
D~stancefmm tunnel axis (m)
accounts for the differences in the results. Figure 8-24; Surface settlement
profiles
3-"
5000
4000
3000
C1
2000
l000
0
Participants
108 to 200
(IcNmlm)
Benchmarking / 355
There were also discrepancies in the predicted bending moments and forces in
the tunnel lining and the positions of the maximum bending moment, as shown on
Figure 8.25. Even if the results of the groups that did not follow the specification
are discounted, there are still significant differences in the predicted magnitudes
and positions of maximum bending moment.
!
!
!
Layer 3
Results
Figure 8.27 shows the predicted 8 0.03
vertical surface displacements behind 2
the wall for the first stage of 3 0.02
X
construction. Horizontal displacements 8
0.01
at the top of the wall at the same stage
in construction are shown on Figure 3
0.00
8.28. Figure 8.29 shows the vertical
displacements behind the wall upon g
completion of the excavation g 6.01,
(construction stage 3).
10
IS
20
25
Comment
A wide scatter of the results is
apparent, even at the first stage in
construction when the wall acted as a
cantilever (see Figures 8.27 and 8.28).
Some participants used beam
elements to model the diaphragm
wall, others used solid elements, and
Panicipants
some groups used interface elements
Figure 8.28: Horizontal
between the ground and the wall.
These differences in modelling displacement of the top of the wall
- stage 7
technique might explain some of the
variations in the results, but they do
not account for all of them. As far as
horizontal wall displacements are
concerned at this stage, only half the
participants show positive
displacements of the top of the
cantilever wall, i.e. movements
towards the excavation. Curiously,
the remaining groups showed
movements away from the excavation % -0.03
5I
10
l
15
I
20
I
25l
(negative) at this stage. At the final
Distance from wall (m)
stage in the construction process there
is significant variation in the results, Figure 8.29:Surface settlement
with the maximum surface
behind the waN - stage 3
settlements predicted being between
l Omm and 50mm.
Benchmarking 1 357
that led to some of the differences that were observed. It is however surprising that
experienced analysts would have submitted results with obvious errors, such as a
cantilever wall moving backwards into the ground, without comment.
8.12
Summary
1. This chapter has reviewed the role of benchmarks in the validation and
verification of computer software and discussed the responsibilities of the
developers and users of a code in this respect. Before a computer program is
used, it should be checked by both the developer and user to ensure that it is
accurate, it is working as intended and it is being used correctly. Any checks
made by a user should be independent of the developer, it is not sufficient
simply to use supplied problems.
2. Despite the developer having tested the program before releasing it for use, it
is unlikely that all the potential 'bugs' will have been identified. Even if this
programme of testing is rigorous, it is difficult to anticipate some of the ways
the program may be used and there may be interface problems with hardware
or operating systems. In commercial terms, the degree of testing by the
developer is likely to be reflected in the cost of the program, Users must be
aware that potentially problems may occur and must take sufficient measures
to ensure that these problems do not affect any analysis that may be undertaken.
The program has to be benchmarked.
3. Ideally a program should be checked against a closed form solution, or an
independent means of calculation, to validate it. A series of simple hand
calculations checking stages of an analysis may be sufficient.
4. Benchmark problems of this form tend to be fairly simple and may not be
sufficiently demanding to expose any potential problems. It may therefore be
necessary to devise more demanding problems which can be checked against
the results of the same analysis being run using at least one other program. For
most users this is not a very satisfactory or practical situation, since they may
not have access to more than one code. There is therefore a need to devise and
publish 'standard' benchmarks for geotechnical problems.
5. The consequences of errors or 'bugs' in computer programs can be serious and
can result in significant economic loss.
6 . A series of examples of benchmark problems have been reviewed. The
INTERCLAY I1 exercise has not only provided a number of valuable examples
of benchmark problems that could be used to validate geotechnical software,
but it also demonstrated the need for programs to be thoroughly tested. During
the course of the exercise potential problems with some of the codes were
exposed. The exercise also highlighted the need for detailed and unambiguous
specifications to be produced before any analysis is run.
Appendix VIII.l:
VIII. 1 . l
Geometry
Not specific, single element may be used.
V111.1.2
Material properties and initial stress conditions
Modified Cam clay model for triaxial sample.
Table VIII. 1:Modified Cam clay parameters
200 kPa
200 kPa
0 kPa
V111.1.3
Loading conditions
Unspecified, but constant, increments of compressive vertical strain up to 20%
under undrained and drained conditions. No change in horizontal total stress.
Appendix V111.2:
V111.2.1
Geometry
The geometry is shown in Figure 8.3.
V111.2.2
Material properties
Total stress analysis using a Tresca soil model.
Table V///.2: Tresca parameters
5000 kPa
5000 kPa
Benchmarking 1 359
VR11.2.3
Loading conditions
Using a total stress analysis a pressure was applied to the internal boundary so as
to unload it. The maximum pressure applied was 3500kPa, giving a residual
pressure of 1500kPa. Incremental load steps were not specified.
Appendix V111.3:
V111.3.1
Geometry
The geometry is shown on Figure 8.6 and the analysis was run as an axi-symmetric
problem.
V111.3.2
A)
Material properties
Note:
V111.3.3
2500 kPa
2500 kPa
2500 kPa
Loading conditions
The total stress analysis used stages 1 to 3 only, the effective stress analysis used
all stages:
1. Advance tunnel face instantaneously by 2m, assuming that the internal pressure
on the excavated face is zero.
2. Insert rigid lining to provide radial support.
3. Repeat stages i and 2 until tunnel heading has reached its full length (i.e. 24m).
4. Allow excess pore water pressures to dissipate with the tunnel lining and face
fully permeable and rigid (around the outer boundary pore water pressure
~ ~ 2 MPa,
. 5 around the boundary of the tunnel p,=O and no longitudinal
displacement of the face).
Appendix V111.4
V111.4.1
Geometry
The geometry is shown in Figures 8.12 and 8.13.
V111.4.2
Material properties
The Mohr-Coulomb parameters for the different materials involved in this analysis
are shown in Table VIII.5.
Benchmarking / 361
For the waste material a constant value of Poissons ratio of 0.2 was specified
for stages 1 to 4 in the analysis, but in stage 5 this was set to zero. All the materials
were drained, apart from the clay. Initially pore water pressures below ground
water level (GWL) were hydrostatic. Above GWL pore water pressures were zero.
V111.4.3
Loading conditions
Stages of analysis:
l.* Construct perimeter wall (wished in place), excavate repository, place
drainage layer (zero thickness), construct repository base slab and construct
partition wall.
2. Allow pore water pressures to dissipate to a steady state condition assuming
that pore water pressures in the drainage layer under the base slab are zero
and that the Upper and Lower sands and gravels are fully recharged.
3.* Fill with waste and place capping layers.
4. Allow excess pore water pressures to dissipate to the same conditions as 2.
5.* Reduce Poisson's ratio of waste to zero. Apply a vertical stress of 125MPa to
upper and lower boundaries of waste to model collapse.
(* No drainage)
Appendix V111.5:
V111.5.1
Geometry
The geometry is shown on Figure
8.20.
Stress (kPa)
50
100
150
200
250
I effective stress
V111.5.2
Material properties
The material properties and initial
stresses are the same as for Example 4.
The initial stresses are as shown on
Figure VIII. l .
V111.5.3
Loading conditions
Stages of analysis:
1.* Excavate to 7.8m below existing
4.
V111.5.4
Additional boundary conditions
1. All vertical boundaries are to have zero horizontal displacement.
2. The base of the mesh (the interface between the sandstone and the Lower sand
and gravel layer) is to have zero vertical and horizontal displacement.
3. All vertical boundaries to have no drainage.
4. The base of the concrete slab is to be undrained in Stages l & 3 and drained
in Stages 2 & 4.
5 . The base of the mesh is to be drained.
Appendix Q111.6:
V111.6.1
Geometry
The geometry of the problem is specified in Figure 8.23.
V111.6.2
Material properties
The Properties for the soil and shotcrete are listed in Table VIII.6.
Table Vlll.6: Material parameters for tunnel construction
* Shotcrete thickness is
Appendix Q111.7:
25cm
Specification of Example 7:
Deep excavation
V111.7.1
Geometry
The geometry of the problem is given in Figure 8.26.
V111.7.2
Material properties
The properties of materials used in this example are given in Table VIII.7.
Benchmarking 1 363
VE11.7.3
Construction stages
The following construction stages are specified:
1.
2.
3.
9.1.
Synopsis
This chapter considers some of the common pitfalls that can arise when performing
finite element analysis of geotechnical problems. Some of the restrictions that may
occur when using this method are also discussed. Most of these arise when trying
to model what is clearly a three dimensional component of a problem in a two
dimensional plane strain analysis. For example, modelling ground anchors to tie
back a retaining wall. Problems associated with the use of the Mohr-Coulomb
model for undrained analysis are also discussed. In particular, it is shown that it is
possible to obtain a failure condition only if the angle of dilation is set to zero. As
noted in Chapter 7 of Volume 1 of this book, the shape of the plastic potential in
the deviatoric plane controls the stress conditions at failure in plane strain analyses.
The implications of this are examined by considering boundary value problems
involving modified Cam clay. Some of the pitfalls in using critical state
constitutive models for undrained analyses, as well as problems associated with
modelling underdrainage and the use of zero thickness interface elements are also
discussed.
9.2.
introduction
From the preceding chapters, both in Volume 1 and this volume of the book, it is
clear that the finite element method, when applied to geotechnical problems, can
be extremely complex. While in principal the method can be used to provide a
solution to most of the problems that we may wish to analyse, there are
approximations which can lead to errors. These approximations can be classified
into two groups. Firstly, there are approximations in the finite element method
itself and secondly, there are approximations arising from the idealisations made
by the user when reducing the real problem to a form which can be analysed.
The two main sources of error involved in the finite element method itself are
associated with the integration of the constitutive equation and the discretisation
of the problem geometry into finite elements. The errors associated with the
integration of the constitutive equations have already been considered in Chapter
9 of Volume 1 of this book, where it has been shown that the errors can be
minimised to an acceptable level by the use of an appropriate solution strategy.
However, the errors due to discretisation have not yet been considered and
9.3
Discretisation errors
Errors arising from the discretisation of the problem geometry into finite elements
is demonstrated by considering the behaviour of smooth rigid strip and circular
surface foundations under vertical loading, see Figure 9.1. The soil is assumed to
be linear elastic perfectly plastic, with a Tresca yield surface (see Chapter 7,
Volume l ) , having a Young's modulus E=10000 kPa, Poisson's ratio p=0.45 and
undrained strength S,,=100 kPa.
Clearly it is important that smaller and therefore more elements are placed
where there are rapid changes in stresses and strains and therefore directions of
movement. Some computer programs will automatically re-generate the finite
element mesh to improve the accuracy of the solution. They work by first
performing a solution with a relatively even mesh. Based on the results of this
analysis a new mesh is generated with more elements in the zones of greatest stress
and strain changes. The analysis is then repeated. In principal, this process of
refinement could be repeated until the results were unaffected by further mesh
refinement.
9.4
9.4.1 I n t r o d u c t i o n
As noted in Section 3.6 of Volume 1, there are several different methods available
to model the behaviour at soil-structure interfaces. Of these alternatives, the use of
zero thickness elements is probably the most popular and consequently details of
their finite element implementation were given in Volume 1. Examples of their use
in finite element analyses of boundary value problems have been discussed in
earlier chapters of this volume.
While these elements are extremely useful, they have limitations. For example,
numerical problems, such as ill-conditioning, poor convergence of the solution and
unstable integration point stresses, have been experienced by several authors in the
past (Wilson (1977), Desai et al. (1984), Bay and Potts (1994)). Some of these
pitfalls/restrictions will be discussed in this section.
9.4.2 Basic t h e o r y
Although a detailed description of the theory behind zero thickness interface
elements is given in Section 3.6.2 of Volume 1, it is instructive to restate here the
main equations for 2D analyses.
The interface stress consists of a normal and a shear component. The normal
stress, D, and the shear stress ,t,are related by the constitutive model to the normal
and tangential element 'strains', E and y:
The interface element 'strain' is defined as the relative displacement of the top and
bottom of the interface element:
ul
= vsina+ucosa
vI = vcosa-usina
and u and v are the global displacements in the X, and y, directions respectively.
Hence:
It is important to note that the element 'strains' are not dimensionless, but have the
same dimensions as the displacements (i.e. length).
The elastic constitutive matrix [D] takes the form:
where K, and K , are the elastic shear stiffness and normal stiffness respectively.
to strains (in
Noting that these stiffnessesrelate stresses (in units of for~e/(length)~)
units of length) via Equation (9.1), implies that they must have units of
force/(length)'). They therefore have different units to the Young's modulus, E, of
the soil andfor structure adjacent to them (i.e. E has the same units as stress for~e/(length)~).
As it is difficult to undertake laboratory tests to determine K,, and
K,, selecting appropriate values for an analysis is therefore difficult.
If a pore fluid exists in the interface, undrained behaviour can be modelled by
in
including the effective bulk stiffness of the pore fluid (again in f~rce/(length)~)
the stiffness matrix, in the same manner as described for solid elements in Section
3.4 of Volumel. It is also possible to allow the interface elements to consolidate
by implementing one dimensional consolidation along their length.
The interface stresses can also be limited by imposing a failure criterion. This
is conveniently achieved by using an elasto-plastic constitutive model. For
example, the Mohr-Coulomb failure criterion can be used to define the yield
surface, F:
F = I zl+oltan p' - cf
(9.7)
where y' is the maximum angle of shearing resistance, c' is the cohesion (see
Figure 9.6) and v is the dilation angle. If the interface moves such that the
maximum normal tensile strength is exceeded (c'/tanyf), the interface is allowed
to subsequently open and close and in a finite element analysis the residual tensile
stress is redistributed via the nonlinear solution algorithm. When the interface is
open, the normal stress remains equal to c'ltanq' and the shear stress remains equal
9.4.3 Ill-conditioning
Ill-conditioning of the global stifhess matrix can occur if the element stiffness
matrices of adjacent elements vary in magnitude by a significant amount. This
results in large off-diagonal terms in the global stiffness matrix and, consequently,
loss of accuracy. The analysis of the overturning of an elastic block will be used
to illustrate problems associated with ill-conditioning (Wilson (1997), Day and
Potts (1994)). Some of the results of Day and Potts (1994) will be presented here.
This problem is crudely representative of
a gravity wall. It consists of an elastic block,
Young's modulus E=106 kPa and Poisson's
ratio p=O, separated from a rigid foundation
along the base AB by interface elements, see
Figure 9.7. For clarity, the interface elements
are shown artificially expanded in Figure 9.7.
Ten 6 noded interface elements and fifty 8
noded 2D solid elements were used to model
the problem. A downwards vertical stress of
200kPa and a shear stress of 5OkPa in the
positive x direction were applied to the top
erface
men&
boundary CD. A large number of analyses
were undertaken and in these the interface
elements were assumed to be elastic, with
stiffnesses varying from K, and K,, =106 to K,
and K,, =lOiOkN/m3.Note that in general K,
Figure 9.7: Overturning of
was not equal to K,,.
elastic block
The integration point stresses in the
interface elements for the analysis with
K,=K,,=107 kN/m3 are plotted in Figure 9.8. The linear normal and shear stress
-5
2
a
E
- - _ _ _ A
[K,]=
i i ~ [ B ] ~ [ D ] [ B ] ~ J ~ ~Figure
s ~ T 9. 70: Reduced ill-
-I -I
(9.10)
2
-
analysis
cause of the large stress oscillations
near the sliding front was due to the
use of Gaussian integration (note the
N=WO~-cotes
lntcgranon
above results were obtained using 3
point Gaussian integration). They $,
reported that if the Newton-Cotes
integration scheme was adopted, the
results were satisfactory. The analyses 4
presented above were therefore
-6
repeated with this integration scheme
(again 3 point integration was used)
Figure 9.73: Sheaf stress in
and the results are also shown on
Figure 9.13.
interface elements
The difference between Gaussian
and Newton-Cotes integration is the location of the integration points and their
corresponding weights (Zienkiewicz (l 977)). The integration points for the 3 point
Newton-Cotes method correspond to the positions of the three nodal pairs (i.e. at
each end and the midpoint) of each interface element. The stress in the element, at
each integration point, is then given by the relative displacement between the pair
of nodes at the integration point. The displacement of the other nodes in the
element has no influence. The element is therefore essentially a linkage element
(Frank et al. (1982), Hermann (1978)). Three point Newton-Cotes integration is
equivalent to Simpson's rule for integration. For Gaussian integration the
integration points are located between the end and the midpoint ofthe element, and
at the midpoint of the element. The relative displacement of all nodes affects the
stress at each integration point.
On first inspection of the results given in Figure 9.13 it appears that the
Newton-Cotes integration scheme greatly improves the behaviour of the interface
elements, as the oscillation of the stresses given by the analyses using Gaussian
3,
-
The problem of oscillating stress is due to the use of elements too large to
model adequately the steep stress gradient that occurs in this problem. Stress
oscillation does not occur when small elements are used. The numerical problems
encountered here are not due to poor performance of the interface element, but are
due to inadequate modelling of the problem at hand. Only high-order elements that
allow for a complex distribution of stress across the element will be able to
accurately model the sliding front with the use of larger elements. Newton-Cotes
integration is unnecessary and perhaps undesirable, as it has the effect of 'glossing
over' or 'smoothing out' the steep stress gradient, thus hiding the real solution.
When sufficiently small elements are used to describe adequately the stress
gradient, Newton-Cotes and Gaussian integration give similar results.
9.5
With many geotechnical problems, while the main feature might be adequately
represented by a plane strain idealisation, other, smaller components of the same
problem, might not be. For example, atypical urban excavation problem, as shown
in Figure 9.1 5, is considered. Here an embedded retaining wall is supported by
ground anchors and a foundation slab is tied down by tension piles. This could
represent a cross section of an excavation for a road. The structural members are
now considered in turn below.
9.5.1 Walls
If the wall is of the concrete diaphragm type,
then it is best modelled using solid finite
elements with the appropriate geometry and
material properties. There are no serious
modelling problems here as the wall satisfies
the plane strain assumption. However, if the
wall is made from secant or contiguous
concrete piles, steel sheet piles, or some
combination of steel columns and sheeting,
then the properties and geometry of the wall
will vary in the out of plane direction and
therefore will not satisfy the requirements of
plane strain (see Section 1.6 in Volume 1).
Modelling of these components in a plane
strain analysis will therefore involve some
Figure 9.15: Typical urban
approximations. Usually it is possible to
exca va tion
estimate the average axial (EA) and bending
stiffness ( E 0 per metre length ofwall. If beam elements are used to model the wall,
then these parameters can be input directly as the material properties, see Section
Axial stiffness:
cE,, = E A
(9.1 1)
Bending stiffness:
where E is the Young's modulus of the wall and A and I are cross sectional area
and moment of inertia per metre length, respectively. It may also be necessary to
calculate some form of average strength, but this will depend on the constitutive
model employed to represent the wall. Clearly, the above procedure only treats the
wall in an approximate manner and it will not be possible to accurately estimate the
details of the stress distribution within the separate components of the wall.
9.5,2 Piles
Modelling the piles below the base slab involves additional assumptions as the
piles are not continuous in the out ofplane direction, but are separated by relatively
large expanses of soil. While it is again possible to estimate average axial and
bending stiffnesses per unit length in the out of plane direction and calculate
equivalent parameters, as was shown for the wall above, modelling the piles with
solid or beam elements implies that the soil is not able to freely move between the
piles, as the piles are essentially modelled as a wall in the out of plane direction,
see Figure 9.16. As a consequence, the lateral movements of the soil below the
excavation will be restricted. It may therefore be more realistic to neglect the
bending stiffness of the piles s s that they provide no resistance to lateral soil
movement. A spring or a series of membrane elements (see Chapter 3 in Volume
1) can then be used to represent the piles.
Real structure
Modelled structure
are involved with the modelling of both the piles and ground anchors. This arises
because these structural elements are not continuous in the out of plane direction.
9.5.4
9.5.5
Structural connections
li
Shear = O
Moment = 0
71-
nodes A and B in
rizontal direction
sup,..
Axial thrust * 0
Moment-0
Axial thrust 0
Shear * 0
Tie nodes A and B in
both horizontal and
n i c a l direction
II
Axial thrust * 0
Shear * 0
Moment * 0
Segments of
lining
Modelled with
beam elements
9.6
be necessary to adjust the value of v between the two phases of the analysis.
Alternatively, a more complex constitutive model which better represents soil
behaviour may have to be employed. However, similar problems will occur with
any constitutive model which does not reproduce critical state conditions.
Consequently, great care must be exercised when selecting the model to use.
9.7
As noted in Section 7.9 of Volume 1, the shape of the plastic potential in the
deviatoric plane can affect the Lode's angle 8 at failure in plane strain analyses.
This implies that it will affect the value of the soil strength that can be mobilised.
In many commercial software packages, the user has little control over the shape
of the plastic potential and it is therefore important that its implications are
understood. This phenomenon is investigated here by considering the modified
Cam clay constitutive model.
Many software packages assume that both the yield and plastic potential
surfaces plot as circles in the deviatoric plane. This is defined by specifying a
constant value of the parameter M,. Such an assumption implies that the angle of
shearing resistance, (p', varies with the Lode's angle, 6. By equating M, to the
expression for g(6) given by Equation (7.41) in Volume 1 and re-arranging, gives
the following expression for (p' in terms of M, and 8:
Figure 9.29 shows the variations of (p' with 8, given by Equation (9.13), for
three values of M,. The values of MJhave been determined from Equation (9.14)
using v', = 20, 25" and 30". If the plastic potential is circular in the deviatoric
plane, it can be shown, see Chapter 7 in Volume I,that plane strain failure occurs
when the Lode's angle 6 = 0". Figure 9.29 indicates that for all values of M, there
=25", then
is a large change in (p' with 8. For example, if M , is set to give 'v,
under plane strain conditions the mobilised (p' value is (p', =34.6". This difference
is considerable and much larger than indicated by laboratory testing. The
and p,' becomes greater the larger the value of M,.
difference between 'p,(,
'O-
45-
3 ,h
OCR = l
g(@ = M,
K<,= 1- sin&
K 1 /Z = 0.1
TE
,,
0.45
$,
,
0.35
0.25~~
-20
10
20
30
,,,
04
OCR = 6
v, = 2.848
2.= 0.161
K
= 0.0322
-30
-20
-10
10
20
30
In one analysis the yield and plastic Figure 9.31: Effect of Bon S, for
the constant 4' formulation
potential surfaces were assumed to be
circular in the deviatoric plane. A
value of M,=0.5 187 was used for this analysis, which is equivalent to yT,'=230. in
the second analysis a constant value of y11=23" was used, giving a Mohr-Coulomb
,,,
9.8
The input parameters to most critical state models are based on drained soil
behaviour and do not involve the undrained shear strength, S,, . Consequently,
undrained analyses can be problematic. For example, if constructing an
embankment or foundation on soft clay, short term undrained conditions are likely
to be critical from a stability point of view. It is therefore important for any
analysis to accurately reproduce the undrained strength that is available. It is also
9.9
KK1
Construction problems
- Adjust the stresses after construction such that they are consistent with the new
I \
9.12 Summary
1.
2.
3.
4.
5.
6.
7.
The two main sources of error involved in the finite element method itself are
associated with the integration of the constitutive equations and with the
discretisation of the geometry into finite elements. The magnitude of the
errors involved in the former have been discussed in Chapter 9 of Volumel,
where it was shown how they could be minimised by the use of a robust stress
point algorithm. In this chapter the magnitude of the discretisation error has
been investigated. This can be significant, but can be reduced by a suitable
choice of finite element mesh. In particular, small elements need to be placed
where there are rapid changes in stresses and strains.
The zero thickness interface element is useful for modelling relative slip and
opening and closing on a pre-defined surface. Numerical problems can
however occur through ill-conditioning of the stiffness matrix and high stress
gradients in the interface elements. In many situations stress gradients are
likely to be high and are increased with increased interface stiffness. The
problem can therefore easily be confused with ill-conditioning.
Ill-conditioning was noticed in the problems analysed when the stiffness of
the interface element was greater then approximately 100 times the Young's
modulus of the surrounding soil. This however depends on the units used in
the problem and the size of the surrounding elements.
The normal and shear stiffness of zero thickness interface elements have
dimensions of force/(length)'. They are therefore different to the bulk and
shear stiffness of solid elements which have dimensions of for~eI(1ength)~.
Newton-Cotes integration tends to improve the numerical behaviour of
interface elements at, possibly, the expense of hiding the true solution. When
sufficiently small elements are used to model the interface behaviour,
Newton-Cotes and Gaussian integration give similar results.
The modelling of structural components in plane strain analysis can be
problematic. While the major component of the problem may render itself to
a plane strain idealisation, many of the structural components may not satisfy
such requirements. Approximations must then be made. This usually involves
some averaging of material properties and is likely to impose some
restrictions on the analysis.
Particular attention must be made where structural elements are modelled
using beam or membrane elements in coupled consolidation analyses. It is
8.
9.
10.
1 1.
12.
13.
14.
very easy to make a mistake and model what are supposed to be impermeable
structural components as permeable in such situations.
Care must also be exercised when modelling the conriections between
structural members.
When using the Mohr-Coulomb model for undrained analysis, it must be
recognised that failure is unlikely to be predicted, unless the angle of dilation
is set to zero.
The shape of the plastic potential in the deviatoric plane can have a major
effect on the strengths mobilised in plane strain analysis. Particular care must
be exercised when using critical state models, as many finite element
programs assume that both the yield and plastic potential surfaces are circular
in the deviatoric plane. This can lead to unrealistic angles of shearing
resistance andlor undrained strengths being mobilised. It is more realistic to
use a model which assumes a Mohr-Coulomb hexagon for the shape of the
yield surface in the deviatoric plane, in conjunction with a plastic potential
which produces a circle.
Although the undrained strength, S,,, is not an input parameter to most critical
state models, it can be evaluated from the input data and initial stress
conditions. It is therefore possible to adjust one or some of these values to
give the desired undrained strength profile. One alternative is to keep all
parameters fixed, except the OCR, which is varied with depth to obtain the
desired S,,profile.
Problems can occur when constructing material in finite element analyses.
This arises due to an incompatibility between the stresses calculated during
construction and the constitutive model that is then assigned after
construction. Several alternatives are available for dealing with this problem.
There is a potential problem associated with changing the boundary condition
for a particular degree of freedom from prescribed to not prescribed during
an analysis. It is important the user knows how the software being used will
treat such a change.
If the initial pore water pressure distribution at the beginning of an analysis
is influenced by underdrainage and a coupled finite element analysis is to be
undertaken, then care must be exercised to ensure compatibility between this
profile and the distribution of permeability. Otherwise, significant errors will
be introduced into the analysis. In such cases a simple 1D column analysis is
recommended before undertaking the main analysis.
References
Abbo A.J. & Sloan S.W. (1996), "An automatic load stepping algorithm with error
control", Int. Jnl. Num. Meth. Engng, Vol. 39, pp 1737-1759
Addenbrooke T. I. (1996), "Numerical analysis of tunnelling in stiff clay", PhD
thesis, Imperial College, University of London
Addenbrooke T.l & Potts D.M. (2001), 'Twin tunnel interaction - surface and
subsurface effects", accepted for publication in the Int. Jnl. Geomech
Addenbrooke T.I., Potts D.M. & Puzrin A.M. (1997), "The influence ofpre-failure
soil stiffness on numerical analysis of tunnel construction",
Geotechnique, Vol. 47, No. 3, pp 693-712
A.G.I. (199 l), "The leaning Tower of Pisa: present situation", Proc. 1Oth Eur. Conf.
Soil Mech. & Found. Eng., Florence
A.G.S (1994), "Validation and use of geotechnical software", Association of
Geotechnical Specialists, London
Anderson K.H. (1998), "Skirted anchors - case histories in cost effectiveness",
NR. 199, Norwegian Geotechnical Institute
Apted J.P. (1977), "Effects of weathering on some geotechnical properties of
London Clay", PhD thesis, Imperial College, University of London
Arthur J.R.F. & Menzies B.K. (1972), "Inherent anisotropy in a sand",
Geotechnique, Vol. 22, No. 1, pp 1 15-128
Barton M.E., Cooper M.R. & Palmer S.N. (1988), "Diagenetic alteration and
micro-structural characteristics of sands: neglected factors in the
interpretationof penetration tests", Proc. Conf. Penetration Testing in the
UK, I.C.E., Birmingham
Bernat S. (1996), "Modelisation du creusement d'un tunnel en terrain meuble",
Qualification sur chantier experimental, PhD thesis, Ecole Central de
Lyon
Binnie G.M. (1978), "The collapse of the Dale Dyke dam in retrospect", Q.J.
Engng.Geol.,Vol. l l, pp 305-324
Bishop A. W. (1955), "The use of the slip circle in the stability analysis of slopes",
Geotechnique, Vol. 5, No. l , pp 7-1 7
Bishop A.W. & Bjerrum L. (1960), "The relevance of the triaxial test to the
solution of stability problems", ASCE Research Conf. Shear Strength of
Cohesive Soils, Boulder, Colorado, pp 437-45 1
Bishop A.W & Henkel D.J. (1962) "Test measurement of soil properties in the
References 1 397
References / 399
Davis E.H. (1968), "Theories of plasticity and the failure of soil masses", Soil
Mechanics - Selected Topics, Edt. I.K. Lee, Butterworths, London, pp
34 1-380
Davis E.H. & Booker J.R. (1973), "The effect of increasing strength with depth on
the bearing capacity of clays", Geotechnique, Vol. 23, No. 4, pp 55 1-565
Day R.A. & Potts D.M. (19941, "Zero thickness interface elements - numerical
stability and application", Int. Jnl. Num. Anal. Geomech., Vol. 18, pp
698-708
Day R.A & Potts D.M. (1998), "The effect of interface properties on retaining wall
behaviour", Int. Jnl. Num. Anal. Meth. Geomech., Vol. 22, pp 1021-1 033
Day R.A. & Potts D.M. (2000), Discussion on "Observations on the computation
of the bearing capacity factor N, by finite elements" by Woodward &
Griffiths, Geotechnique, Vol. 50, No. 3, pp301-303
De Moor E.K. (1994), "An analysis of bored pileldiaphragm wall installation
effects", Geotechnique, Vol. 44, No. 2, pp 341-347
Desai C.S., Zaman M.M, Lightner J.G. & Siriwardane H.J. (1984), "Thin-layer
element for interfaces and joints", Int. Jnl. Num. Anal. Meth. Geomech.,
Vol. 8, pp 19-43
Dounias G.T. (1987), "Progressive failure in embankment dams", PhD thesis,
Imperial College, University of London
Dounias G.T., Potts D.M. & Vaughan P.R. (1996), "Analysis of progressive failure
and cracking in old British dams", Geotechnique, Vol. 46, No. 4, pp 62 1640
Dounias G.T., Potts D.M. & Vaughan, P.R. (1989), "Numerical stress analysis of
progressive failure and cracking in embankment dams", Report to the
Department of the Environment, Contract No.PECD 7171222, Building
Research Establishment
Duncan J.M. (1 992), "Static stability and deformation analysis", Proc. Spec. Conf.
Stability and Performance of Slopes and Embankments - 11, Geot. Eng.
Div., ASCE, Vol. l , (Edt. Seed and Boulanger), pp 222-266
Duncan J.M. (1994), "The role of advanced constitutive relations in practical
applications", Proc. 131h Int. Conf. Soil Mechs. Found. Eng., New Delhi,
Vol. 5, pp 3 1-48
Duncan J.M., Byrne P.M., Wang K.S. & Mabry, P. (1980), "Strength, stress-strain
and bulk modulus parameters for finite element analyses of stresses and
movements in soil masses", Geotechnical Engineering Research Report
No. UCBlGTI8O-0 1, {Jniversity of California, Berkeley
Duncan J.M. & Seed R.B. (1986), "Compaction induced earth pressures under K,
conditions", Jnl. Geotech. Eng., ASCE, Vol. 112, No. 1, pp 1-21
EUR 15285 (1 993), "INTERCLAY I1 project: A co-ordinated benchmark exercise
on the rheology of clays"", The European Commission, Brussels
Fernie R., Kingston P., St John H.D., Higgins K.G. & Potts D.M. (1996), "Case
history of a deep 'stepped box' excavation in soft ground at the seafront Langney Point, Eastl>ourne", Geotechnical Aspects of Underground
References 1 401
Hight D.W. (1993), "A review of sampling effects in clays and sands", Offshore
site investigation and foundation behaviour, Society for Underwater
Technology, Vol. 28, pp 115-146
Hight D.W. & Higgins K.G. (1994), "An approach to the prediction of ground
movements in engineering practice: background and application", Prefailure Deformation of Geomaterials, Edt. Shibuya, Mitachi & Miura,
Balkema, Rotterdam, pp 909-945
Hight D. W., Higgins K. G., Jardine R. J., Potts D. M., Pickles A. R., DeMoor E.
K., & Nyirende, Z. M. (1992), "Predicted and measured tunnel distortions
associated with construction of Waterloo International Terminal",
Predictive Soil Mechanics, Proc. Wroth Memorial Symp., pp 3 17-338
Hight D.W., Jardine R.J. & Gens A. (1987), "The behaviour of soft clays",
Embankments on soft clays, Bulletin of the Public Works Research
Centre, Athens, pp 33-158
Higgins K.G. (1983), "Diaphragm walling, associated ground movements and
analysis of construction", MSc thesis, Imperial College, London
University
Higgins K.G., Fernie R., Potts D.M. & Houston C. (1999), " The use of numerical
methods for the design of base propped retaining walls in a stiff fissured
clay", Geotechnical Aspects of Underground Construction in Soft
Ground, Edt. Kusakabe et al, Balkema, Rotterdam, pp 5 1 1-5 16
Higgins K. G., Potts D. M., and Mai, R. J. (1996), "Numerical modelling of the
influence of the Westminster Station excavation and tunnelling on Big
Ben Clock Tower", Geotechnical Aspects of Underground Construction
in Soft Clay, Edt. R.J. Mair & R. N. Taylor, Balkema, Rotterdam, pp
525-530
Higgins K.G., Potts D.M. & Symonds I.F. (1993), "The use of laboratory derived
soil parameters for the prediction of retaining wall behaviour", Retaining
Structures, Edt. C.R.I. Clayton, Thomas Telford, London, pp 92- 10 1
Hill R. ( 1 950), "The mathematical theory of plasticity", Clarendon Press, Oxford
Hird C.C., Pyrah I.C. & Russell D. (1990), "Finite element analysis of the collapse
of reinforced embankments on soft ground7',Geotechnique, Vol. 40, No.
4, pp 633-640
Horsnell M.R., Norris V.A. & Ims B. (1992), "Mudmat interaction and foundation
analysis", Proc int. Conf. Recent Large Scale Fully Instrumented Pile
tests in Clay, Thomas Telford, London
Huder J. (1 972), "Stability of bentonite slurry trenches with some experience of
Swiss practice", Proc 51h European Conf. Soil Mech. & Found. Eng,
Madrid, pp 5 17-522
Jackson C., Zdravkovic L. & Potts D.M. (1997), "Bearing capacity of pre-loaded
surface foundations on clay", Computer Methods & Advances in
Geomechanics, Edt. J.X. Yuan, Balkema, Rotterdam, Vol. 1, pp 745-750
Jamiolkowski M, Lancellotta K., & Lo Presti D.C.F. (1994), "Remarks on the
stiffness at small strains of six Italian clays", Pre-Failure Deformation of
References / 403
Minister0 dei Lavori Pubblici (1971), "Ricerche e studi sulla Torre di Pisa ed i
fenomeni connessi alle condizione di ambiente, Vol. 3, I.G.M., Florence
Mori R.T. & Pinto N.L.deS. (1988), "Analysis of deformations in concrete face
rockfill dams to improve face movement prediction", Proc. 161hInt. Cong.
Large Dams (ICOLD), San Francisco, Vol. 2, pp 27-34
NAFEMS (1990), "The standard NAFEMS benchmarks", Publication P1 8,
NAFEMS
Nash D.F.T., Sills G.C. & Davison L.R. (1992), "One -dimensional consolidation
testing of soft clay from Bothkennar", Geotechnique, Vol. 42, No. 2, pp
24 1-256
Naylor D.J. (1975). "Numerical models gbr clay core dams, Proc. Int. Symp.
Criteria and Assumptions for Num. Analysis of Dams, Swansea, Edt.
Naylor, Stagg & Zienkiewicz, pp 489-5 14
Naylor D.J. (1991a), "Stress-strain laws and parameter values", Advances in
Rockfill Structures, Edt. Maranha das Neves, pp 269-290
Naylor D.J. (1991 b), "Finite element methods for fills and embankment dams",
Advances in Rockfill Structures, Edt. Maranha das Neves, pp 291-340
Naylor D.J., Tong S.L. & Shahkarami A.A. (1989), "Numerical modelling of
saturation shrinkage", Numerical Models in Geomechanics NUMOG 111,
Edt. Pietruszczak and Pande, pp 636-648
Naylor D.J. (1999), "On the use of the FEM for assessing the stability of cuts and
fills", Numerical Models in Geomechanics, NUMOG VII, Edt.
Pietruszczak & Pande, Balkema, Rotterdam, pp 553-560
NCE (1 983), New Civil Engineer, October Issue, Emap, London
NCE (1984a), New Civil Engineer, January Issue, Emap, London
NCE (1984b), New Civil Engineer, February Issue, Emap, London
NCE (1991), New Civil Engineer, March Issue, Emap, London
NCE (1994), New Civil Engineer, December Issue, Emap, London
Ng C.W.W. (1992), "An evaluation of soil-structure interaction associated with a
multi-propped excavation", PhD thesis, University of Bristol
Nyren R. (1998), "Field measurements above twin tunnels in London Clay", PhD
thesis, Imperial College, University of London
Ong J.C.W. (1997), "Compensation grouting for bored tunnelling in soft clay",
MSc dissertation, Imperial College, University of London
Ong H. L. (1 996), "Numerical analysis for tunnelling and grouting in soft ground",
MSc dissertation, Imperial College, University of London
O'Neil M.W., Ghazzaly 0.1. & Ha H.B. (I977), "Analysis of three dimensional
pile groups with non-linear soil response and pile-soil-pile interaction",
Offshore Technology Conf., Paper 2838, pp 245-256
Padfield C.J. & Mair R.J. (1984), " Design of retaining walls embedded in stiff
clay", CIRIA Report 104
Pagano L. (1998), "Interpretation of mechanical behaviour of earth dams by
numerical analysis", Proc. Symp. Prediction and Performance in
Geotechnical Engineering, Hevelius Edizioni, Napoli, pp 89- 150
References / 405
References 1407
List of symbols
This list contains definitions of symbols and an indication of the section in the
book where they first appear. Because of the large number of parameters that are
used, some symbols represent more than one quantity. To minimise any confusion
this may cause, all symbols are defined in the text, when they are first used. If
some symbols are missing from the list below, they are in the List of symbols in
Volume 1.
C'
c,,'
c,.'
C,
List of symbols 1 41 1
total overburden pressure
pore fluid pressure
pore fluid pressure on boundary
in-situ mean effective stress
deviator stress in triaxial plane
cone resistance
ultimate load
pore pressure ratio
(o;,'+aaf)/2
S
S<' S,,'
Sy
(a',' - ou')/2
element thickness
horizontal displacement
vertical displacement
specific volume
specific volume at unit mean effective stress
area
pore water pressure parameter
diameter of plate
building width
half width of the footing
compression index
swelling index
tunnel diameter
final tunnel diameter
initial tunnel diameter
relative density
hogging deflection ratio
sagging deflection ratio
Young's modulus
effective Young's modulus
undrained Young's modulus
constrained modulus
hammer energy in SPT
free-fall energy in SPT
secant Young's modulus
tangent Young's modulus
small strain Young's modulus
vertical effective Young's modulus
horizontal effective Young's modulus
bending stiffness
axial stiffness
rod energy ratio in SPT
axial force
v
v
VI
A
A
B
B
B
,\
cc
C,
D
D,
D",
D,DRhos
DRsag
E
E'
El,
E,
Er
Ex
E,
E,
E',,,
E,,'
Eh'
EI
EA
ER,
F',
(1.4.5)
(1.3.3)
(4.3.2)
(1.3.3)
(1.3.3)
(1.4.3)
(1.4.2)
(4.3.2)
(1.3.3)
(6.5)
(1.3.3)
(9.4.3)
(1.3.5)
(1.3.5)
(1 3 . 2 )
(5.5.3)
(3.4.5)
(1.3.3)
(1.4.5)
(2.4.1)
(6.5.4)
(1.3.2)
(1.3.2)
(2.4.1)
(2.4.4)
(2.4.4)
(1.4.2)
(2.7.1)
(2.7.1)
(1.3.2)
(1.3.3)
(1.3.3)
(1.3.2)
(1.4.2)
( 1.4.2)
(1.3.3)
(1.3.3)
(1.3.3)
(2.6.3)
(2.6.3)
(2.7.1)
(2.7.1)
(1.4.2)
(1.3.3)
K,,')('
K11
K1
K,
Kt,
M
M,
M,
M/.
M,, My
N
160
Nk
N,, N,,, N,
OCR
P
P,.
Q
Qmax
R
R
Smax
S11
S
ur,uy,U:
List of symbols / 41 3
YXY
Yr2
Y,o
Y20
Y
6
6
6
, ,b
6h
E,,
4.
E11
&,l
9
&hc
&h!
E/,
0
K
R
R
2,
P
P
Pc,rl
V,,'
P
P
P'
02, G3
4"' 0:
01,
or,
0,
G',
(~,,'>b
0,
0,
G,.'
0,
~ 1 1
a;,
G/,<,
o;,,,
o/
5
51,
TV
Tmob
5,.
4
Tr:
6.0
v'
VC,,?
v,,'
E',
Q,
Qy
Index
A
Index 1 417
embedded walls, 113-1 14,114
pin-joint, 105, 106
plain strain analysis, 380-381, 381
support systems, 83, 84,84
tunnel linings, 49-52
constant head permeameter, 22
constitutive model
errors, 365
users' misunderstanding, 365
Construction Industry Research and
Information Association (CIRIA),
338
construction problems, 387-388
construction under water, cut slopes, 162163
convergence-confinementmethod, tunnel
excavation, 46
Coop, M.R., 20
Cording, E.J., 120
counterweight, Leaning Tower of Pisa,
267-27 1
coupled consolidation analyses, 380
Coyle, H.M., 290
CPT see cone penetration test
Crabb, G.I., 156
critical line, tower instability, 272-273, 274
critical state models, undrained analysis,
386387
cut and cover construction
see also earth retaining structures
George Green Tunnel, 91-92,92
cut slopes, 125-163
see also London Clay cut slopes
clay types, 126
construction under water, 162-163
drainage during excavation, 125-126
'non-softening' analyses, 126-140
progressive failure, 141-145
soft clay, 131-140
'softening' analyses, 145-1 63
stability, 125
stiff 'non-softening' clay, 127-131
cuttings
backfill, 96
collapse, 145-146
effect of widening, 158-160, 159, 160
slip, 162
cylinder see thick cylinder
D
Dale Dyke dam, 181-183
collapse, 181
E
earth embankments, 185-1 94
Carsington embankment failure, 189-194
earthfill modelling, 186
London Clay road embankments, 186189
earth pressure at rest see coefficient of earth
pressure at rest
earth pressure balance (EPB) tunnelling, 40,
45
earth retaining structures, 74-121
construction and design, 74-75, 91-92
embedded walls, 76-77, 103-12 1
gravity walls, 75-76, 93-96
bucket foundations, 3 19
Magnus platform, 3 10,3 12,315
pile groups, 290,291,297,298
single piles, 193, 282, 287, 291, 292
Frank, R., 374
Freiseder, M., 105
FSAFEM see Fourier Series Aided finite
element analysis
G
Ganendra, D., 290,29 1,292,294
gap method, tunnel excavation, 45-46
Gens, A., 374
geometry
axi-symmetric, 43, 77
earth-retaining structures, 79-82
tunnel, 91-92, 92
geophysical techniques, 20-21
George Green tunnel, 91-92,92, 109
Geotechnical Consulting Group, 57
geotextile
earth wall reinforcement, 101-102
facing membrane, 97
reinforcement, 200
Germaine, J.T., 20
German Society for Geotechnics
deep excavation exercise, 355-357
tunnel construction exercise, 353-355
Gouwenic, S., 105
granular soils, standard penetration test, 23
gravity walls, 75-76, 75,93-96
cuttings, 96
earth pressure (compaction), 94-95
finite element analysis, 95-96
interface elements, 95
limit states, 94
greenfieid conditions, earth retaining
structures, 88-89
greenfieid settlement, 65
Griffiths, D.V., 126
ground anchors, 83,115-1 16
plain strain analysis, 376, 378-3150
ground loss, tunnelling, 41-43
ground water control, earth retaining
structures, 93
Gun, M.J., 105
H
Hamza, M.M.A.F., 171
Hancock, R.J.R., 57
Head, K.H., 5
Henkel, D.J., 6, 53, 104, 149
I
ICFEP see Imperial College Finite Element
Program
ill-conditioning, interface elements, 370373
Imperial College Finite Element Program
(ICFEP), 282
impounding, rockfill dams, 178-179, 182183
in-situ tests, 23-35
limitations, 23
instability, reinforced earth walls, 98, 101
Institution of Civil Engineers, professional
responsibility, 338
interaction analysis (tunnels), 63-72
Treasury building case study, 66-70
twin tunnels, 70-72
INTERCLAY I1 project, 340,341-342
examples of problems, 342-35 1
objectives, 341
interface elements
bucket foundation, 320-32 1,320
reinforced earth walls, 96-97, 100-102
ring shear test, 16
single piles, 283-285
zero thickness
basic theory, 368-370
ill-conditioning, 370-373
numerical stability, 368-76
steep stress gradients, 373-376
IS0 9000 standard, 338
isotropy see soil isotropy
K
K-G model, rockfill dams, 171
Kalra, J.C., 85, 86
Kenley, R.M., 3 16,3 17
Kim, M.K., 184
Kimmance, J.P., 71
Knights, M.C., 91
Knowles, N.C., 34 I, 356
'Kork' packing, 112
KovaceviC, N., 149, 171, 173, 174, 180
Kuwano, R., 20
Kuzuno, T., 40
L
La Rochelle, P,, 206
laboratory tests, 2-22
scope, 2
Lade, P.V., 173, 184
Lade's model, 85, 173, 175, 176, 179, 184
Lambe, T.W., 168
Lambeth Group Clay and Sands, 57,389
Lancellotta, R., 261
Lane, P.V., 126
Langley Point, Eastbourne, deep
excavation, 119
lateral loading, single piles, 287-289
Lau, C.K., 228,232,233
layered analysis, rockfill dams, 173-1 74
leaning instability, Pisa tower, 256-257
Leaning Tower of Pisa, 252-278
background, 252-253
construction and ground profile, 253254,254
counterweight application, 267-271
critical line, 272-273, 274
finite element analysis, 263-265
history of construction, 254-255
history of tilting, 255-256
model assessment, 277-278
motion of foundations, 256
Index 1 421
finite element analyses, 188
material properties, 187-188
London Underground
Bakerloo Line tunnels, 120, 121
Jubilee line tunnels, 59,61,66,67
tunnel distortion, 120
tunnels beneath parks, 59-60,61, 62,68
Westminster station construction, 89-91,
89, 120
M
Magnus platform, 304-3 17
analyses, 308-309
design of foundations, 309-3 14
environmental loading, 3 14-3 17
pile group displacements, 315
small strain stiffness parameters, 306
soil profile, 305
soil properties and initial conditions,
304-307
turning moments, 311,312,316
Mair, R.J., 30, 74, 195,235
man-made slopes see cut slopes
Marsland, A., 30
Masoswe, J., 10
MCC model, 209,210
mean effective stress, 5
membrane elements
modelling, 378,379
reinforced earth walls, 96-97, 100-102
rockfill dams, 167,180
Menard pressuremeter, 30
Menzies, B.K., 20
Mindlin beam elements, 100
Minister0 dei Lavori Pubblici, 260
MIT-E3 model, 246
anisotropic soil behaviour, 243,244
bucket foundation, 3 18, 326
Champlain clay parameters, 209
embankments, 197,207-208,2 10
material properties, 245
modelling
see also Cam clay model; Lade's model;
MIT-E3 model; Mohr-Coulomb
model
critical state models, 386-387
elastic models, 171,474,478
K-G model, 171
MCC model, 209,2 10
plain strain analysis, 376-381
soil models, 59-63,85-88,215-216
time dependent behaviour, 52-59
underdrainage, 389-394
Mohr-Coulomb model
elastic, 85
elasto-plastic, 86, 146-1 47
footings, 226
lack of understanding, 365
linear elastic, 199
piles, 285,286
plastic, 57,60
shallow waste disposal parameters, 360
soil stiffness, 80
undrained analysis, 365, 382-384
Mori, R.T., 168
motorway cuttings
effect of widening, 158-160, 159, 160
slip, 162
N
NAFEMS see National Agency for Finite
Element Methods and Standards
Nash, D.F.T., 261
National Agency for Finite Element
Methods and Standards
(NAFEMS), 338,340
standard benchmarks, 340-341
National Economic Development Office,
338
Naylor, D.J., 126, 168, 170, 171, 173, 174
NCE see New Civil Engineer
New Civil Engineer (NcE), 334,335,336,
338
Newton-Cotes integration scheme, 374375,376
Newton-Raphson technique, 214,282,343
Ng, C.W.W., 105
NGI see Norwegian Geotechnical Institute
non-homogeneous clay, undrained bearing
capacity, 233-238
'non-softening' analyses, clay, 127-13 1
Norwegian Geotechnical Institute (NGI),
simple shear apparatus, 14, 15
Nyren, R., 60
0
OCR see overconsolidation ratio
oedometer test, 3-6
Bangkok Clay prediction, 133, 134
clay soils, 3-4
rockfill dams, 167-168, 171,176
sands, 5-6
strain, 3
stress, 3
P
Padfield, C.J., 74
Pagano, L., 186
Pancone clay, Pisa, 254,261,265,266,267,
268
peak strength, London Clay, 148
Peck, R.B., 24,34,75, 147
Penman, A.D.M., 169
Pepe, C., 261
permeability, 22
Bangkok Clay, 391
embedded walls, 11 1-1 12
London Clay, 149,389-390
modelling, 53, 55-57
profiles, 392-394
pumping tests, 35
triaxial test, 11
tunnel linings study, 57-59
void ratio, 56
permeameters, 22
Perry, J., 156
piggy-back tunnels, 71-72, 71, 72
pile groups, 280-281,281,289-3 17
analysis, 291
bending moments, 3 12,3 16-3 17
design, 298-304
displacements with depth, 293
load distribution, 294-298,295,296
load vector evaluation, 299-302,300,
301-302
Magnus platform, 304-3 17
matrix formulation, 298-299
rigid pile cap criterion, 294-295,297298
solution displacements and rotations,
302-304
superposition, 291-294,293,299-301
symmetry, 290
three-dimensional analysis, 289-291
turning moments, 3 1 1-3 13
piles, 280-281,281
see also bucket foundations; embedded
walls; pile groups
caissons, 28 1
plain strain analysis, 377-378
single piles, 282-289
interface elements, 283-285
lateral loading, 287-289
load displacement curves, 284-285,
284-287,288,289
vertical loading, 282-286
Pinto, 168
Pisa, Tower of see Leaning Tower of Pisa
pitching, tunnels, 42
plain strain analysis, 376-381
coupled analyses, 380
embedded walls, 106,376-377
ground anchors, 367,378-380
piles, 377-378
segmented tunnel linings, 381,382
structural connections, 380-38 1,381
plain strain geometry, 43
plastic expansion, pressuremeter testing,
3 1-32
plastic model, Mohr-Coulomb, 57,60
plastic potential surfaces, influence in
deviatoric plane, 384-386
plate loading test, 32-34,33
clay, 34
sand, 34
platforms see offshore platforms
Poisson's ratio, hollow cylinder test, 18-19
Polvani Commission (Pisa), 252,260, 261,
267,27 1,276
pore water pressures, 55-56,55,57
Bangkok Clay, 390491,391
clay slopes, 187
Dale Dyke dam, 181,182
London Clay, 151, 152, 153,389-390,
390
Ramsden dam, 183-185
tunnelling study, 57
under-drained, 53-54
~ o r o v iE.,
t 20,245
post shield loss, tunnelling, 42
Potts, D.M., 12, 14, 52, 63,65,70,72, 91,
1 1 1, 115, 146, 171, 172, 189, 223,
255,256,261,267,290,291,292,
293,294,304,3 10
Index / 423
Poulos, H.G., 289, 290
Prandtl, L., 226,236
Prandtl mechanism, failure, 236
pre-loaded strip foundations, undrained
bearing capacity, 238-243
pressuremeter testing, 30-32
coefficient of earth pressure at rest, 32
elastic expansion, 30-3 1
plastic expansion, 31-32
problems in analysis, 364-394
construction problems, 387-388
degrees of freedom removal, 388-389
dicretisation errors, 365-368
modelling underdrainage, 389-394
plain strain analysis, 376-381
undrained analysis, 382-384, 386-387
yield and plastic potential surfaces, 384386
zero thickness interface elements, 3683 76
progessive softening method, tunnel
excavation modelling, 46, 47
progressive failure
compressible beam, 143-145, 144
cut slopes, 141-145
rigid beam, 141-143, 142
props, 83,84,85
see also support systems
puddle clay core dams, 180-185
Dale Dyke dam, 181-183
hydraulic fracture, 180, 182
pumping tests, 35
Puzrin, A., 70
97,100-1 02
reinforcement materials, 101-1 02
soil nails, 98, 98, 103
stability, 100-101
reinforcement
interface properties. 201
soft clay embankments, 200-202
relieving slabs, 116-1 17
residual strength, London Clay, 148, 152
Ridden, J.W., 304, 308
rigid beam, progressive failure, 141-143,
142
rigid foundations, 21 8-2 19
rigid pile cap criterion, 294-295,297-298
ring shear test, 15-1 6
interface testing, 16
soil strength, 15
road tunnel
construction method, 91-92, 92
symmetry, 78, 78
Roadford dam, 175-1 80,175
comparison with observations, 179-180
construction, 178-179
double hardening model parameters, 178
finite element analyses, 177-1 79
material parameters, 175-1 76, 177
Robertson, P.K., 28,29
rockfill dams, 167-185
dilatancy, 172
elasto-plastic models, 171-1 73
hyperbolic model, 170-1 71
impervious earth core, 167
K-G model, 171
layered analysis, 173-1 74
linear elastic analysis, 169-1 70
non-linear elastic model, 171
old puddle clay core dams, 180-1 85
'power law' models, 169-1 70
Roadford dam analysis, 175-1 80, 175
stress paths, 167-168, 167, 168
upstream impervious membrane, 167
Roscoe, K.H., 259
Rowe, R.K., 45,46, 169
rupture surfaces, London Clay, 154-1 55,
1.54
S
St. Alban, Quebec, 206-207
St. John, H.D., 105, 112, 116
Sandroni, S.S., 148, 150
sands
cone penetration test, 28-29
P
Padfield, C.J., 74
Pagano, L., 186
Pancone clay, Pisa, 254, 261, 265,266,267,
268
peak strength, London Clay, 148
Peck, R.B., 24,34,75, 147
Penman, A.D.M., 169
Pepe, C., 261
permeability, 22
Bangkok Clay, 391
embedded walls, 1 1 1-1 12
London Clay, 149,389-390
modelling, 53, 55-57
profiles, 392-394
pumping tests, 35
triaxial test, 11
tunnel linings study, 57-59
void ratio, 56
permeameters, 22
Perry, J., 156
piggy-back tunnels, 71-72, 71, 72
pile groups, 280-28 1,281,289-3 17
analysis, 291
bending moments, 3 12,3 16-3 17
design, 298-304
displacements with depth, 293
load distribution, 294-298,295,296
load vector evaluation, 299-302,300,
301-302
Magnus platform, 304-3 17
matrix formulation, 298-299
rigid pile cap criterion, 294-295, 297298
solution displacements and rotations,
302-304
superposition, 291-294,293,299-301
symmetry, 290
three-dimensional analysis, 289-291
turning moments, 3 11-3 13
piles, 280-281,281
see also bucket foundations; embedded
walls; pile groups
caissons, 28 1
plain strain analysis, 377-378
single piles, 282-289
interface elements, 283-285
lateral loading, 287-289
load displacement curves, 284-285,
284-287,288,289
vertical loading, 282-286
Pinto, 168
Pisa, Tower of see Leaning Tower of Pisa
pitching, tunnels, 42
plain strain analysis, 376-381
coupled analyses, 380
embedded walls, 106,376-377
ground anchors, 367,378-380
piles, 377-378
segmented tunnel linings, 381,382
structural connections, 380-381,381
plain strain geometry, 43
plastic expansion, pressuremeter testing,
31-32
plastic model, Mohr-Coulomb, 57, 60
plastic potential surfaces, influence in
deviatoric plane, 384-386
plate loading test, 32-34,33
clay, 34
sand, 34
platforms see offshore platforms
Poisson's ratio, hollow cylinder test, 18-19
Polvani Commission (Pisa), 252,260,261,
267,27 1,276
pore water pressures, 55-56,55,57
Bangkok Clay, 390-391,391
clay slopes, 187
Dale Dyke dam, 181,182
London Clay, 151, 152, 153,389-390,
390
Ramsden dam, 183-1 85
tunnelling study, 57
under-drained, 53-54
~ o r o v iE.,
t 20,245
post shield loss, tunnelling, 42
Potts, D.M., 12, 14, 52,63,65,70,72, 91,
11 1, 115, 146, 171, 172, 189, 223,
255,256, 261, 267,290,291,292,
293,294,304,3 10
Index 1423
Q
quality assurance procedures, 338
Quarterman, R.S., 30
R
railway cutting slopes, collapse, 145-146
Ramiah, B.K., 268
Ramsden dam, 183-1 85
drawdown displacements, 185
Randolph, M.F., 252,289
Reese, L.C., 290
reinforced earth walls, 76, 96-103
anchors, 97-98,98,103
construction, 96-99,96,97, 101
failure, 98-99, 99, 102
finite element analysis, 99-103
materials stability, 100-1 01
membrane and interface elements, 96-
97,100-102
reinforcement materials, 101-1 02
soil nails, 98, 98, 103
stability, 100-101
reinforcement
interface properties. 201
soft clay embankments, 200-202
relieving slabs, 1 16-1 17
residual strength, London Clay, 148, 152
Ridden, J.W., 304, 308
rigid beam, progressive failure, 141-143,
142
rigid foundations, 2 18-2 19
rigid pile cap criterion, 294-295, 297-298
ring shear test, 15-1 6
interface testing, 16
soil strength, 15
road tunnel
construction method, 9 1-92, 92
symmetry, 78, 78
Roadford dam, 175-1 80, 175
comparison with observations, 179-180
construction, 178-1 79
double hardening model parameters, 178
finite element analyses, 177-179
material parameters, 175-1 76, 177
Robertson, P.K., 28, 29
rockfill dams, 167-185
dilatancy, 172
elasto-plastic models, 171-173
hyperbolic model, 170- 1 7 1
impervious earth core, 167
K-G model, 171
layered analysis, 173-1 74
linear elastic analysis, 169-1 70
non-linear elastic model, 171
old puddle clay core dams, 180-1 85
'power law' models, 169-1 70
Roadford dam analysis, 175-180, 175
stress paths, 167-168, 167, 168
upstream impervious membrane, 167
Roscoe, K.H., 259
Rowe, R.K., 45,46, 169
rupture surfaces, London Clay, 154-155,
154
S
St. Alban, Quebec, 206-207
St. John,H.D., 105, 112, 116
Sandroni, S.S., 148, 150
sands
cone penetration test, 28-29
&
-
"r
r u * & a ; r a u i 4 -
".--&wAuA
Index 1 425
T
tall towers, stability, 256-259
Tavenas, F.A., 132,207
TBM see tunnel boring machines
Tedd, P., 104, 180, 184
Terrace Gravels, 66
Terzhagi, K., 24,34, 147
Thames Gravel, 57,59,389
thick cylinder, analysis, 344345,358-359
time dependent behaviour modelling, 52-59
hydraulic boundary conditions, 54-55
initial conditions, 52-54
permeability, 55-57
tunnel linings, 57-59
under-drained pore pressures, 53-54
time related movements, earth retaining
structures, 92-93
towers, stability, 256-259
Treasury building
settlement trough, 68
tunnel-induced response, 66-70
triaxial test, 6-1 1 , 6
angle of shearing resistance, 8
Bangkok Clay, 133
compressions and swelling, 7-8
idealised analyses, 343-344,358
London Clay, 150
permeability, 1 1
pore pressure, 7
rockfill dams, 168, 171,176
small strain stiffness, 60
soil elasticity, 9
soil stiffness, 8-9, 9
stress-strain diagrams, 8
true triaxial test, 11-1 2
angle of shearing resistance, 11
soil stiffness, 11-12
tunnel, 38-72
see also London Underground; road
tunnel
building stiffness and settlement, 57,6366,64
construction method, 91
modified soil stresses, 89-90, 90
numerical methods, 38-39
parametric study, 57-59
pitching, 42
time dependent behaviour modelling,
52-59
twin tunnel interaction, 70-72
tunnel boring machines (TBM), 40,41
tunnel construction, 39-43
benchmarking exercise, 346-348, 353355,359-360,362
ground response, 4 1-43
interaction analyses, 63-72
manual methods, 39
material parameters, 362
mechanical methods, 40-43
road tunnels, 91-92, 92
shielding, 40
sprayed concrete lining method, 41
support, 3 W O
tunnel construction simulation, 43-52
advancing tunnel heading, 346-348,
359-360
benchmarking exercise, 346-348
choice of soil model, 59-63
FSAFEM, 43,44
important boundary conditions, 45
initial conditions, 44-45
two dimensional modelling, 43
tunnel excavation modelling, 44, 45-48
convergence-confinement method, 46
gap method, 45-46
progessive softening method, 46, 47
soft clay problems, 48
volume loss control method, 46-48,57
tunnel lining modelling, 48-52
bending moments, 354,355
Index 1427
joints, 49-52, 51
moment-free joints, 49,50
permeability study, 57-59
segment connections, 49-52
shell elements, 49, 50
solid elements, 49
s,prayed concrete linings, 52
turning moments
Magnus platform, 311, 312,316
pile groups, 3 1 1-3 13
twin tunnels, interaction, 70-72, 71
U
UCL see University College
UDS see User Defined Subroutines
under-extraction, Leaning Tower of Pisa,
271-275
underdrainage
London area, 53-54,389-390,390
modelling, 389-394
underground tunnels see London
Underground
underpass, pin-joint connection, 105,106
undrained analysis
critical state models, 386-387
Mohr-Coulomb model, 365,382-384
undrained bearing capacity see bearing
capacity
undrained clay
strip footings, 219-223
tower instability, 257
University College (UCL), directional shear
cell, 20
User Defined Subroutines (UDS), 337
users
access to computer codes, 337-338
computer program responsibilities, 338339
constitutive model misunderstanding,
365
v
Vaughan, P.R., 56, 127,146, 184, 187, 194
vertical loading
shallow foundations, 2 19-233
single piles, 282-286
surface foundations, 2 19-233
Vesic, A.S., 25
Viggiani, G., 20
virgin compression line, 4, 5
virgin consolidation line, 5
void ratio, 56
volume loss control method (tunnel
excavation), 46-48,57
Treasury building case study, 68
W
Walbancke, H.J., 146, 187
wall see earth retaining structures
wall stiffness, embedded walls, 109-l l l
wall supports see support systems
Ward, W.H., 71
waste disposal see shallow nuclear waste
disposal
Waterloo International Terminal, 120-1 2 1,
120,121
Watts, K.S., 176
Wesley, L.D., 6
Westminster underground station, site plan,
89-90,89, 120
Wilson, A.C., 175
Wilson, E.L., 368, 370
wind turbine, failure, 335-336
Wood, D.M., 32
Y
yield surfaces, influence in deviatoric plane,
384-386
Young, W.C., 371
z
Zdravkovii, L., 17, 18, 19, 206, 239, 240,
244,318
zero thickness interface elements, 368-376
Zienkiewicz, O.C., 374